P1: OTA FM JWBS037-Chen
February 1, 2011
4:23
Printer Name: Yet to Come
NANOPLATFORM-BASED MOLECULAR IMAGING
P1: O...
16 downloads
1297 Views
57MB Size
Report
This content was uploaded by our users and we assume good faith they have the permission to share this book. If you own the copyright to this book and it is wrongfully on our website, we offer a simple DMCA procedure to remove your content from our site. Start by pressing the button below!
Report copyright / DMCA form
P1: OTA FM JWBS037-Chen
February 1, 2011
4:23
Printer Name: Yet to Come
NANOPLATFORM-BASED MOLECULAR IMAGING
P1: OTA FM JWBS037-Chen
February 1, 2011
4:23
Printer Name: Yet to Come
NANOPLATFORM-BASED MOLECULAR IMAGING
Edited by
Xiaoyuan Chen Laboratory of Molecular Imaging and Nanomedicine National Institute of Biomedical Imaging and Bioengineering National Institutes of Health Bethesda, Maryland
A JOHN WILEY & SONS, INC., PUBLICATION
P1: OTA FM JWBS037-Chen
February 1, 2011
4:23
Printer Name: Yet to Come
C 2011 by John Wiley & Sons, Inc. All rights reserved. Copyright
Published by John Wiley & Sons, Inc., Hoboken, New Jersey. Published simultaneously in Canada. No part of this publication may be reproduced, stored in a retrieval system, or transmitted in any form or by any means, electronic, mechanical, photocopying, recording, scanning, or otherwise, except as permitted under Section 107 or 108 of the 1976 United States Copyright Act, without either the prior written permission of the Publisher, or authorization through payment of the appropriate per-copy fee to the Copyright Clearance Center, Inc., 222 Rosewood Drive, Danvers, MA 01923, (978) 750-8400, fax (978) 750-4470, or on the web at www.copyright.com. Requests to the Publisher for permission should be addressed to the Permissions Department, John Wiley & Sons, Inc., 111 River Street, Hoboken, NJ 07030, (201) 748-6011, fax (201) 748-6008, or online at http://www.wiley.com/go/permission. Limit of Liability/Disclaimer of Warranty: While the publisher and author have used their best efforts in preparing this book, they make no representations or warranties with respect to the accuracy or completeness of the contents of this book and specifically disclaim any implied warranties of merchantability or fitness for a particular purpose. No warranty may be created or extended by sales representatives or written sales materials. The advice and strategies contained herein may not be suitable for your situation. You should consult with a professional where appropriate. Neither the publisher nor author shall be liable for any loss of profit or any other commercial damages, including but not limited to special, incidental, consequential, or other damages. For general information on our other products and services or for technical support, please contact our Customer Care Department within the United States at (800) 762-2974, outside the United States at (317) 572-3993 or fax (317) 572-4002. Wiley also publishes its books in a variety of electronic formats. Some content that appears in print may not be available in electronic formats. For more information about Wiley products, visit our web site at www.wiley.com Library of Congress Cataloging-in-Publication Data: Nanoplatform-based molecular imaging / edited by Xiaoyuan Chen. p. ; cm. Includes bibliographical references and index. ISBN 978-0-470-52115-1 1. Molecular probes. 2. Diagnostic imaging. I. Chen, Xiaoyuan. [DNLM: 1. Molecular Imaging–methods. 2. Molecular Imaging–trends. 3. Molecular Probes–diagnostic use. 4. Nanoparticles–diagnostic use. 5. Nanotechnology–trends. WN 180 N1865 2010] QP519.9.M64N36 2010 616.07 54–dc22 2010007984 Printed in the United States of America eBook ISBN: 978-0-470-76703-0 oBook ISBN: 978-0-470-76704-7 10 9 8 7 6 5 4 3 2 1
P1: OTA FM JWBS037-Chen
February 1, 2011
4:23
Printer Name: Yet to Come
CONTENTS
Preface
ix
Acknowledgments
xi
Contributors
PART I
xiii
BASICS OF MOLECULAR IMAGING AND NANOBIOTECHNOLOGY
1. Basic Principles of Molecular Imaging
3
Sven H. Hausner
2. Synthesis of Nanomaterials as a Platform for Molecular Imaging
25
Jinhao Gao, Jin Xie, Bing Xu, and Xiaoyuan Chen
3. Nanoparticle Surface Modification and Bioconjugation
47
Jin Xie, Jinhao Gao, Mark Michalski, and Xiaoyuan Chen
4. Biodistribution and Pharmacokinetics of Nanoprobes
75
Nagesh Kolishetti, Frank Alexis, Eric M. Pridgen, and Omid C. Farokhzad
PART II
NANOPARTICLES FOR SINGLE MODALITY MOLECULAR IMAGING
5. Computed Tomography as a Tool for Anatomical and Molecular Imaging
107
Pingyu Liu, Hu Zhou, and Lei Xing
6. Carbon Nanotube X-Ray for Dynamic Micro-CT Imaging of Small Animal Models
139
Otto Zhou, Guohua Cao, Yueh Z. Lee, and Jianping Lu
7. Quantum Dots for In Vivo Molecular Imaging
159
Yun Xing
8. Biopolymer, Dendrimer, and Liposome Nanoplatforms for Optical Molecular Imaging
183
David Pham, Ling Zhang, Bo Chen, and Ella Fung Jones v
P1: OTA FM JWBS037-Chen
February 1, 2011
vi
4:23
Printer Name: Yet to Come
CONTENTS
9. Nanoplatforms for Raman Molecular Imaging in Biological Systems
197
Zhuang Liu
10. Single-Walled Carbon Nanotube Near-Infrared Fluorescent Sensors for Biological Systems
217
Jingqing Zhang and Michael S. Strano
11. Microparticle- and Nanoparticle-Based Contrast-Enhanced Ultrasound Imaging
233
Nirupama Deshpande and J¨urgen K. Willmann
12. Ultrasound-Based Molecular Imaging Using Nanoagents
263
Srivalleesha Mallidi, Mohammad Mehrmohammadi, Kimberly Homan, Bo Wang, Min Qu, Timothy Larson, Konstantin Sokolov, and Stanislav Emelianov
13. MRI Contrast Agents Based on Inorganic Nanoparticles
279
Hyon Bin Na and Taeghwan Hyeon
14. Cellular Magnetic Labeling with Iron Oxide Nanoparticles
309
S´ebastien Boutry, Sophie Laurent, Luce Vander Elst, and Robert N. Muller
15. Nanoparticles Containing Rare Earth Ions: A Tunable Tool for MRI
333
C. Rivi`ere, S. Roux, R. Bazzi, J.-L. Bridot, C. Billotey, P. Perriat, and O. Tillement
16. Microfabricated Multispectral MRI Contrast Agents
375
Gary Zabow and Alan Koretsky
17. Radiolabeled Nanoplatforms: Imaging Hot Bullets Hitting Their Target
399
Raffaella Rossin
PART III NANOPARTICLE PLATFORMS AS MULTIMODALITY IMAGING AND THERAPY AGENTS 18. Lipoprotein-Based Nanoplatforms for Cancer Molecular Imaging
433
Ian R. Corbin, Kenneth Ng, and Gang Zheng
19. Protein Cages as Multimode Imaging Agents
463
Masaki Uchida, Lars Liepold, Mark Young, and Trevor Douglas
20. Biomedical Applications of Single-Walled Carbon Nanotubes
481
Weibo Cai, Ting Gao, and Hao Hong
21. Multifunctional Nanoparticles for Multimodal Molecular Imaging
529
Yanglong Hou and Rui Hao
22. Multifunctional Nanoparticles for Cancer Theragnosis Seulki Lee, Ick Chan Kwon, and Kwangmeyung Kim
541
P1: OTA FM JWBS037-Chen
February 1, 2011
4:23
Printer Name: Yet to Come
CONTENTS
23. Nanoparticles for Combined Cancer Imaging and Therapy
vii
565
Vaishali Bagalkot, Mi Kyung Yu, and Sangyong Jon
24. Multimodal Imaging and Therapy with Magnetofluorescent Nanoparticles
593
Jason R. McCarthy and Ralph Weissleder
25. Gold Nanocages: A Multifunctional Platform for Molecular Optical Imaging and Photothermal Treatment
615
Leslie Au, Claire M. Cobley, Jingyi Chen, and Younan Xia
26. Theranostic Applications of Gold Nanoparticles in Cancer
639
Parmeswaran Diagaradjane, Pranshu Mohindra, and Sunil Krishnan
27. Gold Nanorods as Theranostic Agents
659
Alexander Wei, Qingshan Wei, and Alexei P. Leonov
28. Theranostic Applications of Gold Core–Shell Structured Nanoparticles
683
Wei Lu, Marites P. Melancon, and Chun Li
29. Magnetic Nanoparticle Carrier for Targeted Drug Delivery: Perspective, Outlook, and Design
709
R. D. K. Misra
30. Perfluorocarbon Nanoparticles: A Multidimensional Platform for Targeted Image-Guided Drug Delivery
725
Gregory M. Lanza, Shelton D. Caruthers, Anne H. Schmieder, Patrick M. Winter, Tillmann Cyrus, and Samuel A. Wickline
31. Radioimmunonanoparticles for Cancer Imaging and Therapy
755
Arutselvan Natarajan
PART IV TRANSLATIONAL NANOMEDICINE 32. Current Status and Future Prospects for Nanoparticle-Based Technology in Human Medicine
783
Nuria Sanvicens, F´atima Fern´andez, J.-Pablo Salvador, and M.-Pilar Marco
Index
815
P1: OTA FM JWBS037-Chen
February 1, 2011
4:23
Printer Name: Yet to Come
PREFACE
This book focuses on the rational design of water-soluble, biocompatible nanoparticles for the visualization of the cellular function and follow-up of the molecular processes in living organisms without perturbing them. Molecular imaging probes based on nanotechnology hold great potential in diagnosis, imaging guided intervention, and treatment response monitoring of diseases. This book is logically organized by including the basics of molecular imaging, general strategies of particle synthesis and surface chemistry, applications in computed tomography (CT), optical imaging, magnetic resonance imaging (MRI), ultrasound, multimodality imaging, and theranostics, and finally clinical perspectives of nanoimaging. This comprehensive title provides expert opinions on the latest developments in molecular imaging using nanoparticles. This book consists of 32 chapters and was contributed by nearly 100 authors worldwide, who are among the world’s prominent scientists in material science and/or molecular imaging. Part I consists of Chapters 1–4 Chapter 1 describes the basic principles of molecular imaging, how nanoparticles can be applied to different molecular imaging modalities, and challenges in developing nanoparticle-based molecular imaging probes; Chapter 2 highlights the general strategies to produce narrowly dispersed nanomaterials for molecular imaging; Chapter 3 emphasizes the importance of surface modification to render nanoparticles biocompatible and suitable for molecular imaging applications; and Chapter 4 talks about the toxicity and factors such as size, shape, coating, and surface charge that affect the biodistribution and pharmacokinetics of nanoprobes. Part II consists of Chapters 5–17 Chapter 5 illustrates the basic principles of CT, the evolution of CT imaging technology, and the rationale for nanoparticle-based CT contrast agents; Chapter 6 describes the advantages of fascinating carbon nanotube field emission X-ray technology over conventional thermionic X-ray tubes that are used in current X-ray imaging systems; Chapter 7 describes the use of unique optical properties of semiconductor quantum dots (QDs) for near-infrared fluorescence imaging in living animals; Chapter 8 introduces macromolecular nanoconstructs such as biopolymers, dendrimers, and liposomes as carriers for fluorophore conjugation and optical imaging; Chapter 9 summarizes recent progress in developing nanoplatforms for Raman imaging of biological systems; Chapter 10 summarizes the work in using single-walled carbon nanotubes (SWNTs) as near-infrared fluorescent sensors for biomolecule detection; Chapter 11 describes the use of micro- and nanoparticles as ultrasound contrast agents; Chapter 12 proposes the use of metal nanoparticles in ultrasound-based photoacoustic and magnetoacoustic imaging modalities; Chapter 13 reports the progress on magnetic resonance imaging (MRI) contrast agents based on inorganic nanoparticles; Chapter 14 emphasizes the use of iron oxide nanoparticles for cellular labeling followed by T2 - and T∗2 -weighted MRI; Chapter 15 covers the use of rare earth based nanoparticles for MR imaging as positive contrast agents; Chapter 16 reviews the top–down microfabrication technology to synthesize multispectral MRI contrast agents; ix
P1: OTA FM JWBS037-Chen
February 1, 2011
x
4:23
Printer Name: Yet to Come
PREFACE
and Chapter 17 gives an overview of the strategies to label nanoparticles with radionuclides to study in vivo distribution. Part III consists of Chapters 18–31 Chapter 18 introduces techniques to incorporate imaging agents into lipoproteins and to reroute lipoproteins to cancer specific epitopes; Chapter 19 exemplifies the use of protein cages such as virus capsids and ferritins as platforms for MRI contrast agents and fluorescent imaging agents; Chapter 20 provides a comprehensive summary of the state-of-the-art of SWNTs for multimodality biomedical imaging applications; Chapter 21 reviews the progress in the controlled synthesis, surface modification, and multimodality imaging applications of multifunctional nanoparticles in recent years; Chapter 22 argues the use of cancer theranostics as a promising new strategy in cancer management, permitting simultaneous cancer diagnosis, drug delivery, and real-time monitoring of therapeutic efficacy; Chapter 23 provides more examples of multifunctional nanoparticles for combined cancer imaging and therapy (theranostics); Chapter 24 describes the recent progress in modifying magnetic nanoparticles for multimodality imaging as well as targeted treatment of a number of diseases; Chapter 25 introduces gold nanocages as contrast agents for optical bioimaging (such as optical and spectroscopic coherence tomography amd photoacoustic tomography) and photothermal treatment; Chapter 26 describes the biological inertness, ease of manufacture and bioconjugation, and presumed lack of toxicity of gold nanoparticles for simultaneous sensing, imaging, and treatment of tumors; Chapter 27 presents the recent developments in the chemistry and photophysics of gold nanorods and their applications toward biological imaging and photothertmally activated therapies; Chapter 28 describes a number of gold core–shell nanostructures for cancer molecular optical imaging, controlled drug delivery, and photothermal ablation therapy; Chapter 29 describes a novel temperature and pH-responsive magnetic nanocarrier that combines tumor targeting and controlled drug release capabilities; Chapter 30 deals with perfluorocarbon nanoparticles as a multidimensional platform for targeted image-guided drug delivery; and Chapter 31 describes the use of radiolabeled nanoparticles and radiolabeled immunonanoparticles for imaging and therapy. Part IV is the concluding Chapter 32 that highlights some of the nanoparticle-based novel technologies for molecular imaging, diagnosis, and drug delivery formulations. The limitations and future challenges of nanoparticle-based systems are also discussed. Bethesda, Maryland
Xiaoyuan Chen
P1: OTA FM JWBS037-Chen
February 1, 2011
4:23
Printer Name: Yet to Come
ACKNOWLEDGMENTS
The editor thanks the nearly 100 authors throughout the world for their contributions and collaboration on this book project. The editing work of this book was accomplished using a significant amount of the editor’s spare time including family time. Therefore the editor also thanks his wife, Michelle Ji, and his daughter, Grace Chen, for their wonderful support and understanding.
xi
P1: OTA FM JWBS037-Chen
February 1, 2011
4:23
Printer Name: Yet to Come
CONTRIBUTORS
Frank Alexis, Department of Bioengineering, Clemson University, Clemson, South Carolina, USA Leslie Au, Department of Biomedical Engineering, Washington University, St. Louis, Missouri, USA Vaishali Bagalkot, School of Life Sciences, Gwangju Institute of Science and Technology, Gwangju, South Korea R. Bazzi, Laboratoire Physico-Chimie des Electrolytes, Colloides et Sciences Analytiques, Universit´e Pierre et Marie Curie, Paris, France C. Billotey, Laboratoire CREATIS–Animage, Universit´e Claude Bernard, Lyon, France S´ebastien Boutry, Department of General, Organic and Biomedical Chemistry, NMR and Molecular Imaging Laboratory, University of Mons, Mons, Belgium J.-L. Bridot, Service de Chimie G´en´erale, Organique et Biom´edicale, Laboratoire de RMN et d’Imagerie Mol´eculaire, Universit´e de Mons-Hainaut, Mons, Belgium Weibo Cai, Departments of Radiology and Medical Physics, School of Medicine and Public Health, University of Wisconsin–Madison, and University of Wisconsin Carbone Cancer Center, Madison, Wisconsin, USA Guohua Cao, Department of Physics and Astronomy, University of North Carolina at Chapel Hill, Chapel Hill, North Carolina, USA Shelton D. Caruthers, Department of Medicine, Washington University Medical School, St. Louis, Missouri, and Philips Healthcare, Andover, Massachusetts, USA Bo Chen, Center for Molecular and Functional Imaging, Department of Radiology and Biomedical Imaging, University of California, San Francisco, California, USA Jingyi Chen, Department of Biomedical Engineering, Washington University, St. Louis, Missouri, USA Xiaoyuan Chen, Molecular Imaging Program at Stanford and Bio-X Program, Department of Radiology, Stanford University School of Medicine, Stanford, California, and Laboratory for Molecular Imaging and Nanomedicine, National Institute of Biomedical Imaging and Bioengineering, National Institutes of Health, Bethesda, Maryland, USA Claire M. Cobley, Department of Biomedical Engineering, Washington University, St. Louis, Missouri, USA
xiii
P1: OTA FM JWBS037-Chen
February 1, 2011
xiv
4:23
Printer Name: Yet to Come
CONTRIBUTORS
Ian R. Corbin, Department of Medical Biophysics, University of Toronto, Toronto, Ontario, and Division of Biophysics and Bioimaging, Ontario Cancer Institute, Toronto, Ontario, Canada Tillmann Cyrus, Department of Medicine, Washington University Medical School, St. Louis, Missouri, USA Nirupama Deshpande, Department of Radiology and Molecular Imaging Program at Stanford, Stanford University School of Medicine, Stanford, California, USA Parmeswaran Diagaradjane, Department of Radiation Oncology, University of Texas M.D. Anderson Cancer Center, Houston, Texas, USA Trevor Douglas, Department of Chemistry and Biochemistry and Department of Plant Science, Center for Bio-inspired Nanomaterials, Montana State University, Bozeman, Montana, USA Luce Vander Elst, Department of General, Organic and Biomedical Chemistry, NMR and Molecular Imaging Laboratory, University of Mons, Mons, Belgium Stanislav Emelianov, Department of Biomedical Engineering, University of Texas at Austin, Austin, Texas, USA Omid C. Farokhzad, Department of Anesthesiology, Brigham and Women’s Hospital, Harvard Medical School, Boston, Massachusetts, USA F´atima Fern´andez, Applied Molecular Receptors Group, CIBER de Bioingenier´ıa, Biomateriales y Nanotecnolog´ıa, IQAC-CSIC, Barcelona, Spain Jinhao Gao, Molecular Imaging Program at Stanford and Bio-X Program, Department of Radiology, Stanford University School of Medicine, Stanford, California, USA Ting Gao, Tyco Electronics Corporation, Menlo Park, California, USA Rui Hao, Department of Advanced Materials and Nanotechnology, College of Engineering, Peking University, Beijing, China Sven H. Hausner, Department of Biomedical Engineering, University of California– Davis, Davis, California, USA Kimberly Homan, Department of Biomedical Engineering, University of Texas at Austin, Austin, Texas, USA Hao Hong, Departments of Radiology and Medical Physics, School of Medicine and Public Health, University of Wisconsin–Madison, Madison, Wisconsin, USA Yanglong Hou, Department of Advanced Materials and Nanotechnology, College of Engineering, Peking University, Beijing, China Taeghwan Hyeon, National Creative Research Initiative Center for Oxide Nanocrystalline Materials, and School of Chemical and Biological Engineering, Seoul National University, Seoul, South Korea Sangyong Jon, School of Life Sciences, Gwangju Institute of Science and Technology, Gwangju, South Korea
P1: OTA FM JWBS037-Chen
February 1, 2011
4:23
Printer Name: Yet to Come
CONTRIBUTORS
xv
Ella Fung Jones, Center for Molecular and Functional Imaging, Department of Radiology and Biomedical Imaging, University of California, San Francisco, California, USA Kwangmeyung Kim, Biomedical Research Center, Korea Institute of Science and Technology, Seoul, South Korea Nagesh Kolishetti, Department of Anesthesiology, Brigham and Women’s Hospital, Harvard Medical School, Boston, Massachusetts, USA Alan Koretsky, Laboratory of Functional and Molecular Imaging, National Institute of Neurological Disorders and Stroke, National Institutes of Health, Bethesda, Maryland, USA Sunil Krishnan, Department of Radiation Oncology, University of Texas M.D. Anderson Cancer Center, Houston, Texas, USA Ick Chan Kwon, Biomedical Research Center, Korea Institute of Science and Technology, Seoul, South Korea Gregory M. Lanza, Department of Medicine, Washington University Medical School, St. Louis, Missouri, USA Timothy Larson, Department of Biomedical Engineering, University of Texas at Austin, Austin, Texas, USA Sophie Laurent, Department of General, Organic and Biomedical Chemistry, NMR and Molecular Imaging Laboratory, University of Mons, Mons, Belgium Seulki Lee, Biomedical Research Center, Korea Institute of Science and Technology, Seoul, South Korea Yueh Z. Lee, Department of Physics and Astronomy and Department of Radiology, University of North Carolina at Chapel Hill, Chapel Hill, North Carolina, USA Alexei P. Leonov, Department of Chemistry, Purdue University, West Lafayette, Indiana, USA Chun Li, Department of Experimental Diagnostic Imaging, University of Texas M.D. Anderson Cancer, Houston, Texas, USA Lars Liepold, Department of Chemistry and Biochemistry and Department of Plant Sciences, Center for Bio-Inspired Nanomaterials, Montana State University, Bozeman, Montana, USA Pingyu Liu, Palo Alto Unified School District, Palo Alto, California, USA Zhuang Liu, Institute of Functional Nano & Soft Materials, Soochow University, Suzhou, Jiangsu, China Jianping Lu, Department of Physics and Astronomy, Curriculum in Applied Sciences and Engineering, University of North Carolina at Chapel Hill, Chapel Hill, North Carolina, USA Wei Lu, Department of Experimental Diagnostic Imaging, University of Texas M. D. Anderson Cancer Center, Houston, Texas, USA
P1: OTA FM JWBS037-Chen
February 1, 2011
xvi
4:23
Printer Name: Yet to Come
CONTRIBUTORS
Srivalleesha Mallidi, Department of Biomedical Engineering, University of Texas at Austin, Austin, Texas, USA M.-Pilar Marco, Applied Molecular Receptors Group, CIBER de Bioingenier´ıa, Biomateriales y Nanotecnolog´ıa, IQAC-CSIC, Barcelona, Spain Jason R. McCarthy, Center for Molecular Imaging Research, Harvard Medical School and Massachusetts General Hospital, Charlestown, Massachusetts, USA Mohammad Mehrmohammadi, Department of Biomedical Engineering, University of Texas at Austin, Austin, Texas, USA Marites P. Melancon, Department of Experimental Diagnostic Imaging, University of Texas M. D. Anderson Cancer Center, Houston, Texas, USA Mark Michalski, Molecular Imaging Program at Stanford and Bio-X Program, Department of Radiology, Stanford University School of Medicine, Stanford, California, USA R. D. K. Misra, Center for Structural and Functional Materials, University of Louisiana at Lafayette, Lafayette, Louisiana, USA Pranshu Mohindra, Department of Radiation Oncology, University of Texas M.D. Anderson Cancer Center, Houston, Texas, USA Robert N. Muller, Department of General, Organic and Biomedical Chemistry, NMR and Molecular Imaging Laboratory, University of Mons, Mons, Belgium Hyon Bin Na, National Creative Research Initiative Center for Oxide Nanocrystalline Materials, and School of Chemical and Biological Engineering, Seoul National University, Seoul, South Korea Arutselvan Natarajan, Department of Radiology and Molecular Imaging Program at Stanford, Stanford University School of Medicine, Stanford, California, USA Kenneth Ng, Institute of Biomaterials and Biomedical Engineering, University of Toronto, Toronto, Ontario, Canada P. Perriat, Groupe d’Etudes de M´etallurgie Physique et de Physique des Mat´eriaox, Universit´e Claude Bernard, Lyon, France David Pham, Center for Molecular and Functional Imaging, Department of Radiology and Biomedical Imaging, University of California, San Francisco, California, USA Eric M. Pridgen, Department of Chemical Engineering, Massachusetts Institute of Technology, Cambridge, Massachusetts, USA Min Qu, Department of Biomedical Engineering, University of Texas at Austin, Austin, Texas, USA C. Rivi`ere, Laboratoire de Physique de la Mati`ere Condens´ee et Nanostructures, Universit´e de Lyon, Lyon, France Raffaella Rossin, Department of Biomolecular Engineering, Philips Research Europe, Eindhoven, The Netherlands S. Roux, Laboratoire de Physico-Chimie des Mat´eriaux Luminescents, Universit´e de Lyon, Lyon, France
P1: OTA FM JWBS037-Chen
February 1, 2011
4:23
Printer Name: Yet to Come
CONTRIBUTORS
xvii
J.-Pablo Salvador, Applied Molecular Receptors Group, CIBER de Bioingenier´ıa, Biomateriales y Nanotecnolog´ıa, IQAC-CSIC, Barcelona, Spain Nuria Sanvicens, Applied Molecular Receptors Group, CIBER de Bioingenier´ıa, Biomateriales y Nanotecnolog´ıa, IQAC-CSIC, Barcelona, Spain Anne H. Schmieder, Department of Medicine, Washington University Medical School, St. Louis, Missouri, USA Konstantin Sokolov, Department of Biomedical Engineering, University of Texas at Austin, Austin, Texas, and Department of Medical Physics, University of Texas M.D. Anderson Cancer Center, Houston, Texas, USA Michael Strano, Department of Chemical Engineering, Massachusetts Institute of Technology, Cambridge, Massachusetts, USA O. Tillement, Laboratoire de Physico-Chimie des Mat´eriaux Luminescents, Universit´e de Lyon, Lyon, France Masaki Uchida, Department of Chemistry and Biochemistry and Department of Plant Science, Center for Bio-Inspired Nanomaterials, Montana State University, Bozeman, Montana, USA Bo Wang, Department of Biomedical Engineering, University of Texas at Austin, Austin, Texas, USA Alexander Wei, Department of Chemistry, Purdue University, West Lafayette, Indiana, USA Qingshan Wei, Department of Chemistry, Purdue University, West Lafayette, Indiana, USA Ralph Weissleder, Center for Molecular Imaging Research, Harvard Medical School and Massachusetts General Hospital, Charlestown, Massachusetts, USA Samuel A. Wickline, Department of Medicine, Washington University Medical School, St. Louis, Missouri, USA Jurgen K. Willmann, Department of Radiology and Molecular Imaging Program at Stanford, Stanford University School of Medicine, Stanford, California, USA Patrick M. Winter, Department of Medicine, Washington University Medical School, St. Louis, Missouri, USA Younan Xia, Department of Biomedical Engineering, Washington University, St. Louis, Missouri, USA Jin Xie, Molecular Imaging Program at Stanford and Bio-X Program, Department of Radiology, Stanford University School of Medicine, Stanford, California, and Laboratory for Molecular Imaging and Nanomedicine, National Institute of Biomedical Imaging and Bioengineering, National Institutes of Health, Bethesda, Maryland, USA Lei Xing, Department of Radiation Oncology, Stanford University School of Medicine, Stanford, California, USA
P1: OTA FM JWBS037-Chen
February 1, 2011
xviii
4:23
Printer Name: Yet to Come
CONTRIBUTORS
Yun Xing, Department of Material Science and Engineering, University of Dayton, Dayton, Ohio, USA Bing Xu, Department of Chemistry, Brandeis University, Waltham, Massachusetts, USA Mark Young, Department of Chemistry and Biochemistry and Department of Plant Science, Center for Bio-Inspired Nanomaterials, Montana State University, Bozeman, Montana, USA Mi Kyung Yu, School of Life Sciences, Gwangju Institute of Science and Technology, Gwangju, South Korea Gary Zabow, Laboratory of Functional and Molecular Imaging, National Institute of Neurological Disorders and Stroke, National Institutes of Health, Bethesda, Maryland, USA Jingqing Zhang, Department of Chemical Engineering, Massachusetts Institute of Technology, Cambridge, Massachusetts, USA Ling Zhang, Center for Molecular and Functional Imaging, Department of Radiology and Biomedical Imaging, University of California, San Francisco, California, USA Gang Zheng, Department of Medical Biophysics and Institute of Biomaterials and Biomedical Engineering, University of Toronto, Toronto, Ontario, and Division of Biophysics and Bioimaging, Ontario Cancer Institute, Toronto, Ontario, Canada Hu Zhou, Community Cancer Center of Roseburg, Roseburg, Oregon, USA Otto Zhou, Department of Physics and Astronomy, Curriculum in Applied Sciences and Engineering, and Lineberger Comprehensive Cancer Center, University of North Carolina at Chapel Hill, Chapel Hill, North Carolina, USA
P1: OTA/XYZ P2: ABC c01 JWBS037-Chen February 2, 2011
21:44
Printer Name: Yet to Come
PART I
BASICS OF MOLECULAR IMAGING AND NANOBIOTECHNOLOGY
P1: OTA/XYZ P2: ABC c01 JWBS037-Chen February 2, 2011
21:44
Printer Name: Yet to Come
CHAPTER 1
Basic Principles of Molecular Imaging SVEN H. HAUSNER Department of Biomedical Engineering, University of California–Davis, Davis, California, USA
1.1 INTRODUCTION The ability to identify diseased tissue for detection and treatment remains a central goal for medical research. Several noninvasive or minimally invasive diagnostic modalities have been developed which allow one to obtain anatomical, physiological, and molecular information. “Molecular imaging” can be defined as in situ visualization, characterization, and measurement of biological processes in the living organism at the molecular or cellular level. Diagnosis and visualization at the molecular level, that is, detection of a disease in its infancy, may significantly improve treatment and patient care. By combining two or more imaging modalities, each with its different strengths, high-quality complementary (e.g., molecular and anatomical) information can be obtained and analyzed in the context of each other. This has led to the rise of dual- and multimodality imaging approaches. Depending on the modality, imaging probes or contrast agents are required or highly desirable; they can range in size from single atoms to cell-sized constructs. Nanoparticles, that is, entities with dimensions in the range of several tens of nanometers, can display desirable pharmacokinetic properties and permit the combination of different clinically relevant moieties (e.g., targeting groups, molecular beacons, and contrast agents for different modalities, surface coatings, enclosed payload) in a single unit. The inclusion of a therapeutic component yields “theranostics.” Taken together, nanotechnology-based molecular probes offer the promise for tailor-made clinical tools required for “personalized medicine.” This chapter provides an introductory overview of molecular imaging, major imaging modalities, and imaging probes, with particular focus on the promises and challenges of nanoparticle-based compounds.
1.2 IMAGING IN MEDICINE Most areas of clinical practice require identification and localization of diseased tissue for detection and treatment. Ideally, reliable, specific, and noninvasive high-contrast Nanoplatform-Based Molecular Imaging Edited by Xiaoyuan Chen C 2011 John Wiley & Sons, Inc. Copyright
3
P1: OTA/XYZ P2: ABC c01 JWBS037-Chen February 2, 2011
4
21:44
Printer Name: Yet to Come
BASIC PRINCIPLES OF MOLECULAR IMAGING
whole-body evaluations would allow physicians to detect serious abnormalities before patients present with symptoms, thus permitting early intervention, thereby increasing the chance for cure or, at a minimum, allow for better patient management and improved quality of life. Given these incentives, it is clear that practical (i.e., minimally inconvenient for the patient) and affordable (i.e., overall cost-saving to the health care system and society) diagnostic approaches are highly desirable. Ever since Wilhelm R¨ontgen’s first use in 1895 of the then newly discovered X-rays to noninvasively image the interior of the body, the keen interest in medical imaging has been met by increasingly sophisticated technologies (Fig. 1.1). While R¨ontgen’s X-ray image was a grainy two-dimensional anatomical projection, physicians nowadays have access to tomographic (three-dimensional) imaging modalities with, depending on the technique, submillimeter resolution, which allows visualization of anatomical, physiological, and, increasingly, molecular (cellular) biological information. Since diseases often arise from changes on the molecular and cellular levels, long before manifesting themselves in detectable large-scale physiological or anatomical changes, molecular imaging is gaining increasing attention. If a disease can be diagnosed and visualized at the molecular level, that is, detected in its infancy, it can be treated at a much earlier stage, the treatment’s efficacy can be determined much sooner and, if necessary, the treatment plan can be adjusted accordingly. This benefits the individual patient and society as a
FIGURE 1.1 (Left) Wilhelm R¨ontgen’s (1845–1923) first X-ray image, depicting the hand of his wife, Anna, taken on 22 December 1895. (Right) A slice of a modern whole-body multimodality positron emission tomography/computed tomography (PET/CT) scan showing glucose metabolism within the body, including a large, metabolically active tumor (arrow). (PET/CT image courtesy of Dr. Cameron Foster and Dr. Ramsey Badawi, UC Davis Medical Center, Davis, California.)
P1: OTA/XYZ P2: ABC c01 JWBS037-Chen February 2, 2011
21:44
Printer Name: Yet to Come
IMAGING IN MEDICINE
5
whole. Molecular biology is discovering a growing number of disease-specific cellular targets and is determining their distribution in patient populations [1]. For certain diseases this has already had significant effects on determining beforehand which patients will benefit from a certain treatment (“patient stratification”). A prime example is testing for the expression of HER2/neu in breast cancer for prognosis, as well as for selection and monitoring of treatment: expression has been linked to aggressiveness of the disease, but it also provides a target for highly effective treatment with antibodies (Trastuzumab, Herceptin® ) [2, 3]. Similarly, monitoring glucose metabolism with the imaging agent 18 F-fluorodeoxyglucose (18 F-FDG) has proved itself to be the preferred approach for staging, restaging, and evaluation of response to treatment for several cancers [4]. Concurrently with the advances in molecular biology, engineers and physicists are developing increasingly sophisticated imaging instrumentation capable of localizing imaging agents in the body at high sensitivity and high resolution in short acquisition time [5]. By bridging the clinical and engineering worlds, research in imaging agents plays a central role. To that end, the development of target-specific (and disease-specific) nanoparticle-based molecular probes draws on research in several fields including biology, molecular biology, medicine, chemistry, and biomedical engineering. 1.2.1 Molecular Imaging Rather than relying only on intrinsic large-scale differences of tissue characteristics (e.g., density) or passive accumulation of administered probes to reveal disease in vivo, molecular imaging strives to make use of disease-specific (“targeted”) interactions of imaging probes with the target tissue on a molecular and a cellular level. The goal is the real-time in situ visualization of biological processes in the living organism. This focus is also reflected in the Society of Nuclear Medicine’s definition of molecular imaging as “an array of non-invasive, diagnostic imaging technologies that can create images of both physical and functional aspects of the living body. It can provide information that would otherwise require surgery or other invasive procedures to obtain. Molecular imaging differs from microscopy, which can also produce images at the molecular level, in that microscopy is used on samples of tissue that have been removed from the body, not on tissues still within a living organism. It differs from X-rays and other radiological techniques in that molecular imaging primarily provides information about biological processes (function) while [computed tomography] CT, X-rays, [magnetic resonance imaging] MRI and ultrasound, image physical structure (anatomy)” [6]. As stated above, the information obtained is linked to which imaging modality is chosen. Individual imaging modalities can be grouped by the energy spectrum and energy type evaluated (X-ray, photons, sound; positrons), the resolution that can be achieved, and the type of information obtained (anatomical, physiological, cellular/molecular) (Table 1.1). Widely used clinical imaging modalities include magnetic resonance imaging, ultrasound (US), computed tomography, as well as positron emission tomography (PET) and single photon emission computed tomography (SPECT). All of these modalities allow for the noninvasive imaging of living subjects. Although the first three imaging modalities are primarily anatomical and not molecular, the two types of modalities can be combined for dual- or multimodality imaging. In addition, MRI, US, and CT can be used with molecular imaging probes, especially as part of nanoplatforms. In addition, a number of more specialized optical modalities are being used or are under investigation, including endoscopic methods [12].
6 M, P
M, P
M, P
Positron emission tomography (PET)
Single photon emission computed tomography (SPECT)
Typeb
Optical imaging (fluorescence and bioluminescence)
Imaging Modalitya −15
Sensitivity (Concentration of Imaging Probe/Contrast Agent)
Photon emitted by radioactive isotope of imaging probe.
∼10−11 mole/L [7]
∼10−11 –10−12 mole/L
Depth
0.5–2 mm (preclinical) 10–15 mm (clinical)
1–2 mm (preclinical) 4–8 mm (clinical)
No limit
No limit
Yes
Yes
∼1 to ∼10 mm Centimeters (Yes, limited quantification possible)
Resolution
Quantitative Modality
Minutes to tens of minutes
Minutes to tens of minutes
Seconds to minutes
Typical Scan Acquisition Time
Clinical and preclinical. High cost. Versatile imaging probe chemistry. Possibility to distinguish different radioisotopes based on photon energy.
Clinical and preclinical. High cost. Versatile imaging probe chemistry.
Preclinical; limited clinical translation (close to skin or requiring endoscopic approaches) Low cost. Depth limitation based on wavelength-dependent absorption by tissue. Resolution is depth dependent. Two-dimensional (surface) image.
Other
21:44
511-keV Photons generated during annihilation of positron emitted by radioactive isotope of imaging probe.
Fluorescence: External As low as ∼10 mole/L excitation light absorbed by fluorochrome of imaging probe and reemitted at longer wavelength. Bioluminescence: Chemiluminescence of enzymatic reaction.
Basis for Detection
TABLE 1.1 Widely Used Imaging Modalities
P1: OTA/XYZ P2: ABC c01 JWBS037-Chen February 2, 2011 Printer Name: Yet to Come
b
Less than ∼50 m (preclinical) <500 m (clinical) [9]
<100 m (preclinical) <1 mm (clinical) [9]
Absorption of focused Low (gram amounts <10 m external X-rays by of contrast agent (preclinical) tissue (or contrast required) agent).
Primary molecular imaging modalities are listed in bold. A, anatomical; M, molecular/cellular; P, physiological. Source: Adapted from Willmann [11] and Weissleder [12].
a
A, P
A, P (M) Echoes of tissue (or High (single imaging probe) microbubbles— generated by volume ∼0.004 high-frequency pL—can be (∼1–40 MHz) detected) [8] sound waves propagating through tissue.
A, P (M) Interaction of external 10−3 –10−4 mole/L magnetic field and radiofrequencies with atomic nuclear spins (of tissue or contrast agent) depending on environment of nuclei.
No limit
Up to ∼25 cm [10]
No limit
Yes
Yes
Yes
Minutes
Seconds to minutes
Minutes to hour
Clinical and preclinical. Excellent bone and lung contrast. Poor soft-tissue contrast.
Clinical and preclinical. Low cost. Frequency used determines resolution and (inversely) penetration depth. Targeted imaging limited to vasculature. High operator dependency.
Clinical and preclinical. High cost. Long scan acquisition time. Excellent soft tissue contrast.
21:44
Computed tomography (CT)
Ultrasound imaging (US)
Magnetic resonance imaging (MRI)
P1: OTA/XYZ P2: ABC c01 JWBS037-Chen February 2, 2011 Printer Name: Yet to Come
7
P1: OTA/XYZ P2: ABC c01 JWBS037-Chen February 2, 2011
8
21:44
Printer Name: Yet to Come
BASIC PRINCIPLES OF MOLECULAR IMAGING
Regardless of the imaging modality chosen, quantifiable high-resolution images and reasonable acquisition times are highly desired, and if modalities are combined they should yield relevant additional (e.g., anatomical plus molecular) information. Molecular imaging per se is complementary to primarily anatomical imaging (Table 1.1). This is the motivation behind the ongoing push toward dual-/multimodality imaging where molecular imaging data are collected at the same time as anatomical imaging data (Fig. 1.1). This synergistic approach, in which superimposed tomographic images are analyzed, allows the physician interpretation of the molecular imaging data within the anatomical context. The tremendous benefits of this diagnostic approach have also been recognized by the manufacturers of clinical imaging equipment. This has led to the rapid spread of integrated hybrid PET/CT and SPECT/CT scanners in recent years. Dual-modality scanners are now becoming the norm rather than the exception in the clinic [5, 13]. Similarly, hybrid PET/MR scanners are now becoming available; they are eagerly awaited for tasks where the molecular imaging data have to be interpreted in the context of soft tissue, such as, for example, within the brain. Engineering and technical challenges are largely the reason that the availability of hybrid-MR systems has been lagging behind their CT counterparts [14, 15]. The instrumentation for the various modalities has also been adapted for preclinical applications [16]. By using mice, rats, nonhuman primates, or other animal models, specialized small animal scanners allow dedicated imaging in a preclinical research setting. Spatial resolution is generally higher because the subjects can be moved closer to the detectors and the instruments are specifically designed for the reduced dimension required. Because of their small body size, whole-body imaging is easily possible for several species with many of the imaging modalities.
1.3 MAJOR IMAGING MODALITIES 1.3.1 Optical Imaging (Fluorescence and Bioluminescence) Optical imaging is finding increasing clinical use in several specialized applications, largely using endoscopic (or similar fiberoptic intravital) methods or in regions with limited tissue thickness (e.g., the breast) [12, 17]. Still, the major application of optical imaging lies in preclinical use for small animal studies, chiefly thanks to relatively low cost and simple setup: the subject is placed in a light-tight box and imaged with a highly sensitive charge-coupled device (CCD) camera. A considerable number of optical probes and tags are commercially available, making optical imaging the most popular preclinical imaging modality [5]. For fluorescence imaging the subject is typically illuminated by an external source with excitation light that is absorbed by the fluorophore of an imaging probe. The fluorophore then emits light at lower energies (longer wavelengths) that is detected by the camera. Ideally, the light involved should be in the near-infrared range (∼650–900 nm), where absorbance by blood is minimal. For bioluminescence imaging, no external excitation is required; rather, the faint light emitted by certain biological processes is measured directly. In laboratory studies, this can be achieved by linking a “reporter gene” encoding for a luminescent protein (usually luciferase) to the gene of interest and genetically transferring them into the animal before the study. After administration of an exogenous substrate (e.g., luciferin) light is generated only at sites where the genes are expressed. A similar approach can also be used for modified fluorescence imaging. In this case, the gene for green fluorescent protein (GFP) or one of its derivatives is commonly used as the reporter. Under
P1: OTA/XYZ P2: ABC c01 JWBS037-Chen February 2, 2011
21:44
Printer Name: Yet to Come
MAJOR IMAGING MODALITIES
9
illumination, locally expressed GFP emits light that is detected by the CCD camera. An advantage of this approach is the possibility of longitudinal studies because injection of a substrate is not necessary for visualization, whereas the useful window for bioluminescence after a luciferin injection is usually only about 5–30 minutes. Owing to the fact that a carefully designed optical probe can be switched on and off in vivo as a result of chemical or physicochemical transformations, “activatable” or “smart” fluorescent probes have been developed that can respond to the presence and level of biological markers at sites within the body [18]. This has been used in preclinical tumor models to monitor treatment response using a near-infrared fluorophore (NIRF)-based imaging probe responsive to the level of matrix metalloproteinase (MMP)-2. Treatment reduced the level of MMP-2 expressed by the tumor, which was reflected in a reduced signal emitted by the imaging probe. Several challenges exist for optical imaging. For fluorescence imaging, they include high background signals caused by tissue autofluorescence [19] and limited stability (photobleaching) of many small-molecule fluorophores. Bioluminescence does not have the same problems, but researchers face the tasks of genetically engineering the animal model and detecting very faint signals. Both approaches are constrained by depth limitations due to scattering and absorbance by overlying tissue and the concomitant difficulties with exact quantification. If spatial resolution is not a major concern, whole-body optical imaging is possible for small rodents (especially mice) since scattering and absorption are limited because of the small body size [20]. Perhaps more than for other imaging modalities, a notable number of new approaches based on different technologies are being investigated for optical imaging [12]. Fluorescence lifetime imaging (FLIM), photoaccoustic imaging, multispectral imaging [21], self-illuminating fluorescent imaging probes [19], Raman microscopy techniques, and tomographic fluorescence systems are among the exciting approaches currently under development [12]. Some of them rely entirely on endogenous contrast and do not require the administration of any exogenous probes. Two examples are coherent anti-Stokes Raman scattering (CARS) and optical coherence tomography (OCT). CARS is a nonlinear Raman technique that measures the vibrational spectra of light scattered from illuminated biological specimens. Analysis of the spectra allows conclusions about the constituents of the tissue close to the surface. It has been used in vivo to map lipid compartments, protein clusters, and water distribution at subcellular resolution [22]. OCT is a technique based on light scattering that can be described as an optical version of ultrasound (see below). Despite a shallow penetration depth of only about 2–3 mm, it is attractive since it yields real-time very-high resolution (1–15 m) “optical biopsy” images that are comparable to conventional histopathology. It is finding applications in ophthalmic, gastrointestinal, and intravascular imaging using noninvasive or minimally invasive instrumentation such as handheld probes, endoscopes, catheters, laparoscopes, or needles [23]. 1.3.2 Radionuclide-Based Imaging Modalities: Positron Emission Tomography (PET) and Single Photon Emission Computed Tomography (SPECT) Because of high sensitivity and absence of depth limitations, PET and SPECT are the two molecular imaging modalities that have risen to prominence in both the clinical and preclinical settings. They require the administration of a positron- or single-photonemitting radioisotope, usually attached to a larger molecule. Examples are [18 F]fluorine in
P1: OTA/XYZ P2: ABC c01 JWBS037-Chen February 2, 2011
10
21:44
Printer Name: Yet to Come
BASIC PRINCIPLES OF MOLECULAR IMAGING
2-[18 F]fluoro-2-deoxy-glucose (18 F-FDG), [123 I]iodine in [123 I]metaiodobenzylguanidine (123 I-MIBG), or radioactive metal isotopes captured by a chelator (e.g., [64 Cu]copper or [111 In]indium in chelator-bearing proteins and antibodies). As such, both imaging modalities rely completely on exogenous probes for imaging. For both imaging modalities, the availability, the chemistry, and the radioactive half-life of the chosen isotope have to be considered. This is illustrated by comparing the two popular PET isotopes [11 C]carbon and [18 F]fluorine. [11 C]Carbon is an attractive isotope because it can directly replace a nonradioactive carbon without changing the molecular structure of a compound. However, it has a half-life (t1/2 ) of only 20.4 min, necessitating production in an on-site cyclotron and limiting preparation of the imaging probe to a handful of very fast chemical reactions. By contrast, the nearly 2-h half-life of [18 F]fluorine allows a much wider range of chemistries and even some degree of shipment of the imaging probe from central production facilities to outlying hospitals by ground or air. Since the fluorine atom usually takes the place of another element (often a hydrogen atom), possible effects on pharmacokinetics have to be evaluated during drug development. Regardless of which radionuclide-based imaging modality is employed, it is important to use a radioisotope whose physical (radioactive) half-life is matched to the pharmacokinetics of the imaging probe to ensure a sufficiently high signal-to-noise ratio at the time of imaging [24]. Since most nanoparticles have long blood circulation times, it may take up to a few days before the level in the target tissue has risen significantly over background levels. In order to match the long biological half-life of the probe, long-lived radioisotopes are often required. Fortunately, many such radioisotopes are available. For example, the SPECT isotopes [123 I]iodine, [99m Tc]technetium, and [111 In]indium have half-lives of 13.2 h, 6.0 h, and 67.3 h, respectively, and long-lived PET isotopes include [64 Cu]copper, [124 I]iodine, and [89 Zr]zirconium (t1/2 = 12.7 h, 100.2 h, 78.4 h, respectively). PET, in particular, distinguishes itself through its high sensitivity combined with the ability to image effectively without depth limitation [25]. As mentioned earlier, especially 18 F-FDG has helped clinical PET to play a prominent role in cancer detection and monitoring of response to treatment because it allows the visualization of glucose hypermetabolism associated with many malignancies and whole-body PET scans permit the detection of distant metastases (Fig. 1.1). Delicate biological systems (e.g., the brain) can be imaged with minimal disturbance of the molecular processes investigated thanks to the extremely low amount of imaging probe required. The signal detected by the scanner originates from the radioactive decay of a positron-emitting radioisotope prepared in a cyclotron prior to incorporation into the imaging probe. The positron loses energy by scattering through the tissue until undergoing annihilation with an electron, resulting in the emission of two 511-keV photons at an angle of nearly 180◦ . The pair of photons is detected by a cylindrical array of scintillators connected to photomultiplier tubes (PMTs). Image quality is greatly improved by only accepting valid coincidences and rejecting random events stemming from background radiation: only signals obtained in opposite detectors within a narrow time window, commonly 2–5 ns, are accepted as originating from the same positron-decay event. Resolution-limiting factors are the average range positrons travel before undergoing annihilation (“positron range”), the noncollinearity of the two photons emitted, and detector geometry [26, 27]. The positron range is isotope specific as it depends on the energy with which the positrons are emitted. It can range from <1 mm to >5 mm for common PET isotopes; for [18 F]fluorine it is approximately 0.7 mm [26, 28]. Clinical PET scanners have a typical resolution on the order of several millimeters. Submillimeter resolution is possible
P1: OTA/XYZ P2: ABC c01 JWBS037-Chen February 2, 2011
21:44
Printer Name: Yet to Come
MAJOR IMAGING MODALITIES
11
with dedicated small animal scanners, allowing detailed studies of mouse, rat, and monkey models, including brain imaging [25, 29]. Contrary to PET, SPECT directly measures single gamma-ray photons emitted by the decay of the chosen radioisotope. The scanner consists of scintillator/PMT detector heads that slowly rotate around the object in a stepwise fashion, collecting data every few degrees for a certain time. Since only single photons are measured, collimators are used to obtain directional information. Only photons traveling along the axis of the collimator are captured by the detector. This allows SPECT to reach resolutions comparable to PET (Table 1.1), but only at the cost of reduced sensitivity. Depending on the collimator used, sensitivity is reduced by at least about one order of magnitude [7, 11]. On the other hand, since the isotopes commonly used for SPECT have relatively long half-lives on the order of several hours to days, longitudinal studies are greatly facilitated and shipping from distant production facilities is possible. Because of the chemistry of the isotopes involved, straightforward preparation (radiometal chelation) protocols exist for many SPECT imaging probes. In addition, different isotopes can be distinguished based on the isotope-specific energies of the emitted photons (e.g., [123 I]iodine: 159 keV; [99m Tc]technetium: 140 keV; or [111 In]indium: 171, and 245 keV). 1.3.3 Magnetic Resonance Imaging (MR Imaging or MRI) MRI is based on the principles of nuclear magnetic resonance (NMR). The nuclei of certain isotopes, most notably naturally occurring hydrogen (i.e., protons), possess nuclear spins that can be aligned along strong external magnetic fields such as those produced inside an MRI scanner. Alignment can be either parallel or antiparallel to the field. Similar to spinning top toys, the spins are not totally aligned with the field but precess around the field vector, resulting in what can be viewed as cone-shaped distributions. Following the application of transverse radiofrequency (rf ) pulses to perturb both the longitudinal and transverse alignments, the nuclear spins return to their previous (ground) state, generating a faint signal detected by a receiver coil. Longitudinal (T1 recovery) and transverse (T2 decay) relaxations are measured by the scanner and form the basis for the tomographic image. Relaxation times are dependent on the local environment of the nuclei and on their density within the tissue. Diseased tissue may therefore be detected by its effect on the local density and environment. Furthermore, the rf pulses and their sequence can be modified in order to weight images toward primarily detecting differences in T1 or T2. Even though protons are virtually ubiquitous throughout the body as part of water and fat molecules, physical limitations, most notably the fact that only a miniscule fraction of spins are aligned properly along the field to generate a signal, result in the relatively low sensitivity of MRI, coupled with long acquisition times (Table 1.1) [11]. Still, MRI is widely used as it permits imaging of soft tissue at high resolution without depth limitation. While images can be acquired by solely relying on native protons (endogenous contrast), the image quality can be greatly improved by administration of certain contrast-enhancing agents. After accumulation in the tissue, these agents change the protons’ local environment, thereby shortening their relaxation times (T1 and T2, measured as relaxivities R1 = 1/T1, and R2 = 1/T2). Gadolinium-chelate-based paramagnetic agents are frequently used to increase T1 contrast, resulting in local brightening in the tomographic image. Newly developed T1 contrast agents include long-circulating manganese oxide (MnO) nanoparticles [30] and gadolinium-containing micelles and liposomes amendable to functionalization [31].
P1: OTA/XYZ P2: ABC c01 JWBS037-Chen February 2, 2011
12
21:44
Printer Name: Yet to Come
BASIC PRINCIPLES OF MOLECULAR IMAGING
To obtain negative contrast enhancement (darkening) in T2-weighted images, superparamagnetic particles such as iron oxide nanopraticles are typically used. Biocompatibility and ease of synthesis have made colloidal superparamagnetic iron oxide (SPIO) and ultrasmall superparamagnetic iron oxide (USPIO) particles particularly popular. In addition, the continued pursuit of improved magnetic properties has yielded metal-doped (MnFe2 O4 , FeFe2 O4 , CoFe2 O4 , NiFe2 O4 ), metallic (coated iron, cobalt, nickel), and bimetallic (FePt, FeCo) nanoparticles [32]. Composition, size, surface characteristics, and potential aggregation of the particles, as well as external experimental parameters, all can have an effect on relaxivity and thus the MRI image [32]. Because of the abundance of hydrogen atom within the body, clinical MRI still to a large extent relies on the detection of protons, using contrast agents or high-field MRI scanners to improve contrast [11]. Another way to increase signal strength is to employ hyperpolarized agents. Optical pumping can greatly enrich the excited spin state of certain isotopes (e.g., [3 He]helium, [129 Xe]xenon), resulting in signal enhancements of up to 106 [11]. However, because the hyperpolarized state lasts only a few seconds, the agents need to be prepared immediately prior to imaging and, even then, the time window available for imaging is extremely limited. In addition to the isotopes mentioned so far, certain isotopes of other elements can be detected as well, provided their natural abundance is high enough (e.g., [23 Na]sodium, [31 P]phosphorus) or they can be administered in sufficient quantities (e.g., [17 O]oxygen as water, [13 C]carbon as glucose, [19 F]fluorine as fluorocarbons). Because MRI is based on NMR detection, it is possible to tune special scanners to specifically detect spectra of different nuclei. This forms the basis for magnetic resonance spectroscopy (MRS), which is beginning to open the door for quantitative or semiquantitative physiological and molecular MRI by tracing the distribution of individual compounds (e.g., cancer-related metabolites) [33]. To map brain activity, functional MRI (fMRI) detects blood-oxygenlevel-dependent changes in the environment of the iron contained in hemoglobin, a measure for the hemodynamic response linked to neural activity. Other novel functional imaging techniques include diffusion-weighted imaging (DWI) and diffusion tensor imaging (DTI) as well as perfusion MRI [33]. DWI and DTI assess the movement of free water molecules in tissue. Perfusion MRI measures blood-transport dynamics by following signal changes with special imaging protocols (dynamic susceptibility contrast MRI, DSC-MRI; dynamic contrast enhanced MRI, DCE-MRI). 1.3.4 Ultrasound (US) Imaging Ultrasound imaging records the reflections of high-frequency sound waves by internal body structures. Focused sound waves, originating from a transducer placed on the skin, propagate through the body and generate echoes as they encounter tissues of different density. The echoes, in turn, are captured again by the transducer and the data are processed to generate digital images. By performing Doppler sonography, velocity can be measured; this approach is commonly employed for blood-flow examinations. Thus physiology as well as anatomy can be imaged with US [8]. Resolution can be increased by choosing higher frequencies, which, however, lead to a concomitant decrease in penetration depth. Since US imaging does not involve the use of ionizing radiation, it is considered a particularly “safe” imaging modality. Therefore, and because of relatively low cost, ease of use, and wide availability, US is the most popular clinical imaging modality [9]. The image quality depends on sound frequency, sound speed, sound attenuation, sound backscatter, data processing, and
P1: OTA/XYZ P2: ABC c01 JWBS037-Chen February 2, 2011
21:44
Printer Name: Yet to Come
MOLECULAR IMAGING PROBES
13
transducer handling [11]. Because of the latter, US imaging is more dependent on the experience and skill of the operator than the other imaging modalities. Although not necessary for US imaging, contrast agents can be used. Because of their ability to significantly increase the signal intensity, gas-containing microbubbles, along with echogenic liposomes and perofluoro-emulsion nanoparticles, ranging in size from about 100 nm to 8 m, are most commonly used [8]. They allow for greatly improved imaging of small blood vessels that are otherwise difficult to distinguish from surrounding tissues. Their use, however, is in fact largely limited to imaging the vascular compartment because their particle size precludes escape from blood vessels (“extravasation”). Despite this considerable limitation, clinically approved US contrast agents have been developed and targeted US imaging has been demonstrated with surface-modified microbubbles for diseases that result in molecular changes to the vascular compartment. For example, RGDpeptide modified microbubbles have been used to target the integrin ␣v 3 , a cell-surface receptor linked to tumor angiogenesis [8]. US contrast agents such as microbubbles can also be used for drug delivery. It has been shown that drug-carrying bubbles can be fragmented by high-intensity ultrasound, leading to the local release of a therapeutic agent [34]. 1.3.5 Computed Tomography (CT) For CT a series of X-ray images is acquired and reconstructed into a tomographic image. Focused X-rays emerge from a source rotating stepwise around the subject and, after traveling through the body, are captured by a detector situated opposite the source. Contrast is based on differences in attenuation of the X-rays as they pass through the body. Dense tissue (bones) absorbs significantly more energy than soft tissue, while air (lung) absorbs less. As a result, CT is primarily an anatomical imaging modality that is especially useful for bone and lung imaging but does not show particularly good soft tissue contrast. For contrast-enhanced CT X-ray absorbing agents are administered. Typically, large amounts of iodine-based compounds have to be administered intravenously to achieve an appreciable soft tissue contrast. The compounds have a largely nonspecific distribution and fast clearance kinetics, although work on improving these characteristics is ongoing [9]. For contrast-enhanced CT of the gastrointestinal tract, insoluble contrast agents based on barium sulfate are given orally or administered as an enema. Recently, lanthanide-bearing compounds, bismuth sulfide nanoparticles, and targeted gold nanorods have been studied as potential future CT contrast agents with increased image contrast and longer circulation times than the standard iodine agents [35–37]. Although CT is just at the verge of also being explored as a molecular imaging modality, it is already extremely useful in providing a high-resolution anatomical framework in which to interpret molecular imaging (e.g., PET or SPECT) data (Fig. 1.1) [13]. The use of PET/CT and SPECT/CT hybrid scanners has improved diagnostic accuracy in the clinic and facilitated preclinical drug evaluation [5].
1.4 MOLECULAR IMAGING PROBES Even when ex vivo diagnostic tests (e.g., blood, urine, Pap smear) exist for a disease [1], the ability to obtain data on location and local amount of a disease-specific indicator, and thereby the disease itself, within the body is what makes molecular imaging such a powerful tool. Nowhere is this perhaps more evident than in oncology where tumor heterogeneity, possible involvement of lymph nodes, and the potential of metastases create an extremely
P1: OTA/XYZ P2: ABC c01 JWBS037-Chen February 2, 2011
14
21:44
Printer Name: Yet to Come
BASIC PRINCIPLES OF MOLECULAR IMAGING
challenging diagnostic target. Detection based on disease-related changes in the local environment provides one possible avenue. This includes imaging of metabolic rate (e.g., glucose consumption), cell proliferation (i.e., DNA synthesis), or a change in the microenvironment (e.g., local pH, concentration of oxygen or proteases) typically associated with cancers. Indeed, a number of small-molecule compounds have been developed into highly successful probes for these conditions, as exemplified by the PET-probes 18 F-FDG (for glucose consumption), 18 F-FLT (for cell proliferation), and 18 F-FMISO or 64 Cu-ATSM (for hypoxia) [38]. Even though these probes for local conditions provide very useful information, they still tend to be relatively nonspecific and may miss slow-growing neoplasms [38] or tumors in certain locations such as excretory organs that have a high nonspecific background signal. Since upregulation of intracellular markers and cell-surface receptors is a hallmark of disease states, aiming for these disease-specific targets is highly attractive, provided that suitable ligands are available. For imaging, cell-surface receptors may be favored over intracellular targets because they do not require internalization into the cell and therefore can be accessed more readily, particularly by nanoparticle-based imaging probes. If the nanoparticle is designed for concomitant drug or gene delivery, subsequent internalization into the cells becomes important. Giving examples from oncology, Table 1.2 illustrates the range of receptor targets and targeting ligands presently available. The examples are taken from MRI imaging [39], but are equally representative for other molecular imaging modalities [12, 18, 38, 40]. Most imaging probes are administered intravenously and allowed to distribute through the body; Notable exceptions include experimental implantable microprobes containing nanoparticles [41], intrinsically produced optical imaging probes (i.e., fluorescent or bioluminescent proteins [42]) expressed by reporter genes, and locally injected imaging probes for sentinel lymph node mapping during cancer surgery [43]. Typically, as the individual probes circulate through the body they can accumulate in tissues, be excreted from the body, and/or be metabolized. To be clinically useful, molecular imaging probes should be based on high-affinity high-specificity ligands and show a relatively low background signal [44]. The ratio of imaging probe in the target tissue versus in the surrounding tissues (i.e., the signal-to-noise ratio) determines how much image contrast can be achieved. In other words, the image depends on how well the imaging probe has been taken up in the target tissue and how well it has cleared from other regions of the body at the time the image is acquired. Therefore some uptake period is frequently required between injection and imaging. Depending on the desired information and the pharmacokinetics of the imaging probe, it can range from a very short time (e.g., seconds to minutes for [15 O]oxygen-water in blood flow studies [45] to over tens of minutes (e.g., for 18 F-FDG) to days (e.g., for antibodies). Size has a significant impact on the time imaging probes circulate in the body. Generally, small molecules are eliminated fast, resulting in a rapidly decreasing background signal, while larger compounds, including nanoparticles [46], have longer circulation times. Thus, as outlined earlier, for radionuclide-based imaging the radioactive half-life of the radioisotope has to be matched to the pharmacokinetic profile of the imaging probe to ensure that a sufficient radioactive signal remains by the time an adequate amount of the imaging probe has accumulated in the target tissue in order to obtain statistically significant contrast. For nanoparticles this usually requires radioisotopes with half-lives on the order of several hours to days. Algorithms help to choose radioisotopes with an appropriate half-life: by calculating a suitable “imaging figure of merit” (IFOM) that takes into account parameters such as the size of the target organ (or tumor), the measured decay corrected organ uptake (% injected
15
Source: Peng [39] and references therein.
Magnetism-engineered iron oxide (MEIO) nanoparticles PEG-IO
SPIO USPIO PEG-SPIO Streptavidin-SPIO
Chlorotoxin
RGD4C Luteinizing hormone releasing hormone (LHRH) CREKA peptide Arg-Gly-Asp (RGD) Folic acid Antibody to prostate-specific membrane antigen (PSMA) Herceptin
Folic acid F3 peptide
Monoclonal antibody A7 Herceptin Methotrexate Chlorotoxin peptide
Folic acid
Monoclonal antibody-610 Antibody to carcinoembryonic antigen (CEA) Monoclonal antibody L6 Transferrin Monoclonal antibody-Her-2/neu EPPT peptide
Targeting Ligands
Membrane-bound matrix metalloproteinase-2 (MMP-2)
Her-2/neu receptors
Clotted plasma proteins ␣v 3 integrins Folate receptor PSMA
Rat glioma
NIH3T6.7
In vivo
In vivo
In vivo In vivo In vivo In vitro
In vitro In vivo
Melanoma cells Breast cancer Breast cancer Human epidermoid carcinoma Human epithelial mouth carcinoma Human prostate cancer cells
In vitro In vivo
In vivo In vivo In vitro In vitro
In vitro
In vivo
In vivo In vivo In vitro
In vitro In vivo
Experimental Conditions
Human cervical cancer cells Rat glioma
Colorectal carcinoma NIH3T6.7 Human cervical cancer cells Rat glioma
Breast, colon, pancreas, and lung cancer cell lines Human epithelial mouth carcinoma
Underglycosylated mucin-1 antigen (uMUC-1) Folate receptor
Colorectal tumor antigen Her-2/neu receptors Folate receptor Membrane-bound matrix metalloproteinase-2 (MMP-2) Folate receptor Surface-localized tumor vasculature ␣v 3 integrins LHRH receptor
Intracranial tumor LX-1 Rat mammary carcinoma Breast cancer
Colon carcinoma cell lines Colon tumor
Tumor
Surface antigen Transferrin receptor Her-2/neu receptors
Surface antigen CEA
Targets
21:44
Biofunctional PEG-SPIO SPIO encapsulated with photodynamic agent HFn-IO SPIO
Dextran-coated superparamagnetic maghemite (␥ -Fe2 O3 ) nanocrystals Ferumoxides (SPIO) Iron oxide nanocrystals (Fe3 O4 ) SPIO SPIO
CLIO-NH2
MINO USPIO Streptavidin-conjugated SPIO
USPIO SPIO
Iron Oxide Nanoparticles
TABLE 1.2 Targeted Iron Oxide Nanoparticles for Tumor Imaging with MRI
P1: OTA/XYZ P2: ABC c01 JWBS037-Chen February 2, 2011 Printer Name: Yet to Come
P1: OTA/XYZ P2: ABC c01 JWBS037-Chen February 2, 2011
16
21:44
Printer Name: Yet to Come
BASIC PRINCIPLES OF MOLECULAR IMAGING
dose/gram, % ID/g) or organ activity (%ID/organ), the area under the curve (AUC) and their respective changes over time, predictions for an optimal imaging time, and thus radioactive half-life are possible [24, 47]. Similar calculations can be performed to obtain a “therapeutic figure of merit” (TFOM). Allometric analysis is used to obtain predictions for clinical trials of a novel probe based on preclinical data from animal models. Most imaging modalities can yield quantitative data after careful calibration (Table 1.1). The previous paragraph illustrates the importance of quantification during early probe development and for clinical translation. Once an imaging probe has reached this step, quantitative image-based data analysis becomes even more relevant. Newly developed imaging probes need to be evaluated in patients, and sensitivity (ideally 100% = no falsenegative examinations) and specificity (ideally 100% = no false-positive examinations) need to be established and compared to standard test methods (i.e., a “gold standard” such as biopsy/histopathology). Since molecular imaging is founded on the ability to quantify specific disease-related molecular profiles and target densities, the clinical goal is to go beyond mere detection of disease and to permit staging, restaging, and evaluation of response to treatment. To that end, image-derived quantitative parameters for evaluating the progression of a disease have been developed. For example, in oncology the “standard uptake value” (SUV) of a PET probe within a chosen “region of interest” (ROI)—the tumor—can be used as prognostic indicator [4, 48] and response to treatment can be assessed based on the change in SUV, often soon after the treatment since molecular changes precede changes detectable by anatomical or physiological imaging [49]. Regardless of the imaging modality, important parameters of an imaging protocol include (1) the imaging probe and its formulation, (2) the type and model of the scanner/detector, (3) patient selection and preparation, (4) data acquisition, (5) data processing, and (6) data analysis. All parameters need to be carefully controlled, recorded, and analyzed to yield results that are comparable over time and between different hospitals [50]. This is a prerequisite for meaningful results from clinical trials evaluating new imaging probes or new therapy regimens, and for subsequent clinical use based on reliable image acquisition and data handling protocols approved by regulatory agencies [49]. Ultimately, to be successful in clinical use, a new imaging probe needs to pass several tests: Is it better than competing diagnostic approaches? Does it have a cost-effective role in clinical management? Is the interpretation of the imaging data easy [4]? Once established, a well-studied imaging probe such as 18 F-FDG can then also play an important role in future drug development [51]. 1.4.1 Nanoparticle-Based Probes Molecular imaging can be performed with imaging probes of vastly different sizes. Entities can be the size of single atoms (e.g., a 18 F fluoride ion), small molecules (e.g., the modified glucose 18 F-FDG, the radiometal complex 99m Tc-sestamibi), peptides, antibodies, or constructs approaching the dimension of small cells (e.g., micrometer-sized microbubbles [34]). Typical “nanoparticles” fall somewhere in the middle of this range. As outlined below, to be clinically useful the preferred size of nanoparticles is roughly between 10 and 100 nm [52]. Thus this classification covers compounds being similar in size to hemoglobin (6.5 nm), an antibody (12 nm), a hepatitis virus (45 nm), or an influenza virus (130 nm), although the constructs approaching the micrometer range are also commonly included in the definition. The nanoparticles themselves can extend in one to three dimensions, as depicted schematically in Figure 1.2, and show unique size- and shape-dependent properties that can be beneficial for imaging.
P1: OTA/XYZ P2: ABC c01 JWBS037-Chen February 2, 2011
21:44
Printer Name: Yet to Come
MOLECULAR IMAGING PROBES
Quantum Dot
Nanoshell
Gold
Liposome
Iron Oxide
FeCo
Perfluorocarbon
Nanotube
Dendrimer
17
FIGURE 1.2 Schematic depictions of representative nanoparticles for targeted molecular imaging in living subjects. (Reproduced with permission from Cai [9].)
Nanoparticle-based probes are receiving considerable attention because the particles form a convenient platform for nanodiagnostics and nanomedicines that can combine different clinically relevant properties in a single unit. Each particle can carry highly specific targeting ligands, antibodies (for targeting or treatment), different imaging probes, drugs, or various combinations there of (Fig. 1.3). Depending on the nature of the particle, it itself can assume some of these roles. Examples include fluorescent quantum dots detectable by optical imaging, iron-based MRI contrast agents, carbon nanotubes used for targeted thermal ablation [53], and drug-carrying bubbles for ultrasound imaging. When an imaging
FIGURE 1.3 Schematic depiction of a targeted multifunctional nanoparticle carrying imaging probes, drugs, and antibodies. (Reproduced with permission from Rao [52].)
P1: OTA/XYZ P2: ABC c01 JWBS037-Chen February 2, 2011
18
21:44
Printer Name: Yet to Come
BASIC PRINCIPLES OF MOLECULAR IMAGING
probe, that is, a diagnostic entity, is combined with a drug, the resulting compound is also called a “theranostic” or “theragnostic” [40, 54, 55]. Ideally, this integrated theranostic approach will permit diagnosis, targeted delivery of therapy, and monitoring of response to treatment Sumer 2008 [56]. By choosing the right combination of targeting moieties and therapeutic components, the theranostic can home in on multiple markers and target cell subpopulations, thereby providing a tool to address disease heterogeneity and adaptive resistance. The treatment can, if necessary, quickly be adjusted based on the rapid feedback obtained. Although still in its early stages, the hope is that this approach will allow true “personalized medicine” based on individualized, molecular diagnosis within a heterogeneous disease and patient population, something that cannot be accomplished by taking small biopsy samples or following standardized “one-dose-fits-all” protocols. As already alluded to in the previous paragraphs, nanoparticle-based probes offer several advantages over standard probes Sanvicens 2008 [57, 58]. First, the size of nanoparticles allows them to carry a large number of targeting ligands. The binding behavior is influenced by the ligands’ affinities, their density, and by the combination of the chosen ligands. Because of the resulting multivalent binding (“avidity”), a nanoparticle can exhibit high-affinity binding even though low-affinity ligands were employed [59] and different receptors may be targeted simultaneously. Second, moieties for multiple imaging modalities can be combined onto a single particle. This makes the nanoparticles amendable to dual-/multimodality imaging and allows one to draw on the strengths of different imaging modalities (e.g., high resolution and high sensitivity, Table 1.1). For a comparable small-molecule probe, achieving desirable signal strengths for each of the different imaging modalities can be challenging because the ratio of the different imaging moieties (e.g., an optical dye molecule and a radioisotope-bearing prosthetic group) cannot be easily changed [60] and the introduction of additional groups can significantly affect pharmacokinetics. By contrast, the ratio of the different moieties can be tailored more easily on nanoparticles, be it to achieve simultaneous detection in all modalities, or to use a high-sensitivity “beacon” (e.g., a PET-isotope) to trace low-sensitivity particles (e.g., MRI particles) at very low concentrations [61, 62]. Third, nanoparticles used as theranostics can carry large numbers of therapeutic entities like small molecules or peptides. Depending on the type of nanoparticle used, these payloads can be stowed away in the core, eliminating deleterious interference with pharmacokinetics and biodistribution. In addition, different types of drug molecules can be combined in a single particle. Fourth, drug-delivery kinetics can be tuned for a particular target and monitored using molecular imaging. A “controlled release” after delivery to the target helps to protect the rest of the body from harmful side effects, while at the same time increasing the “effective dose” in the target tissue. The drugs may be released by a microenvironmental stimulus such as pH or enzymes [63] or by external stimuli. An example for the latter are drugcarrying microbubbles that can be tracked by conventional ultrasound and fragmented using destructive high-intensity US pulses, releasing the drug in the target region [34]. When drug delivery directly into the cells is desired [64], multimodality nanoparticles can be modified for efficient endocytosis by decoration with cell-penetrating groups such as Tat-peptides [65, 66] and tracked by molecular imaging. Furthermore, there is evidence that nanoparticles may be able to sidestep multidrug resistance involving protein efflux pumps [57]. Fifth, nanoparticle pharmacokinetics can be modified with surface modifications. To improve the pharmacokinetic properties and in vivo stability, the particles are oftentimes
P1: OTA/XYZ P2: ABC c01 JWBS037-Chen February 2, 2011
21:44
Printer Name: Yet to Come
MOLECULAR IMAGING PROBES
19
coated with biocompatible polymers such as polyethylene glycol (PEG) or the polysaccharide dextran. This coating effectively “hides” the particles from detection by the reticuloendothelial system (RES), which would otherwise rapidly eliminate them by opsonization and phagocytosis [63]. All these modifications have to be accomplished while keeping the hydrodynamic (HD) size roughly between 10 and 100 nm. If particles fall below the lower limit, renal filtration provides a rapid way of elimination, while, on the other hand, hepatic and splenic elimination becomes increasingly efficient for particles greater than 100 nm [46]. For solid tumors, effective delivery to the target tissue beyond the vascular compartment is another reason for the upper size limit: since the gaps in leaky vasculature are about 400 nm, extravasation is most efficient for particles well below this limit [67]. This fact also explains why microbubble-based US imaging is largely confined to the vascular compartment; for other—smaller—nanoparticles, this passive targeting phenomenon, termed “enhanced permeation and retention” (EPR) effect, together with a lack of efficient lymphatic drainage, represents an important way of tumor-directed delivery, particularly for smaller tumors [68]. Besides the hydrodynamic size, the charge of the nanoparticle can also affect pharmacokinetics and biodistribution. Ideally, a nanoparticle should exhibit a near neutral charge since positively charged particles can form aggregates with negatively charged plasma proteins, while negatively charged particles demonstrate increased liver uptake [69]. The positive effects of nanoparticle coatings on pharmacokinetics include shielding of surface charges and surface chemistry. The commercial availability of PEG in a wide range of lengths and modifications has made it particularly popular for fine-tuning of these pharmacokinetic properties. For example, one study showed that after relatively small, nontargeted quantum dots with identical 3.2-nm diameter InAs(ZnS) core were coated with PEG-chains of various lengths (PEG2 to PEG22), resulting in HD sizes ranging from 5 to 16 nm, the biodistribution and clearance characteristics varied widely [70]. Expectedly, PEGylation improved hydrophilicity, protected from opsonization, and circulation times in the blood increased with increasing chain length. In addition, by simply increasing the length of the PEG units, major organ uptake shifted from liver over kidneys and bladder to pancreas, intestine, and lymph nodes. However, it has to be underscored that in the study cited above, the quantum dots did not bear any specific targeting moiety. Fortunately for targeted nanoparticles, recent data indicate that the presence of targeting moieties may minimize the influences of other nanoparticle properties in vivo [71]: a preclinical study compared different carriers (linear polymer, dendrimer, and liposome) carrying an anticancer drug (paclitaxel) and/or an imaging probe (Cy5.5) in conjunction with a targeting moiety (a LHRH peptide). Notably, the authors conclude “that the architecture, composition, size and molecular mass of the receptor-targeted drug nanocarriers can be selected based on other than anticancer efficacy considerations (cost, type of active ingredients, difficulties in production, stability, patient compliance, etc.) ensuring that the high efficacy and low adverse side effects could be achieved automatically by tumor targeting” [emphasis added] [71]. If similar results are found in other studies, the focus might shift toward careful selection of the targeting ligand(s) while allowing a high degree of flexibility for the underlying platform [72]. 1.4.2 Challenges for Nanoparticle-Based Imaging Probes The body of work produced to date has demonstrated the promising potential of nanoparticle-based molecular imaging agents [9, 11, 12]. At the same time, the studies
P1: OTA/XYZ P2: ABC c01 JWBS037-Chen February 2, 2011
20
21:44
Printer Name: Yet to Come
BASIC PRINCIPLES OF MOLECULAR IMAGING
have shown some of the accompanying challenges Sanvicens 2008 [57]. They chiefly fall within three basic areas: compound preparation, toxicity, and pharmacokinetics/ pharmacodynamics. Understanding and addressing these challenges will also be highly important to gain the regulatory approval necessary for rapid and successful translation into the clinic. Some of the challenges are described next. 1. Size Control and Batch-to-Batch Reproducibility. In contrast to small-molecule probes, achieving good control over the size and precise composition of nanoparticles is more challenging. But since pharmacokinetics and, for theranostics, pharmacodynamics can be significantly influenced by these factors, gaining a thorough understanding of these manufacturing parameters and their in vivo consequences is crucial [46, 69, 71]. Taking magnetic nanoparticles used for MRI as an example, it is well known that the (superpara)magnetic properties are size dependent as surface effects play an increasing role for smaller particles and that magnetic properties can vary significantly depending on production chemistry [32, 73]. 2. Particle Stability During Storage and Formulation. Because of their size and surface properties, nanoparticles may aggregate over time. This and other deleterious changes such as decomposition and oxidation have to be evaluated. 3. Toxicity and Immune Response. The composition and shape of the nanoparticles can potentially have toxic effects. For example, many quantum dots popular for optical fluorescence imaging contain cadmium, lead, or selenium [74]. Magnetic nanoparticles used for MRI may contain gadolinium, cobalt, or nickel [31, 75]. Even if these toxic elements are shielded from direct contact with the in vivo environment by protective coatings, they may eventually leach out if residence times in the body are long. For carbon nanotubes, their shape rather than their composition could be cause for concern. Similar to quartz and asbestos, they have been shown to be capable of causing chronic inflammation and fibrosis [76]. In addition, while coatings are useful for shielding the core of the nanoparticles from the immune system, if antibodies are used for targeting, they necessarily need to be exposed to the in vivo environment, thus posing a risk for an immune response. Fortunately, less immunogenic antibody-derived targeting moieties may be available and biodegradable low-toxicity nanoparticles are being developed [72]. 4. Control Over Pharmacokinetics/Pharmacodynamics. Nanoparticles demonstrate long blood circulation times and, because of their size, show a tendency to undergo opsonization and clear the body via the liver and spleen. Clearance through the reticuloendothelial system is slower than the excretion in the urine encountered for many small molecules [69]. The longer residence time in excretory organs may not be a significant concern for imaging probes administered in trace amount, but could become pharmacodynamically relevant for theranostics carrying highly potent drug molecules or therapeutic radioisotopes. In the end, solving some of these challenges could be easier than it appears at present. As the recent study by Saad and co-workers has indicated [71], depending on which information is desired about the target, choosing the imaging modality (or modalities) and targeting moiety may be the most important factors to consider when designing the imaging agent, while leaving considerable flexibility for the underlying platform. Specifically, the authors found “that tumor-specific targeting minimized the differences between nanocarriers of distinct architecture, size, mass, and composition” [71]. If these findings represent a
P1: OTA/XYZ P2: ABC c01 JWBS037-Chen February 2, 2011
21:44
Printer Name: Yet to Come
REFERENCES
21
general principle, they could have wide-ranging implications for the rapidly growing field of nanoparticle-based diagnostic and treatment tools. Thus despite the seemingly overwhelming number of potential combinations of sizes, shapes, compositions, surface coatings, targeting moieties, imaging modalities, functional groups, and potential drug molecules, in the future the challenge may not be so much trying to discover a molecular imaging agent for a particular target, but rather to assemble the best combination of components from the choices available in the toolbox.
REFERENCES 1. Ludwig, J. A.; Weinstein, J. N. Biomarkers in cancer staging, prognosis and treatment selection. Nat. Rev. Cancer 2005, 5, 845–856. 2. Tan, A. R.; Swain, S. M. Ongoing adjuvant trials with trastuzumab in breast cancer. Semin. Oncol. 2003, 30, 54–64. 3. Menard, S.; Casalini, P.; Campiglio, M.; Pupa, S. M.; Tagliabue, E. Role of HER2/neu in tumor progression and therapy. Cell. Mol. Life Sci. 2004, 61, 2965–2978. 4. Hoh, C. K. Clinical use of FDG PET. Nucl. Med. Biol. 2007, 34, 737–742. 5. Cherry, S. R. Multimodality in vivo imaging systems: twice the power or double the trouble? Annu. Rev. Biomed. Eng. 2006, 8, 35–62. 6. Society of Nuclear Medicine, Glossary of Molecular Imaging Terms. Available at http:// www.molecularimagingcenter.org/index.cfm?PageID=7834 (accessed April 3, 2009) 7. Baert, A. L. In Encyclopedia of Diagnostic Imaging. Springer-Verlag: New York, 2008, p. 1618. 8. Bloch, S. H.; Dayton, P. A.; Ferrara, K. W. Targeted imaging using ultrasound contrast agents. IEEE Eng. Med. Biol. Mag. 2004, 23, 18–29. 9. Cai, W. B.; Chen, X. Y. Nanoplatforms for targeted molecular imaging in living subjects. Small 2007, 3, 1840–1854. 10. Ferrara, K. W. Driving delivery vehicles with ultrasound. Adv. Drug Deliv. Rev. 2008, 60, 1097–1102. 11. Willmann, J. K.; van Bruggen, N.; Dinkelborg, I. M.; Gambhir, S. S. Molecular imaging in drug development. Nat. Rev. Drug Discov. 2008, 7, 591–607. 12. Weissleder, R.; Pittet, M. J. Imaging in the era of molecular oncology. Nature 2008, 452, 580–589. 13. Shah, C.; Patton, J. A.; Sandler, M. P. How much CT is needed in nuclear medicine. Eur. J. Nucl. Med. Mol. Imaging 2008, 35, 1759–1760. 14. von Schulthess, G. K.; Schlemmer, H. P. W. A look ahead: PET/MR versus PET/CT. Eur. J. Nucl. Med. Mol. Imaging 2009, 36, 3–9. 15. Moser, E.; Stadlbauer, A.; Windischberger, C.; Quick, H. H.; Ladd, M. E. Magnetic resonance imaging methodology. Eur. J. Nucl. Med. Mol. Imaging 2009, 36, 30–41. 16. Pichler, B. J.; Wehrl, H. F.; Judenhofer, M. S. Latest advances in molecular imaging instrumentation. J. Nucl. Med. 2008, 49, 5S–23S. 17. Ntziachristos, V.; Chance, B. Probing physiology and molecular function using optical imaging: applications to breast cancer. Breast Cancer Res. 2001, 3, 41–46. 18. Rudin, M.; Weissleder, R. Molecular imaging in drug discovery and development. Nat. Rev. Drug Discov. 2003, 2, 123–131. 19. So, M. K.; Xu, C. J.; Loening, A. M.; Gambhir, S. S.; Rao, J. H. Self-illuminating quantum dot conjugates for in vivo imaging. Nat. Biotechnol. 2006, 24, 339–343.
P1: OTA/XYZ P2: ABC c01 JWBS037-Chen February 2, 2011
22
21:44
Printer Name: Yet to Come
BASIC PRINCIPLES OF MOLECULAR IMAGING
20. Mather, S. Molecular imaging with bioconjugates in mouse models of cancer. Bioconjug. Chem. 2009, 20, 631–643. 21. Levenson, R. M.; Mansfield, J. R. Multispectral imaging in biology and medicine: slices of life. Cytometry A 2006, 69A, 748–758. 22. Evans, C. L.; Potma, E. O.; Puoris’haag, M.; et al. Chemical imaging of tissue in vivo with videorate coherent anti-Stokes Raman scattering microscopy. Proc. Natl. Acad. Sci. U.S.A. 2005, 102, 16807–16812. 23. Fujimoto, J. G. Optical coherence tomography for ultrahigh resolution in vivo imaging. Nat. Biotechnol. 2003, 21, 1361–1367. 24. Williams, L. E.; Lopatin, G.; Kaplan, D. D.; Liu, A.; Wong, J. Y. C. Update on selection of optimal radiopharmaceuticals for clinical trials. Cancer Biother. Radiopharm. 2008, 23, 797–806. 25. Cherry, S. R. The 2006 Henry N. Wagner lecture: of mice and men (and positrons)—advances in PET imaging technology. J. Nucl. Med. 2006, 47, 1735–1745. 26. Levin, C. S.; Hoffman, E. J. Calculation of positron range and its effect on the fundamental limit of positron emission tomography system spatial resolution. Phys. Med. Biol. 1999, 44, 781–799. 27. Lecomte, R. Technology challenges in small animal PET imaging. Nucl. Instrum. Methods Phys. Res. A—Accelerators Spectrometers Detectors and Associated Equipment 2004, 527, 157–165. 28. Nayak, T. K.; Brechbiel, M. W. Radioimmunoimaging with longer-lived positron-emitting radionuclides: potentials and challenges. Bioconjug. Chem. 2009, 20, 825–841. 29. Ambrosini, V.; Quarta, C.; Nanni, C.; et al. Small animal PET in oncology: the road from bench to bedside. Cancer Biotherapy Radiopharm. 2009, 24, 277–285. 30. Na, H. B.; Lee, J. H.; An, K. J.; et al. Development of a T1 contrast agent for magnetic resonance imaging using MnO nanoparticles. Angew. Chem. Int. Ed. 2007, 46, 5397–5401. 31. Castelli, D. D.; Gianolio, F.; Crich, S. G.; Terreno, E.; Aime, S. Metal containing nanosized systems for MR—molecular imaging applications. Coord. Chem. Rev. 2008, 252, 2424–2443. 32. Sun, C.; Lee, J. S. H.; Zhang, M. Q. Magnetic nanoparticles in MR imaging and drug delivery. Adv. Drug Deliv. Rev. 2008, 60, 1252–1265. 33. Jenkinson, M. D.; du Plessis, D. G.; Walker, C.; Smith, T. S. Advanced MRI in the management of adult gliomas. Br. J. Neurosurg. 2007, 21, 550–561. 34. Ferrara, K.; Pollard, R.; Borden, M. Ultrasound microbubble contrast agents: fundamentals and application to gene and drug delivery. Annu. Rev. Biomed. Eng. 2007, 9, 415–447. 35. Zielhuis, S. W.; Nijsen, J. F.; Seppenwoolde, J. H.; et al. Lanthanide bearing microparticulate systems for multi-modality imaging and targeted therapy of cancer. Curr. Med. Chem. Anticancer Agents 2005, 5, 303–313. 36. Rabin, O.; Perez, J. M.; Grimm, J.; Wojtkiewicz, G.; Weissleder, R. An X-ray computed tomography imaging agent based on long-circulating bismuth sulphide nanoparticles. Nat. Mater. 2006, 5, 118–122. 37. Popovtzer, R.; Agrawal, A.; Kotov, N. A.; et al. Targeted gold nanoparticles enable molecular CT imaging of cancer. Nano Lett. 2008, 8, 4593–4596. 38. Gambhir, S. S. Molecular imaging of cancer with positron emission tomography. Nat. Rev. Cancer 2002, 2, 683–693. 39. Peng, X. H.; Qian, X. M.; Mao, H.; et al. Targeted magnetic iron oxide nanoparticles for tumor imaging and therapy. Int. J. Nanomed. 2008, 3, 311–321. 40. Del Vecchio, S.; Zanneti, A.; Fonti, R.; Pace, L.; Salvatore, M. Nuclear imaging in cancer theranostics. Q. J. Nucl. Med. Mol. Imaging 2007, 51, 152–163. 41. Daniel, K. D.; Kim, G. Y.; Vassiliou, C. C.; et al. Implantable diagnostic device for cancer monitoring. Biosensors & Bioelectronics 2009, 24, 3252–3257.
P1: OTA/XYZ P2: ABC c01 JWBS037-Chen February 2, 2011
21:44
Printer Name: Yet to Come
REFERENCES
23
42. Gross, S.; Piwnica-Worms, D. Spying on cancer: molecular imaging in vivo with genetically encoded reporters. Cancer Cell 2005, 7, 5–15. 43. Barrett, T.; Choyke, P. L.; Kobayashi, H. Imaging of the lymphatic system: new horizons. Contrast Media & Mol. Imaging 2006, 1, 230–245. 44. Hausner, S. H.; Abbey, C. K.; Bold, R. J.; et al. Targeted in vivo imaging of integrin alpha(v)beta(6) with an improved radiotracer and its relevance in a pancreatic tumor model. Cancer Res. 2009, 69, 5843–5850. 45. Ito, H.; Inoue, K.; Goto, R.; et al. Database of normal human cerebral blood flow measured by SPECT: I. Comparison between I-123-IMP, Tc-99m-HMPAO, and Tc-99m-ECD as referred with O-15 labeled water PET and voxel-based morphometry. Ann. Nucl. Med. 2006, 20, 131–138. 46. Alexis, F.; Pridgen, F.; Molnar, L. K.; Farokhzad, O. C. Factors affecting the clearance and biodistribution of polymeric nanoparticles. Mol. Pharm. 2008, 5, 505–515. 47. Williams, L. E.; Wu, A. M.; Yazaki, P. J.; et al. Numerical selection of optimal tumor imaging agents with application to engineered antibodies. Cancer Biother. Radiopharm. 2001, 16, 25–35. 48. Weber, W. A. Positron emission tomography as an imaging biomarker. J. Clin. Oncol. 2006, 24, 3282–3292. 49. Krause, B. J.; Herrmann, K.; Wieder, H.; zum Buschenfelde, C. M. 18 F-FDG PET and 18 F-FDG PET/CT for assessing response to therapy in esophageal cancer. J. Nucl. Med. 2009, 50(Suppl 1), 89S–96S. 50. McLennan, G.; Clarke, L.; Hohl, R. J. Imaging as a biomarker for therapy response: cancer as a prototype for the creation of research resources. Clin. Pharmacol. Ther. 2008, 84, 433–436. 51. Weber, W. A.; Czernin, J.; Phelps, M. F.; Herschman, H. R. Technology insight: novel imaging of molecular targets is an emerging area crucial to the development of targeted drugs. Nat. Clin. Practice Oncol. 2008, 5, 44–54. 52. Rao, J. H. Shedding light on tumors using nanoparticles. ACS Nano 2008, 2, 1984–1986. 53. Kim, K. Y. Nanotechnology platforms and physiological challenges for cancer therapeutics. Nanomed. Nanotechnol. Biol. Med. 2007, 3, 103–110. 54. Warner, S. Diagnostics plus therapy = theranostics. Scientist 2004, 18, 38–39. 55. Pene, F.; Courtine, E.; Cariou, A.; Mira, J. P. Toward theragnostics. Crit. Care Med. 2009, 37, S50–S58. 56. Sumer, B.; Gao, J. M. Theranostic nanomedicine for cancer. Nanomedicine 2008, 3, 137–140. 57. Sanvicens, N.; Marco, M. P. Multifunctional nanoparticles—properties and prospects for their use in human medicine. Trends Biotechnol. 2008, 26, 425–433. 58. Davis, M. E.; Chen, Z.; Shin, D. M. Nanoparticle therapeutics: an emerging treatment modality for cancer. Nat. Rev. Drug Discov. 2008, 7, 771–782. 59. Hong, S.; Leroueil, P. R.; Majoros, I. J.; et al. The binding avidity of a nanoparticle-based multivalent targeted drug delivery platform. Chem. Biol. 2007, 14, 107–115. 60. Edwards, W. B.; Akers, W. J.; Ye, Y. P.; et al. Multimodal imaging of integrin receptor-positive tumors by bioluminescence, fluorescence, gamma scintigraphy, and single-photon emission computed tomography using a cyclic RGD peptide labeled with a rear-infrared fluorescent dye and a radionuclide. Mol. Imaging 2009, 8, 101–110. 61. Devaraj, N. K.; Keliher, F. I.; Thurber, G. M.; Nahrendorf, M.; Weissleder, R. F-18 labeled nanoparticles for in vivo PET-CT imaging. Bioconjug. Chem. 2009, 20, 397–401. 62. Lee, S.; Chen, X. Y. Dual-modality probes for in vivo molecular imaging. Mol. Imaging 2009, 8, 87–100. 63. Gullotti, E.; Yeo, Y. Extracellularly activated nanocarriers: a new paradigm of tumor targeted drug delivery. Mol. Pharm. 2009, 6, 1041–1051.
P1: OTA/XYZ P2: ABC c01 JWBS037-Chen February 2, 2011
24
21:44
Printer Name: Yet to Come
BASIC PRINCIPLES OF MOLECULAR IMAGING
64. Bae, Y. H. Drug targeting and tumor heterogeneity. J. Control. Release 2009, 133, 2–3. 65. Koch, A. M.; Reynolds, F.; Kircher, M. F.; et al. Uptake and metabolism of a dual fluorochrome tat-nanoparticle in HeLa cells. Bioconjug. Chem. 2003, 14, 1115–1121. 66. Bullok, K. E.; Gammon, S. T.; Violini, S.; et al. Permeation peptide conjugates for in vivo molecular imaging applications. Mol. Imaging 2006, 5, 1–15. 67. Gabizon, A. A.; Shmeeda, H.; Zalipsky, S. Pros and cons of the liposome platform in cancer drug targeting. J. Liposome Res. 2006, 16, 175–183. 68. Duncan, R. The dawning era of polymer therapeutics. Nat. Rev. Drug Discov. 2003, 2, 347–360. 69. Li, S. D.; Huang, L. Pharmacokinetics and biodistribution of nanoparticles. Mol. Pharm. 2008, 5, 496–504. 70. Choi, H. S.; Ipe, B. I.; Misra, P.; et al. Tissue- and organ-selective biodistribution of NIR fluorescent quantum dots. Nano Lett. 2009, 9, 2354–2359. 71. Saad, M.; Garbuzenko, O. B.; Ber, E.; et al. Receptor targeted polymers, dendrimers, liposomes: Which nanocarrier is the most efficient for tumor-specific treatment and imaging? J. Controll. Release 2008, 130, 107–114. 72. Park, J. H.; Gu, L.; von Maltzahn, G.; et al. Biodegradable luminescent porous silicon nanoparticles for in vivo applications. Nat. Mater. 2009, 8, 331–336. 73. Lu, A. H.; Salabas, E. L.; Schuth, F. Magnetic nanoparticles: synthesis, protection, functionalization, and application. Angew. Chem. Int. Ed. 2007, 46, 1222–1244. 74. Resch-Genger, U.; Grabolle, M.; Cavaliere-Jaricot, S.; Nitschke, R.; Nann, T. Quantum dots versus organic dyes as fluorescent labels. Nat. Methods 2008, 5, 763–775. 75. Shultz, M. D.; Calvin, S.; Fatouros, P. P.; Morrison, S. A.; Carpenter, E. E. Enhanced ferrite nanoparticles as MRI contrast agents. J. Magn. Magn. Mater. 2007, 311, 464–468. 76. Lam, C. W.; James, J. T.; McCluskey, R.; Arepalli, S.; Hunter, R. L. A review of carbon nanotube toxicity and assessment of potential occupational and environmental health risks. Crit. Rev. Toxicol. 2006, 36, 189–217.
P1: OTA/XYZ P2: ABC c02 JWBS037-Chen January 25, 2011
21:45
Printer Name: Yet to Come
CHAPTER 2
Synthesis of Nanomaterials as a Platform for Molecular Imaging JINHAO GAO Molecular Imaging Program at Stanford and Bio-X Program, Department of Radiology, Stanford University School of Medicine, Stanford, California, USA
JIN XIE Molecular Imaging Program at Stanford and Bio-X Program, Department of Radiology, Stanford University School of Medicine, Stanford, California, and Laboratory for Molecular Imaging and Nanomedicine, National Institute of Biomedical Imaging and Bioengineering, National Institutes of Health, Bethesda, Maryland, USA
BING XU Department of Chemistry, Brandeis University, Waltham, Massachusetts, USA
XIAOYUAN CHEN Molecular Imaging Program at Stanford and Bio-X Program, Department of Radiology, Stanford University School of Medicine, Stanford, California, and Laboratory for Molecular Imaging and Nanomedicine, National Institute of Biomedical Imaging and Bioengineering, National Institutes of Health, Bethesda, Maryland, USA
Recent successful syntheses of narrowly dispersed nanomaterials have offered a unique opportunity to probe biological interactions and visualize biological structures. This chapter highlights the general strategies to produce novel nanomaterials for molecular imaging, such as quantum dots as fluorescent markers, metal oxide nanoparticles as magnetic resonance imaging (MRI) contrast enhancement agents, metallic nanoparticles as optical contrast agents, and multifunctional nanomaterials for multimodality imaging. Well-established and thoroughly characterized nanomaterials as a general platform may offer many prominent biomedical applications with high sensitivity and high affinity via molecular imaging. 2.1 INTRODUCTION Nanotechnology refers to the control, manipulation, and study of structures and devices with length scales in the range of 1–100 nanometers, which usually falls at atomic, molecular, and macromolecular scales. Recently, the successful syntheses of narrowly dispersed Nanoplatform-Based Molecular Imaging Edited by Xiaoyuan Chen C 2011 John Wiley & Sons, Inc. Copyright
25
P1: OTA/XYZ P2: ABC c02 JWBS037-Chen January 25, 2011
26
21:45
Printer Name: Yet to Come
SYNTHESIS OF NANOMATERIALS AS A PLATFORM FOR MOLECULAR IMAGING
nanomaterials have offered a unique opportunity to probe biological interactions and visualize biological structures. A new emerging field that combines nanotechnology and biotechnology, so called nanobiotechnology, is expanding new frontiers of knowledge [1, 2]. One of the most attractive topics in the nanobiotechnology field is molecular imaging based on the nanomaterials, that is, using nanomaterials as a platform to perform molecular imaging [3]. Molecular imaging has the ability to detect the biological and biochemical processes in living subjects in a noninvasive manner, which would assist many applications in biomedicine, such as drug development, diagnosis, and therapy [4, 5]. The properties and functions conferred by nanomaterials, for example, their high volume/surface ratio, surface tailorability, improved solubility, and multifunctionality, are expanding many new possibilities for molecular imaging. The intrinsic optical, magnetic, and biological properties of the nanomaterials offer remarkable opportunities to detect and study the complex biological processes for biomedical applications in an unprecedented manner. Multifunctionality is the key advantage of nanoplatform over traditional approaches. One can integrate targeting ligands, therapeutic drugs, imaging labels, and other agents into the nanoplatform to allow targeted molecular imaging and molecular therapy. One of the most advanced and exciting forefronts of nanobiotechnology is the use of quantum dots (QDs) as fluorescent dyes in biology and medicine [6]. In the past two decades, the advances in the synthesis of QDs significantly promoted fluorescent imaging for in vitro and in vivo applications. Another rapidly evolving nanoplatform-based molecular imaging is magnetic resonance imaging (MRI), that is, using nanomaterials (e.g., iron oxide nanoparticles, MnO nanoparticles) or the nanomaterials functionalized by paramagnetic entities (e.g., Gd3+ ) as MRI contrast agents. Many research groups have intensively pursued the development of controllable uniform metal oxide nanoparticles as MRI contrast enhancement agents. Indeed, chemists developed many synthetic strategies and provided assays of nanomaterials for the potential applications in molecular imaging. Based on these well-established nanomaterials, biologists and doctors can extend the potential biomedical applications of the nanomaterials and further improve them for clinical applications. After all, the interdisciplinary collaboration in the fields of material science, chemistry, biology, and clinical medicine is very important for the development of the nanoplatform-based molecular imaging. In this chapter, we give an overview on the synthetic strategies of nanomaterials used for molecular imaging in living subjects. First, we summarize the synthesis of QDs and metal oxide nanoparticles. We then give a brief introduction to the synthesis of metallic nanomaterials (e.g., Ag, Au). Finally, we discuss the design and synthesis of multifunctional nanomaterials for multimodality molecular imaging.
2.2 SYNTHESIS OF QUANTUM DOTS AS FLUORESCENT DYES QDs find many applications as fluorescent labels in biology because the mere change of their diameters significantly shifts their emission maxima. As now well recognized, besides their narrow, tunable, and symmetric emission spectra, QDs have much greater temporal stability and resistance to photobleaching than fluorescent dyes do. Over the past three decades, the syntheses of QDs with different materials have been well developed. In the early 1980s, researchers started to focus on the synthesis of semiconductor nanostructures because they realized the photophysical properties of semiconductor structures were important for a broad range of computer and electronic applications. Henglein [7] and Rossetti et al [8].
P1: OTA/XYZ P2: ABC c02 JWBS037-Chen January 25, 2011
21:45
Printer Name: Yet to Come
SYNTHESIS OF QUANTUM DOTS AS FLUORESCENT DYES
27
reported the first method for synthesizing colloidal QDs. They mixed the cadmium and sulfide salts in an aqueous buffer to produce CdS QDs. Researchers in the late 1980s and early 1990s attempted to improve the overall optical qualities of these QDs (e.g., improving particle monodispersity and size tunability) by investigating the effect of different reaction conditions (e.g., solvents, salts, precursor ratios, pH, and temperature) on their morphological and optical properties. These early fundamental research findings provided a guide for the current state-of-the-art in QD synthesis, an organometallic synthetic procedure developed by Murray et al [9]. in 1993 for making monodispersed QDs, such as CdSe nanocrystals. For the typical synthesis of CdSe QDs by an organometallic procedure, dimethylcadmium (Cd(CH3 )2 ) and selenium were initially dissolved in the organic solvent trin-octylphosphine (TOP) at defined ratios and injected in a hot coordinating solvent of tri-n-octylphosphine oxide (TOPO; 350 ◦ C) under argon gas protection. Nucleation and growth of the nanocrystals of CdSe happened after injection and lowering of the temperature from 350 to 300 ◦ C, indicated by a change in the color of the solution (from clear to light yellow to orange to red). According to this method, CdSe QDs that have different shapes can be produced by tuning the reaction conditions [10, 11]. However, Cd(CH3 )2 precursor is extremely toxic, pyrophoric, expensive, unstable at room temperature, and explosive at elevated temperatures by releasing a large amount of gas. In 2001, Peng and co-workers reported the synthesis of CdS, CdSe, and CdTe QDs using CdO as precursor [12, 13]. Compared with Cd(CH3 )2 , CdO is much safer and cheaper. This “greener” method quickly gained in popularity. The emission wavelength of Cd-based QDs can only reach about 700 nm in terms of their bandgap energy. Alivisatos and co-workers reported the synthesis of InAs nanocrystals with the emission wavelength from 700 to 1300 nm by tuning the size of the nanoparticles from 3 to 7 nm [14]. As shown in Figure 2.1, the emission wavelength of the colloidal QDs made of ZnS, CdSe, CdTe, PbS, PbSe, and InAs covers the UV to infrared range. Because of their wide absorption spectra, QDs with different emission wavelengths excited by a single light source can emit various color fluorescence; therefore QDs have a great potential for multicolor molecular imaging. However, these synthetic processes may need refinement for biological applications as issues of reproducible synthesis and inorganic passivation remain. The breakthrough in synthesizing high-quality colloidal QDs came with the reports of CdSe/CdS core–shell QDs with high quality and strong luminescence [15, 16]. Because of the toxicity and instability of the CdSe and InAs QDs surface, it is extremely important
FIGURE 2.1 Representative QDs with different materials scaled as a function of their emission wavelength superimposed over the spectrum.
P1: OTA/XYZ P2: ABC c02 JWBS037-Chen January 25, 2011
28
21:45
Printer Name: Yet to Come
SYNTHESIS OF NANOMATERIALS AS A PLATFORM FOR MOLECULAR IMAGING
to passivate or cap the CdSe and InAs QDs with a layer of ZnS or ZnSe. The ZnS or ZnSe layer improves the fluorescence quantum yield of the QDs and protects them against photo-oxidation (which is important for minimizing cytotoxicity and for enhancing photostability). The ZnS shell has a larger bandgap energy than CdSe, eliminating the core’s surface defect states. Also, the ZnS shell has a similar bond length to that of CdSe, minimizing crystal–lattice strain and allowing for epitaxial growth. The strategy of using a ZnS or ZnSe shell to cap QDs has become a popular approach. In order to meet the requirements of in vivo biological imaging applications, the fluorescent emission wavelength of the QDs ideally should be in a region of the spectrum where blood and tissue absorb minimally but still be detectable by the instruments. Thus the QDs should emit at approximately 700–900 nm in the NIR region to minimize the problems of indigenous fluorescence of tissues. Because of their interest and high demand, the development of NIR QDs has progressed rapidly. Recently, Bawendi’s group developed several types of QDs with near-infrared emission, such as CdTe@CdSe [17], InAs@InP@ZnSe [18], and InAs@ZnSe [19] QDs with emission wavelengths between 700 and 850 nm. Peng and co-workers also reported several types of NIR QDs based on an InAs core [20, 21]. Due to the potential cytotoxicity of cadmium- and arsenic-containing QDs, it is imperative to replace the toxic metals by other more benign elements. CuInSe QDs, promising candidates, have recently been reported by Bawendi’s group [22]. The high-temperature organic phase synthetic strategy produces QDs that are nonpolar and insoluble in aqueous solvents. In order to render the QDs water soluble, surface modification of the nanoparticles is necessary, such as chemical exchange and hydrophobic–hydrophobic interaction, which sometimes needs complicated manipulations and yet with low quantum yield. The water-dispersed QDs can be synthesized directly via an aqueous method, which is simpler, cheaper, and less toxic, for example, thiol-capped CdTe QDs [23, 24]. Recently, Wang and co-workers reported the rapid synthesis of waterdispersed CdTe, CdTe@CdS core–shell, and CdTe@CdS@ZnS core–shell–shell QDs in aqueous phase assisted by microwave irradiation [25–27]. This method may open up a new route for the synthesis of water-dispersed QDs.
2.3 SYNTHESIS OF METAL OXIDE NANOPARTICLES AS MRI CONTRAST AGENTS Metal oxide nanoparticles, especially iron oxide nanoparticles, have shown many potential applications in biology and medicine. The most prominent biomedical application of iron oxide nanoparticles is their usage as MRI contrast enhancement agents, which have already received some clinical acceptance [28]. Among many types of iron oxides, superparamagnetic iron oxide (SPIO) nanoparticles have become the front runner for gaining regulatory approval. Recent advances in the synthesis of various metal oxide nanoparticles by using colloidal chemical approaches offer the capability to precisely control the size of iron oxide nanoparticles from 2 to 20 nm in diameter [29–31]. As shown in Table 2.1, besides iron oxide nanoparticles, there are several other types of metal oxide nanoparticles that have been developed as MRI contrast agents, such as manganese oxide [32] and gadolinium oxide (Gd2 O3 ) [33]. Sun and co-workers developed the synthesis of monodisperse magnetite (Fe3 O4 ) nanoparticles by a high-temperature solution phase reaction of Fe(acac)3 and 1,2hexadecanediol in the presence of oleic acid and oleylamine as surfactants [34, 35]. Particle
P1: OTA/XYZ P2: ABC c02 JWBS037-Chen January 25, 2011
21:45
Printer Name: Yet to Come
SYNTHESIS OF METAL OXIDE NANOPARTICLES AS MRI CONTRAST AGENTS
29
TABLE 2.1 Several Types of Metal Oxide Nanoparticles as Potential MRI Contrast Enhancement Agents Metal Oxide Magnetite (Fe3 O4 ) ␥ -Fe2 O3 Manganese oxide Gadolinium oxide (Gd2 O3 )
Methods
Reference
Thermal decomposition Decomposition and oxidation Thermal decomposition Thermal decomposition, polyol synthesis
34, 35 36–38 32, 38 33, 39–41
diameter can be tuned from 3 to 20 nm by varying reaction conditions or by seed-mediated growth. For example, 4-nm Fe3 O4 nanoparticle seeds were synthesized by refluxing a reaction mixture composed of Fe(acac)3 , diphenyl ether, 1,2-hexadecanediol, oleic acid, and oleylamine. By controlling the quantity of the nanoparticle seeds and reaction conditions, they generated various larger sized Fe3 O4 nanoparticles. Figure 2.2 shows the TEM images of different sizes of Fe3 O4 nanoparticles synthesized by the seed-mediated growth. Magnetic measurements indicate that the nanoparticles are superparamagnetic at room temperature (no hysteresis); that is, the thermal energy can overcome the anisotropy energy barrier of a single particle to result in zero net magnetization of the nanoparticle assemblies in the absence of an external field. The mixture of metal (Co, Mn, and Ni) precursors in the reaction resulted in metal-doped iron oxide nanoparticles, which also show good MRI contrast enhancement [35], especially for Mn-doped iron oxide nanoparticles [42]. However, elements other than iron may cause cytotoxicity, thus limiting their in vivo applications. In the past ten years, Hyeon’s group has contributed several excellent strategies on the synthesis of monodisperse iron oxide nanoparticles. In 2001, they reported on the synthesis of monodisperse and highly crystalline maghemite (␥ -Fe2 O3 ) nanoparticles [36]. Monodisperse iron nanoparticles were first synthesized and then were transformed to monodisperse ␥ -Fe2 O3 nanocrystals by controlled oxidation using trimethylamine N-oxide ((CH3 )3 NO) as a mild oxidant. Particle size could be varied from 4 to 20 nm by altering the experimental conditions. The dominating size-controlling factor was the molar ratio of Fe(CO)5 to oleic acid. Figure 2.3 shows the TEM images of ␥ -Fe2 O3 nanoparticles with the sizes of 7, 11, and 13 nm in diameter. A subsequent study by Hyeon and co-workers [37] implied that the oxidation state of iron in these particles also depended on the sizes of the nanoparticles. Using this method, which is similar to the seed-mediated growth, they succeeded in producing monodisperse iron oxide nanoparticles with a continuous size spectrum of 6–13 nm in diameter without size-selection processes [37]. Based on the thermal decomposition of a metal–surfactant complex, Hyeon and coworkers further developed large-scale syntheses of monodisperse iron oxide nanocrystals [38]. Instead of using toxic and expensive organometallic compounds such as Fe(CO)5 , they prepared the iron–oleate complex by reacting inexpensive and environmentally friendly compounds, namely, metal chlorides and sodium oleate. They generated monodisperse iron oxide nanocrystals by heating 1-octadecene (ODE) solution with iron–oleate complex to 320 ◦ C and keeping that temperature for 30 minutes. The amount of the separated nanocrystals produced was up to 40 g with a yield higher than 95%. The iron oxide nanocrystals can easily redisperse in various organic solvents including hexane and toluene. Moreover, the size of the nanoparticles could be controlled with reasonable precision by simply varying the experimental conditions, with a range from 5 to 22 nm.
P1: OTA/XYZ P2: ABC c02 JWBS037-Chen January 25, 2011
30
21:45
Printer Name: Yet to Come
SYNTHESIS OF NANOMATERIALS AS A PLATFORM FOR MOLECULAR IMAGING
(a)
20 nm
(b)
20 nm
(c)
20 nm
FIGURE 2.2 TEM images of Fe3 O4 nanoparticles with a diameter of (a) 6 nm, (b) 10 nm, and (c) 12 nm. (Reproduced with permission from Sun et al [35]. Copyright © 2004 American Chemical Society.)
Recently, Sun’s group reported the synthesis of iron nanoparticles with high chemical stability and good dispersion by controlled oxidation of the iron surface to crystalline Fe3 O4 shell [43]. The crystalline Fe3 O4 shell offered more robust protection to the metallic
P1: OTA/XYZ P2: ABC c02 JWBS037-Chen January 25, 2011
21:45
Printer Name: Yet to Come
SYNTHESIS OF METAL OXIDE NANOPARTICLES AS MRI CONTRAST AGENTS
31
(a)
20 nm
(b)
20 nm
(c)
20 nm
FIGURE 2.3 TEM images of ␥ -Fe2 O3 nanoparticles with a diameter of (a) 7 nm, (b) 11 nm, and (c) 13 nm. (Reproduced with permission from Hyeon et al. [36]. Copyright © 2001 American Chemical Society.)
core of iron, and the core/shell-structured Fe@Fe3 O4 nanoparticles were stable in hexane or water dispersion for a certain period of time. So it is possible to produce small Fe nanoparticles (<10 nm in radius) with the desired stability for highly efficient biological
P1: OTA/XYZ P2: ABC c02 JWBS037-Chen January 25, 2011
32
21:45
Printer Name: Yet to Come
SYNTHESIS OF NANOMATERIALS AS A PLATFORM FOR MOLECULAR IMAGING
separation [44], drug delivery, and high sensitivity in biological detection [45]. Sun and coworkers generated the crystalline Fe3 O4 shells by controlled oxidation of the as-synthesized nanoparticles using the oxygen transferring agent (CH3 )3 NO. The thickness of the shell was tunable by controlling the amount of (CH3 )3 NO added into the reaction mixture. For example, they produced the 2.5-nm/5-nm Fe@Fe3 O4 core/shell nanoparticles (Fig. 2.4a) by oxidizing as-synthesized Fe nanoparticles with (CH3 )3 NO. Magnetic measurements of the 2.5-nm/5-nm Fe@Fe3 O4 nanoparticles show that they are superparamagnetic with the magnetic moment reaching 61.6 emu/g. Moreover, the gap between the core and shell shown in Figure 2.4a suggests the Kirkendall effect [46–48]. The further oxidization of Fe@Fe3 O4 nanoparticles results in the hollow Fe3 O4 nanoparticles (Fig. 2.4b) [49]. Hollow Fe3 O4 magnetic nanoparticles may exhibit advantages in MR relaxation enhancement due to their unique geometry. The interface between the iron oxide hollow nanoparticles and
(a)
10 nm
(b)
20 nm
FIGURE 2.4 TEM images of (a) Fe@Fe3 O4 core–shell nanoparticles and (b) Fe3 O4 hollow nanoparticles synthesized due to the Kirkendall effect.
P1: OTA/XYZ P2: ABC c02 JWBS037-Chen January 25, 2011
21:45
Printer Name: Yet to Come
SYNTHESIS OF METALLIC NANOMATERIALS
33
the water phase is much larger than that of solid nanoparticles for unit weight or mass, which may contribute to the stronger MR contrast enhancement per unit weight [50, 51]. The thermal decomposition method developed recently has become the most effective approach for preparing high-quality iron oxide nanoparticles. However, most of the products made by the thermal decomposition approach only disperse well in organic solvents, and postpreparative procedures and modifications are necessary to transfer the hydrophobic iron oxide nanoparticles to the water phase before further biomedical applications. It is thus worthwhile to develop a direct synthesis (i.e., one-pot synthesis) method of water-dispersible iron oxide nanoparticles. Recently, Gao and co-workers reported the one-pot synthesis of water-dispersible and biocompatible Fe3 O4 nanoparticles in strong polar 2-pyrrolidone solvent using Fe(acac)3 or hydrated ferric salts as iron precursors [52–54]. The method provides a new way to prepare water-soluble Fe3 O4 nanoparticles, although the size distribution of the resulting nanomaterials has yet to be improved. Besides iron oxide nanoparticles, there are many other metal oxide nanoparticles that can act as MRI contrast enhancement agents. Hyeon and co-workers reported that the waterdispersible and biocompatible MnO nanoparticles can serve as T1 MRI contrast agent [32]. The thermolysis of Mn–oleate complex in organic solvent with high boiling point can generate MnO nanoparticles with a very narrow size distribution (Fig. 2.5a) [38]. Cao [33] reported the solution synthesis of plate-shaped Gd2 O3 nanocrystals (Fig. 2.5b) by thermal decomposition of gadolinium acetate in the presence of oleylamine and oleic acid. The high-quality Gd2 O3 nanocrystals have potential as a T1 MRI contrast agent if the toxicity of Gd(III) could be managed.
2.4 SYNTHESIS OF METALLIC NANOMATERIALS Metallic (e.g., Ag, Au, and Pt) nanomaterials have fascinated scientists for several centuries, partly because of the amazing colorful colloidal solution of metallic nanomaterials, in which resonant electron oscillations on the surface of noble metal nanomaterials create the surface plasmon resonance (SPR) that greatly enhances the absorption and Rayleigh (Mie) scattering of light by these nanomaterials. Because of their inherent ultrahigh density, enhanced resonance absorption and scattering properties, and strong Raman scattering, the metallic nanoparticles are useful for some biomedical applications, such as optical contrast agents, multimodal sensors combining optical imaging and scattering imaging, photothermal therapy, and Raman probes [55–57]. The chemical, optical, and thermal properties of the metallic nanomaterials heavily depend on the shape of the nanomaterials [58, 59]. For example, adjusting the size and shape of the Au particles from spheres to rods can tune the SPR absorption from the visible to the NIR region. In this section, we summarize the synthesis of Ag and Au nanomaterials with different shapes from nanoparticles to nanorods and nanowires, from nanocubes to nanocages. The Ag nanoparticles can easily grow in aqueous solution containing ethylene glycol and poly(vinyl pyrrolidone) (PVP) polymer to prevent aggregation, the so-called polyol synthesis. In most solution-phase syntheses of Ag nanoparticles, the results are the mixture of single-crystalline and twinned structures. In order to produce single-crystalline nanoparticles at high yields without the cumbersome separation steps, it is necessary to eliminate the twinned nanoparticles by selective etching [60]. Because PVP interacts more strongly with <100> than <111> facets and passivates silver atoms on the <100> facets, the silver
P1: OTA/XYZ P2: ABC c02 JWBS037-Chen January 25, 2011
34
21:45
Printer Name: Yet to Come
SYNTHESIS OF NANOMATERIALS AS A PLATFORM FOR MOLECULAR IMAGING
(a)
50 nm
(b)
10 nm
FIGURE 2.5 TEM images of (a) MnO nanoparticles and (b) self-assembly of Gd2 O3 nanoplates.
nanowires along the <100> direction can be formed by reducing silver nitrate with hot ethylene glycol in the presence of PVP [61]. There are many methods to produce Au nanoparticles with uniform size and structure in either the aqueous phase or organic phase by using different surfactants. For example, started by Faraday in the 1850s, using tetrachloroauric acid (HAuCl4 ) as precursor and sodium citrate as surfactant, Au nanoparticles with different sizes (>2 nm) were formed in the aqueous phase [62]. Using HAuCl4 as precursor, sodium borohydride (NaBH4 ) as the reductant, and quaternary ammonium bromide salts (R4 N+ Br− ) as phase-transfer agents, Au nanoparticles about 8 nm in diameter can be formed in toluene by using a two-phase system [63]. In single organic phase (e.g., toluene), the large-scale synthesis of Au nanoparticles can also be easily achieved using oleylamine as reductant and surfactant [64]. Moreover, the well-established surface chemistry of Au nanomaterials with organic ligands containing the thiol group [65] generated several strategies for the synthesis of
P1: OTA/XYZ P2: ABC c02 JWBS037-Chen January 25, 2011
21:45
Printer Name: Yet to Come
DESIGN AND SYNTHESIS OF MULTIFUNCTIONAL NANOMATERIALS FOR MULTIMODALITY IMAGING
35
Au nanomaterials. In particular, thiol-derivatized Au nanomaterials can be prepared in one-step, two-phase reactions [66–68]. The stability of Au nanomaterials conferred by the attachment of thiols in self-assembled structures is remarkable; that is, the solutions show no sign of decomposition or loss of solubility even after several months of storage at room temperature [66]. One-dimensional metallic nanostructures such as nanorods and nanowires are of tremendous interest for sensing and imaging applications [69–71]. Shape anisotropy introduces new optical properties in gold and silver nanomaterials, such as longitudinal plasmon resonance bands in the visible and NIR portions of the spectrum. The synthesis of nanorods using surfactants can be achieved by electrochemical, photochemical, and seed-mediated methods [72–74]. Typically, a metal salt such as AgNO3 or HAuCl4 is reduced in the presence of a surfactant to yield one dimensional (1D) nanostructures. The aqueous metal salt solutions are reduced by the passage of an applied current, by irradiation with photons, or by the use of a chemical reductant, such as hydrazine, citrate, or ascorbic acid. The surfactant is critical to the generation of 1D nanostructures. One simple chemical reduction technique often used for the preparation of Au and Ag nanorods and nanowires is the seed-mediated method, which has been well developed by Murphy and co-workers [74–77]. First, a metal salt of gold or silver is initially reduced by a strong reducing agent such as NaBH4 to form spherical nanoparticles in water. Then, these spherical nanoparticles can be used as seeds to grow nanorods. In the presence of more metal salt, surfactants, and additional but milder reducing agents, such as ascorbic acid, seeds grow into nanorods or nanowires. Using this strategy, Au nanorods with various aspect ratios have been prepared, as shown in Figure 2.6. Among a large variety of metallic nanostructures, Ag nanocubes and Au nanocages are two impressive and interesting structures. Sun and Xia [78] reported the large-scale synthesis of Ag nanocubes with high quality (Fig. 2.7a) by reducing AgNO3 with ethylene glycol (polyol synthesis) in the presence of PVP as the surfactant. The presence of PVP and its molar ratio relative to AgNO3 both play important roles in determining the geometric shape and size of the product mostly due to the selective adsorption of PVP on the various crystallographic planes of the silver nanocrystalline seeds. Using Ag nanocubes with truncated corners as the sacrificial template, they further synthesized Au nanocages with well-controlled pores at the corners (Fig. 2.7b) by galvanic replacement reaction [79]. The resulting Au nanocages have large absorption cross sections because of the discrete dipole approximations, and they also display a large photothermal effect with absorbed photons being converted into phonons (i.e., lattice vibrations) to produce a localized temperature jump. Moreover, the Au nanocages show strong SPR absorption in the NIR region, together with potential computed tomography (CT) contrast enhancement of Au nanomaterials; therefore they are particularly attractive for biomedical applications that require the selective absorption of light at great depths [80, 81].
2.5 DESIGN AND SYNTHESIS OF MULTIFUNCTIONAL NANOMATERIALS FOR MULTIMODALITY IMAGING Nanostructures provide the platform to integrate different functional nanocomponents into one single nano-entity to exhibit multifunctional properties. To assemble different nanoparticles into a single entity offers a unique building block of materials. Based on the QDs, metal oxide nanoparticles, and metallic nanoparticles, the combination of two of them
P1: OTA/XYZ P2: ABC c02 JWBS037-Chen January 25, 2011
36
21:45
Printer Name: Yet to Come
SYNTHESIS OF NANOMATERIALS AS A PLATFORM FOR MOLECULAR IMAGING
(a)
(b)
(c)
100 nm
FIGURE 2.6 TEM images of Au nanorods with aspect ratios of (a) ∼4.6, (b) ∼13, and (c) ∼18. (Reproduced with permission from Jana et al. [75]. Copyright © 2001 American Chemical Society.)
will produce several kinds of multifunctional nanomaterials as dual-mode imaging agents. There are already several review papers that summarize and discuss the synthesis of this type of hybrid nanostructures [82–87]. Several useful applications, in the study of subcellular processes of fundamental importance in biology, have highlighted the potential of QDs in nanobiotechnology [88–90]. Such exciting and attractive properties of the QDs have inspired the fabrication of hybrid nanostructures based on QDs, such as Co@CdSe core–shell nanocomposites [91], Fe2 O3 –CdS heterojunctions [92], FePt–CdX (X = S or Se) heterodimer or core–shell nanoparticles [93, 94], and CdSe–Au heteronanorods [95].
P1: OTA/XYZ P2: ABC c02 JWBS037-Chen January 25, 2011
21:45
Printer Name: Yet to Come
DESIGN AND SYNTHESIS OF MULTIFUNCTIONAL NANOMATERIALS FOR MULTIMODALITY IMAGING
37
(a)
(b)
200 nm
FIGURE 2.7 SEM images of (a) Ag nanocubes and (b) Au nanocages (insets, TEM images). (Reproduced with permission from Chen et al. [79]. Copyright © 2006 American Chemical Society.)
Based on the FePt magnetic nanoparticles and semiconducting chalcogenide nanocomponents, the systematic studies show that the reaction condition variables control the formation of different hybrid nanostructures. In the one-pot reaction, the sequential growth of CdX (X = S or Se) onto the FePt nanoparticles under the lower reaction temperature will result in the formation of FePt@CdX core–shell nanoparticles (Fig. 2.8a,b). However, use of a higher boiling point solvent will result in FePt-CdX heterodimer nanoparticles (Fig. 2.8c,d). The formation of heterodimer nanoparticles at the higher temperature is probably due to the difference in phase transition temperatures between the FePt and CdX components. The CdX components may melt at the higher temperature and induce dewetting from the FePt cores, resulting in the formation of heterodimeric nanostructures [93]. The synthesis of these core–shell nanoparticles and heterodimer nanoparticles is highly reproducible and general. Although the optical properties of these nanostructures must be improved for higher quantum yields, this rather simple approach indeed opens an avenue to design and synthesis of sophisticated and multifunctional nanostructures. The combination of superparamagnetism and fluorescence at the nanometer scale should help expand the biological applications of multifunctional nanomaterials. Although the fluorescence of QDs can be partially quenched by metallic nanoparticles in FePt–CdX hybrid nanostructures, the replacement of the metallic nanoparticle part with a metal oxide nanoparticle has resulted in a higher quantum yield. For example, the Fe3 O4 –CdSe heterodimer nanoparticles show an emission wavelength peak at 610 nm with quantum yield of about 38% [96]. The resulting
P1: OTA/XYZ P2: ABC c02 JWBS037-Chen January 25, 2011
38
21:45
Printer Name: Yet to Come
SYNTHESIS OF NANOMATERIALS AS A PLATFORM FOR MOLECULAR IMAGING
Core-Shell
Heterodimer (c)
(a)
5 nm
20 nm (b)
5 nm (d)
5 nm
20 nm
5 nm
FIGURE 2.8 The formation of core–shell or heterodimer nanostructures by FePt and CdX (X = S or Se) nanocomponents. The core–shell structures of (a) FePt@CdS nanoparticles and (b) FePt@CdSe nanoparticles (insets, HRTEM images). The heterodimer structures of (c) FePt@CdS nanoparticles and (d) FePt@CdSe nanoparticles.
fluorescent magnetic nanoparticles bear two attractive features, superparamagnetism and fluorescence, which allow their intracellular movements to be controlled using magnetic force and to be monitored using a fluorescent microscope. Using these fluorescent magnetic nanoparticles, it is straightforward to assess the intracellular manipulation of nanoparticles by a magnetic force. It also promises a good model for dual-functional molecular imaging (i.e., combining MRI and fluorescence imaging) [97]. Based on CdSe dots and rods, Banin and co-workers reported on the growth process of the formation of symmetric or asymmetric heterostructures [95, 98]. They found that one-sided growth of a gold tip on the CdSe dots or rods is preceded by the two-sided growth. As shown in Figure 2.9, at lower gold precursor (AuCl3 ) concentrations, two-sided growth is evident in the simulation (Fig. 2.9a) and was confirmed by experimental result (Fig. 2.9b). As the gold concentration is increased, a marked change happens, instead of two-sided growth, the one-sided growth process becomes dominant and results in heterodimer nanostructures (Fig. 2.9c,d). Experimental analysis and theoretical modeling show that a ripening process drives gold from one end to the other. Ripening is therefore occurring effectively on the nanostructure itself, leading to a phase-segregated structure and thus extending the realm of ripening phenomena and their importance in nanostructures. The combination of metallic nanoparticles and magnetic nanoparticles certainly promises many applications in nanoplatform multimodality molecular imaging. Recently,
P1: OTA/XYZ P2: ABC c02 JWBS037-Chen January 25, 2011
21:45
Printer Name: Yet to Come
DESIGN AND SYNTHESIS OF MULTIFUNCTIONAL NANOMATERIALS FOR MULTIMODALITY IMAGING
39
(c)
(a)
(b)
(d)
100 nm
100 nm
FIGURE 2.9 The formation of CdSe–Au heterostructures at (a, b) low gold concentrations and (c, d) high gold concentrations.
there have been several strategies developed for the design and synthesis of hybrid nanostructures containing magnetic nanoparticles and metallic nanoparticles. The simple and efficient method to form magnetic nanoparticle-based heterodimer structures is the sequential growth of metallic components (e.g., Ag or Au) onto the “colloidosome,” that is, the self-assembly of nanoparticles at a liquid–liquid interface [99]. As shown in Figure 2.10a, the metallic nucleation reaction takes place on the exposed surface of the magnetic nanoparticles and produces heterodimers of two distinct nanospheres. TEM and HRTEM images (Fig. 2.10b,c) shows the Fe3 O4 –Ag heterodimer nanoparticles with uniform size and structure, indicating that the method of liquid–liquid interface heterogeneous growth is a general way to make heterodimer nanoparticles, such as FePt–Ag and Au–Ag heterodimer nanoparticles. Recently, Sun and co-workers reported another way to fabricate Fe3 O4 –Au heterodimer nanoparticles with controllable size of each part of the component [100]. As shown in Figure 2.10d, in the homogeneous organic solvent (e.g., 1-octadecene), the thermal decomposition of Fe(CO)5 on the surface of the Au nanoparticles and the following oxidation of the intermediate resulted in the uniform Fe3 O4 –Au heterodimer nanoparticles (Fig. 2.10e,f). Based on the optical properties of metallic nanoparticles and magnetic nanoparticles as potential MRI contrast agents, these types of heterodimer nanoparticles show great potential in multimodal biomedical applications, especially for molecular imaging [101–103]. Recently, Xu and co-workers developed FePt@CoS2 yolk–shell and FePt@Fe2 O3 yolk–shell nanoparticles as a potential nanodevice in controlled drug release [104]. The unique structure and properties of FePt@CoS2 yolk–shell nanoparticles give a good example for the investigation of a novel drug delivery system that uses the magnetic nanomaterials to directly encapsulate the potential anticancer drug. The premise is that
P1: OTA/XYZ P2: ABC c02 JWBS037-Chen January 25, 2011
40
21:45
Printer Name: Yet to Come
SYNTHESIS OF NANOMATERIALS AS A PLATFORM FOR MOLECULAR IMAGING
(a)
(b)
Ag+
(c)
Ag
organic droplet
Fe3O4 10 nm
aqueous phase (d)
(e)
Au Fe
Ag
1 nm (f)
Fe(CO)5
Fe3O4(111) 0.485 nm
organic phase oxidation
Au(111) 0.24 nm
Au Fe3O4
24 nm
2 nm
FIGURE 2.10 The two approaches for the synthesis of Fe3 O4 –M (M = Ag or Au) heterodimer nanoparticles. (a) The formation of Fe3 O4 –Ag heterodimer nanoparticles at a liquid–liquid interface. (b) TEM and (c) HRTEM images of Fe3 O4 –Ag heterodimer nanoparticles. (d) The formation of Fe3 O4 –Au heterodimer nanoparticles by the decomposition of Fe(CO)5 on the surface of the Au nanoparticles followed by oxidation in organic solvent. (e) TEM and (f) HRTEM images of Fe3 O4 –Au heterodimer nanoparticles.
platinum-containing nanoparticles (e.g., FePt) without any surface coating may be used as a potential anticancer drug. The sequential growth of CoS2 porous nanoshells by the Kirkendall effect [46–48] provides the feasibility to produce the “naked” FePt nanoyolks and the special interface between the outside and inside of the shells (Fig. 2.11a). The bifunctional FePt@Fe2 O3 yolk–shell nanoparticles exhibit high cytotoxicity and strong MR contrast enhancement [50]. Recent development on the synthesis of iron oxide hollow nanoparticles [49, 105] facilitates the design of FePt@Fe2 O3 yolk–shell nanoparticles. The direct injection of Fe(CO)5 into the refluxing solution of 1-octadecene containing oleylamine and FePt nanoparticles forms FePt@Fe core–shell nanoparticles as the intermediate. The further oxidation of as-prepared FePt@Fe core–shell nanoparticles in 1-octadecene under ambient atmosphere in the presence of O2 produces FePt@Fe2 O3 yolk–shell nanoparticles with uniform size and structure (Fig. 2.11b). Given the capability of surface functionalization of these yolk–shell nanoparticles using disease-specific molecules (e.g., antibodies), one could develop the multifunctional nanoparticles that target a specific tissue for delivering the therapeutic agents to tumors and monitoring the transformation of the tumor by MRI. Based on the multimodal features, multifunctional nanomaterials provide many new tools and opportunities for the exploration of life science because of their integrated functions. For example, based on the magnetic nanoparticles as MRI contrast agents, one of
P1: OTA/XYZ P2: ABC c02 JWBS037-Chen January 25, 2011
21:45
Printer Name: Yet to Come
SUMMARY AND PERSPECTIVE
41
(a)
5 nm
(b)
5 nm
FIGURE 2.11 TEM images of (a) FePt@CoS2 yolk–shell nanoparticles and (b) FePt@Fe2 O3 yolk–shell nanoparticles.
the most exciting applications is the multimodal molecular imaging that integrates optical imaging (e.g., dyes, QDs, and Au nanoparticles) or positron emission tomography (PET) imaging (e.g., isotopes) with MRI [106–110]. Multimodality molecular imaging will likely become a dominant imaging tool for future biomedical research because it has the potential to provide more accurate information that is critical for clinical diagnostics and therapeutics [111–113].
2.6 SUMMARY AND PERSPECTIVE In this chapter, we outlined the syntheses of the nanomaterials that can serve as a platform for molecular imaging. Based on QDs, magnetic nanoparticles, and metallic nanomaterials, multifunctional nanomaterials combine at least two of these components for multimodality imaging. There are several synthetic strategies for different kinds of nanomaterials depending on different purposes. Most of the synthetic strategies are organic solvent-based reactions, which result in nonpolar soluble nanomaterials, so further surface modification
P1: OTA/XYZ P2: ABC c02 JWBS037-Chen January 25, 2011
42
21:45
Printer Name: Yet to Come
SYNTHESIS OF NANOMATERIALS AS A PLATFORM FOR MOLECULAR IMAGING
and functionalization are needed for biological molecular imaging applications. The purity, dispersibility, and stability of the nanomaterials in the physiological environment are highly important. Recent successful syntheses of narrowly dispersed nanomaterials have offered abundant opportunities for molecular imaging applications as nanoprobes. The designs and syntheses of sophisticated and multifunctional nanomaterials open up a new avenue for multimodality molecular imaging. Future nanomedicines mostly lie in the multifunctional nanoplatformbased materials, which combine therapeutic components and nanoprobes. Therefore it is necessary and important to further study and explore multifunctional nanomaterials for creating successful nanobiotechnology to meet increasing demands in therapeutic and diagnostic applications.
ACKNOWLEDGMENT We gratefully acknowledge Stanford University and Brandeis University for financial support.
REFERENCES 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. 14. 15. 16. 17.
18.
Whitesides, G. M. Nat. Biotechnol. 2003, 21, 1161. Yang, H.; Xia, Y. N. Adv. Mater. 2007, 19, 3085. Cai, W. B.; Chen, X. Y. Small 2007, 3, 1840. Weissleder, R. Science 2006, 312, 1168. Willmann, J. K.; van Bruggen, N.; Dinkelborg, L. M.; Gambhir, S. S. Nat. Rev. Drug Discov. 2008, 7, 591. Bruchez, M.; Moronne, M.; Gin, P.; Weiss, S.; Alivisatos, A. P. Science 1998, 281, 2013. Henglein, A. J. Phys. Chem. 1982, 86, 2291. Rossetti, R.; Nakahara, S.; Brus, L. E. J. Chem. Phys. 1983, 79, 1086. Murray, C. B.; Norris, D. J.; Bawendi, M. G. J. Am. Chem. Soc. 1993, 115, 8706. Peng, X. G.; Manna, L.; Yang, W. D.; Wickham, J.; Scher, E.; Kadavanich, A.; Alivisatos, A. P. Nature 2000, 404, 59. Peng, Z. A.; Peng, X. G. J. Am. Chem. Soc. 2002, 124, 3343. Peng, Z. A.; Peng, X. G. J. Am. Chem. Soc. 2001, 123, 183. Qu, L. H.; Peng, X. G. J. Am. Chem. Soc. 2002, 124, 2049. Guzelian, A. A.; Banin, U.; Kadavanich, A. V.; Peng, X. G.; Alivisatosa, A. P. Appl. Phys. Lett. 1996, 69, 1432. Dabbousi, B. O.; Rodriguez-Viejo, J.; Mikulec, F. V.; Heine, J. R.; Mattoussi, H.; Ober, R.; Jensen, K. F.; Bawendi, M. G. J. Phys. Chem. 1997, 101, 9463. Peng, X. G.; Schlamp, M. C.; Kadavanich, A. V.; Alivisatos, A. P. J. Am. Chem. Soc. 1997, 119, 7019. Kim, S.; Lim, Y. T.; Soltesz, E. G.; De Grand, A. M.; Lee, J.; Nakayama, A.; Parker, J. A.; Mihaljevic, T.; Laurence, R. G.; Dor, D. M.; Cohn, L. H.; Bawendi, M. G.; Frangioni, J. V. Nat. Biotechnol. 2004, 22, 93. Kim, S. W.; Zimmer, J. P.; Ohnishi, S.; Tracy, J. B.; Frangioni, J. V.; Bawendi, M. G. J. Am. Chem. Soc. 2005, 127, 10526.
P1: OTA/XYZ P2: ABC c02 JWBS037-Chen January 25, 2011
21:45
Printer Name: Yet to Come
REFERENCES
43
19. Zimmer, J. P.; Kim, S. W.; Ohnishi, S.; Tanaka, E.; Frangioni, J. V.; Bawendi, M. G. J. Am. Chem. Soc. 2006, 128, 2526. 20. Xie, R. G.; Peng, X. G. Angew. Chem., Int. Ed. 2008, 120, 7791. 21. Xie, R. G.; Battaglia, D.; Peng, X. G. J. Am. Chem. Soc. 2007, 129, 15432. 22. Allen, P. M.; Bawendi, M. G. J. Am. Chem. Soc. 2008, 130, 9240. 23. Gaponik, N.; Talapin, D. V.; Rogach, A. L.; Hoppe, K.; Shevchenko, E. V.; Kornowski, A.; Eychmuller, A.; Weller, H. J. Phys. Chem. B 2002, 106, 7177. 24. Li, L.; Qian, H. F.; Fang, N. H.; Ren, H. C. J. Lumines. 2006, 116, 59. 25. He, Y.; Lu, H. T.; Sai, L. M.; Lai, W. Y.; Fan, Q. L.; Wang, L. H.; Huang, W. J. Phys. Chem. B 2006, 110, 13370. 26. He, Y.; Sai, L. M.; Lu, H. T.; Hu, M.; Lai, W. Y.; Fan, Q. L.; Wang, L. H.; Huang, W. Chem. Mater. 2007, 19, 359. 27. He, Y.; Lu, H. T.; Sai, L. M.; Su, Y. Y.; Hu, M.; Fan, C. H.; Huang, W.; Wang, L. H. Adv. Mater. 2008, 20, 3416. 28. Harisinghani, M. G.; Barentsz, J.; Hahn, P. F.; Deserno, W. M.; Tabatabaei, S.; van de Kaa, C. H.; de la Rosette, J.; Weissleder, R. N. Engl. J. Med. 2003, 348, 2491. 29. Hyeon, T. Chem. Commun. 2003, 927. 30. Park, J.; Joo, J.; Kwon, S. G.; Jang, Y.; Hyeon, T. Angew. Chem., Int. Ed. 2007, 46, 4630. 31. Kwon, S. G.; Hyeon, T. Acc. Chem. Res. 2008, 41, 1696. 32. Na, H. B.; Lee, J. H.; An, K. J.; Park, Y. I.; Park, M.; Lee, I. S.; Nam, D. H.; Kim, S. T.; Kim, S. H.; Kim, S. W.; Lim, K. H.; Kim, K. S.; Kim, S. O.; Hyeon, T. Angew. Chem., Int. Ed. 2007, 46, 5397. 33. Cao, Y. C. J. Am. Chem. Soc. 2004, 126, 7456. 34. Sun, S. H.; Zeng, H. J. Am. Chem. Soc. 2002, 124, 8204. 35. Sun, S. H.; Zeng, H.; Robinson, D. B.; Raoux, S.; Rice, P. M.; Wang, S. X.; Li, G. X. J. Am. Chem. Soc. 2004, 126, 273. 36. Hyeon, T.; Lee, S. S.; Park, J.; Chung, Y.; Bin Na, H. J. Am. Chem. Soc. 2001, 123, 12798. 37. Park, J.; Lee, E.; Hwang, N. M.; Kang, M. S.; Kim, S. C.; Hwang, Y.; Park, J. G.; Noh, H. J.; Kini, J. Y.; Park, J. H.; Hyeon, T. Angew. Chem., Int. Ed. 2005, 44, 2872. 38. Park, J.; An, K. J.; Hwang, Y. S.; Park, J. G.; Noh, H. J.; Kim, J. Y.; Park, J. H.; Hwang, N. M.; Hyeon, T. Nat. Mater. 2004, 3, 891. 39. Hifumi, H.; Yamaoka, S.; Tanimoto, A.; Citterio, D.; Suzuki, K. J. Am. Chem. Soc. 2006, 128, 15090. 40. Bridot, J. L.; Faure, A. C.; Laurent, S.; Riviere, C.; Billotey, C.; Hiba, B.; Janier, M.; Josserand, V.; Coll, J. L.; Vander Elst, L.; Muller, R.; Roux, S.; Perriat, P.; Tillement, O. J. Am. Chem. Soc. 2007, 129, 5076. 41. Fortin, M. A.; Petoral, R. M.; Soderlind, F.; Klasson, A.; Engstrom, M.; Veres, T.; Kall, P. O.; Uvdal, K. Nanotechnology 2007, 18, 9. 42. Lee, J. H.; Huh, Y. M.; Jun, Y.; Seo, J.; Jang, J.; Song, H. T.; Kim, S.; Cho, E. J.; Yoon, H. G.; Suh, J. S.; Cheon, J. Nat. Med. 2007, 13, 95. 43. Peng, S.; Wang, C.; Xie, J.; Sun, S. H. J. Am. Chem. Soc. 2006, 128, 10676. 44. Xu, C. J.; Xu, K. M.; Gu, H. W.; Zheng, R. K.; Liu, H.; Zhang, X. X.; Guo, Z. H.; Xu, B. J. Am. Chem. Soc. 2004, 126, 9938. 45. Gu, H. W.; Ho, P. L.; Tsang, K. W. T.; Wang, L.; Xu, B. J. Am. Chem. Soc. 2003, 125, 15702. 46. Yin, Y. D.; Rioux, R. M.; Erdonmez, C. K.; Hughes, S.; Somorjai, G. A.; Alivisatos, A. P. Science 2004, 304, 711. 47. Gao, J. H.; Zhang, B.; Zhang, X. X.; Xu, B. Angew. Chem., Int. Ed. 2006, 45, 1220.
P1: OTA/XYZ P2: ABC c02 JWBS037-Chen January 25, 2011
44
21:45
Printer Name: Yet to Come
SYNTHESIS OF NANOMATERIALS AS A PLATFORM FOR MOLECULAR IMAGING
48. Fan, H. J.; G¨osele, U.; Zacharias, M. Small 2007, 3, 1660. 49. Peng, S.; Sun, S. H. Angew. Chem., Int. Ed. 2007, 46, 4155. 50. Gao, J. H.; Liang, G. L.; Cheung, J. S.; Pan, Y.; Kuang, Y.; Zhao, F.; Zhang, B.; Zhang, X. X.; Wu, E. X.; Xu, B. J. Am. Chem. Soc. 2008, 130, 11828. 51. Piao, Y.; Kim, J.; Bin Na, H.; Kim, D.; Baek, J. S.; Ko, M. K.; Lee, J. H.; Shokouhimehr, M.; Hyeon, T. Nat. Mater. 2008, 7, 242. 52. Li, Z.; Chen, H.; Bao, H.; Gao, M. Chem. Mater. 2004, 16, 1391. 53. Li, Z.; Sun, Q.; Gao, M. Y. Angew. Chem., Int. Ed. 2005, 44, 123. 54. Li, Z.; Wei, L.; Gao, M. Y.; Lei, H. Adv. Mater. 2005, 17, 1001. 55. Jain, P. K.; El-Sayed, I. H.; El-Sayed, M. A. Nano Today 2007, 2, 18. 56. Jain, P. K.; Huang, X. H.; El-Sayed, I. H.; El-Sayed, M. A. Acc. Chem. Res. 2008, 41, 1578. 57. Skrabalak, S. E.; Chen, J. Y.; Sun, Y. G.; Lu, X. M.; Au, L.; Cobley, C. M.; Xia, Y. N. Acc. Chem. Res. 2008, 41, 1587. 58. El-Sayed, M. A. Acc. Chem. Res. 2001, 34, 257. 59. Xia, Y. N.; Halas, N. J. MRS Bull. 2005, 30, 338. 60. Wiley, B.; Sun, Y. G.; Chen, J. Y.; Cang, H.; Li, Z. Y.; Li, X. D.; Xia, Y. N. MRS Bull. 2005, 30, 356. 61. Sun, Y. G.; Mayers, B.; Herricks, T.; Xia, Y. N. Nano Lett. 2003, 3, 955. 62. Enustun, B. V.; Turkevich, J. J. Am. Chem. Soc. 1963, 85, 3317. 63. Fink, J.; Kiely, C. J.; Bethell, D.; Schiffrin, D. J. Chem. Mater. 1998, 10, 922. 64. Hiramatsu, H.; Osterloh, F. E. Chem. Mater. 2004, 16, 2509. 65. Love, J. C.; Estroff, L. A.; Kriebel, J. K.; Nuzzo, R. G.; Whitesides, G. M. Chem. Rev. 2005, 105, 1103. 66. Brust, M.; Walker, M.; Bethell, D.; Schiffrin, D. J.; Whyman, R. J. Chem. Soc., Chem. Commun. 1994, 801. 67. Wilcoxon, J. P.; Williamson, R. L.; Baughman, R. J. Chem. Phys. 1993, 98, 9933. 68. Leff, D. V.; Ohara, P. C.; Heath, J. R.; Gelbart, W. M. J. Phys. Chem. 1995, 99, 7036. 69. Xia, Y. N.; Yang, P. D.; Sun, Y. G.; Wu, Y. Y.; Mayers, B.; Gates, B.; Yin, Y. D.; Kim, F.; Yan, Y. Q. Adv. Mater. 2003, 15, 353. 70. Rosi, N. L.; Mirkin, C. A. Chem. Rev. 2005, 105, 1547. 71. Patolsky, F.; Timko, B. P.; Yu, G. H.; Fang, Y.; Greytak, A. B.; Zheng, G. F.; Lieber, C. M. Science 2006, 313, 1100. 72. Kim, F.; Song, J. H.; Yang, P. D. J. Am. Chem. Soc. 2002, 124, 14316. 73. Esumi, K.; Matsuhisa, K.; Torigoe, K. Langmuir 1995, 11, 3285. 74. Murphy, C. J.; Sau, T. K.; Gole, A.; Orendorff, C. J. MRS Bull. 2005, 30, 349. 75. Jana, N. R.; Gearheart, L.; Murphy, C. J. J. Phys. Chem. B 2001, 105, 4065. 76. Jana, N. R.; Gearheart, L.; Murphy, C. J. Adv. Mater. 2001, 13, 1389. 77. Nikoobakht, B.; El-Sayed, M. A. Chem. Mater. 2003, 15, 1957. 78. Sun, Y. G.; Xia, Y. N. Science 2002, 298, 2176. 79. Chen, J. Y.; McLellan, J. M.; Siekkinen, A.; Xiong, Y. J.; Li, Z. Y.; Xia, Y. N. J. Am. Chem. Soc. 2006, 128, 14776. 80. Chen, J. Y.; Wang, D. L.; Xi, J. F.; Au, L.; Siekkinen, A.; Warsen, A.; Li, Z. Y.; Zhang, H.; Xia, Y. N.; Li, X. D. Nano Lett. 2007, 7, 1318. 81. Yang, X. M.; Skrabalak, S. E.; Li, Z. Y.; Xia, Y. N.; Wang, L. H. V. Nano Lett. 2007, 7, 3798. 82. Cozzoli, P. D.; Pellegrino, T.; Manna, L. Chem. Soc. Rev. 2006, 35, 1195. 83. Salgueirino-Maceira, V.; Correa-Duarte, M. A. Adv. Mater. 2007, 19, 4131.
P1: OTA/XYZ P2: ABC c02 JWBS037-Chen January 25, 2011
21:45
Printer Name: Yet to Come
REFERENCES
84. 85. 86. 87. 88. 89. 90. 91. 92. 93. 94. 95. 96. 97. 98. 99. 100. 101. 102. 103. 104. 105. 106. 107. 108. 109. 110. 111. 112. 113.
45
Zeng, H.; Sun, S. H. Adv. Func. Mater. 2008, 18, 391. Kim, J.; Piao, Y.; Hyeon, T. Chem. Soc. Rev. 2009, 38, 372. Suh, W. H.; Suh, Y. H.; Stucky, G. D. Nano Today 2009, 4, 27. Gao, J. H.; Xu, B. Nano Today 2009, 4, 37. Michalet, X.; Pinaud, F. F.; Bentolila, L. A.; Tsay, J. M.; Doose, S.; Li, J. J.; Sundaresan, G.; Wu, A. M.; Gambhir, S. S.; Weiss, S. Science 2005, 307, 538. Smith, A. M.; Duan, H. W.; Mohs, A. M.; Nie, S. M. Adv. Drug Deliv. Rev. 2008, 60, 1226. Klostranec, J. M.; Chan, W. C. W. Adv. Mater. 2006, 18, 1953. Kim, H.; Achermann, M.; Balet, L. P.; Hollingsworth, J. A.; Klimov, V. I. J. Am. Chem. Soc. 2005, 127, 544. Kwon, K. W.; Shim, M. J. Am. Chem. Soc. 2005, 127, 10269. Gu, H. W.; Zheng, R. K.; Zhang, X. X.; Xu, B. J. Am. Chem. Soc. 2004, 126, 5664. Gao, J. H.; Zhang, B.; Gao, Y.; Pan, Y.; Zhang, X. X.; Xu, B. J. Am. Chem. Soc. 2007, 129, 11928. Mokari, T.; Sztrum, C. G.; Salant, A.; Rabani, E.; Banin, U. Nat. Mater. 2005, 4, 855. Gao, J. H.; Zhang, W.; Huang, P. B.; Zhang, B.; Zhang, X. X.; Xu, B. J. Am. Chem. Soc. 2008, 130, 3710. Park, J. H.; von Maltzahn, G.; Ruoslahti, E.; Bhatia, S. N.; Sailor, M. J. Angew. Chem., Int. Ed. 2008, 47, 7284. Mokari, T.; Rothenberg, E.; Popov, I.; Costi, R.; Banin, U. Science 2004, 304, 1787. Gu, H. W.; Yang, Z. M.; Gao, J. H.; Chang, C. K.; Xu, B. J. Am. Chem. Soc. 2005, 127, 34. Yu, H.; Chen, M.; Rice, P. M.; Wang, S. X.; White, R. L.; Sun, S. H. Nano Lett. 2005, 5, 379. Choi, J. S.; Jun, Y. W.; Yeon, S. I.; Kim, H. C.; Shin, J. S.; Cheon, J. J. Am. Chem. Soc. 2006, 128, 15982. Xu, C. J.; Xie, J.; Ho, D.; Wang, C.; Kohler, N.; Walsh, E. G.; Morgan, J. R.; Chin, Y. E.; Sun, S. H. Angew. Chem., Int. Ed. 2008, 47, 173. Jiang, J.; Gu, H. W.; Shao, H. L.; Devlin, E.; Papaefthymiou, G. C.; Ying, J. Y. Adv. Mater. 2008, 20, 4403. Gao, J. H.; Liang, G. L.; Zhang, B.; Kuang, Y.; Zhang, X. X.; Xu, B. J. Am. Chem. Soc. 2007, 129, 1428. Cabot, A.; Puntes, V. F.; Shevchenko, E.; Yin, Y.; Balcells, L.; Marcus, M. A.; Hughes, S. M.; Alivisatos, A. P. J. Am. Chem. Soc. 2007, 129, 10358. Selvan, S. T.; Patra, P. K.; Ang, C. Y.; Ying, J. Y. Angew. Chem., Int. Ed. 2007, 46, 2448. Choi, J. S.; Park, J. C.; Nah, H.; Woo, S.; Oh, J.; Kim, K. M.; Cheon, G. J.; Chang, Y.; Yoo, J.; Cheon, J. Angew. Chem., Int. Ed. 2008, 47, 6259. Lee, H. Y.; Li, Z. B.; Chen, K.; Hsu, A. R.; Xu, C. J.; Xie, J.; Sun, S. H.; Chen, X. Y. J. Nucl. Med. 2008, 49, 1371. Cai, W. B.; Chen, K.; Li, Z. B.; Gambhir, S. S.; Chen, X. Y. J. Nucl. Med. 2007, 48, 1862. Medarova, Z.; Pham, W.; Farrar, C.; Petkova, V.; Moore, A. Nat. Med. 2007, 13, 372. Cheon, J.; Lee, J. H. Acc. Chem. Res. 2008, 41, 1630. McCarthy, J. R.; Weissleder, R. Adv. Drug Deliv. Rev. 2008, 60, 1241. Cai, W. B.; Chen, X. Y. J. Nucl. Med. 2008, 49, 113S.
P1: OTA/XYZ P2: ABC c03 JWBS037-Chen August 29, 2010
13:6
Printer Name: Yet to Come
CHAPTER 3
Nanoparticle Surface Modification and Bioconjugation JIN XIE Molecular Imaging Program at Stanford and Bio-X Program, Department of Radiology, Stanford University School of Medicine, Stanford, California, and Laboratory for Molecular Imaging and Nanomedicine, National Institute of Biomedical Imaging and Bioengineering, National Institutes of Health, Bethesda, Maryland, USA
JINHAO GAO and MARK MICHALSKI Molecular Imaging Program at Stanford and Bio-X Program, Department of Radiology, Stanford University School of Medicine, Stanford, California, USA
XIAOYUAN CHEN Molecular Imaging Program at Stanford and Bio-X Program, Department of Radiology, Stanford University School of Medicine, Stanford, California, and Laboratory of Molecular Imaging and Nanomedicine, National Institute of Biomedical Imaging and Bioengineering, National Institutes of Health, Bethesda, Maryland, USA
The last decade has witnessed fast progress in nanotechnology. Especially, the marriage of nanotechnology and conventional imaging facilities has brought forth a new research realm that holds great potential in bringing revolutionary changes to the current diagnostic technologies. The interface that connects nanomaterials and functional biomolecules plays a critical role in determining the performance of nanoconjugates in living subjects, and a lot of effort has been directed to this area. This chapter attempts to review the current status of biomolecule–nanoparticle conjugate preparation and use, with an emphasis on introducing the methodologies that have been developed to achieve a physiologically stable and functioning extendable nanoparticle coating.
3.1 INTRODUCTION Nanoparticle-based contrast agents, with their unique size (several to hundreds of nanometers in diameter) and physical properties (magnetic, optical, acoustic, etc.), hold great promise for revolutionizing current radiologic methodologies by offering more accurate, specific, and detailed imaging at both the cellular and subcellular levels. Recent Nanoplatform-Based Molecular Imaging Edited by Xiaoyuan Chen C 2011 John Wiley & Sons, Inc. Copyright
47
P1: OTA/XYZ P2: ABC c03 JWBS037-Chen August 29, 2010
48
13:6
Printer Name: Yet to Come
NANOPARTICLE SURFACE MODIFICATION AND BIOCONJUGATION
breakthroughs in nanoparticle preparation have permitted researchers unparalleled control over nanoparticle size, shape, and composition, all of which have implications for this technology’s translatability to clinical practice [1–3]. Despite these advances in nanoparticle synthesis, progression of nanoparticle technology to the clinical arena remains challenging, in part because of the poor native biocompatibility of these particles. Many of these particles, as-synthesized, bear an unfavorable layer, which causes them to aggregate in the physiology environment and severely compromises their in vivo applications. Thus particle modification, through the alteration of the particle surface coating layer to achieve sufficient hydrophilicity, has been a critical focus area within nanoparticle research. Beyond providing biocompatibility, postsynthesis modifications are important for achieving the potential nanoparticles integrated with disease-targeting agents; much of the impetus behind developing the nanoparticles discussed here is their ability to effectively transduce the presence of disease in a targeted fashion, to serve as an in vivo biosensor or delivery agent. The conventional nanoparticle technology very often relies on the enhanced permeability and retention (EPR) effect, a nonspecific pathophysiologic property of neoplastic tissue to naturally retain and concentrate particles of a certain size, to passively target and enhance tumors [4, 5]. While this has been a useful approach in the past, highly sensitive, specific pathology targeting ultimately requires the addition of specific targeting biomolecules to illuminate biomarkers present in various disease states. Such targeting biomolecule–particle conjugates can be formed through various means, including electrostatic adsorption, biological interaction, or, more commonly, through covalent linkage. Forming robust conjugates can be a challenging process, largely because standard chemical linkage techniques often cannot be used due to the fragility of the candidate biomolecules, most of which can only withstand mild reaction conditions (pH 5–10, 4 ◦ C to room temperature). Whatever the functional agent, successful conjugation usually depends on effective nanoparticle surface modification, adding a degree of complexity when attempting to render a nanoparticle biocompatible. In this chapter, we describe standard practices and some recent developments in nanoparticle surface modification and conjugation. Surface modification is highly dependent on the nature of the native nanoparticle surface and is therefore highly variable among different types of particles. In contrast, bioconjugation techniques are more broadly applicable: amine-terminal molecules, for example, can be linked through forming an amide bond with any nanoparticle with a carboxyl-containing coating layer, independent of the nanoparticle’s core composition. Because of their interchangeability, we first summarize applicable conjugation techniques that are routinely utilized to link functional biomolecules with nanomaterials. Surface modification methods for the most common nanoparticle varieties are then discussed in the second section.
3.2 CONJUGATION TECHNIQUES In this section, we briefly introduce the most commonly utilized methods for conjugation. For a complete list of the conjugation methods and additional details, readers are referred to Bioconjugate Techniques written by Greg Hermanson [6] and Recent Advances of Bioconjugation Chemistry in Molecular Imaging by Xiaoyuan Chen [7]. Here we separate existing techniques into three broad categories: zero-length crosslinking, crosslinker modulated linking, and conjugation through streptavidin–biotin interaction.
P1: OTA/XYZ P2: ABC c03 JWBS037-Chen August 29, 2010
13:6
Printer Name: Yet to Come
CONJUGATION TECHNIQUES
49
3.2.1 Zero-Length Crosslinking Zero-length crosslinks are chemical bonds that are formed directly between chemical groups on particles surfaces and biomolecules, without the use of a mediating linker. Typically, some chemical reagents are added to catalyze the linking between these two entities. One commonly utilized method within this class is 1-ethyl-3-(3-dimethylaminopropyl) carbodiimide hydrochloride (EDC) mediated carboxylate-amine coupling. EDC is a watersoluble derivative of N,N -dicyclohexylcarbodiimide (DCC), which converts carboxylate into an O-acylisourea intermediate that is reactive toward amine groups and results in forming a amide bond. However, this intermediate is not stable and tends to hydrolyze in water. Therefore N-hydroxysuccinimide (NHS), or its more water-soluble derivative N-hydroxysulfosuccinimide (sulfo-NHS), is usually employed in combination with EDC, for resulting in much more stable NHS–ester intermediates while retaining coupling capacity in water (Fig. 3.1). A very common strategy is to preactivate the carboxyl with NHS and then store the reagent in a solid state or in anhydrous solution. Linking can be achieved later by simply incubating it with the desired aminated species. Another example of zerolength crosslinking is the conjugation between aldehyde and amine groups. Aldehydes can form Schiff bases with amines, although these bonds are labile and easily decompose. Elevating the environmental pH can enhance but cannot completely stabilize these bonds; however, the addition of sodium cyanoborohydride can reduce the intermediate species to a secondary amine, and result in a stable linkage (Fig. 3.1). “Click chemistry” has arrived as a recent alternative to these strategies and has been widely applied in forming biomolecule–nanoparticle couplings, due to its dependability, speed, and efficiency. The most common reaction within the click chemistry paradigm, the Azide–Alkyne Huisgen Cycloaddition, is a 1,3-dipolar cycloaddition between azide and alkyne groups, resulting in a stable 1,2,3-triazole product (Fig. 3.1). This reaction is typically catalyzed by Cu(I) or, in some cases, Cu(II) and a reducing agent (e.g., sodium ascorbate). The reaction is fast, spontaneous, and irreversible, with perfect transformation
1. N
COOH
C
N
+
– Cl NH
O 2.
O
O S
– O
N HO O sulfo-NHS
EDC
O C N O O
O S O – O
H2N
C N H O
O H NH2
NaCNBH3
N
N3 N Cu(I) or Cu(II) & sodium ascorbate
N
N
FIGURE 3.1 Examples of zero-length crosslinking in bioconjugation.
HN
P1: OTA/XYZ P2: ABC c03 JWBS037-Chen August 29, 2010
50
13:6
Printer Name: Yet to Come
NANOPARTICLE SURFACE MODIFICATION AND BIOCONJUGATION
yield. However, since most biomolecules and particle coating materials contain no azide and alkyne groups, both elements need to be preactivated, which can be a nontrivial task and has been the main disadvantage of click chemistry-based bioconjugation. Nonetheless, several reports describing click chemistry-mediated particle conjugation have been documented [8–10]. 3.2.2 Conjugation with Crosslinkers The natural chemical compositions of nanoparticles and biomolecules are not necessarily sufficient for direct conjugation; in these cases, a crosslinker intermediate can be used. Crosslinkers are composed of a spacer with two terminal chemical groups that bind complementary groups on the nanoparticle and the biomolecule to be conjugated. One example of such a crosslinker is BS3 (bis(sulfosuccinimidyl) suberate) with sulfo-NHS groups at each end, which functions to link two amine-possessing species (Fig. 3.2). When two complementary groups are different, a heterobifunctional crosslinker can be used. Figure 3.2 gives examples, where amine-containing particles and thiol-containing molecules (and vice versa) are coupled by applying a crosslinker with NHS ester and maleimide on opposing ends (such as sulfo-succinimidyl-(4-N-maleimidomethyl) cyclohexane-1-carboxylate (sulfo-SMCC) and 4-maleimidobutyric acid N-hydroxysuccinimide ester) [11, 12]. Notably, since forming conjugations with homobifunctional crosslinkers may inevitably cause crosslinking among the same species, it is very often necessary to preconvert the active group of one of the to-be-coupled moieties and thereafter use heterobifunctional crosslinkers to achieve linking. For instance, instead of using BS3 to achieve coupling between two amine-containing species, it is advisable to thiolate one component and modulate the coupling with sulfo-SMCC. 3.2.3 Streptavidin–Biotin Interaction Besides using chemical coupling methods, specific biological interactions can also be utilized to form particle–biomolecule conjugates, the most popular of which is the streptavidin–biotin interaction. The streptavidin–biotin interaction is one of the strongest known non–covalent bonds, and is stable in a wide variety of conditions. Because streptavidin has four biotin binding sites, it can work as a bridge to link two or more biotinylated species (Fig. 3.3). As compared with chemical reactions, the streptavidin–biotin interaction O O NaO S O
N
O O
O
N
O
O O
O BS3
O S O ONa H N
NH
O
O O
O O
N
H N
O S O ONa
O O
O O
O S OH O
O O
HN
O O
O
N O
N O NH
sulfo-SMCC
HS
H N O O
O N
H N O O
O
FIGURE 3.2 Examples of crosslinker-mediated bioconjugation.
O N O
S
N H
P1: OTA/XYZ P2: ABC c03 JWBS037-Chen August 29, 2010
13:6
Printer Name: Yet to Come
QUANTUM DOTS
51
Streptavidin
FIGURE 3.3 Forming conjugation via streptavidin–biotin interaction.
is highly specific and less likely to alter biomolecules in a way that might result in a functional loss. Despite the considerable differences in primary structure, avidin, which was found in egg whites, is able to bind to biotin with a similar affinity as that of streptavidin. One major reason that makes it less applied in bioconjugation, however, is the severe nonspecific binding issue of avidin. With a highly basic pI of 10 (as compared to that of 5–6 for streptavidin), avidin can interact with molecules via ionic interaction; moreover, as a glycoprotein, it has the potential concern of binding to carbohydrate receptors that are expressed on many types of cell surfaces. Addressing that, a deglycosylated version of avidin, NeutrAvidin, was later developed and employed. By removing a carbonhydrate section, not only is the lectin binding risk of avidin avoided, but the protein pI also has been decreased to a nearly neutral level (pI 6.3), making it a popular alternative to streptavidin. In this section, we briefly introduce the most common techniques for nanoparticle– biomolecule conjugation. For each nanoparticle–biomolecule pair there are usually several options available to achieve conjugation; the choice of which specific method to use is based on the expense, effectiveness, particle stability, and bioactivity loss associated with each approach. Next, we introduce the most common methods used for nanoparticle surface modification. Specific considerations for surface modification of several types of nanoparticles, including quantum dots, iron oxide nanoparticles, gold nanoparticles, carbon nanotubes, silica particles, and nanocomposites, are discussed.
3.3 QUANTUM DOTS Quantum dots (QDs) refer to a large, varied class of nanostructures made from semiconductor materials, which, at nanoscale, demonstrate unique size-dependent fluorescent emissions. Much brighter and more photostable than traditional organic dyes, QDs are
P1: OTA/XYZ P2: ABC c03 JWBS037-Chen August 29, 2010
52
13:6
Printer Name: Yet to Come
NANOPARTICLE SURFACE MODIFICATION AND BIOCONJUGATION
expected to serve as the next generation optical tags for molecular imaging. Accordingly, intensive studies have been conducted on QDs, from both the synthetic and surface modification perspectives [13]. Commonly used QDs have cores made of materials like CdSe, CdTe, PbS, and InAs, with tunable emissions spanning most of the visible and near-infrared regions. Although there are reports of aqueous phase-based QD synthesis, most of the QDs utilized for bioapplications are made through pyrolysis in organic solution, which generally produces higher quality crystals. Described briefly, this synthesis is performed by heating an appropriate organometallic precursor in a high boiling point organic solvent in the presence of certain surfactants (such as trioctylphosphine (TOP) and trioctylphosphine oxide (TOPO)), which passivate the particle surface and limit their growth. Typically, a coating step is subsequently applied, in which a layer of semiconductor material with a wider bandgap than the core, such as ZnS, is introduced to the particle surface. Such a coating is essential for increasing the particles’ photoluminescent quantum efficiencies and reducing the particle toxicity by suppressing the dissolution of cadmium ions [14–16]. As synthesized, the QDs are intrinsically water insoluble due to the hydrophobic TOPO/TOP coating. In order to convert the particles to a water-soluble form and to allow subsequent conjugation, an extra surface modification step is needed. Numerous approaches have been developed for this purpose, but despite their variability, most of them can generally be divided into two categories: ligand exchange and ligand addition. The former approach refers to the method of entirely replacing the original organic surface coating, while the latter refers to introducing a layer of new materials on top of the existing coating to render the particles hydrophilic. 3.3.1 Ligand Exchange For ligand exchange to be successful, the new hydrophilic ligand needs to have comparable or higher affinity toward the particle surface than the original surfactant, so that when incubated in high concentration, the nanoparticles’ original coating will be substituted by the new ligand. Thanks to the ZnS protection layer, thiolated molecues can easily be introduced to the QD suface, which form dative thiol bonds with the exterior S atoms. A host of small biofunctional molecules, including mercaptoacetic acid, mercaptopropionic acid, mercaptosuccinic acid, dithiothreitol, glutathione, histidine, DTT, and cysteine are capable of rendering QDs water-soluble [14] (Fig. 3.4). For example, Bawendi and co-workers used cysteine to coat CdSe/ZnS nanoparticles, resulting in water-soluble conjugates with a hydrodynamic diameter less than 5.5 nm [17]. Owing to their zwitterionic nature and ultrasmall size, such particles have proved stable in blood circulation and are inhibited to serum adsorption. When intravenously injected, those particles were found to be rapidly excreted via renal clearance, which is a size-exclusive excretion route with a size-cut around 7 nm. More sophisticated molecules, such as thiolated DNA oligos [18], peptides [19], or even proteins [20], can be anchored onto QDs in the same manner (Fig. 3.4). For example, thiolated biotinylated bovine serum albumin (bBSA), which is developed by treating bBSA with 2-iminothiolane, can be linked to QD surfaces, rendering them water-soluble. However, there are concerns about the longevity of dative thiol-based surface modifications in in vivo environments [14], considering the fragility of the disulfide bond between R-SH and ZnS. To protect against this, a crosslinking step can be used to reduce the risk of dissolution of the protective cap. Peng and co-workers anchored thiolated dendrimers to
P1: OTA/XYZ P2: ABC c03 JWBS037-Chen August 29, 2010
13:6
Printer Name: Yet to Come
QUANTUM DOTS
53
DNA oligos
Proteins S H N
S S
S
N Poly-His tagged proteins/peptides
O O HN
Zn
NH
QD
N N H
S
S
S
S
R
Polydendate ligands S S R Thiolated compounds
FIGURE 3.4 Commonly used molecues for QD surface modification.
the surface of QDs [21], producing multiple hydroxyl groups. Those hydroxyl groups were further crosslinked with second-generation dendrons, which improved particle stability while providing multiple amines for use as binding sites [22]. Alternatively, polydentate ligands, such as DHLA [23], oligomeric phosphines [24], or peptides with multiple cysteinerich domains [19], are also widedly utilized, which are expected to have higher affinity to the particle surface than monothiolated species (Fig. 3.4). Although it is most common to target S from the ZnS shell, it is also possible to utilize Zn for surface modification. For instance, proteins or peptides containing polyhistidine tag have been found to easily anchor onto QDs through Zn(II)–His interaction [25, 26] (Fig. 3.4), with a binding strength that is much stronger than most antibody–antigen interactions [16]. 3.3.2 Ligand Addition Ligand addition is mostly performed by encapsulating the original particles with a layer of amphiphilic ligand. When mixing in water, the new ligand self-assembles onto the particle surface by having its hydrophobic section intercalating into the TPOP/TOP layer, while leaving its hydrophilic tail protruding into water. Compared with those modified by ligand exchange, QDs rendered through ligand addition are generally more resistant to agglomeration that may otherwise occur during conjugation, long-term storage, or exposure to complex physiologic environments. This added stability, however, comes with a cost. The ligand addition usually leads to a significant hydrodynamic size increase, which could potentially compromise the particle’s in vivo mobility and penetrability. Nevertheless, a variety of amphiphilic compounds have been utilized for this purpose, including phospholipids [27, 28], amphiphilic saccharides [29], acrylic acid polymers [30, 31], and others. For example, Ballou [32] utilized polyacrylic acid and its PEGylated derivatives to coat QDs; the resulting adducts showed potential as sentinel lymph node imaging agents [32]. In another effort, Nie and colleagues utilized a triblock copolymer (consisting of a polybutylacrylate segment, a polyethylacrylate segment, a polymethacrylic acid segment,
P1: OTA/XYZ P2: ABC c03 JWBS037-Chen August 29, 2010
54
13:6
Printer Name: Yet to Come
NANOPARTICLE SURFACE MODIFICATION AND BIOCONJUGATION
(a)
(b)
(c)
FIGURE 3.5 (a) Schematic illustration of triblock copolymer coated QDs. (b) The chemical composition of the triblock copolymer (c) The postulated mechanism of the QD tumor retention, which is believed to involve both passive targeting (left) and active targeting (right). (Adapted from Gao et al. [33] with permission.)
and a hydrophobic hydrocarbon side chain) for QD modification (Fig. 3.5). Such polymercoated QDs were conjugated with a prostate-specific membrane antigen (PSMA) targeting antibody through EDC/NHS chemistry and were administrated into prostate cancer bearing mice [33]. Fluorescence optical imaging confirmed successful particle accumulation in the tumor area, attributed to both the EPR effect and specific antibody–antigen interaction.
P1: OTA/XYZ P2: ABC c03 JWBS037-Chen August 29, 2010
13:6
Printer Name: Yet to Come
IRON OXIDE NANOPARTICLES
55
3.3.3 Protein Encapsulation Proteins can be utilized as biomaterials to be adsorbed onto QDs to render the particle watersoluble. For instance, dihydrolipoic acid-modified QDs are highly negatively charged but are not sufficiently stable in water. In order to improve particle solubility, a layer of avidin, which is positively charged at neutral pH, can be adsorbed onto the particles directly without performing chemical conjugations. The resultant particles are highly stable in water and can easily be linked with other biomolecules through biotin–avidin interaction [34]. Another example of protein-based surface modification is the use of GroEL. GroEL is a cylinder-like chaperonin protein, which has a hydrophobic cavity with an inner diameter of about 4.5 nm. The as-synthesized QDs, being hydrophobic, can be accommodated in the GroEL cavity, resulting in a protein–QD hybrid. An interesting feature of this hybrid is that the release of QD particles can be governed by the addition of ATP, which has potential implications for biosensor applications [35]. Expected to be the next-generation fluorescent labels for bioapplications, the biocompatibility of quantum dots has been a great concern. Previously utilized QD formulas are mainly constituted of CdSe, CdTe, and PbS, which contain extremely toxic metals such as Cd and Pb, and therefore have limited prospects in clinics. Addressing such issue, recently, several non-Cd-based QDs have successfully been developed by Peng and several other groups. Especially, InAs/InP/ZnSe core/shell/shell QDs have been synthesized by a one-pot approach and were converted to be water soluble with mercaptopropionic acid (MPA) [36]. The success of commercially available options, like Qdots from Invitrogen Inc., has removed some of the impetus for developing improved coating strategies, since much research is now conducted with these particles. Nevertheless, due to the large hydrodynamic size (50 nm) and unclear surface chemical composition of such particles, individual surface modification is still preferred in many cases.
3.4 IRON OXIDE NANOPARTICLES Iron oxide nanoparticles (IONPs) are well known for their superior magnetic properties. This, together with their native biocompatibility and inexpensiveness, make IONPs important actors in molecular imaging, specifically as contrast agents for MRI [37–42]. As compared with other particles, the additional magnetic interaction among IONPs makes the stabilization and functionalization of them especially challenging. 3.4.1 Surface Modification of IONPs Synthesized from Aqueous Solution IONPs are conventionally made in aqueous solution by coprecipitating Fe(II) and Fe(III) precursors in the presence of a base, such as ammonium hydroxide and sodium hydroxide [39, 43, 44]. Due to their high surface energy and magnetic interactions, bare IONPs tend to form aggregations, so additives, in most cases hydrophilic polymers, are usually applied in the synthesis, which helps modulate particle growth and provide electrostatic and steric forces to counteract the IONP attractive forces. Conventionally applied polymers include polyvinylpyrrolidone (PVP), dendrimer, polyaniline, and dextran [39, 45, 46]. Among them, Feridex particles (AMAG Pharmaceuticals), which are essentially dextran-coated IONPs, are an FDA approved agent for use in MR imaging for the detection of lesions
P1: OTA/XYZ P2: ABC c03 JWBS037-Chen August 29, 2010
56
13:6
Printer Name: Yet to Come
NANOPARTICLE SURFACE MODIFICATION AND BIOCONJUGATION
in liver and spleen. With an overall size above 50 nm, Feridex particles are dominantly removed from circulation through the reticuloendothelial system (RES) uptake and are selectively concentrated in liver and spleen [47, 48]. Combidex particles, another AMAG product currently in clinical trials, have a similar iron oxide core but a shorter dextran for coating. With a small hydrodynamic size, Combidex particles possess a longer circulation time and aim for lymph node imaging [49]. Also, Schering AG is promoting Resovist as a liver-specific contrast agent in MRI, which basically is carbon carboxydextran-coated iron oxide nanoparticles [50, 51]. Biomolecule immobilization on the polymer-coated IONPs is achieved by making use of the functional groups on the surface polymers. For instance, dendrimer-coated IONPs, having terminal amine groups, can form conjugation with bovine serum albumin using glutaraldehyde as the homobifunctional crosslinkers [52]. In cases when more than one kind of chemical group is available, it is possible to load multiple functional entities onto the particles. One such example is polyaspartic acid (PASP)-coated nanoparticles, which have both carboxyl and amine groups available on the particle surfaces. Lee and co-workers covalently linked both a 64 Cu-DOTA chelate and the ␣v 3 integrin targeting peptide c(RGDyK) onto the IONPs. The chelate was covalently bound to the carboxyl group through EDC/NHS mediated chemistry, while the peptide, in order to avoid the self-crosslinking issue, was prethiolated and subsequently coupled with the amines from PASP-IONPs using 4-maleimidobutyric acid N-hydroxysuccinimide ester as the bridge linker [53]. Notably, this was one of the first examples of an integrated MRI/PET probe with specific targeting capacity, which is significant given the ongoing clinical efforts of building integrated MRI/PET detectors. For easy conjugation, sometimes the nanoparticle coatings need to be pretreated before linking with functional biomolecules. For instance, bioconjugation with dextran-coated IONPs is usually preceded by treatment with epichlorohydrin and ammonia, which converts the dextran hydroxyls to amines. The resulting crosslinked iron oxide (CLIO) nanoparticles have been a very effective platform and have been integrated with various kinds of biomolecules, including transferrin, annexin V, anti-VCAM monoclonal antibody (mAb), anti-E-Selectin mAb, oligonucleotides, and TAT peptides [54–58]. While surface modifications are usually introduced during particle synthesis, stabilizing agents can be grafted onto IONPs postsynthesis as well. For example, Zhang and co-workers grafted PEGylated silanes onto the IONPs to generate stable conjugates. The silane moiety can anchor onto the IONP surface by forming Si–O–Fe bonds with the surface Fe atoms, and the long PEG chains help to avoid particle aggregation. Further conjugation can be derived from the amine groups on the ends of the PEG chain (Fig. 3.6) [59–61]. Similarly, Jon and co-workers utilized a PEG-silane copolymer poly(TMSMA-r-PEGMA) to coat IONPs. Multiple silanes make the polymer adhere strongly to the particle surface, and the resultant conjugates are very stable in vivo, with long circulation half-lives [62, 63].
3.4.2 Surface Modification of IONPs Synthesized from Pyrolysis Although widely utilized, there are several drawbacks for the IONPs produced in waterbased synthesis, such as poor crystallinity, suboptimal size and shape control, and unfavorable magnetization [64, 65]. These concerns have facilitated the progress of pyrolysis-based IONP synthesis, as commonly used in the QD synthesis described previously. Fe(acac)3 , Fe(CO)5 , and, more recently, Fe-oleateare are the most utilized of the iron precursors, which
P1: OTA/XYZ P2: ABC c03 JWBS037-Chen August 29, 2010
13:6
Printer Name: Yet to Come
IRON OXIDE NANOPARTICLES
57
FIGURE 3.6 Using PEGylated silane to modify and stabilize IONPs. The resulting particles can be covalently linked with functional entities, such as methotrexate (MTX) in this case. (Adapted from Kohler et al. [60] with permission.)
decompose, nucleate, and grow to desired size in high boiling solvent with the presence of oleic acid/oleylaimne [66, 67].
Ligand Addition As is the case with TOPO/TPO coating for QD passivation, oleic acid/oleylamine forms a hydrophobic layer on the IONP surface during the pyrolysis-based synthesis. The particles can be made water soluble with some single-chained surfactants, such as dioctyl sodium sulfosuccinate, but due to the low stability, surfactants with multiple hydrophobic tails are preferred [68]. Phospholipid analogs, particularly their PEGylated derivatives, are widely used as effective phase transfer ligands. A variety of phospholipids with a wide selection of functional terminal groups are now commercially available from outside vendors, like Avantilipids Inc. Hutman and co-workers used maleimide-terminated PEG 2000 phospholipids to coat IONPs, which were subsequently conjugated with RT1 anti-MHC II antibodies. These conjugates showed targeting specificity for renal medulla of the rat, a section of the kidney in which MHC Class II is specifically expressed [69]. Larger polymers possessing multiple hydrocarbon chains have also been reported for IONP modification. In one such example, poly(maleic anhydride alt-1-tetradecene) was first coated onto IONPs with the multiple alkyl chains intercalating into the oleic acid/oleylamine coating layer (Fig. 3.7). Next, bis(6-aminohexyl) amine was added, which induced crosslinking among the polymer anhydride groups. In an aqueous environment, the unreacted anhydride
P1: OTA/XYZ P2: ABC c03 JWBS037-Chen August 29, 2010
58
13:6
Printer Name: Yet to Come
NANOPARTICLE SURFACE MODIFICATION AND BIOCONJUGATION
FIGURE 3.7 Poly(maleic anhydride alt-1-tetradecene)-based nanoparticle surface modification. After coating and crosslinking, biomolecules can be linked to the particles by forming amide bonds with the surface terminal carboxyls. (Adapted from Pellegrino et al. [70] with permission.)
groups were subjected to hydrolysis and were converted to carboxyls, ensuring the particle water solubility while creating new bonding sites for future bioconjugation [70].
Ligand Exchange While many of the previous ligand addition strategies are interchangeable between QDs and IONPs, ligand exchange approaches are generally exclusive. For most IONPs, it is the iron–carboxyl interaction that needs to be overcome by the newly added ligands. It is possible to achieve this by simply utilizing highly concentrated carboxyl-terminated compounds [71], but the efficacy of this approach is unsurprisingly low so that ligands with higher surface affinity are still highly desirable. Recent research has demonstrated that bidentate compounds, which can bind the surface Fe atom in chelating fashion, are far more effective as IONP surface-exchange ligands. One such example is meso-2,3-dimercaptosuccinic acid (DMSA), which is a small compound with two carboxyls at adjacent positions [12, 65]. By incubating it with IONPs overnight, such a ligand
P1: OTA/XYZ P2: ABC c03 JWBS037-Chen August 29, 2010
13:6
Printer Name: Yet to Come
59
IRON OXIDE NANOPARTICLES
Fe3O4
COOH
Fe3O4
COOH
O
COOH
X
O
O
OH
X
n
O
Oleylamine & oleic acid OH
OH
NH
FIGURE 3.8 Surface modification of IONPs with PEGylated dopamine. Further function docking can be done with the terminal carboxyl. (Adapted from Xie et al. [73] with permission.)
can bind to the particle surface, while leaving the thiols exposed on the outer surface. A portion of these surface thiols can form disulfide bonds among each other, which helps to stabilize the particle coating. The remaining unreacted thiols, on the other hand, can be utilized for further bioconjugation purposes. It is worth noting that such chelation is formed with two carboxylates binding to one Fe atom, which is different from the case of using dithiol ligands for QD modification, where thiols bind with S from the ZnS coating. Similarly, dopamine and its derivatives, which contain two adjacent hydroxyl groups from the catechol moiety, can also serve to form a chelate complex with the Fe on the particle surface [72]. Particle stability can then be further established by PEGylating the dopamine conjugates with PEG diacids. The resulting particle conjugates are very stable in aqueous solution and have terminal carboxyl groups that can be used for further functionalization (Fig. 3.8) [73].
3.4.3 Polymer and Protein Adsorption Multidendate ligands such as PEI have also been applied as IONP capping ligands. Containing multiple primary, secondary, and tertiary amines, PEI has strong affinity toward the IONP surface and can easily form conjugates with IONPs. The resulting adducts are highly positively charged in aqueous solution and have been exploited as cell labeling agents [74] and as transfection agents [46, 75]. Compared with PEI alone, these PEI–IONP conjugates may have enhanced cellular permeability, if used in conjunction with a vertically oriented magnetic field [76]. Being highly positively charged, the PEI coating may also work as the foundation layer to induce a layer-by-layer coating, where materials
P1: OTA/XYZ P2: ABC c03 JWBS037-Chen August 29, 2010
60
13:6
Printer Name: Yet to Come
NANOPARTICLE SURFACE MODIFICATION AND BIOCONJUGATION
FIGURE 3.9 BSA-coated manganese ferrite nanoparticles. The surface amines from BSA were modified with SPDP, followed by coupling with HS-PEG-RGD and HS-siRNA-Cy5. The resulting nanoconjugates proved to have both MR imaging and gene regulation capacities. (Adapted from Lee et al. [78] permission.)
with alternating charge natures are deposited onto the particles in succession. For instance, a copolymer, poly(ethylene oxide)-block-poly(glutamic acid) (PEO-PGA), was adsorbed onto the PEI–IONPs through the electrostatic interaction between PEI and PGA [77]. More recently, Cheon and his colleagues used bovine serum albumin (BSA) to coat manganese-doped IONPs, which yielded stable nanoconjugates. With N-succinimidyl-3(2-pyridyldithio) propionate (SPDP) as the crosslinker, they successfully coupled thiolated RGD and thiol-siRNA onto the nanoparticles, resulting in an all-in-one target-specific nanoplatform with both MR imaging and siRNA delivery capacities (Fig. 3.9) [78].
3.5 GOLD NANOPARTICLES Gold nanoparticles have the potential to become effective probes in an array of imaging modalities, such as CT, photoacoustic, and surface-enhanced Raman spectroscopy (SERS). From the synthesis perspective, modulating the shape of gold nanoparticles is well established, with those in the form of spheres [2, 79, 80], cubes [81], rods [82, 83], cages [81, 84], and wires [85] already developed. Such shape control is important for manipulating the nanostructure’s spectral absorption profile, which affects their roles as imaging probes. Common 10-nm spherical gold particles have characteristic surface plasmonic absorption at around 520 nm. Tuning the particle size may alter the spectra, but to a limited degree. For example, the maximum absorption of gold nanoparticles with diameters of 48.3 and 99.3 nm are 533 nm and 575 nm, respectively [86], both in a spectral region where light has limited tissue penetrating depth. Changing the nanoparticle configuration to a nanorod structure, on the other hand, can dramatically shift the absorption spectrum to the nearinfrared region (650–900 nm), which potentiates their role as probes in photoacoustics and mediators in photothermal nanotherapeutics (Fig. 3.10) [81, 82, 84, 87]. Due to the strong interaction between thiol and gold, the surface modification of gold nanostructures is overwhelmingly conducted with thiolated species. Typically, a bifunctional thiolated compound is utilized, such as carboxyl-ended alkanethiol with various lengths of hydrocarbon chains [79]. When applied, the thiol group will anchor onto the particle surface, and the terminal carboxyl group can be utilized for forming conjugates. Alternatively, biomolecules can be prethiolated, after which they can be directly loaded onto the gold particle surface [88–90]. Thiolated DNA oligos have consistently been utilized for
P1: OTA/XYZ P2: ABC c03 JWBS037-Chen August 29, 2010
13:6
Printer Name: Yet to Come
61
GOLD NANOPARTICLES
A
B
C
D
500
500
E
100
100
100
1.2 1
Abs.
0.8
e
0.6
b
c
d
0.4 0.2
Seed
a
a
b
c
d
e
Increasing Aspect Ratio
0 400
500
600 700 Wavelength (nm)
800
900
FIGURE 3.10 Transmission electron micrographs (top), optical spectra (bottom left), and photographs (bottom right) of gold nanorods of various aspect ratios. (Adapted from Murphy et al. [87] with permission.)
Au nanoparticle functionalization, and the resultant particle conjugates have been found to be useful in many aspects of biodiagnostics [91]. Unlike QDs, which may prefer multidendate compounds for the sake of stable binding, in most scenarios, a monothiol compound is sufficient to result in stable gold nanoconjugates. In fact, monothiol compounds can be advantageous over multidentatethiol compounds for affording better loading capacity. In one such demonstration, two oligonucleotides were loaded onto 13-nm gold particles: one was tetrathiol modified while the other was monothiol modified. The two oligos share the same sequence, which aims at inhibiting EGFR expression. It was found that, for each 13-nm Au nanoparticle, 45–50 tetrathiol strands can be loaded, whereas the number for the monothiol oligo is 110–120. In vitro experiments showed that, while both particles were able to penetrate cell membrane and suppress the EGFR expression, the monothiol group was 35 times more efficient in inducing knockdown, presumably due to an improved oligo loading efficacy [92]. Tuning the anisotropy of Au nanostructures, aside from altering the particle absorption spectrum, also changes the particle’s surface chemistry. To date, most gold nanorod synthesis is conducted in water with cetrimonium bromide (CTAB) as the capping ligand, as compared to citric acid or alkyl acid/alkyl amine, which are commonly used in spherical gold nanoparticle production. The CTAB forms a thick double-layer coating over the growing gold rods during synthesis, which constrains the particle radial growth and makes the particle sidewall hard to access. The ends of the rods are relatively less protected and can serve as the binding sites for thiolated compounds. Figure 3.11 is one good demonstration of this model: CTAB-coated gold nanorods were first modified with thioctic acid and were subsequently conjugated with anti-mouse IgG through EDC-mediated coupling. Afterwards, mouse IgG, with two binding sites, was added, which was recognized by the anti-mouse IgG and resulted in crosslinking among the nanorods. The fact that the particle aggregation was formed in a head-to-tail configuration confirms that only at the
P1: OTA/XYZ P2: ABC c03 JWBS037-Chen August 29, 2010
62
13:6
Printer Name: Yet to Come
NANOPARTICLE SURFACE MODIFICATION AND BIOCONJUGATION
FIGURE 3.11 Head-to-tail assembly of Au nanorods, which is caused by the interaction between IgG and the anti-mouse IgG that was immobilized at the ends of the Au nanorods. (Adapted from Chang et al. [94] with permission.)
ends of the nanorods did the ligand exchange and the subsequent antibody immobilization occur [93]. Besides ligand exchange, gold nanorod surface modification can also be done through a layer-by-layer deposition approach. This takes advantage of the highly positively charged CTAB coating, which can absorb one layer of negatively charged species, thereafter making it ready for another round of absorption but with positively species, and so on. For example, poly(sodium-4-styrenesulfonate) PSS(−) and poly(diallyldimethyl-ammoniumchloride) PDADMAC(+) were alternatively applied to generate a multilayer coating on the gold nanorod surface [95, 96]. One good feature of this approach is that functional molecules such as antibodies can easily be immobilized onto the outer layer via electrostatic interaction rather than through chemical conjugation [97]. Recently, a green process was developed by Katti and colleagues to prepare gold nanoparticles. In brief, without applying conventional chemicals, soybean extract was utilized in the particle preparation. Multiple components of the extract are believed to play a role in particle formation, either as reductants (sucrose, stachyose, glycinin, -conglycinin, trypsin inhibitors, soybean vacuolar protein p34) or stabilizers (raffinose, isoflavones, daidzein, glycitin, and various proteins), or both. The resulting gold nanoparticles are nontoxic and highly stable and can be coupled with other species by taking advantage of the multiple amines from the surface proteins [98]. Using a similar strategy, the same group also utilized gum arabic, a plant-derived biocompatible protein, to synthesize gold nanoparticles [99, 100].
P1: OTA/XYZ P2: ABC c03 JWBS037-Chen August 29, 2010
13:6
Printer Name: Yet to Come
CARBON NANOTUBES
63
3.6 CARBON NANOTUBES Unlike the aforementioned particles, carbon nanotubes (CNTs) are not yielded from wet chemistry routes; therefore no ligands passivate the surface. However, the missing of a surface coating does not necessarily facilitate the surface modification of CNTs. Rather, the hydrophobic nature, the propensity to form bundles, and the poor solubility even in organic solvents make the modification and bioconjugation of CNTs a nontrival one. 3.6.1 Direct Surface Modification The intact CNTs have sidewalls structurally resembling graphite, which is inert and is unfavorable for chemical modifications. In order to generate binding sites, some extreme oxidative conditions are applied to the CNTs, which rusults in defects and generates carboxyl groups on the CNT surfaces [101]. For example, after treating single wall nanotubes (SWNTs) in refluxing 2.5 M HNO3 for two 36-h periods with an interval of 30-minute sonication, the SWNTs can be made water-soluble. At pH 7, these modified CNTs showed a surface potential of −75 mV, attributable to multiple surface carboxyl groups [102]. Alternatively, cooking CNTs with succinic or glutaric acid acyl peroxide at 80–90 ◦ C leads to the immobilization of 2-carboxyethyl or 3-carboxypropyl on the nanotube surface [103]. Subsequently, conjugation chemistry can be applied to covalently link molecules such as fluorescein, biotin, sugar moieties, oligonucleotides, and proteins onto the nanotubes [102–109]. Besides strong acid treatment, another common approach of CNT modification is via 1,3-dipolar cycloaddition with azomethine ylide and its derivatives, which anchor onto the CNT surface by forming a pyrrolidine ring. Reported functional molecules that have been conjugated to CNT via this means include organic dyes, peptides, and even antibiotics [110]. The two surface modification strategies can be applied simultaneously, and by doing so, more than one type of binding site can be generated on the CNT surface, permitting the loading of multiple entities. In one such demonstration, amphotericin B, a commonly used antifungal agent, and an organic dye, FITC, were both covalently linked to the CNTs. The former was coupled to the end of the aminotriethylene glycol chain that was preanchored onto the CNTs through a 1,3-dipolar cycloaddition, while the later was coupled to the short PEG chain that was derived from the carboxyl groups generated from acid treatment (Fig. 3.12) [111]. 3.6.2 Ligand Addition Due to the hydrophobic nature of the surface, CNTs can be modified as well with amphiphilic compounds. Many types of surfactants have been documented for doing this job, including sodium dodecyl sulfate (SDS), triton-X-100, CTAB, sodium dodecylbenzene sulfonate (SDBS), and sodium dodecane sulphonic acid (SDSA) [101, 113–115]. With sonication, these surfactants can have their hydrophobic hydrocarbon chain interact with the graphite surface and achieve an effective CNT dispersion. One interesting finding is that SDBS is much more efficient at making CNT suspension than its analog SDS. It could be due to the longer hydrocarbon chain of SDBS compared to SDS. But more importantly, it is attributed to the possession of an aromatic ring in SDBS, which allows for the additional - interaction with the CNT surface that further stabilizes the nanostructure. It
P1: OTA/XYZ P2: ABC c03 JWBS037-Chen August 29, 2010
64
13:6
Printer Name: Yet to Come
NANOPARTICLE SURFACE MODIFICATION AND BIOCONJUGATION
FIGURE 3.12 Both amphotericin B and FITC were covalently immobilized onto CNTs but via different routes. (Adapted from Wu et al. [112] with permission.)
was supported by a separate study, in which trimethyl-(2-oxo-2-phenylethyl)-ammonium bromide, trimethyl-(2-naphthalene-2-yl-2-oxo-ethyl)-ammonium bromide, trimethyl-(2oxo-2-phenan-threne-9-yl-ethyl)-ammonium bromide, and trimethyl-(2-oxo-2-pyrene-1yl-ethyl)-ammonium bromide were utilized to solubilize CNTs. Although the later two compounds are suboptimal in forming micelles themselves, they were found far more effective in achieving CNT water-suspension than the former two, suggesting that their association with CNTs is mainly a - interaction. To achieve even stronger ligand–CNT interaction, polycyclic aromatics, such as 1pyrenebutanoic acid succinimidyl ester, have been applied to modify CNTs [116]. Also, a diblock polymer, PS-b-PAA, which has multiple aromatic styrene groups, was immobilized on the CNTs. After crosslinking with diamine linkers, the resulting CNTs were reported to have good solubility in water (> 0.5 mg/L) [117]. The layer-by-layer deposition can be applied on the foundation laid by the aromatic compounds. For instance, the CNTs could be rendered positively charged first with 1-pyrenepropylamine hydrochloride, followed by
P1: OTA/XYZ P2: ABC c03 JWBS037-Chen August 29, 2010
13:6
Printer Name: Yet to Come
CARBON NANOTUBES
65
FIGURE 3.13 One example of using a layer-by-layer approach to modify CNTs. (Adapted from Artyukhin et al. [119] with permission.)
alternative depositions of negatively charged polystyrene sulfonate (PSS) and positively charged poly (diallyldimethylammonium chloride) (PDDA) to achieve water-solubility (Fig. 3.13) [118]. Among the many ligands that are suitable for ligand addition, PEGlyated phospholipids might be one of the best options, given their effectiveness, ease of functionalization, and biocompatibility. In a recent study conducted by Gambhir and co-workers, a group of mice were intravenously administrated phospholipid–CNT conjugates and were carefully monitored for 4 months [120]. Even though some CNTs were found trapped in the liver and spleen without degradation, evidence of toxicity was found neither from the survival, clinical, and laboratory parameters nor from the following necropsy and tissue histology studies. This result, while confirming the biosafety of phospholipid–CNT hybrids, suggests that the toxicity of CNTs that was previously observed was likely due to the coating materials, rather than the CNTs themselves [121]. The Dai group has led efforts to utilize phospholipid–CNT conjugates for several applications, including tumor imaging, drug delivery, and gene delivery [122–126]. Specifically, phospholipid-branched-PEG was used to modify CNTs and to load paclitexal (PTX); the resultant conjugates showed improved therapeutic effects compared to PTX alone [125]. 3.6.3 Direct Adsorption Biomolecules can also be loaded onto CNTs through direct physical interaction. For example, single strand DNA (ssDNA) can be adsorbed spontaneously onto SWNT surfaces, with its sugar–phosphate backbones exposed to the aqueous environment to render water solubility of the CNTs. The driving force underlying the conjugate formation is attributed to the - stacking between the bases and the graphite surface. Due to the electrostatic and torsional interaction within the sugar–phosphate backbones, the adsorbed DNAs
P1: OTA/XYZ P2: ABC c03 JWBS037-Chen August 29, 2010
66
13:6
Printer Name: Yet to Come
NANOPARTICLE SURFACE MODIFICATION AND BIOCONJUGATION
encompass the CNT in a helical manner [127–129]. Since chemical manipulation and polymerization of DNA is well understood, DNA–CNT hybrids can easily be modified for function addition. Erie and co-workers utilized thiolated DNA to modify CNTs, followed by the addition of CdSe/ZnS QDs. Atomic force microscopy (AFM) confirmed the formation of a CNT–DNA–QD hybrid and showed a distance of 14 nm between the adjacent QDs on the CNTs, implying an end-to-end arrangement of the DNAs on the CNTs [130]. Proteins can also be adsorbed directly onto CNTs. For instance, streptavidin was found to be easily adsorbed onto the CNT surface in a well-ordered self-assembly [131]. Although the mechanism is not yet fully understood, it is believed to involve a baseline hydrophobic interaction and aromatic stacking and may also result from the amino–carbon interaction between proteins and CTNs [132]. The adsorbed streptavidin retains its functionality and still actively binds to biotin and its derivatives. Double stranded DNA fragments, when biotinylated, can be immobilized onto such a streptavidin–CNT hybrid through biotin–streptavidin interaction. It is worth mentioning that both the DNA–CNT adducts and the protein–CNT adducts are highly stable and can sustain dialysis without forming aggregations, which is rarely observed for other CNT formulas [3].
3.7 SILICA PARTICLES Silica nanoparticles are another type of biomaterial that has been extensively studied. Interesting features of silica particles include (1) their biocompatibility, (2) the sophisticated size control (from 20 nm to several m) [133], (3) good water-solubility, (4) mesoporosity [134], and (5) the ready incorporation of imaging modalities into the matrices. The main precursor of the silica particle is tetraethyl orthosilicate (TEOS), which hydrolyzes and coagulates during synthesis to form the main body of the particles. However, lacking conjugation-friendly chemical groups on the surface, the silica particles are not readily functionalizable. To facilitate bioconjugation, aminopropyltrimethoxysilane (APS) and mercaptopropylmethoxysilane (MPS) are usually co-coagulated with TEOS, which introduces amines or thiols to the particle surface. If some functional molecules are precoupled with APS/MPS and co-coagulated with TEOS, then the functions can be imported into the silica particles. For example, organic dyes and Gd–DTPA complex were preconjugated with APS and mixed with TEOS as particle precursors. The resulting silica particles may possess thousands of dyes or Gd chelates in the matrices and are optically or magnetically active [135–137]. Aside from accommodating small molecules, various nanoparticles can also be incorporated into silica matrices as well. In other words, silica coagulation can serve as a coating strategy, which imports water solubility and facile functionalizability to the particles. Reports on employing silica to coat IONPs, gold nanoparticles, and QDs have been well documented [133, 138–142]. A great virtue of utilizing such a coating strategy is that it allows the encapsulation of more than one component in a fast and even quantitative way [143–148]. Nie and co-workers utilized silica to encapsulate both IONPs and QDs, creating a hybrid that retained both magnetic and optical properties [146]. Koole et al. [149] reported a dual-function core–shell–shell (CSS) nanoparticle, which were basically silica nanoparticles loaded with QDs and Gd complexes. While the QDs were encapsulated within the silica matrices via the above-mentioned method, the Gd chelates were loaded on the surface of the particles, which was achieved by making the particle surface hydrophobic
P1: OTA/XYZ P2: ABC c03 JWBS037-Chen August 29, 2010
13:6
Printer Name: Yet to Come
NANOCOMPOSITES
67
FIGURE 3.14 Preparation of multifunctional silica nanoparticles embedded with Gd chelates and QDs. (Adapted from Koole et al. [149] with permission.)
with otcadecanol followed by the adsorption of a hybrid phospholipid layer constituted of PEG-DSPE, Gd-DTPA-DSA, and Mal-PEG-DSPE. (Fig. 3.14) [149].
3.8 NANOCOMPOSITES A current effort in molecular imaging is the development of complex systems with more than one imaging entity, which encourages multimodality imaging. One way to achieve this goal is to encapsulate different nanostructures into one unit, as is the case when utilizing a silica coating. Alternatively, integration can be achieved at the chemical synthesis stage, which is advantageous in affording better ratiometric control and size limitations. Many efforts have recently been dedicated to the synthesis of composite nanoparticles, which resulted in a burst of reports on novel heterodimer particles [150]. Such particles typically possess two side-by-side substructures and therefore two different surfaces. In some cases, the two surfaces can be modified with the same ligand. For example, FePt–Au nanoparticles can be modified with PEGylated DHLA, which can bind with both Pt and Au atoms [151]. More commonly, however, two different ligands are needed, each targeting one moiety of the heterodimeric particles. For instance, Sun and co-workers utilized thiolated PEG and PEGylated dopamine to modify Fe3 O4 –Au nanoparticles, where the former ligand was anchored on the Au side and the latter one resided on the Fe3 O4 side (Fig. 3.15) [152]. For some particles, the two composites are assembled in a core–shell instead of a sideby-side manner. In these cases, the imbedded moiety has no contribution to the surface and
P1: OTA/XYZ P2: ABC c03 JWBS037-Chen August 29, 2010
68
13:6
Printer Name: Yet to Come
NANOPARTICLE SURFACE MODIFICATION AND BIOCONJUGATION
FIGURE 3.15 (a) Schmatic illustration of the Au–Fe3 O4 dumbell nanoparticle surface modification. (b, c) TME images of Au–Fe3 O4 nanoparticle in hexane and in water, respectivelly. (Adapted from Xu et al. [152] with permission.)
the surface modification is just about the moiety that is exposed. For instance, Fe3 O4 @Au nanocomposites can be made water-soluble by sodium citrate, which is known to have a high affinity toward the Au surface but not the Fe3 O4 surface [153]. For a more detailed introduction to recent progress on nanocomposites, readers are referred to several other excellent review papers [150, 154, 155].
3.9 CONCLUSION In summary, surface modification is essential for the development of nanoparticle-based imaging probes. It makes the nanoparticles water-soluble and affords them an interface for biomolecule conjugation. With the increased availability of animal facilities worldwide, more and more experiments are conducted at the in vivo level. In this setting, better control over the particle pharmacokinetics is of critical importance, which is highly dependent on the nanoparticle surface modification and bioconjugation.
REFERENCES 1. Laurent, S.; Forge, D.; Port, M.; Roch, A.; Robic, C.; Vander Elst, L.; Muller, R. N. Chem. Rev. 2008, 108, 2064. 2. Biju, V.; Itoh, T.; Anas, A.; Sujith, A.; Ishikawa, M. Anal. Bioanal. Chem. 2008, 391, 2469. 3. Lin, Y.; Taylor, S.; Li, H. P.; Fernando, K. A. S.; Qu, L. W.; Wang, W.; Gu, L. R.; Zhou, B.; Sun, Y. P. J. Mater. Chem. 2004, 14, 527. 4. Maeda, H.; Wu, J.; Sawa, T.; Matsumura, Y.; Hori, K. J. Control. Release 2000, 65, 271. 5. Moghimi, S. M.; Hunter, A. C.; Murray, J. C. Pharmacol. Rev. 2001, 53, 283. 6. Hermanson, G. T. Bioconjug. Techniques; Academic Press: San Diego, 1996. 7. Chen, X. Recent Advances of Bioconjugation Chemistry in Molecular Imaging; Research Signpost: Kerala, India, 2008. 8. Fischler, M.; Sologubenko, A.; Mayer, J.; Clever, G.; Burley, G.; Gierlich, J.; Carell, T.; Simon, U. Chem. Commun. 2008, 169.
P1: OTA/XYZ P2: ABC c03 JWBS037-Chen August 29, 2010
13:6
Printer Name: Yet to Come
REFERENCES
69
9. Lu, J.; Shi, M.; Shoichet, M. S. Bioconjug. Chem. 2009, 20, 87. 10. White, M. A.; Johnson, J. A.; Koberstein, J. T.; Turro, N. J. J. Am. Chem. Soc. 2006, 128, 11356. 11. Cai, W.; Shin, D. W.; Chen, K.; Gheysens, O.; Cao, Q.; Wang, S. X.; Gambhir, S. S.; Chen, X. Nano Lett. 2006, 6, 669. 12. Jun, Y. W.; Huh, Y. M.; Choi, J. S.; Lee, J. H.; Song, H. T.; Kim, S.; Yoon, S.; Kim, K. S.; Shin, J. S.; Suh, J. S.; Cheon, J. J. Am. Chem. Soc. 2005, 127, 5732. 13. Yu, W. W.; Chang, E.; Drezek, R.; Colvin, V. L. Biochem. Biophy. Res. Commun. 2006, 348, 781. 14. Bailey, R. E.; Smith, A. M.; Nie, S. M. Phys. E Low-Dimensional Syst. Nanostructures 2004, 25, 1. 15. Green, M. Angew. Chem. Int. Ed. Engl. 2004, 43, 4129. 16. Medintz, I. L.; Uyeda, H. T.; Goldman, E. R.; Mattoussi, H. Nat. Mater. 2005, 4, 435. 17. Choi, H. S.; Liu, W.; Misra, P.; Tanaka, E.; Zimmer, J. P.; Ipe, B. I.; Bawendi, M. G.; Frangioni, J. V. Nat. Biotechnol. 2007, 25, 1165. 18. Mitchell, G. P.; Mirkin, C. A.; Letsinger, R. L. J. Am. Chem. Soc. 1999, 121, 8122. 19. Pinaud, F.; King, D.; Moore, H. P.; Weiss, S. J. Am. Chem. Soc. 2004, 126, 6115. 20. Willard, D. M.; Carillo, L. L.; Jung, J.; Van Orden, A. Nano Lett. 2001, 1, 469. 21. Wang, Y. A.; Li, J. J.; Chen, H. Y.; Peng, X. G. J. Am. Chem. Soc. 2002, 124, 2293. 22. Guo, W. Z.; Li, J. J.; Wang, Y. A.; Peng, X. G. Chem. Mater. 2003, 15, 3125. 23. Kim, S.; Bawendi, M. G. J. Am. Chem. Soc. 2003, 125, 14652. 24. Kim, S.; Bawendi, M. G. J. Am. Chem. Soc. 2003, 125, 14652. 25. Clapp, A. R.; Medintz, I. L.; Mauro, J. M.; Fisher, B. R.; Bawendi, M. G.; Mattoussi, H. J. Am. Chem. Soc. 2004, 126, 301. 26. Medintz, I. L.; Konnert, J. H.; Clapp, A. R.; Stanish, I.; Twigg, M. E.; Mattoussi, H.; Mauro, J. M.; Deschamps, J. R. Proc. Natl. Acad. Sci. U.S.A. 2004, 101, 9612. 27. Carion, O.; Mahler, B.; Pons, T.; Dubertret, B. Nat. Protoc. 2007, 2, 2383. 28. Dubertret, B.; Skourides, P.; Norris, D. J.; Noireaux, V.; Brivanlou, A. H.; Libchaber, A. Science 2002, 298, 1759. 29. Osaki, F.; Kanamori, T.; Sando, S.; Sera, T.; Aoyama, Y. J. Am. Chem. Soc. 2004, 126, 6520. 30. Ballou, B.; Lagerholm, B. C.; Ernst, L. A.; Bruchez, M. P.; Waggoner, A. S. Bioconjug. Chem. 2004, 15, 79. 31. Mattheakis, L. C.; Dias, J. M.; Choi, Y. J.; Gong, J.; Bruchez, M. P.; Liu, J. Q.; Wang, E. Anal. Biochem. 2004, 327, 200. 32. Ballou, B.; Ernst, L. A.; Andreko, S.; Harper, T.; Fitzpatrick, J. A.; Waggoner, A. S.; Bruchez, M. P. Bioconjug. Chem. 2007, 18, 389. 33. Gao, X.; Cui, Y.; Levenson, R. M.; Chung, L. W.; Nie, S. Nat. Biotechnol. 2004, 22, 969. 34. Goldman, E. R.; Balighian, E. D.; Mattoussi, H.; Kuno, M. K.; Mauro, J. M.; Tran, P. T.; Anderson, G. P. J. Am. Chem. Soc. 2002, 124, 6378. 35. Ishii, D.; Kinbara, K.; Ishida, Y.; Ishii, N.; Okochi, M.; Yohda, M.; Aida, T. Nature 2003, 423, 628. 36. Xie, R.; Chen, K.; Chen, X.; Peng, X. Nano Res. 2008, 1, 457. 37. Chouly, C.; Pouliquen, D.; Lucet, I.; Jeune, J. J.; Jallet, P. J. Microencapsul. 1996, 13, 245. 38. Corot, C.; Violas, X.; Robert, P.; Gagneur, G.; Port, M. Invest. Radiol. 2003, 38, 311. 39. Duguet, E.; Vasseur, S.; Mornet, S.; Devoisselle, J. M. Nanomedicine 2006, 1, 157. 40. Edelstein, A. S.; Cammarata, R. C. Institute of Physics (Great Britain). Nanomaterials: Synthesis, Properties, and Applications. Institute of Physics Publishing: Bristol, UK, 1996.
P1: OTA/XYZ P2: ABC c03 JWBS037-Chen August 29, 2010
70
13:6
Printer Name: Yet to Come
NANOPARTICLE SURFACE MODIFICATION AND BIOCONJUGATION
41. Fleige, G.; Seeberger, F.; Laux, D.; Kresse, M.; Taupitz, M.; Pilgrimm, H.; Zimmer, C. Invest. Radiol. 2002, 37, 482. 42. Grancharov, S. G.; Zeng, H.; Sun, S. H.; Wang, S. X.; O’Brien, S.; Murray, C. B.; Kirtley, J. R.; Held, G. A. J. Phys. Chem. B. 2005, 109, 13030. 43. Ito, A.; Shinkai, M.; Honda, H.; Kobayashi, T. J. Biosci. Bioeng. 2005, 100, 1. 44. Lu, A. H.; Salabas, E. L.; Schuth, F. Angew. Chem. Int. Ed. 2007, 46, 1222. 45. Kang, Y. S.; Risbud, S.; Rabolt, J. F.; Stroeve, P. Chem. Mater. 1996, 8, 2209. 46. Gupta, A. K.; Gupta, M. Biomaterials 2005, 26, 3995. 47. Lanza, G. M.; Winter, P. M.; Caruthers, S. D.; Morawski, A. M.; Schmieder, A. H.; Crowder, K. C.; Wickline, S. A. J. Nucl. Cardiol. 2004, 11, 733. 48. Mornet, S.; Vasseur, S.; Grasset, F.; Duguet, E. J. Mater. Chem. 2004, 14, 2161. 49. Harisinghani, M. G.; Barentsz, J.; Hahn, P. F.; Deserno, W. M.; Tabatabaei, S.; van de Kaa, C. H.; de la Rosette, J.; Weissleder, R. N. Engl. J. Med. 2003, 348, 2491. 50. Wang, Y. X.; Hussain, S. M.; Krestin, G. P. Eur. Radiol. 2001, 11, 2319. 51. Corot, C.; Robert, P.; Idee, J. M.; Port, M. Adv. Drug. Deliv. Rev. 2006, 58, 1471. 52. Pan, B. F.; Gao, F.; Gu, H. C. J. Colloid Interface Sci 2005, 284, 1. 53. Lee, H.-Y.; Li, Z.; Chen, K.; Hsu, A. R.; Xu, C.; Xie, J.; Sun, S.; Chen, X. J. Nucl. Med. 2008, 49, 1371. 54. Pirko, I.; Johnson, A.; Ciric, B.; Gamez, J.; Macura, S. I.; Pease, L. R.; Rodriguez, M. FASEB. J. 2003, 17, 179. 55. Zhao, M.; Kircher, M. F.; Josephson, L.; Weissleder, R. Bioconjug. Chem. 2002, 13, 840. 56. Kang, H. W.; Josephson, L.; Petrovsky, A.; Weissleder, R.; Bogdanov, A. Bioconjug. Chem. 2002, 13, 122. 57. Josephson, L.; Tung, C. H.; Moore, A.; Weissleder, R. Bioconjug. Chem. 1999, 10, 186. 58. Hogemann, D.; Ntziachristos, V.; Josephson, L.; Weissleder, R. Bioconjug. Chem. 2002, 13, 116. 59. Kohler, N.; Fryxell, G. E.; Zhang, M. Q. J. Am. Chem. Soc. 2004, 126, 7206. 60. Kohler, N.; Sun, C.; Fichtenholtz, A.; Gunn, J.; Fang, C.; Zhang, M. Q. Small 2006, 2, 785. 61. Kohler, N.; Sun, C.; Wang, J.; Zhang, M. Q. Langmuir 2005, 21, 8858. 62. Lee, H.; Lee, E.; Kim do, K.; Jang, N. K.; Jeong, Y. Y.; Jon, S. J. Am. Chem. Soc. 2006, 128, 7383. 63. Yu, M. K.; Jeong, Y. Y.; Park, J.; Park, S.; Kim, J. W.; Min, J. J.; Kim, K.; Jon, S. Angew. Chem. Int. Ed. Engl. 2008, 47, 5362. 64. Jun, Y. W.; Seo, J. W.; Cheon, A. Acc. Chem. Res. 2008, 41, 179. 65. Lee, J. H.; Huh, Y. M.; Jun, Y.; Seo, J.; Jang, J.; Song, H. T.; Kim, S.; Cho, E. J.; Yoon, H. G.; Suh, J. S.; Cheon, J. Nat. Med. 2007, 13, 95. 66. Sun, S.; Zeng, H. J. Am. Chem. Soc. 2002, 124, 8204. 67. Zeng, H.; Rice, P. M.; Wang, S. X.; Sun, S. J. Am. Chem. Soc. 2004, 126, 11458. 68. Yoshino, T.; Hirabe, H.; Takahashi, M.; Kuhara, M.; Takeyama, H.; Matsunaga, T. Biotechnol. Bioeng. 2008, 101, 470. 69. Hultman, K. L.; Raffo, A. J.; Grzenda, A. L.; Harris, P. E.; Brown, T. R.; O’Brien, S. ACS Nano 2008, 2, 477. 70. Pellegrino, T.; Manna, L.; Kudera, S.; Liedl, T.; Koktysh, D.; Rogach, A. L.; Keller, S.; Radler, J.; Natile, G.; Parak, W. J. Nano Lett. 2004, 4, 703. 71. Song, H. T.; Choi, J. S.; Huh, Y. M.; Kim, S.; Jun, Y. W.; Suh, J. S.; Cheon, J. J. Am. Chem. Soc. 2005, 127, 9992.
P1: OTA/XYZ P2: ABC c03 JWBS037-Chen August 29, 2010
13:6
Printer Name: Yet to Come
REFERENCES
71
72. Xu, C. J.; Xu, K. M.; Gu, H. W.; Zheng, R. K.; Liu, H.; Zhang, X. X.; Guo, Z. H.; Xu, B. J. Am. Chem. Soc. 2004, 126, 9938. 73. Xie, J.; Xu, C.; Kohler, N.; Hou, Y.; Sun, S. Adv. Mater. 2007, 19, 3648. 74. Li, W. Z.; Nesselmann, C.; Zhou, Z. H.; Ong, L. L.; Ori, F.; Tang, G. P.; Kaminski, A.; Lutzow, K.; Lendlein, A.; Liebold, A.; Stamm, C.; Wang, J.; Steinhoff, G.; Ma, N. J. Magn. Magn. Mater. 2007, 311, 336. 75. Neuberger, T.; Schopf, B.; Hofmann, H.; Hofmann, M.; von Rechenberg, B. J. Magn. Magn. Mater. 2005, 293, 483. 76. Scherer, F.; Anton, M.; Schillinger, U.; Henkel, J.; Bergemann, C.; Kruger, A.; Gansbacher, B.; Plank, C. Gene. Ther. 2002, 9, 102. 77. Thunemann, A. F.; Schutt, D.; Kaufner, L.; Pison, U.; Mohwald, H. Langmuir 2006, 22, 2351. 78. Lee, J. H.; Lee, K.; Moon, S. H.; Lee, Y.; Park, T. G.; Cheon, J. Angew. Chem. Int. Ed. Engl. 2009, 48, 4174. 79. Daniel, M. C.; Astruc, D. Chem. Rev. 2004, 104, 293. 80. Niemeyer, C. M. Angew. Chem. Int. Ed. 2001, 40, 4128. 81. Hu, M.; Chen, J. Y.; Li, Z. Y.; Au, L.; Hartland, G. V.; Li, X. D.; Marquez, M.; Xia, Y. N. Chem. Soc. Rev. 2006, 35, 1084. 82. Murphy, C. J.; San, T. K.; Gole, A. M.; Orendorff, C. J.; Gao, J. X.; Gou, L.; Hunyadi, S. E.; Li, T. J. Phys. Chem. B 2005, 109, 13857. 83. Murphy, C. J.; Sau, T. K.; Gole, A.; Orendorff, C. J. MRS Bull. 2005, 30, 349. 84. Chen, J.; Saeki, F.; Wiley, B. J.; Cang, H.; Cobb, M. J.; Li, Z. Y.; Au, L.; Zhang, H.; Kimmey, M. B.; Li, X. D.; Xia, Y. Nano Lett. 2005, 5, 473. 85. Wang, C.; Hu, Y.; Lieber, C. M.; Sun, S. J. Am. Chem. Soc. 2008, 130, 8902. 86. Link, S.; El-Sayed, M. A. J. Phys. Chem. B 1999, 103, 4212. 87. Murphy, C. J.; Sau, T. K.; Gole, A. M.; Orendorff, C. J.; Gao, J.; Gou, L.; Hunyadi, S. E.; Li, T. J. Phys. Chem. B 2005, 109, 13857. 88. Dixit, V.; Van den Bossche, J.; Sherman, D. M.; Thompson, D. H.; Andres, R. P. Bioconjug. Chem. 2006, 17, 603. 89. Oyelere, A. K.; Chen, P. C.; Huang, X.; El-Sayed, I. H.; El-Sayed, M. A. Bioconjug. Chem. 2007, 18, 1490. 90. Reynolds, A. J.; Haines, A. H.; Russell, D. A. Langmuir 2006, 22, 1156. 91. Rosi, N. L.; Mirkin, C. A. Chem. Rev. 2005, 105, 1547. 92. Rosi, N. L.; Giljohann, D. A.; Thaxton, C. S.; Lytton-Jean, A. K.; Han, M. S.; Mirkin, C. A. Science 2006, 312, 1027. 93. Chang, J. Y.; Wu, H. M.; Chen, H.; Ling, Y. C.; Tan, W. H. Chem. Commun. 2005, 1092. 94. Chang, J. Y.; Wu, H.; Chen, H.; Ling, Y. C.; Tan, W. Chem. Commun. (Camb.) 2005, 1092. 95. Hauck, T. S.; Ghazani, A. A.; Chan, W. C. Small 2008, 4, 153. 96. Gole, A.; Murphy, C. J. Chem. Mater. 2005, 17, 1325. 97. Huang, X. H.; El-Sayed, I. H.; Qian, W.; El-Sayed, M. A. Nano Lett. 2007, 7, 1591. 98. Shukla, R.; Nune, S. K.; Chanda, N.; Katti, K.; Mekapothula, S.; Kulkarni, R. R.; Welshons, W. V.; Kannan, R.; Katti, K. V. Small 2008, 4, 1425. 99. Kannan, R.; Rahing, V.; Cutler, C.; Pandrapragada, R.; Katti, K. K.; Kattumuri, V.; Robertson, J. D.; Casteel, S. J.; Jurisson, S.; Smith, C.; Boote, E.; Katti, K. V. J. Am. Chem. Soc. 2006, 128, 11342. 100. Kattumuri, V.; Katti, K.; Bhaskaran, S.; Boote, E. J.; Casteel, S. W.; Fent, G. M.; Robertson, D. J.; Chandrasekhar, M.; Kannan, R.; Katti, K. V. Small 2007, 3, 333.
P1: OTA/XYZ P2: ABC c03 JWBS037-Chen August 29, 2010
72
13:6
Printer Name: Yet to Come
NANOPARTICLE SURFACE MODIFICATION AND BIOCONJUGATION
101. Liu, J.; Rinzler, A. G.; Dai, H.; Hafner, J. H.; Bradley, R. K.; Boul, P. J.; Lu, A.; Iverson, T.; Shelimov, K.; Huffman, C. B.; Rodriguez-Macias, F.; Shon, Y. S.; Lee, T. R.; Colbert, D. T.; Smalley, R. E. Science 1998, 280, 1253. 102. Shi Kam, N. W.; Jessop, T. C.; Wender, P. A.; Dai, H. J. Am. Chem. Soc. 2004, 126, 6850. 103. Peng, H.; Alemany, L. B.; Margrave, J. L.; Khabashesku, V. N. J. Am. Chem. Soc. 2003, 125, 15174. 104. Jiang, K. Y.; Schadler, L. S.; Siegel, R. W.; Zhang, X. J.; Zhang, H. F.; Terrones, M. J. Mater. Chem. 2004, 14, 37. 105. Baker, S. E.; Cai, W.; Lasseter, T. L.; Weidkamp, K. P.; Hamers, R. J. Nano Lett. 2002, 2, 1413. 106. Huang, W. J.; Taylor, S.; Fu, K. F.; Lin, Y.; Zhang, D. H.; Hanks, T. W.; Rao, A. M.; Sun, Y. P. Nano Lett. 2002, 2, 311. 107. Pompeo, F.; Resasco, D. E. Nano Lett. 2002, 2, 369. 108. Nguyen, C. V.; Delzeit, L.; Cassell, A. M.; Li, J.; Han, J.; Meyyappan, M. Nano Lett. 2002, 2, 1079. 109. Katz, E.; Willner, I. Chemphyschem 2004, 5, 1084. 110. Bianco, A.; Kostarelos, K.; Partidos, C. D.; Prato, M. Chem. Commun. (Camb.) 2005, 571. 111. Wu, W.; Wieckowski, S.; Pastorin, G.; Benincasa, M.; Klumpp, C.; Briand, J. P.; Gennaro, R.; Prato, M.; Bianco, A. Angew. Chem. Int. Ed. Engl. 2005, 44, 6358. 112. Wu, W.; Wieckowski, S.; Pastorin, G.; Benincasa, M.; Klumpp, C.; Briand, J. P.; Gennaro, R.; Prato, M.; Bianco, A. Angew. Chem. Int. Ed. 2005, 44, 6358. 113. Islam, M. F.; Rojas, E.; Bergey, D. M.; Johnson, A. T.; Yodh, A. G. Nano Lett. 2003, 3, 269. 114. Nakashima, N.; Tomonari, Y.; Murakami, H. Chem. Lett. 2002, 638. 115. Moore, V. C.; Strano, M. S.; Haroz, E. H.; Hauge, R. H.; Smalley, R. E.; Schmidt, J.; Talmon, Y. Nano Lett. 2003, 3, 1379. 116. Chen, R. J.; Zhang, Y. G.; Wang, D. W.; Dai, H. J. J. Am. Chem. Soc. 2001, 123, 3838. 117. Kang, Y.; Taton, T. A. J. Am. Chem. Soc. 2003, 125, 5650. 118. Artyukhin, A. B.; Bakajin, O.; Stroeve, P.; Noy, A. Langmuir 2004, 20, 1442. 119. Artyukhin, A. B.; Bakajin, O.; Stroeve, P.; Noy, A. Langmuir 2004, 20, 1442. 120. Schipper, M. L.; Nakayama-Ratchford, N.; Davis, C. R.; Kam, N. W. S.; Chu, P.; Liu, Z.; Sun, X. M.; Dai, H. J.; Gambhir, S. S. Nat. Nanotechnol. 2008, 3, 216. 121. Schipper, M. L.; Nakayama-Ratchford, N.; Davis, C. R.; Kam, N. W.; Chu, P.; Liu, Z.; Sun, X.; Dai, H.; Gambhir, S. S. Nat. Nanotechnol. 2008, 3, 216. 122. Kam, N. W. S.; O’Connell, M.; Wisdom, J. A.; Dai, H. J. Proc. Nat. Acad. Sci. U.S.A. 2005, 102, 11600. 123. Liu, Z. A.; Li, X. L.; Tabakman, S. M.; Jiang, K. L.; Fan, S. S.; Dai, H. J. J. Am. Chem. Soc. 2008, 130, 13540. 124. De La Zerda, A.; Zavaleta, C.; Keren, S.; Vaithilingam, S.; Bodapati, S.; Liu, Z.; Levi, J.; Smith, B. R.; Ma, T. J.; Oralkan, O.; Cheng, Z.; Chen, X. Y.; Dai, H. J.; Khuri-Yakub, B. T.; Gambhir, S. S. Nat. Nanotechnol. 2008, 3, 557. 125. Liu, Z.; Chen, K.; Davis, C.; Sherlock, S.; Cao, Q. Z.; Chen, X. Y.; Dai, H. J. Cancer Res. 2008, 68, 6652. 126. Liu, Z.; Davis, C.; Cai, W. B.; He, L.; Chen, X. Y.; Dai, H. J. Proc. Nat. Acad. Sci. U.S.A. 2008, 105, 1410. 127. Zheng, M.; Jagota, A.; Semke, E. D.; Diner, B. A.; Mclean, R. S.; Lustig, S. R.; Richardson, R. E.; Tassi, N. G. Nat. Mater. 2003, 2, 338. 128. Zheng, M.; Jagota, A.; Strano, M. S.; Santos, A. P.; Barone, P.; Chou, S. G.; Diner, B. A.; Dresselhaus, M. S.; McLean, R. S.; Onoa, G. B.; Samsonidze, G. G.; Semke, E. D.; Usrey, M.; Walls, D. J. Science 2003, 302, 1545.
P1: OTA/XYZ P2: ABC c03 JWBS037-Chen August 29, 2010
13:6
Printer Name: Yet to Come
REFERENCES
73
129. Johnson, R. R.; Johnson, A. T. C.; Klein, M. L. Nano Lett. 2008, 8, 69. 130. Campbell, J. F.; Tessmer, I.; Thorp, H. H.; Erie, D. A. J. Am. Chem. Soc. 2008, 130, 10648. 131. Balavoine, F.; Schultz, P.; Richard, C.; Mallouh, V.; Ebbesen, T. W.; Mioskowski, C. Angew. Chem. Int. Ed. 1999, 38, 1912. 132. Kong, J.; Dai, H. J. J. Phys. Chem. B 2001, 105, 2890. 133. Jana, N. R.; Earhart, C.; Ying, J. Y. Chem. Mater. 2007, 19, 5074. 134. Slowing, I. I., Vivero-Escoto, J. L.; Wu, C. W.; Lin, V. S. Adv. Drug. Deliv. Rev. 2008, 60, 1278. 135. Ow, H.; Larson, D. R.; Srivastava, M.; Baird, B. A.; Webb, W. W.; Wiesner, U. Nano Lett. 2005, 5, 113. 136. Tapec, R.; Zhao, X. J.; Tan, W. J. Nanosci. Nanotechnol. 2002, 2, 405. 137. Hsiao, J. K.; Tsai, C. P.; Chung, T. H.; Hung, Y.; Yao, M.; Liu, H. M.; Mou, C. Y.; Yang, C. S.; Chen, Y. C.; Huang, D. M. Small 2008, 4, 1445. 138. Wolcott, A.; Gerion, D.; Visconte, M.; Sun, J.; Schwartzberg, A.; Chen, S. W.; Zhang, J. Z. J. Phys. Chem. B 2006, 110, 5779. 139. Gerion, D.; Pinaud, F.; Williams, S. C.; Parak, W. J.; Zanchet, D.; Weiss, S.; Alivisatos, A. P. J. Phys. Chem. B 2001, 105, 8861. 140. Kang, K.; Choi, J.; Nam, J. H.; Lee, S. C.; Kim, K. J.; Lee, S. W.; Chang, J. H. J. Phys. Chem. B 2009, 113, 536. 141. Wang, C. G.; Chen, Y.; Wang, T. T.; Ma, Z. F.; Su, Z. M. Adv. Functional Mater. 2008, 18, 355. 142. Lu, C. W.; Hung, Y.; Hsiao, J. K.; Yao, M.; Chung, T. H.; Lin, Y. S.; Wu, S. H.; Hsu, S. C.; Liu, H. M.; Mou, C. Y.; Yang, C. S.; Huang, D. M.; Chen, Y. C. Nano Lett. 2007, 7, 149. 143. Kim, J.; Lee, J. E.; Lee, J.; Yu, J. H.; Kim, B. C.; An, K.; Hwang, Y.; Shin, C. H.; Park, J. G.; Hyeon, T. J. Am. Chem. Soc. 2006, 128, 688. 144. Rieter, W. J.; Kim, J. S.; Taylor, K. M.; An, H.; Lin, W.; Tarrant, T. Angew. Chem. Int. Ed. Engl. 2007, 46, 3680. 145. Salgueirino-Maceira, V.; Correa-Duarte, M. A.; Spasova, M.; Liz-Marzan, L. M.; Farle, M. Adv. Functional Mater. 2006, 16, 509. 146. Sathe, T. R.; Agrawal, A.; Nie, S. Anal. Chem. 2006, 78, 5627. 147. Selvan, S. T.; Patra, P. K.; Ang, C. Y.; Ying, J. Y. Angew. Chem. Int. Ed. Engl. 2007, 46, 2448. 148. Yi, D. K.; Selvan, S. T.; Lee, S. S.; Papaefthymiou, G. C.; Kundaliya, D.; Ying, J. Y. J. Am. Chem. Soc. 2005, 127, 4990. 149. Koole, R.; van Schooneveld, M. M.; Hilhorst, J.; Castermans, K.; Cormode, D. P.; Strijkers, G. J.; de Mello Donega, C.; Vanmaekelbergh, D.; Griffioen, A. W.; Nicolay, K.; Fayad, Z. A.; Meijerink, A.; Mulder, W. J. Bioconjug. Chem. 2008, 19, 2471. 150. Gao, J.; Gu, H.; Xu, B. Acc. Chem. Res. 2009, 42, 1097. 151. Choi, J. S.; Jun, Y. W.; Yeon, S. I.; Kim, H. C.; Shin, J. S.; Cheon, J. J. Am. Chem. Soc. 2006, 128, 15982. 152. Xu, C.; Xie, J.; Ho, D.; Wang, C.; Kohler, N.; Walsh, E. G.; Morgan, J. R.; Chin, Y. E.; Sun, S. Angew. Chem. Int. Ed. 2008, 47, 173. 153. Xu, Z. C.; Hou, Y. L.; Sun, S. H. J. Am. Chem. Soc. 2007, 129, 8698. 154. Rhyner, M. N.; Smith, A. M.; Gao, X.; Mao, H.; Yang, L.; Nie, S. Nanomedicine 2006, 1, 209. 155. Liong, M.; Lu, J.; Kovochich, M.; Xia, T.; Ruehm, S. G.; Nel, A. E.; Tamanoi, F.; Zink, J. I. ACS Nano 2008, 2, 889.
P1: OTA/XYZ P2: ABC c04 JWBS037-Chen January 25, 2011
8:49
Printer Name: Yet to Come
CHAPTER 4
Biodistribution and Pharmacokinetics of Nanoprobes NAGESH KOLISHETTI Department of Anesthesiology, Brigham and Women’s Hospital, Harvard Medical School, Boston, Massachusetts, USA
FRANK ALEXIS Department of Bioengineering, Clemson University, Clemson, South Carolina, USA
ERIC M. PRIDGEN Department of Chemical Engineering, Massachusetts Institute of Technology, Cambridge, Massachusetts, USA
OMID C. FAROKHZAD Department of Anesthesiology, Brigham and Women’s Hospital, Harvard Medical School, Boston, Massachusetts, USA
4.1 INTRODUCTION The field of medicine has made significant progress in the areas of diagnosis and therapy due to recent advances in nanotechnology. Early disease diagnosis substantially increases patient survival. Therefore, significant research efforts have been aimed at the development of new imaging agents and imaging systems. Nanoprobes are a class of imaging agents that consist of metal and inorganic materials such as quantum dots (QDs) [1–6], magnetic nanoparticle (MNPs) [7–11], carbon nanotubes (CNTs) [12, 13], and gold nanoparticles (GNPs) [14–19]. These agents are synthesized using a “bottom–up” approach. . . for precise control over physicochemical properties, allowing particles to be engineered with small sizes, long blood circulation times, and a functionalized surface for targeting capabilities. The ability to tailor the properties of nanoprobes combined with the discovery of new disease markers or pathological defects have led to increases in the specificity and sensitivity of these agents. Clinical translation of nanoprobes follows a path similar to drugs, with the evaluation of efficacy, stability, toxicity, purity, and pharmacokinetics. However, because of their larger size and slower excretion rate, national agencies have had to create specific guidelines
Nanoplatform-Based Molecular Imaging Edited by Xiaoyuan Chen C 2011 John Wiley & Sons, Inc. Copyright
75
P1: OTA/XYZ P2: ABC c04 JWBS037-Chen January 25, 2011
76
8:49
Printer Name: Yet to Come
BIODISTRIBUTION AND PHARMACOKINETICS OF NANOPROBES
to test short- and long-term interactions of nanoprobes with the biological environment. Research into the synthesis and screening of nanoprobes to improve their safety profiles is a very active area due to their clinical potential. More recently, progress has been made in the development of theragnostics [20–29], a term denoting the fusion of therapeutics and diagnostics. Nanoprobes are being combined with drug delivery systems to enable the simultaneous imaging and treatment of diseases, further enhancing their potential use in biomedical applications. Theragnostic agents are receiving increasing attention as pharmacogenomics increases in clinical use [30]. In this chapter, nanoprobe technologies, specific toxicological characteristics, clinical developments, pharmacokinetics, and multifunctional nanoprobes for theranostic applications will be reviewed.
4.2 NANOPROBE TECHNOLOGIES Various imaging techniques developed over the past few decades have contributed to the advancement of cancer and cardiovascular research and their translation to present medical practice [31–33]. Imaging techniques have aided the understanding of cancer biology as well as the detection and monitoring of tumors. Recent reviews have covered various advances in imaging technologies and their applications for cancer diagnosis [31, 32, 34]. These technologies are clinically used based on their sensitivity, spatial resolution, and tissue penetration. For example, X-rays are used for imaging hard tissues such as bones while ultrasound, magnetic resonance imaging (MRI), and computed tomography (CT) are used to image soft tissues such as the heart, lungs, and tumors. [31]. These techniques provide information about tissue defects or pathologies. Their spatial resolution varies from microns to millimeters. Recently, various microscopic and other intravital optical techniques like tomography or macroscopic reflectance have been developed to study genetic, molecular, and cellular events in vivo [31]. Numerous microversions of imaging modalities such as microCT, microMRI, and microPET (positron emission tomography), as well as microscopy techniques such as fluorescence, multiphoton, and electron microscopic techniques, are primarily used in cellular studies [31, 32]. Depending on the imaging technique, the contrast agents/molecular probes could be proteins, dyes, iodinated molecules, microbubbles, or radiolabeled molecules. Several nanomaterials are promising candidates (Fig. 4.1) as imaging agents (nanoprobes) such as iron oxide-based magnetic nanoparticles, quantum dots, carbon nanotubes, and gold nanoparticles. Unlike conventional imaging agents and therapeutics, nanoprobes are stable in the biological environment, have enhanced bioavailability, display absorption at various wavelengths, and provide better contrast or magnetic moment. In general, nanomaterials provide higher sensitivity and specificity than molecular imaging agents [31, 35]. 4.2.1 Quantum Dots A unique property of quantum dots (QDs) is their increased absorbance with increasing separation between excitation and emission wavelengths. Attenuation of signal during imaging is a major concern due to tissue thickness [36]. Absorbance, scattering, and waterto-hemoglobin ratio can affect the performance of QDs [37, 38]. One of the advantages of QDs is their ability to emit in the visible and near-infrared (NIR) emission wavelengths, where the biological tissues can transmit the light [39]. QDs are >10 times brighter than
P1: OTA/XYZ P2: ABC c04 JWBS037-Chen January 25, 2011
8:49
Printer Name: Yet to Come
NANOPROBE TECHNOLOGIES
77
FIGURE 4.1 Examples of promising nanoprobes: quantum dots (QDs), gold nanoparticles (GNPs), magnetic nanoparticles (MNPs), and carbon nanotubes (CNTs).
fluorescent dyes and significantly less sensitive to photobleaching. QDs are used to image coronary vasculature, blood vessels, and the lymphatic system due to their superior fluorescence properties [40]. For example, NIR QDs have been used to image the coronary vasculature of a rat [37]. Coronary vasculature of thickness 282 m in a rat was imaged with a signal-to-background ratio of 5 using an inexpensive white light of 2 mW/cm2 after intravenous (IV) injection of 2.5 nmol QDs, while conventional fluorophores like IRDye78CA needed 12.5 mW/cm2 to get similar resolution. QDs with higher emission wavelengths (>655 nm) were used to image blood vessels of chick chorioallantoic membrane (CAM) [41], a model for studying angiogenesis. These QDs eliminated all chick-derived autofluorescence and improved depth of field imaging. In another example, green light-emitting QDs remained fluorescent and detectable in the capillaries of adipose tissue and skin of a living mouse following IV injection [42]. Conventional FITC-dextran was unable to provide equivalent resolution even after using 5 times more imaging power, indicating that QDs have higher resolution when imaging was performed at a depth of 250 m in a live mouse. Similarly, Gao et al. [38] reported targeted QDs for cancer imaging in prostate cancer tumor xenograft mice and estimated that tumor imaging sensitivity was increased to a detection limit of 10 to 100 cancer cells. In addition, cadmium selinide (CdSe)–zinc sulfide (ZnS) core–shell QDs were shown to be present in the liver and spleen for 4 months post-injection suggesting high in vivo stability over a long period of time. All of these reports demonstrate that QDs are excellent fluorescent probes for imaging various blood vessel tissues. They are extremely bright and photostable, have low background interference, and an improved depth-of-imaging over conventional fluorescent dye conjugates. Recently, major research and development efforts have focused on large-scale synthesis and size-controlling synthetic schemes to produce nanoparticles with high monodispersity
P1: OTA/XYZ P2: ABC c04 JWBS037-Chen January 25, 2011
78
8:49
Printer Name: Yet to Come
BIODISTRIBUTION AND PHARMACOKINETICS OF NANOPROBES
for precise control of emission characteristics. QDs are an emerging class of nanoprobes for in vivo optical imaging [1, 43, 44] but have not reached the clinical phase of development mainly due to toxicological concerns.
4.2.2 Carbon Nanotubes Carbon nanotubes (CNTs) are rolls of graphene sheets, with unparalleled physical, mechanical, and chemical properties. CNTs can easily be synthesized in large quantities via chemical vapor deposition, arc discharge, and laser ablation processes. The two main types of CNTs are single-walled carbon nanotubes, SWNTs (0.4–2 nm diameter and 20–1000 nm length), and multiwalled carbon nanotubes, MWNTs (1–100 nm diameter and 1000–5000 nm length). MWNTs are distinguished by having multiple layers of graphene rolled on top of each other, resulting in strnger mechanical properties and chemical resistance. Due to the sharp densities of electronic states, SWNTs have distinctive optical properties [45]. A recent report provided detailed information about various methods used for both covalent and noncovalent functionalization of CNTs [46]. CNTs offer new opportunities in biomedical research because of their large surface area, which could allow the efficient loading of multiple molecules. CNTs have been used for in vivo animal imaging and have potential applications in various optical imaging techniques such as Raman imaging, photoablation, and photoacoustic molecular imaging of cancer [46–48]. The intrinsic near-infrared fluorescence makes SWNTs a promising optical agent [49]. For example, gadolinium-loaded SWNTs have been shown to have greater relaxivity than conventional MRI agents and therefore might become a promising contrast agent for MRI in the near future [50]. However, the clinical translation of CNTs will depend on how well-controllled the synthesis processes are, toxicological studies in various animal models, and interactions between the administered CNTs and the immune complement system.
4.2.3 Gold Nanoparticles Gold nanoparticles (GNPs) are interesting candidates for imaging/diagnostic and theragnostic applications due to their biocompatibility, inert nature, resistance to oxidation, and strong absorption of NIR radiation. Recent studies on colloidal gold nanoparticles have focused on CT applications in place of traditional iodinated molecules dissolved in liquids [51], or as X-ray and MRI contrast agents [51]. Their tunable plasmon resonance enables them to be specifically engineered to a desired wavelength for a particular application [52]. One distinct advantage of GNPs is the ease of surface functionalization using either thiol/disulfide sulfydryl or click chemistry approaches [52]. GNPs systems are currently being developed for high-sensitivty assays that require functionalization of the surface with biological molecules. For example, Hirsch et al. [52] developed a gold nanoparticle-based immunoassay to detect immunoglobulins in whole blood with a detection limit in the range of ng/mL. Mirkin’s group [15, 53–55] and others [14, 56] have developed quantitative multiple nucleic acid detection technologies, including a colorimetric sensing system based on the formation of nanoparticle networks and changes in color due to different distances between each nanoparticle. In addition to detection applications, GNPs have great potential to become a future theragnostic agent by combining the imaging capabilities with drug delivery and/pr photothermal therapy.
P1: OTA/XYZ P2: ABC c04 JWBS037-Chen January 25, 2011
8:49
Printer Name: Yet to Come
NANOPROBE TOXICITY
79
4.2.4 Magnetic Nanoparticles Iron oxide magnetic nanoparticles (MNPs) are the most commonly used metallic nanoparticles due to their superior biocompatibility. Magnetic properties of iron oxide MNPs are characterized by the shortening of the T2 relaxation times of water, allowing imaging and diagnosis based on MRI [57]. MNPs do not possess magnetic properties outside an external magnetic field in contrast to other magnetic particles [58]. Unlike large magnetic particles with different domains of spins, MNPs have only a single domain of mutually aligned spins with a very large magnetic moment, making them especially effective contrast agents [59]. Because of the large magnetic moment, iron oxide MNPs are also known as superparamagnetic iron oxide (SPIO). MNPs are hydrophobic and require surface modification to disperse them in aqueous solutions. Multiple approaches are used to solubilize MNPs including polymer, surfactant, lipid-based, protein, and metal or oxide coatings [21, 22, 31, 34, 60–64]. Coating improves the nanoparticles’ stability, prevents immunogenic response, reduces opsonization, and has a limited effect on the nanoparticles’ magnetic properties. Polymeric coatings of magnetic nanoparticle have been shown to be the most promising approach and formulations used in the clinic or in clinical development are based on polymer coatings with affinity to iron oxide crystals [11]. SPIO-based MNPs are already used in medical imaging, mostly for imaging liver tissues due to their high accumulation in Kupffer cells, while ultrasmall paramagnetic iron oxide (USPIO) nanoparticles are in clinical development for lymph node imaging due to their longer circulation time [59, 65]. 4.2.5 Summary Nanoprobe imaging technologies have been developed based on the optical properties of materials, although the materials have not been optimized for biomedical applications. Most of the efforts have been in the synthesis of monodisperse nanoparticles with reduced impurities and surface modifications to improve their stability and biocompatibility. Thus, in the past 20 years, numerous imaging approaches using nanoprobes have been developed to improve resolution, sensitivity, and therapeutic applications but their possible clinical translation is limited by their immunogenicity and cellular and tissue toxicity.
4.3 NANOPROBE TOXICITY Nanoprobe toxicity is a concern for biomedical and clinical applications [66]. Different toxicity profiles are addressed depending on the material used for imaging or residual impurities from synthesis. Multiple factors such as aggregation, formation of reactive oxygen species, and nonspecific binding to proteins and cellular membranes influence the mechanism of toxicity in vivo. In general, surface coating, inert nanomaterials and/or biodegradability properties, and small size have been shown to decrease cell and tissue toxicity [67]. However, a universal toxicity assay to assess nanoprobes is difficult due to difference in cell types and assays used. Toxicities associated with different nanomaterials include stability, oxidative stress, genotoxicity, platelet coagulation, tissue-specific toxicity, hepatocyte toxicity, and immunogenicity [68, 69]. Therefore, in order to determine the potential toxicity of new nanomaterial formulations, the National Cancer Institute, NCI established the Nanotechnology Characterization Laboratory (NCL) to test and accelerate the transfer of technologies to the clinic through physical, in vitro, and in vivo characterization.
P1: OTA/XYZ P2: ABC c04 JWBS037-Chen January 25, 2011
80
8:49
Printer Name: Yet to Come
BIODISTRIBUTION AND PHARMACOKINETICS OF NANOPROBES
Specifically, nanoprobes are tested for accuracy, sensitivity, and specificity. The NCL serves as a national resource of knowledge to facilitate the regulatory process and study preclinical toxicology parameters. The NCL has developed a series of characterization tests, including toxicity, that have been incorporated into the American Society for Testing and Materials (ASTM). The three new ASTM standards are based on the evaluation of nanomaterial toxicity in kidney and liver cells (E2526), hemolysis (E2524), and stimulation or inhibition of bone marrow macrophages (E2525). In addition, Ralph Weissleder’s group [70] developed an in vitro screening platform based on multiple cellular assays and cell types to analyze nanoprobe toxicity profiles in vitro and predict their in vivo behavior. A hierarchical clustering method has been used to create an identification map of nanomaterials with various surface and core properties that correlates to biological activity. However, it is accepted that three-dimensional (3D) toxicity assays of drugs represent a better structural framework to reproduce the complex interactions in tissues [71, 72]. Recently, Kotov and co-workers tested nanoparticle liver toxicity using a 3D cell culture system based on polyacrylamide hydrogels and demonstrated significant differences when compared to two-dimensional (2D) assays [72]. Because the 3D toxicity assays are emerging technologies and not yet ready for industrial applications and standardization, only toxicity data based on 2D assays and in vivo studies will be discussed. 4.3.1 Quantum Dots Toxic heavy metals like cadmium, mercury, lead, and arsernic are part of the nanocrystalline core of QDs covered by biologically inert zinc sulfide. Two major factors contributing to QD toxicity include their colloidal instability and heavy metal constituents leaching out, leading to metal poisoning [67, 68, 73]. In general, divalent cations such as cadmium or gadolinium have a long half-life in humans and are toxic in the body due to their accumulation in the placenta and kidneys, possibly leading to nephrotoxicity and kidney failure. Due to the large variety of QDs and disparity in experimental conditions to assay their toxicity, such as varied concentration and exposure times, it is only possible to draw conclusions based on their stability and surface coatings. In vitro studies with QDs were used to relate cytotoxic events to the release of potentially toxic elements and the size, shape, surface, and cellular uptake of QDs. Several groups investigated the release of cadmium ions from QDs into culture media by ICP-MS or fluorimetric assays. All these studies concluded that the release of cadmium ions correlates with cytotoxicity due to oxidative stress [74–76]. Yang et al. [77] studied the concentration of cadmium in the liver and kidneys after intravenous administration of 40 pmol of cadmium-based QDs containing ZnS shell and a hydrophilic surface coating (PEG-5000, Molecular weight, MW = 5 kDa). It was observed from the ICP-MS analysis that cadmium concentration increased over the course of 28 days and reached 40% in the liver and 40% in the kidneys with respect to the injected dose [77]. These results were confirmed using fluorescence microscopic analysis showing the presence of intact QDs in both the liver and kidneys. Free cadmium may be redistributed to the kidneys via hepatic production of metallothionein, which can lead to malignant transformation of cells and ultimately induce cancer [78]. In addition, oxidative degradation of QDs by exposure to air and UV irradiation released cadmium ions, leading to in vitro and in vivo toxicity [76, 79, 80]. To protect the body from metal poisoning, QD surfaces must be capped with a passivating agent, typically zinc sulfide (ZnS), cadmium sulfide (CdS), or bovine serum albumin (BSA). These core/shell QDs were prepared in nonpolar solvents and transferred to water using
P1: OTA/XYZ P2: ABC c04 JWBS037-Chen January 25, 2011
8:49
Printer Name: Yet to Come
NANOPROBE TOXICITY
81
thiolated ligands. CdSe is used over cadmium telluride (CdTe) due to its higher stability toward oxidation. Hence these QDs were more stable and reported to be less cytotoxic. However, concentrations in the range of 10 g/mL and long-term exposure (∼24 h) to cells were found to decrease cellular metabolic activity. The critical factors contributing to the toxicity of core/shell QDs are not yet clear, but it suggests a possible role of the Cd2+ leakage and QD aggregation [81]. In addition, carboxylated QDs were synthesized in aqueous solution with core CdTe and stabilized by mercapto acids (mercaptoacetic acid or mercaptopropio acid) due to their ease of synthesis, low cost, and immediate utility in biological buffers. The cytotoxicity of these thiolate-stabilized QDs was reported to be due to aggregation and degradation [74, 82]. Oxidative release of cadmium ions from the CdSe QDs by exposure to air or ultraviolet irradiation was investigated by Bhatia and co-workers [76]. Under the same conditions, CdSe QD cores coated with small thiolated ligands were also found to be toxic. Another report showed that cadmium ions release through oxidative degradation of the QDs and bind to sulfhydryl groups on a variety of intracellular proteins, causing decreased functionality in many subcellular organelles [83]. QDs were found to induce reactive oxygen species (ROS), leading to oxidative stress, inhibition of cell proliferation and induction of cell differentiation. The ROS toxicity was cell-dependent and oxidation of intracellular DNA and protein led to apoptosis. In the presence of ultraviolet radiation, QDs were found to catalyze the formation of ROS [82, 84]. Coating QDs with a thick ZnS shell was found to reduce the ROS production [85, 86]. Aldana et al. [87] showed photo-instability led to the degradation of thiolated coatings and aggregation of 3.2-nm size QDs. More stable colloidal QDs with core–shell structures (CdSe–ZnS) were prepared by encapsulation of the QDs in amphiphilic polymers and crosslinked silica [88]. These core–shell QDs with polymer coatings were found to be optically stable and less toxic than their counterparts coated in small ligands like trioctylphosphine oxide ligand [38, 39, 88, 89]. However, high concentrations (>5 × 109 QDs/cell or >2.3 M) of these QDs were shown to cause toxic or inflammatory responses in Xenopus embryos for reasons that are not fully understood [89, 90]. Another report showed that colloidal instability causing precipitation of nanoparticles on cells led to the detachment of cells from cell culture substrates [83]. Other studies suggest that the toxicity is controlled by the size of the QDs. For example, cystamine-coated QDs (2 nm) were shown to accumulate in the nucleus. Small QDs (below 3 nm) were found to adsorb nonspecifically to intracellular proteins and impair cellular functions by invading the nucleus and binding to histones or nucleosomes, causing DNA damage [43, 91–93]. Biotinylated CdSe/ZnS QDs incubation with supercoiled DNA for 60 min showed 56% and 29% of DNA damage in the presence of UV exposure and dark, respectively [91, 93]. DNA damage was due to the free radicals generated by light and oxidation. In addition, Geys et al. [94] reported higher toxicity of negatively charged nanoparticles change to period and end sentence. They observed that negatively-charged nanoparticles caused pulmonary vascular thrombosis and suggested that the negative charge contributed to the activation of the coagulation through aggregation with plasma proteins. A current problem for in vivo applications is that a major fraction of the QDs remain at the site of injection for a very long time [95] and the long-term fate of these QDs in the body is yet to be clearly understood. Recently, Yang et al. [77] reported a long-term study (over 7 weeks) in mice showing that signal was found in the intestinal tract, suggesting fecal as the primary excretion pathway. Tissue histology and neurological evaluations did not show apparent changes 100 days post-injection. The long-term toxicity of QDs coated with crosslinked polymers does not show harmful effects at dosages of ∼10 mg/kg, but
P1: OTA/XYZ P2: ABC c04 JWBS037-Chen January 25, 2011
82
8:49
Printer Name: Yet to Come
BIODISTRIBUTION AND PHARMACOKINETICS OF NANOPROBES
distress was reported at 42 mg/kg. This dose-dependent toxicity was previously reported for QDs containing mercury, which is known to have major safety concerns. These results suggest that the potential release of cytotoxic elements from QDs and their distribution in specific organs, tissues, cell types, and subcellular locations must be well understood before they can be used in humans. Due to their enhanced optical properties, QDs might be injected at doses below their cytotoxicty threshold, although the minimum dose for precise diagnosis remains to be determined. QDs with reduced or no heavy metal toxicity and colloidal stability may find an important role as clinical contrast agents in future optical imaging techniques. 4.3.2 Gold Nanoparticles Gold is known to be an inert material and it is reported that gold nanoparticles (GNPs) (size range 10–250 nm) can be taken up by various human cells without any cytotoxic effects at up to 4000 particles per cell [96, 97]. Murphy and co-workers did not observe any in vitro cytotoxic effect using gold nanoparticles of various sizes (4, 12, and 18 nm) and surface coatings (anionic, neutral, and cationic) on leukemia cells [98, 99]. The purification process for gold nanoparticles was a key step because residual chemical impurities such as surfactants were shown to be toxic to the cells at the nanomolar level. Ofek et al. [100] showed that zebrafish embryos treated with gold nanoparticles at sizes ranging from 3 to 100 nm did not show significant increase in mortality with increasing concentrations up to 250 mol of gold in contrast to silver nanoparticles. The difference in toxicity between silver and gold nanoparticles was thought to be due to the chemical composition and residual impurities of silver nitrates [100]. Leonov et al. [101] showed that polystyrenesulfonate sodium salt (PSS, 70 kDa) could efficiently remove residual surfactant from gold nanorods. However, cytotoxicity profiles using three different cell lines (porcine kidney, human liver carcinomas, and human nasopharyngeal carcinomas) [101] showed unexpected results indicating that purified gold nanorods had higher toxicity than the surfactant-coated nanorods. These results indicated that PSS adsorption on the gold nanorods is not stable enough, leading to complexes with residual surfactants [101]. Several other studies have examined the effect of gold nanoparticle size on in vitro and in vivo toxicity [67, 102]. Hainfeld et al. [51] observed no in vivo toxicity after 30 days for 1.9-nm sized gold nanoparticles using mice as an animal model (dose = 2.7 g/kg) and the LD50 was found to be 3.2 g/kg. The nanoparticles were quickly cleared through kidneys without significant accumulation in the liver and spleen. Others have shown that gold nanoparticles with hydrodynamic diameter smaller than 2 nm are more toxic than larger particles. Maria et al. [103] showed that small gold nanoclusters (Au55 ) with a size of 1.4 nm are toxic to various human cancer and healthy cell lines in contrast to larger gold nanoparticles. Toxicity studies with 18-nm and 1.4-nm GNPs showed that larger NPs were less toxic [104], with the results attributed to the intercalation of GNPs in the DNA major groove. The smaller GNPs were also found to easily translocate in significant amounts through the blood barrier of the respiratory tract. Positively-charged, small nanoparticles were shown to increase cytotoxicity due to electrostatic interactions with cell membranes and/or DNA. No significant cytotoxicity differences have been reported due to shape when comparing spherical and nanorod gold nanoparticles. However, most of the surfactants used for stabilizing gold nanoparticles have been found to be cytotoxic. Therefore significant efforts are being devoted to using biocompatible materials. These results were consistent with data showing that gold nanorods
P1: OTA/XYZ P2: ABC c04 JWBS037-Chen January 25, 2011
8:49
Printer Name: Yet to Come
NANOPROBE TOXICITY
83
coated with polyethylene glycol were less toxic than surfactant-coated nanorods based on cellular micromotility assays. No general conclusion can be drawn at present, although most of the research has clearly shown the effect of the size on the increased cytotoxicity profile of GNPs. Thus, these results suggest that more in vivo studies need to be carried out to understand the size, charge, and shape-dependent toxicity of GNPs. 4.3.3 Carbon Nanotubes Carbon nanotubes are known to cause significant cytotoxic responses such as induction of oxidative stress, inhibition of cell proliferation, and apoptosis in different cell lines [105]. Carbon nanotubes have also demonstrated major cytotoxicity in the lungs upon inhalation [106–109]. Lam et al. [106] showed pulmonary toxicity of SWNTs in mice after intratracheal instillation. All the mice showed epithelioid granulomas and interstitial inflammation after 7 and 90 days. Similar pulmonary toxicities were observed in rats by another group [110]. The gene expression of macrophages that had taken up SWNTs showed activation of oxidative stress and an inflammatory response [111]. For example, interleukin-6 (IL-6) was overexpressed (25-fold increase) upon incubation of macrophage with SWNTs [109]. The potential hazard of SWNTs strongly depends on the metal content and the size of the agglomerates [112]. CNT products contain impurities such as amorphous carbon and metals such as, Co, Fe, Ni, and Mo. The amount and type of impurities depends on the manufacturer and synthetic methods [112, 113]. Elemental analysis of raw SWNTs (uncoated) using ICP-MS showed substantial content of iron (10%), sodium (0.03%), and nickel (0.02%). A large number of studies over the past several years reported varied toxicity levels suggesting a dependence on the type of nanotube materials used and functionalization strategies. In vitro and in vivo toxicological studies have suggested reduced toxicity due to surface functionalization of carbon nanotubes [108, 114–116]. For example, glycodendrimercoated SWNTs were found to be nontoxic to HEK292 cells at a concentration of 100 g/mL, while the non-functionalized SWNTs greatly inhibited cell growth [116]. In another report, biotinylated SWNTs were found to be nontoxic up to concentrations of 0.05 mg/mL in HL60 cells [114]. In mice, 151 and 47 mg of PEG-functionalized SWNTs was found to be nontoxic even after a 4-month period [108]. SWNTs suspended in Tween-80 exhibited toxicities due to accumulation in the liver and lungs after 3 months at higher dose (40 mg/kg), while no detectable toxicity was observed at 2 mg/kg [117]. However, SWNTs functionalized with phenyl-carboxylic groups were shown to be less cytotoxic than surfactant-coated (Pluronic F108) SWNTs. Functionalized CNTs with biocompatible surface coatings have been shown to be nontoxic with greater renal clearance and insignificant reticuloendothelial system (RES) uptake and are promising candidates for future studies [46, 50, 108, 117–119]. For example, PEGylated SWNTs administered through IV injection in mice at a concentration of 3 mg/kg showed normal blood chemistry and histological observations after 4 months [108]. This study did not show significant systemic toxicity of SWNTs after a single dose injection but white blood cell concentrations decreased by 50%. In another example, 20 g of linear and branched PEG functionalized SWNTs showed no toxic side effects, low RES uptake, and near complete clearance two months after IV injection [119]. Although no acute toxicity is reported using histological assessment of tissues, multiple reports suggested increased levels of oxidative stress. Overall, SWNT have been shown to be more toxic than MWNTs, possibly due to their geometry [67]. Poland et al. [120] showed that non-functionalized long MWNTs pose a carcinogenic risk in mice.
P1: OTA/XYZ P2: ABC c04 JWBS037-Chen January 25, 2011
84
8:49
Printer Name: Yet to Come
BIODISTRIBUTION AND PHARMACOKINETICS OF NANOPROBES
The report systematically studied various commercially available MWNTs with different lengths. Upon administration at a concentration of 100 mg/mL via peritoneal injection in mice, MWNTs demonstrated asbestos-like, length-dependent, pathogenic behavior. The longer length of MWNTs is believed to be the reason for the inflammatory response and granuloma formation within a week post-injection due to less aggregation and increased macrophage uptake. 4.3.4 Magnetic Nanoparticles The chemical toxicity of iron and its derivatives is well studied [121, 122]. Iron oxides have been shown to degrade in vivo according to natural iron pathways. Normal human tissues contain iron or iron oxides in the form of hemosiderin, ferritin, and transferrin [123]. Therefore, numerous iron oxide-based magnetic nanoparticles (MNPs) have been reported and studied for MRI applications over the last two decades [60–62]. SPIOs must be functionalized during the synthesis process for biocompatibility before medical use [59]. A variety of methods have been developed to encapsulate SPIO particles within a sheath of benign polymers such as dextran [124], polysaccharides [125], PEG, and polyethylene oxide [126]. Weissleder and co-workers reported the proof of concept profiling 50 different iron oxide based nanoprobes. The in vitro assays addressed the effect of the core composition, coating, and surface functionalization on the interactions of nanoparticles with the biological environment [70]. All the nanoparticle formulations were organized into clusters based on their biological activity and compared to Feridex as an FDA approved control [70]. The data from the four cell toxicity assays, five different cell types, and four different doses and incubation times showed that the core composition had a strong contribution to the biological effects. For example, nanoparticle formulations with carboxylic or ethylenediamine surface coatings had similar biological acitivty. The analysis of the data showed that crosslinked iron oxide (CLIO)-amine was in the same cluster as dextran-coated Feridex even though their surface coatings were significantly different. More importantly, results showed that these two formulations did not affect monocytes in the blood and spleen in contrast to the control sample that was clustered in a different group based on in vitro assays. In addition, Jain et al. [127] did not observe significant differences in the liver enzyme activity in vivo compared to the group of mice injected with saline. Overall, the histology and enzymatic activity data suggested that the oxidative stress was minor and did not affect cellular and tissue integrity. Iron oxide concentration in liver did not exceed 300 g/g of tissue during the experimental period of 3 weeks, which included the redistribution of iron through protein binding. Iron oxide concentrations were 10 times lower than the toxic concentration reported to develop cirrhosis and hepatocellular carcinoma and 4 times lower than the dose used for Feridex in humans (2.6 vs. 10 mg/kg, respectively). Many iron oxide-based MNPs are thought to have suitable toxicity profiles and are now being evaluated in clinical trials. The promising results of the Phase I clinical trial using ultrasmall superparamagnetic iron oxide colloids (USPIOs) BMS180549 showed mild to moderate side effects for 45% of the volunteers [128, 129] and the most common adverse event was urticaria. This is possibly due to the dosage of iron oxide NPs used for diagnostic imaging (1–2 mg Fe/kg bodyweight), which was less than the normal iron store and dose required to develop chronic iron toxicity. The reported iron concentration for cirrhosis and hepatocellular carcinoma is over 4 mg of Fe per gram of wet liver [63]. These SPIO formulations passed the standard toxicological and pharmacological requirements, most
P1: OTA/XYZ P2: ABC c04 JWBS037-Chen January 25, 2011
8:49
Printer Name: Yet to Come
NANOPROBE CLINICAL TRANSLATION
85
likely due to the low doses used in humans and their use as a single-dose contrast agent. However, patients injected with SPIO generated anti-dextran antibodies and developed allergic reactions. 4.4 NANOPROBE CLINICAL TRANSLATION Two SPIOs and one USPIO are currently approved by the FDA for clinical use. In 1996, Feridex I.V. became the first FDA approved nanoprobe for imaging liver lesions based on the differential amount of Kupffer cells in metastatic tumors in the liver, allowing detection of small metastases using magnetic resonance imaging (MRI). Feridex is a SPIO nanoparticle (120–180 nm) coated with dextran (MW = 10 kDa) given at a dose of 0.56 mg Fe/kg and has a circulation half-life of ∼1 hour in humans. In addition, Resovist, which is an SPIO formulation coated with carboxydextran (Molecular weight = 20 kDa) dosed at ∼2.6 mg/kg, was approved in 2001 for the European market. The advantage of Resovist is safety because it allows rapid injection without cardiovascular side effects and lumbar pain. This is due to its smaller size (∼60 nm) and increased cell internalization. In June 2009, the FDA approved the first USPIO nanoprobe, Ferumoxytol, which is used for the treatment of iron-deficiency anemia in adult patients with chronic kidney disease. Clinical data showed that patients treated with Ferumoxytol demonstrated greater hemoglobin enhancement compared to iron given orally. In addition, many USPIO magnetic nanoparticles (Table 4.1) are currently in different clinical trials as contrast agents for several different applications. In TABLE 4.1 Various Nanoprobes Under Clinical or Preclinical Investigations Trade Name
Composition
Status
Indication/Application
Size (nm)
Circulation Time (h)
Feridex I.V. (Ferumoxides AMI-25) Combidex (Ferumoxtran10, AMI-227, BMS-180549) Gastromark (Ferumoxsil AMI-121) Resovist (Ferucarbotran SHU-555A) Supravist (Ferucarbotran SHU-555C) Ferumoxytol code 7228
Dextran-coated SPIO
Approved
Liver imaging, cellular labeling
120–180
2
Dextran-coated USPIO
Phase IV
15–30
24–36
Siloxane-coated SPIO
Approved
Lymph nodes, RES directed liver diseases, macrophage imaging, cellular labeling Bowel marking, oral GI imaging
300
Oral
Carboxydextrancoated SPIO
Approved
Liver imaging, cellular labeling
60
2.4–3.6
Carboxydextrancoated SPIO
Phase III
Blood pool agent, cellular labeling
21
6
Cyclomethyl dextran-coated USPIO Citrate-coated VSSPIO Pegylated colloidal gold nanoparticles
Phase II
Macrophage imaging, blood pool agent
30
10–14
Phase II
Blood pool agent, cellular labeling Tumor therapy and tumor vasculature imaging
7
0.5–1.5
30
2–6
VSOP-C184 Aurimune (CYT-6091)
Phase I
P1: OTA/XYZ P2: ABC c04 JWBS037-Chen January 25, 2011
86
8:49
Printer Name: Yet to Come
BIODISTRIBUTION AND PHARMACOKINETICS OF NANOPROBES
general, USPIOs have longer circulation times and are mostly used for cancer metastases diagnosis. For example, Ferumoxtran-10 was shown to be a very sensitive and specific agent in human cancer trials allowing accurate diagnosis of metastatic lymph nodes with a reported circulation half-life of ∼24 h [130]. Ferumoxtran-10 is currently being tested in 8 clinical trials for tissue neoplasm (brain and bladder), aortic aneurysm, and prostate metastases diagnosis (www.clinicaltrials.gov) in Europe. Aurimune (CYT-6091) is the only system based on materials other than iron oxide that has reached clinical trials. Aurimune delivers TNF-␣ bound to PEG-coated gold nanoparticles (∼27 nm) for solid tumor therapy [131] and possible imaging abilities. TNF-␣ is a potent cytokine with antitumor cytotoxicity which requires incorporation into a nanocarrier formulation to reduce systemic toxicity. The toxicity data from the Phase I trial showed that gold nanoparticles could deliver 3 times more TNF-alpha systemically than a free dose of TNF-alpha. The TNF-alpha dose was reported to be 1.2 mg and considered to be in the therapeutic range used previously in isolated limb perfusion (ILB).
4.5 FACTORS AFFECTING THE BIODISTRIBUTION AND PHARMACOKINETICS Nanoparticle-based technology is being increasingly employed in drug delivery and imaging applications. For most applications, intravenous administration of nanoprobes is required. Compared to in vitro studies, different challenges arise in vivo due to the increased complexity of the organism. The unique physiochemical properties of nanoprobes include size, shape, chemical composition, surface chemistry, porosity, and agglomeration state. These properties control their interactions with the biological environment. In addition to toxicity, understanding the fate of nanoprobes in a complex biological environment is very important for potential clinical translation. Biodistribution and pharmacokinetic studies give a detailed knowledge about the fate of nanoprobes in the body, including circulation time, excretion pathway, degradation time, and plasma protein binding. We discuss the main factors affecting nanoprobe biodistribution and pharmacokinetics such as size, shape, surface functionality/coating, and charge. 4.5.1 Effect of Size In general, nanoparticles enter the body through the vein, skin, lungs, or gastrointestinal adsorption. Circulating nanoparticles in the body will differentially distribute/accumulate in major organs according to cell interaction properties and pathological defects [132].For example, filtration in the spleen and liver is dependent on the opsonization effect and related to the size of the nanoparticles. The endothelial barrier controls and prevents the diffusion of circulating nanoparticles out of the bloodstream. Blood vessel defects can form under pathological conditions, leading to enhanced permeability and differential accumulation of nanoparticles at specfic site in the body. This phenomenon can be taken advantage of to facilitate nanoparticle (10-500 nm) accumulation in cancer tissue due to the relatively leaky tumor vasculature and absence of a lymphatic network, resulting in the enhanced permeability and retention effect (EPR). Hence the effect of particle size on interactions with plasma components, blood cells, endothelium, and distribution in various tissues are of great interest.
P1: OTA/XYZ P2: ABC c04 JWBS037-Chen January 25, 2011
8:49
Printer Name: Yet to Come
FACTORS AFFECTING THE BIODISTRIBUTION AND PHARMACOKINETICS
87
There are only a few reports of QD biodistribution and pharmacokinetics parameters. However but it is well established that QDs accumulate significantly in the reticuloendothelial system (RES), including the liver, spleen, and the lymphatic system [135–137]. Kim et al. [39] reported a critical size of ∼15–19 nm for the retention of QDs in the sentinel lymph node during imaging of the lymphatic system. The injection of QDs of different colors at different intradermal locations allowed imaging of their drain to common nodes [138, 139]. Similarly, it was reported that QDs smaller than 9 nm could migrate into the lymphatic system with up to 5 nodes showing fluorescence [140]. QDs with size ∼9 nm could be entirely eliminated from the kidneys and directly extravasate out of blood vessels into interstitial fluid [140]. Another study found that the renal clearance of QDs was closely related to hydrodynamic diameter and the renal size threshold was defined as ∼5–6 nm [141]. The blood half-life of QDs varied from 48 min to 20 h for sizes ranging from 4.4 to 8.7 nm, respectively, demonstrating rapid urinary excretion [141]. Smaller QDs (∼5 nm) were found in detectable amounts only in the liver (4.5%) and kidneys (2.6%). Larger QDs (∼8 nm) showed higher uptake in the liver (26.5%), lungs (9.1%), and spleen (6.3%) 4 h post-intravenous injection. This was possibly due to longer circulation times allowing higher accumulation in tissues [141]. However, more studies need to be carried in vivo to determine the effects of QD size. Many reports on the biodistribution of MNPs examined the effect of particle size [142–144]. In vivo biodistribution data for five neutral particles with different sizes ranging from 30 to 90 nm was reported. [142] and the results showed that smaller nanoparticles (<50 nm) have longer blood circulation time (>10 min for 50% clearance) and less uptake by the liver (∼20% for small particles and >50% for larger particles). After 20 min, it was found that the liver uptake increased as the size of the NPs increased. Pharmacokinetic profiles were studied with nanoparticles of three different sizes ranging from 46 to 75 nm to determine the clearance time. It was found that increasing the particle size decreased the plasma half-life, with 46-nm particles showing 50% clearance from the bloodstream in the first 10 min in contrast to only 5 min for 90-nm particles [142]. Neuberger et al. [143] reported that 200-nm MNPs were cleared faster than MNPs of sizes 30–100 nm. It was also demonstrated that for iron oxide particles smaller than 40 nm in diameter, both biodistribution and blood half-life were mostly controlled by the coating material rather than the mean particle size [11, 145]. For example, circulation half-life of two magnetic nanoparticles with the same polymeric surface coating, Ferumoxtran-10, had a longer circulation time compared to Feridex due to its smaller size (∼15–30 nm vs. ∼120 nm, respectively). The slower clearance of Ferumoxtran-10 was believed to be due to larger particle size and/or surface charge affecting degradation and biodistribution. Blood half-lives of magnetic nanoparticles in human are reported to be up to 2 days and USPIOs were found to have the longest circulation time partially due to their smaller size and lower accumulation in the liver. Functionalizing the surface of MNPs with targeting ligands can significantly affect the circulation half-life and biodistribution. Reddy et al. [146] found that MNPs of 40 nm showed increased tumor contrast half-life in mice from 39 min to 123 min by attaching a targeting F3 ligand. Montet et al. [147] showed that the MNP–RGD conjugates with sizes 28 and 36 nm were found to have blood half-lives of 180 and 207 min in rats. Biodistribution studies 24 h after IV administration showed accumulations in the liver, spleen, skin, kidneys heart, and tumor [147]. According to the reported data, the overall size of MNPs should be smaller than 200 nm to evade rapid splenic filtration [148], but larger than ∼5 nm to avoid renal clearance [141].
P1: OTA/XYZ P2: ABC c04 JWBS037-Chen January 25, 2011
88
8:49
Printer Name: Yet to Come
BIODISTRIBUTION AND PHARMACOKINETICS OF NANOPROBES
Carbon nanotubes have been shown to penetrate into the lungs and cross the blood–brain barrier, leading to neurotoxicity. The first study of CNTs in animals to compare tissue distribution characteristics after subcutaneous, oral, and intravenous administration was performed using iodine-labeled hydroxyl-functionalized carbon nanotubes (125 I-CNT) with a length of 300 nm. It was found that CNT accumulated in the stomach, kidneys, and bone similar to small molecules. The results showed no tissue damage and excretion through urine [149]. Later, this same group reported that the functionalized CNTs showed persistent liver accumulation after IV injection [150]. Additional biodistribution data using indiumlabeled carbon nanotubes (111 I-CNT) also showed accumulation in kidneys, bone, muscle, and skin. The circulation half-life was found to be ∼3 h with rapid clearance through urine. In a recent report, yttrium-labeled carbon nanotubes (86 Y-CNT) with 42 ± 17 nm length were shown to be cleared from the bloodstream within 3 h when injected intravenously in mice [151]. The PET images indicated that the major sites of accumulation of 86 Y-CNT were the liver (18%), spleen (14%), kidneys (8%), and to a less extent bone (2%). The uptake of CNTs in the kidneys, liver, and spleen was slightly lower when the CNTs were administered via intraperitoneal (IP) injection, suggesting a relatively slow egress from the intraperitoneal compartment into the vascular compartment. It was observed that the clearance of CNTs from the kidneys was much faster than from the spleen and liver. Elgrabli et al. [152] used commercial uncoated MWNTs with a diameter of 20–50 nm and length of 500–2000 nm to study the biodistribution in rats. The results showed CNT bio-persistence and clearance 6 months after respiratory administration. The MWNTs were intratracheally instillated and 53% of the initial dose accumulated in the lungs after 24 h and 16% of MWCNTs were still found in the lungs after 6 months. However, most of the biodistribution studies were carried out using a radiolabel tracking method. Therefore, the dissociation of the label could affect the measurements. There are few reports describing the size-dependent cellular uptake and biodistribution of gold NPs over the last decade. Chithrani et al. [96] reported in vitro cellular uptake of GNPs with different sizes in HeLa cells. It was found from the uptake kinetic studies that the uptake half-lives of 14-, 50-, and 74-nm colloidal gold nanoparticles were 2.1, 1.9, and 2.24 h, respectively. De Jong and co-workers studied size-dependent (10-, 50-, 100-, and 250-nm) biodistribution of spherical gold nanoparticles after intravenous administration in rats [97]. A clear difference was observed between the biodistribution of the 10-nm particles and the larger NPs. Small gold nanoparticles (10 nm) were found in various organs including the blood, kidneys, testes, thymus, heart, lungs, and brain, whereas larger particles (>10 nm) were only detected in the blood, liver, and spleen after 24 h [97]. However, Wolfgang and co-workers studied the biodistribution of gold NPs with 1.4- and 18-nm sizes in rats after intravenous injection and found no clear differences between the nanoparticle sizes tested [104]. Similar accumulation and biodistribution of the NPs was observed in the liver, blood, carcass, spleen, skin, kidneys, urine, and feces 24 h post-injection. Intratracheal administration of the same nanoparticles using rats as an animal model showed 99.8% and 91.5% accumulation in the lungs after 24 h for 18-nm and 1.4-nm GNPs, respectively. Small amounts of the instilled 1.4-nm NPs (0.6–3.3% I.D.) were detected in the blood, urine, skin, and carcass, whereas no detectable amounts of larger nanoparticles (18 nm) were found [104]. Hillyer and Albrecht [153] showed in mice that metallic colloidal gold nanoparticles of different sizes (4, 10, 28, and 58 nm) distributed to other organs after oral administration in mice. The 4-nm GNPs showed differential accumulation in the kidneys, liver, spleen, lungs, and even in the brain in contrast to the 58-nm particles detected solely inside the gastrointestinal tract. Hainfeld et al. [51] subcutaneously injected 1.9-nm GNPs
P1: OTA/XYZ P2: ABC c04 JWBS037-Chen January 25, 2011
8:49
Printer Name: Yet to Come
FACTORS AFFECTING THE BIODISTRIBUTION AND PHARMACOKINETICS
89
into mice and examined their long-term distribution (11 and 30 days) in various organs. The biodistribution of the gold nanoparticles in an EMT-6 syngeneic mammary carcinoma xenograft mouse model showed very low retention of GNPs in the liver and spleen due to elimination by the kidneys after 24 h. Pharmacokinetic studies revealed a biphasic clearance, with a drop of 50% between 2 min and 10 min post-injection followed by a slower decrease of another 50% between 15 min and 1.4 h. Accumulation of the NPs in the kidneys (10.6% I.D.), tumor (4.2% I.D.), liver (3.6% I.D.), and muscle (1.2% I.D.) was evaluated 15 min post-injection. In another report, Vijaya Kattumuri et al. [154] used 15–20-nm size gum arabic stabilized GNPs to study biodistribution in larger animals such as pigs. Following intravenous administration of the GNPs (0.8 mg/kg), identical amounts (∼50 ppm) of Au were observed in the lungs and liver after 24 h, while 40 ppm of gold was measured in the spleen. After 72 h, 43–69% of the administered dose was retained in the liver while only 1–2% was observed in the kidneys, indicating slow clearance of GNPs. 4.5.2 Effect of Shape Most NPs exhibit a spherical shape as a result of surface energy minimization during their synthesis. The advent of mimicking nature with nanoparticles of different shapes such as a virus or bacterium is now possible due to the development of new fabrication processes. The ability to make particles with shapes other than spheres is opening a path to new design solutions for systemically administered particulates. Decuzzi et al. [155] reported that the intravascular journey of the particle can be broken down into three events: (1) margination dynamics, (2) firm adhesion, and (3) control of internalization. This report compared the predictions of mathematical models and showed that the particle geometry plays an important role in all three events [155]. In addition, recent studies performed using polystyrene nanoparticles of different shapes showed that the phagocytosis pathway by macrophages exhibits a strong dependence on shape [156, 157]. A recent report used a physiologically based pharmacokinetic (PBPK) model along with experimental data of biodistribution of QDs with various shapes, sizes, and exposure routes, to predict the biodistribution of other QDs [158]. The model estimated partition coefficients for various tissue concentrations of QDs in the blood, kidneys, liver, muscle, and skin using the experimental data from commercial QD705. The ellipsoid particles showed increases in the steady-state concentration of QDs in the liver and kidneys, similar to experimental observations. This model considered that the QDs initially distributed to the kidneys, then redistributed to other compartments with higher partition coefficients but much slower flow rates per mass of tissue as these compartments approach steady-state concentrations. Gold nanoparticles can also be made in different shapes such as rods and pyramids with an aspect ratio of 1:1 to 1:5 [96]. The cellular uptake in HeLa cells was studied with various shapes and found that spherical particles have a higher probability of uptake than rod-shaped nanoparticles with aspect ratios 3 and 5. Niidome et al. [159] studied the biodistribution of PEG(5000)-functionalized gold nanorods (65-nm length and 11-nm width) and surfactant-stabilized gold nanorods in mice. The results showed that the pegylated gold nanorods had a blood circulation time of 30 min (∼50% I.D.), while the residual CTABfunctionalized nanorods in blood were significantly lower (∼5%) [159]. No detectable amounts of gold were present in the blood, while 35% was accumulated in the liver and very small amounts were in the lungs, spleen, and kidneys, showing the clearance of gold nanorods after 72 h. In contrast, CNTs, which are known to have high aspect ratios, were shown to have a rapid blood clearance (circulation half-life = 1 h) in rabbits due to their
P1: OTA/XYZ P2: ABC c04 JWBS037-Chen January 25, 2011
90
8:49
Printer Name: Yet to Come
BIODISTRIBUTION AND PHARMACOKINETICS OF NANOPROBES
rapid renal excretion and accumulation in the liver [49]. Another study showed a blood circulation half-life of 3 h for the water-soluble DTPA-functionalized SWNTs in mice after IV administration [160]. 4.5.3 Effect of Surface Coating and Charge Other than size and shape, the nanoparticle surface coating and charge can modulate the in vivo fate of nanoprobes. NP coatings affect surface properties such as surface charge (zeta potential), surface functionalities, and degree of hydrophilicity. Surface modification of QDs is necessary due to their low solubility in polar solvents. There are many reports addressing the surface modification of QDs with a variety of coatings such as small molecules or encapsulation in amphiphilic polymers to increase surface hydrophilicity, reduce nonspecific binding, and prolong blood circulation [38, 89, 161, 88, 162]. QD structure and surface properties have been found to strongly impact plasma halflife. The half-life of anionic, carboxylated QDs in the bloodstream of mice was significantly increased from 4.6 min to 71 min by coating QDs with PEG polymer chains of MW 5000 [135]. Schipper et al. [163] evaluated the biodistribution of commercially available PEGcoated CdSe QD525 and QD800 in mice. Both types of QDs rapidly accumulated in the liver within 6 min postinjection. In mice, PEG-coated CdSe/ZnS QDs were rapidly removed from the bloodstream into organs of the RES with detection possible for up to 4 months post-injection using fluorescence imaging [135]. TEM studies on the tissues revealed that these QDs retained their morphology, suggesting QDs are stable in vivo given the proper coating. The blood circulation half-life of these QDs was significantly increased by coating QDs with poly(acrylic) or a short PEG(750), while the half-life increased to 72 min with an increase in the length of PEG(5000) as determined from venipuncture studies [135]. Multiple approaches have been used to attach ligands to the surface. For example, heterofunctional ligands have been attached to the surface of QDs with the hydrophilic group exposed [164, 165]. In addition, biological molecules such as monoclonal antibodies (J591) or EGF receptor (erbB1) have been attached to the surface, serving the dual purpose of increasing surface hydrophilicity and acting as a targeting ligand [161, 166, 167]. Fc fragments of antibodies conjugated to QDs were reported to have longer circulation times due to the reduction of nonspecific interactions [168]. Another report demonstrated slower removal of albumin-coated QDs from the bloodstream before accumulation in the liver compared to QDs without albumin [137]. Surface charge will lead to different mechanisms of interaction with macrophages and may affect the intracellular fate of internalized NPs. Even though cationic NPs show increased cell internalization, supramagnetic particles with negative surface charge were found to exhibit a high but nonspecific affinity for the plasma membrane, favoring adsorption and endocytosis in endosomal compartments of HeLa cells [169]. In another report, rpositively-charged MNPs were prepared by the coating of monocrystalline iron oxide nanoparticles(MIONs) with cationic poly-l-lysine. The blood half-life was found to be only 1–2 min in comparison to 2–3 h for the uncharged variant [170]. Small neutral MION particles distributed to lymph nodes, whereas the positively-charged particles of similar size were found to be rapidly taken up by the liver [63]. Jallet and co-workers carried out a systematic study to analyze the effect of charge on MNPs as contrast agents for MRI [142]. When the surface charge of these nanoparticles was varied from neutral to negative, there was a slight decrease in the particle size while the change to a positive charge led to an increase in the size of the MNPs. The biodistribution results showed that charged
P1: OTA/XYZ P2: ABC c04 JWBS037-Chen January 25, 2011
8:49
Printer Name: Yet to Come
FACTORS AFFECTING THE BIODISTRIBUTION AND PHARMACOKINETICS
91
particles had greater accumulation in the liver relative to neutral particles. Twenty minutes after injection of neutral MNPs into the bloodstream via intravenous injection, 40% of the NPs remained in the bloodstream. In contrast, only 10% of the negatively charged particles remained in the blood. Uptake of positively charged particles (∼60%) was higher than neutral particles (40%), but lower than the negatively charged MNPs (∼75%) in the liver. The uptake of MNPs in the liver was almost 3 times higher (∼75%) for negatively charged MNPs compared to neutral MNPs [142]. MNPs of 100 nm showed 20% more uptake when the surface charge was positive compared to neutral [142]. Citrate-coated iron oxide NPs (8.7 nm) had a blood half-life of 1 h in humans due to their highly anionic surface charge, while a commercially available Ferumoxtran-10 had a blood half-life of 24–36 h in humans and 2–3 h in rats [11]. Similarly detectable amounts of charged particles were observed in the carcass (bones, skin, muscles, and the whole head). Overall, this study concluded that the negative charge on the particle surface resulted in enhanced liver uptake [142], while the MNPs with neutral surfaces were taken up the least by the liver. More importantly, coated magnetic nanoparticles remained in the liver for a few months with sustained contrast properties and were shown to be mainly eliminated through feces in human studies. CNTs were functionalized with various groups on the surface to improve water solubility. Yuliang and co-workers used hydroxyl-functionalized, iodine-labeled carbon nanotubes (125 I-CNT), which showed blood clearance and accumulation in the stomach, kidneys, and bones [149]. Amine-functionalized CNTs with indium labeling cleared the blood compartment [160] and all tissues within 3 h except for the kidneys, liver, and spleen. Rapid clearance of indium-labeled CNTs and significant accumulation in the kidneys, liver, spleen, and, to a small extent, bone was reported within 1 h after administration. Wenxin and co-workers reported water-soluble MWNTs by surface functionalization with glucosamine (MWNT-G) [171]. MWNT-G as injected into mice by intraperitoneal injection quickly spread to measurable levels in the blood, heart, lungs, liver, spleen, kidneys, stomach, intestines, coat, muscle, enterogastric area, and feces, behaving like a small molecule. A significant amount of total activity was retained throughout the 24-h study, particularly in the stomach. Samples collected from urine and feces showed >70% activity after 24 h, indicating MWNT-G was excreted predominantly via urine and feces. The pharmacokinetic profiles indicated a blood circulation half-life of about 5.5 h. Another study by Liu et al. [172] showed PEG-wrapped SWNTs can escape the RES for a blood circulation half-life up to 2 h. A more recent study revealed the increase in blood circulation time from 2 to 15 h when branched (MW = 7 kDa) PEG is used instead of linear (MW ranging from 2 Da to 12 kDa, methoxy end group) [119]. When the branched PEG is covalently bound to SWNTs, the blood circulation time is prolonged to 15.3 h and a low hepatic uptake was observed in the biodistribution study [118]. Most of the SWNTs were shown to accumulate in the spleen and liver in contrast to other studies using a radiolabeled tracking method. PEG-SWNTs were found to be distributed throughout most of the organs within 1 h and there were still considerable amounts of PEG-SWNTs present in the liver and spleen after 50 days. However, this study confirmed earlier reports of feces as the major excretion path. McDevitt et al. [173] showed significant difference in biodistribution and pharmacokinetics for targeted and nontargeted CNTs. The targeted CNTs showed more accumulation in the spleen and liver while the nontargeted CNTs showed more accumulation in the kidneys. Gold nanoparticles tend to agglomerate in a few hours without proper stabilizers or coatings. Typical stabilizers are surfactants, organic carboxylates, long-chain amines or thiols, sugars, carbohydrates, and proteins [154]. Coating gold nanoparticles with gum
P1: OTA/XYZ P2: ABC c04 JWBS037-Chen January 25, 2011
92
8:49
Printer Name: Yet to Come
BIODISTRIBUTION AND PHARMACOKINETICS OF NANOPROBES
arabic, a hydroxyproline-rich arabinogalactan, has been found to have enhanced kinetic inertness and high in vivo stability. These nanoparticles were found in the lungs, liver, and spleen within 30 min [154] and significant amounts of NPs were present in the liver 72 h after intravenous administration. Another report showed negatively charged GNPs obtained by coating with water-soluble sulfonated phosphane accumulated in the lungs (>90%) 24 h after intratracheal instillation, while 47% were found in the liver when administered via intravenous injection [104]. However, more systematic studies of biodistribution of gold nanoparticles with variations in surface charge are needed before they are tested in the clinic.
4.6 THERAGNOSTIC NANOPROBES Nanoparticle delivery systems are becoming increasingly recognized as a potential therapeutic vehicle with multiple examples approved by the FDA for therapy. The similarities to nanoprobes have led to the design of multifunctional nanoparticle systems able to perform imaging and therapy simultaneously. In general, the therapeutic agent adsorbed on the surface of the nanoprobe is directly exposed to the biological environment and delivered at a dose controlled by the surface loading, which is reported to be in the range of 10–100 molecules in most cases. There are two promising therapeutic approaches (Fig. 4.2) that have been combined with imaging. The first approach is based on the combination of imaging and hyperthermia. The second approach is based on the combination of imaging and delivery of cytotoxic drugs. We discuss both strategies for the combination of therapy and imaging. Currently, Aurimune (CYT-6091) is the only multifunctional nanoparticle with imaging and delivery properties that has reached the clinic. CytImmune Sciences Inc. developed a nanoparticle system with the ability to bind TNF-␣ on the surface of gold nanoparticles through electrostatic interactions. Preliminary SEM micrographs of nanoparticles accumulated in breast tumor tissue sections in contrast to healthy tissues showed targeting of the nanoparticles by the EPR effect. Other formulations are still in the discovery stage using combinations of drugs such as TNF-␣ with paclitaxel, doxorubicin, or interleukin–12. Although the load of therapeutic agent is reported to be several hundreds of molecules,
FIGURE 4.2 Various promising nanoprobes used for theragnostics (see Refs. 175–183).
P1: OTA/XYZ P2: ABC c04 JWBS037-Chen January 25, 2011
8:49
Printer Name: Yet to Come
THERAGNOSTIC NANOPROBES
93
recent data from a Phase I clinical trial reported that gold nanoparticles could deliver a therapeutic dose (www.cytimmune.com). Most other theragnostic nanoprobes are in early discovery/development stages, but the early clinical results are clearly demonstrating the potential opportunities for the development of theragnostic agents. Gold nanoprobes have been used mostly as a hyperthermia therapeutic agent by heating the tumor through the application of a magnetic field. The local increase in temperature kills the surrounding cells. In vivo results [174] of nanoshell-mediated NIR (near-infrared) thermal therapy showed that the nanoparticles induced irreversible cancer tissue damage at a temperature ∼40 ◦ C in a human breast cancer xenograft. However, the distribution of nanoparticle in the tumor might significantly affect the control of the local temperature. Similarly, Schwartz et al. [175] reported efficient therapy in large animals for brain cancer. The results showed that the temperature reached ∼70 ◦ C in tumor tissues and ∼50 ◦ C in normal white and grey matter, which is expected to significantly damage nondiseased areas of the brain. In addition, Ross and co-workers reported targeted iron oxide based nanoparticles as a delivery system (∼40 nm) of photofrin, which can be activated. . . with wavelength light (∼630 nm) to treat brain cancer [146]. Light irradiation activates the photosensitizer to produce a single oxygen leading to apoptotic and necrotic cytotoxicity. Photofrin loading on the nanoparticles was ∼4% w/w and was administered intravenously at a dose of 7 mg/kg using a rat 9L tumor xenograft model implanted into the animal brain. Tumor size was significantly reduced and the 50% survival time was extended to at least double that observed in the nontargeted nanoparticle control group [146]. MRI was used to monitor changes in the tumor and the nanoparticles showed magnetic resonance contrast of the targeted tumor and therapeutic efficacy of photofrin [146]. Similarly, Weissleder and coworkers have reported a macrophage-targeted theragnostic system based on photodynamic therapy for cardiovascular diseases [176, 177]. A potent sensitizer, 5-(4-carboxyphenyl)10,15,20-tryphenyl-2,3-dihydroxychlorin (TPC), was covalently attached to the surface of the iron oxide nanoparticle. The load was estimated to be three photosensitizer molecules per nanoparticle and the results showed efficient killing of macrophages [177]. Others have combined imaging and therapeutic functions using drug–material interactions as a loading strategy. For example, Farokhzad and co-workers reported the loading of doxorubicin, a potent chemotherapeutic agent, through intercalation between base pairs of aptamers conjugated to the surface of quantum dots [178]. In this case, aptamers were used as a specific targeting ligand to cancer cells and molecular carrier of the chemotherapeutic drug [178]. This system was precisely designed to have doxorubicin loaded on the surface quenching the quantum dot fluorescence. Therefore, the quantum dot lit up only when doxorubicin was released, allowing simultaneous delivery of the chemotherapeutic drug into specific cells and monitoring of the delivery using fluorescent imaging [178]. Adair’s group [179, 180] reported intracellular imaging and delivery using biocompatible calcium phosphate nanoparticles (∼27 nm). This system provided controlled drug delivery through pH dissolution of calcium phosphate in the acidic tumor environment. In vitro studies showed high uptake of the nanoparticles in bovine aortic endothelial cells and efficient inhibition of human vascular smooth cells using hexanoyl-ceramide (Cer-6) as a model drug. This technology is now being developed by Keystone Nano for imaging and delivery of therapeutic agents. Recently, Perez and co-workers used poly(acrylic acid) (PAA)-coated iron oxide nanoparticles loaded with paclitaxel, a potent chemotherapeutic agent, and an NIR fluorescent dye to combine optical imaging, magnetic resonance imaging, and drug delivery [181]. The nanoparticle was functionalized with a targeting
P1: OTA/XYZ P2: ABC c04 JWBS037-Chen January 25, 2011
94
8:49
Printer Name: Yet to Come
BIODISTRIBUTION AND PHARMACOKINETICS OF NANOPROBES
ligand to specifically bind cancer cells. Nanoparticle size was characterized to be ∼90 nm and taxol loading was 11 molecules per nanoparticle. Taxol release kinetics were accelerated under esterase and acidic pH -mediated degradation. In addition, the results showed multimodal imaging capabilities using NIR and MRI in vitro. Other groups have reported quantum dots delivering nucleic base therapeutics, although endosomal escape and unpacking still remain a challenge. Recently, Gao and co-workers reported the surface modification of quantum dots with tertiary amine groups for the intracellular delivery of siRNA [182]. The results showed uptake and distribution of quantum dots in the cell cytoplasm. Gene silencing efficiency using formulations suspended with and without serum was better than lipofectamine and no significant toxicity of the vehicle was reported. The quantum dot gene silencing efficiency is reported to be dependent on the ratio of carboxylic group/tertiary amine groups, which control endosomal escape. Quantum dots have also been developed to monitor RNAi delivery and improve gene silencing using cationic liposome [183].
4.7 CONCLUSION Early detection is a critical component of improving patient survival in many diseases. The development of new imaging systems and imaging agents has significantly improved the ability to detect disease earlier in its progression. Nanoprobes are an emerging technology that has the potential to increase the sensitivity and specificity of imaging systems. Early designs of nanoprobes have shown great promise, but there are significant concerns over the toxicity of these probes since they are nanomaterials composed of metallic or inorganic components. Despite toxicity concerns, nanoprobes have already reached the clinic for the diagnosis of liver lesions and cancer metastases in lymph nodes. As synthesis and modification techniques improve and a greater understanding of interactions between these nanomaterials and tissues in the body is reached, the capability to design safer nanoprobes should be possible. The ability to target nanoprobes should allow their clinical applications to expand from liver lesions to other metastatic sites, such as the brain and bones. Another promising application of nanoprobes is as theragnostic agents, combining imaging and therapeutic modalities. These systems have the potential to simultaneously image and treat disease, as well as exploit synergistic cytotoxicity by combining approaches such as hyperthermia therapy with chemotherapeutic agents. Advanced designs of nanoprobes and theragnostic agents are anticipated to significantly improve patient care and compliance. The multidisciplinary approach of combining chemistry and bioengineering with toxicology and clinical research has the potential to lead to novel agents with advanced, multifunctional properties.
ACKNOWLEDGEMENT The authors acknowledge funding from National Institutes of Health Grants CA119349 and EB003647, and the David-Koch-Prostate Cancer Foundation Award in Nanotherapeutics. EMP is supported by a National Defense Science and Engineering Graduate Fellowship (NDSEG). OCF has a financial interest in BIND Biosciences and Selecta Biosciences.
P1: OTA/XYZ P2: ABC c04 JWBS037-Chen January 25, 2011
8:49
Printer Name: Yet to Come
REFERENCES
95
REFERENCES 1. Smith, A. M.; Gao, X.; Nie, S. Quantum dot nanocrystals for in vivo molecular and cellular imaging. Photochem. Photobiol. 2004, 80, 377–385. 2. Medintz, I. L.; Uyeda, H. T.; Goldman, E. R.; Mattoussi, H. Quantum dot bioconjugates for imaging, labelling and sensing. Nat. Mater. 2005, 4, 435–446. 3. Pinaud, F.; Michalet, X.; Bentolila, L. A.; Tsay, J. M.; Doose, S.; Li, J. J.; Iyer, G.; Weiss, S. Advances in fluorescence imaging with quantum dot bio-probes. Biomaterials 2006, 27, 1679–1687. 4. Ballou, B. Quantum dot surfaces for use in vivo and in vitro. Curr. Top. Dev. Biol. 2005, 70, 103–120. 5. Kaji, N.; Tokeshi, M.; Baba, Y. Single-molecule measurements with a single quantum dot. Chem. Rec. (N.Y.) 2007, 7, 295–304. 6. Gao, X.; Cui, Y.; Levenson, R. M.; Chung, L. W.; Nie, S. In vivo cancer targeting and imaging with semiconductor quantum dots. Nat. Biotechnol. 2004, 22, 969–976. 7. Osaka, T.; Matsunaga, T.; Nakanishi, T.; Arakaki, A.; Niwa, D.; Iida, H. Synthesis of magnetic nanoparticles and their application to bioassays. Anal. Bioanal. Chem. 2006, 384, 593– 600. 8. Lu, A. H.; Salabas, E. L.; Schuth, F. Magnetic nanoparticles: synthesis, protection, functionalization, and application. Angew. Chem. Int. Ed. 2007, 46, 1222–1244. 9. Gupta, A. K.; Naregalkar, R. R.; Vaidya, V. D.; Gupta, M. Recent advances on surface engineering of magnetic iron oxide nanoparticles and their biomedical applications. Nanomedicine (London) 2007, 2, 23–39. 10. Ito, A.; Shinkai, M.; Honda, H.; Kobayashi, T. Medical application of functionalized magnetic nanoparticles. J. Biosci. Bioeng. 2005, 100, 1–11. 11. Corot, C.; Robert, P.; Idee, J. M.; Port, M. Recent advances in iron oxide nanocrystal technology for medical imaging. Adv. Drug Deliv. Rev. 2006, 58, 1471–1504. 12. de Jonge, N.; Bonard, J. M. Carbon nanotube electron sources and applications. Philos. Trans. 2004, 362, 2239–2266. 13. Hafner, J. H.; Cheung, C. L.; Woolley, A. T.; Lieber, C. M. Structural and functional imaging with carbon nanotube AFM probes. Prog. Biophys. Mol. Biol. 2001, 77, 73–110. 14. Boisselier, E.; Astruc, D. Gold nanoparticles in nanomedicine: preparations, imaging, diagnostics, therapies and toxicity. Chem. Soc. Rev. 2009, 38, 1759–1782. 15. Baptista, P.; Pereira, E.; Eaton, P.; Doria, G.; Miranda, A.; Gomes, I.; Quaresma, P.; Franco, R. Gold nanoparticles for the development of clinical diagnosis methods. Anal. Bioanal. Chem. 2008, 391, 943–950. 16. Daniel, M. C.; Astruc, D. Gold nanoparticles: assembly, supramolecular chemistry, quantumsize-related properties, and applications toward biology, catalysis, and nanotechnology. Chem. Rev. 2004, 104, 293–346. 17. Huang, X.; Jain, P. K.; El-Sayed, I. H.; El-Sayed, M. A. Gold nanoparticles: interesting optical properties and recent applications in cancer diagnostics and therapy. Nanomedicine (London) 2007, 2, 681–693. 18. Jennings, T.; Strouse, G. Past, present, and future of gold nanoparticles. Adv. Exp. Med. Biol. 2007, 620, 34–47. 19. Sperling, R. A.; Rivera Gil, P.; Zhang, F.; Zanella, M.; Parak, W. J. Biological applications of gold nanoparticles. Chem. Soc. Rev. 2008, 37, 1896–1908. 20. Sharma, P.; Brown, S.; Walter, G.; Santra, S.; Moudgil, B. Nanoparticles for bioimaging. Adv. Colloid Interface Sci. 2006, 123–126, 471–485.
P1: OTA/XYZ P2: ABC c04 JWBS037-Chen January 25, 2011
96
8:49
Printer Name: Yet to Come
BIODISTRIBUTION AND PHARMACOKINETICS OF NANOPROBES
21. McCarthy, J. R.; Weissleder, R. Multifunctional magnetic nanoparticles for targeted imaging and therapy. Adv. Drug Deliv. Rev. 2008, 60, 1241–1251. 22. Shubayev, V. I.; Pisanic, T. R., 2nd; Jin, S. Magnetic nanoparticles for theragnostics. Adv. Drug Deliv. Rev. 2009, 61, 467–477. 23. Pene, F.; Courtine, E.; Cariou, A.; Mira, J. P. Toward theragnostics. Crit. Care Med. 2009, 37, S50–S58. 24. Liu, Y.; Miyoshi, H.; Nakamura, M. Nanomedicine for drug delivery and imaging: a promising avenue for cancer therapy and diagnosis using targeted functional nanoparticles. Int. J. Cancer 2007, 120, 2527–2537. 25. Pauwels, E. K.; Erba, P. Towards the use of nanoparticles in cancer therapy and imaging. Drug News & Perspect. 2007, 20, 213–220. 26. Sun, C.; Lee, J. S.; Zhang, M. Magnetic nanoparticles in MR imaging and drug delivery. Adv. Drug Deliv. Rev. 2008, 60, 1252–1265. 27. Namdeo, M.; Saxena, S.; Tankhiwale, R.; Bajpai, M.; Mohan, Y. M.; Bajpai, S. K. Magnetic nanoparticles for drug delivery applications. J. Nanosci. Nanotech. 2008, 8, 3247–3271. 28. Han, G.; Ghosh, P.; Rotello, V. M. Functionalized gold nanoparticles for drug delivery. Nanomedicine (London) 2007, 2, 113–123. 29. Han, G.; Ghosh, P.; Rotello, V. M. Multi-functional gold nanoparticles for drug delivery. Adv. Exp. Med. Biol. 2007, 620, 48–56. 30. Ozdemir, V.; Williams-Jones, B.; Glatt, S. J.; Tsuang, M. T.; Lohr, J. B.; Reist, C. Shifting emphasis from pharmacogenomics to theragnostics. Nat. Biotechnol. 2006, 24, 942–946. 31. Weissleder, R.; Pittet, M. J. Imaging in the era of molecular oncology. Nature 2008, 452, 580–589. 32. Kherlopian, A.; Song, T.; Duan, Q.; Neimark, M.; Po, M.; Gohagan, J.; Laine, A. A review of imaging techniques for systems biology. BMC Systems Biol. 2008, 2, 74. 33. Douma, K.; Prinzen, L.; Slaaf, D. W.; Reutelingsperger, C. P.; Biessen, E. A.; Hackeng, T. M.; Post, M. J.; van Zandvoort, M. A. Nanoparticles for optical molecular imaging of atherosclerosis. Small (Germany) 2009, 5, 544–557. 34. Weissleder, R. Scaling down imaging: molecular mapping of cancer in mice. Nat. Rev. Cancer 2002, 2, 11–19. 35. Jaffer, F. A.; Weissleder, R. Molecular imaging in the clinical arena. JAMA 2005, 293, 855–862. 36. Weissleder, R. A clearer vision for in vivo imaging. Nat. Biotechnol. 2001, 19, 316–317. 37. Lim, Y. T.; Kim, S.; Nakayama, A.; Stott, N. E.; Bawendi, M. G.; Frangioni, J. V. Selection of quantum dot wavelengths for biomedical assays and imaging. Mol. Imaging 2003, 2, 50–64. 38. Gao, X.; Cui, Y.; Levenson, R. M.; Chung, L. W.; Nie, S. In vivo cancer targeting and imaging with semiconductor quantum dots. Nat. Biotechnol. 2004, 22, 969–976. 39. Kim, S.; Lim, Y. T.; Soltesz, E. G.; De Grand, A. M.; Lee, J.; Nakayama, A.; Parker, J. A.; Mihaljevic, T.; Laurence, R. G.; Dor, D. M.; Cohn, L. H.; Bawendi, M. G.; Frangioni, J. V. Nearinfrared fluorescent type II quantum dots for sentinel lymph node mapping. Nat. Biotechnol. 2004, 22, 93–97. 40. Smith, A. M.; Duan, H.; Mohs, A. M.; Nie, S. Bioconjugated quantum dots for in vivo molecular and cellular imaging. Adv. Drug Deliv. Rev. 2008, 60, 1226–1240. 41. Smith, J. D.; Fisher, G. W.; Waggoner, A. S.; Campbell, P. G. The use of quantum dots for analysis of chick CAM vasculature. Microvasc. Res. 2007, 73, 75–83. 42. Larson, D. R.; Zipfel, W. R.; Williams, R. M.; Clark, S. W.; Bruchez, M. P.; Wise, F. W.; Webb, W. W. Water-soluble quantum dots for multiphoton fluorescence imaging in vivo. Science 2003, 300, 1434–1437.
P1: OTA/XYZ P2: ABC c04 JWBS037-Chen January 25, 2011
8:49
Printer Name: Yet to Come
REFERENCES
97
43. Nabiev, I.; Mitchell, S.; Davies, A.; Williams, Y.; Kelleher, D.; Moore, R.; Gun’ko, Y. K.; Byrne, S.; Rakovich, Y. P.; Donegan, J. F.; Sukhanova, A.; Conroy, J.; Cottell, D.; Gaponik, N.; Rogach, A.; Volkov, Y. Nonfunctionalized nanocrystals can exploit a cell’s active transport machinery delivering them to specific nuclear and cytoplasmic compartments. Nano Lett. 2007, 7, 3452–3461. 44. Diagaradjane, P.; Orenstein-Cardona, J. M.; Colon-Casasnovas, N. E.; Deorukhkar, A.; Shentu, S.; Kuno, N.; Schwartz, D. L.; Gelovani, J. G.; Krishnan, S. Imaging epidermal growth factor receptor expression in vivo: pharmacokinetic and biodistribution characterization of a bioconjugated quantum dot nanoprobe. Clin. Cancer Res. 2008, 14, 731–741. 45. Tans, S. J.; Devoret, M. H.; Dai, H.; Thess, A.; Smalley, R. E.; Geerligs, L. J.; Dekker, C. Individual single-wall carbon nanotubes as quantum wires. Nature 1997, 386, 474–477. 46. Liu, Z.; Tabakman, S.; Welsher, K.; Dai, H. Carbon nanotubes in biology and medicine: in vitro and in vivo detection, imaging and drug delivery. Nano Res. 2009, 2, 85–120. 47. Guldi, D. M.; Rahman, G. M.; Sgobba, V.; Ehli, C. Multifunctional molecular carbon materials—from fullerenes to carbon nanotubes. Chem. Soc. Rev. 2006, 35, 471–487. 48. Tasis, D.; Tagmatarchis, N.; Bianco, A.; Prato, M. Chemistry of carbon nanotubes. Chem. Rev. 2006, 106, 1105–1136. 49. Cherukuri, P.; Gannon, C. J.; Leeuw, T. K.; Schmidt, H. K.; Smalley, R. E.; Curley, S. A.; Weisman, R. B. Mammalian pharmacokinetics of carbon nanotubes using intrinsic near-infrared fluorescence. Proc. Natl. Acad. Sci. U.S.A. 2006, 103, 18882–18886. 50. Longmire, M.; Choyke, P. L.; Kobayashi, H. Clearance properties of nano-sized particles and molecules as imaging agents: considerations and caveats. Nanomedicine 2008, 3, 703–717. 51. Hainfeld, J. F.; Slatkin, D. N.; Focella, T. M.; Smilowitz, H. M. Gold nanoparticles: a new X-ray contrast agent. Br. J. Radiol. 2006, 79, 248–253. 52. Hirsch, L.; Gobin, A.; Lowery, A.; Tam, F.; Drezek, R.; Halas, N.; West, J. Metal nanoshells. Ann. Biomed. Eng. 2006, 34, 15–22. 53. Lytton-Jean, A. K.; Han, M. S.; Mirkin, C. A. Microarray detection of duplex and triplex DNA binders with DNA-modified gold nanoparticles. Anal. Chem. 2007, 79, 6037–6041. 54. Elghanian, R.; Storhoff, J. J.; Mucic, R. C.; Letsinger, R. L.; Mirkin, C. A. Selective colorimetric detection of polynucleotides based on the distance-dependent optical properties of gold nanoparticles. Science (N.Y.) 1997, 277, 1078–1081. 55. Hurst, S. J.; Hill, H. D.; Macfarlane, R. J.; Wu, J.; Dravid, V. P.; Mirkin, C. A. Synthetically programmable DNA binding domains in aggregates of DNA-functionalized gold nanoparticles. Small 2009, 5, 2156–2161. 56. Murphy, C. J.; Gole, A. M.; Hunyadi, S. E.; Stone, J. W.; Sisco, P. N.; Alkilany, A.; Kinard, B. E.; Hankins, P. Chemical sensing and imaging with metallic nanorods. Chem. Commun. 2008, 544–557. 57. Ohgushi, M.; Nagayama, K.; Wada, A. Dextran-magnetite: a new relaxation reagent and its application to T2 measurements in gel systems. J. Magn. Reson. 1978, 29, 599–601. 58. Bonnemain, B. Superparamagnetic agents in magnetic resonance imaging. Physicochemical characteristics and clinical applications. A review. J. Drug Targeting 1998, 6, 167–174. 59. Hartman, K. B.; Wilson, L. J.; Rosenblum, M. G. Detecting and treating cancer with nanotechnology. Mol. Diagn. Ther. 2008, 12, 1–14. 60. Weissleder, R.; Stark, D. D.; Compton, C. C.; Wittenberg, J.; Ferrucci, J. T. Ferrite-enhanced MR imaging of hepatic lymphoma: an experimental study in rats. Am. J. Roentgenol. 1987, 149, 1161–1165. 61. Weissleder, R.; Hahn, P. F.; Stark, D. D.; Rummeny, E.; Saini, S.; Wittenberg, J.; Ferrucci, J. T. MR imaging of splenic metastases: ferrite-enhanced detection in rats. Am. J. Roentgenol. 1987, 149, 723–726.
P1: OTA/XYZ P2: ABC c04 JWBS037-Chen January 25, 2011
98
8:49
Printer Name: Yet to Come
BIODISTRIBUTION AND PHARMACOKINETICS OF NANOPROBES
62. Weissleder, R.; Elizondo, G.; Wittenberg, J.; Lee, A. S.; Josephson, L.; Brady, T. J. Ultrasmall superparamagnetic iron oxide: an intravenous contrast agent for assessing lymph nodes with MR imaging. Radiology 1990, 175, 494–498. 63. Weissleder, R.; Bogdanov, A.; Neuwelt, E. A.; Papisov, M. Long-circulating iron oxides for MR imaging. Adv. Drug Deliv. Rev. 1995, 16, 321–333. 64. Sosnovik, D. E.; Nahrendorf, M.; Weissleder, R. Magnetic nanoparticles for MR imaging: agents, techniques and cardiovascular applications. Basic Res. Cardiol. 2008, 103, 122– 130. 65. Benderbous, S.; Corot, C.; Jacobs, P.; Bonnemain, B. Superparamagnetic agents: physicochemical characteristics and preclinical imaging evaluation. Acad. Radiol. 1996, 3(Suppl 2), S292–S294. 66. Vega-Villa, K. R.; Takemoto, J. K.; Yanez, J. A.; Remsberg, C. M.; Forrest, M. L.; Davies, N. M. Clinical toxicities of nanocarrier systems. Adv. Drug Deliv. Rev. 2008, 60, 929–938. 67. Aillon, K. L.; Xie, Y.; El-Gendy, N.; Berkland, C. J.; Forrest, M. L. Effects of nanomaterial physicochemical properties on in vivo toxicity. Adv. Drug Deliv. Rev. 2009, 61, 457–466. 68. Monteiro-Riviere, N. A.; Inman, A. O.; Zhang, L. W. Limitations and relative utility of screening assays to assess engineered nanoparticle toxicity in a human cell line. Toxicol. Appl. Pharmacol. 2009, 234, 222–235. 69. Jones, C. F.; Grainger, D. W. In vitro assessments of nanomaterial toxicity. Adv. Drug Deliv. Rev. 2009, 61, 438–456. 70. Shaw, S. Y.; Westly, E. C.; Pittet, M. J.; Subramanian, A.; Schreiber, S. L.; Weissleder, R. Perturbational profiling of nanomaterial biologic activity. Proc. Natl. Acad. Sci. U.S.A. 2008, 105, 7387–7392. 71. Toh, Y. C.; Lim, T. C.; Tai, D.; Xiao, G.; van Noort, D.; Yu, H. A microfluidic 3D hepatocyte chip for drug toxicity testing. Lab on a Chip 2009, 9, 2026–2035. 72. Lee, J.; Lilly, G. D.; Doty, R. C.; Podsiadlo, P.; Kotov, N. A. In vitro toxicity testing of nanoparticles in 3D cell culture. Small (Germany) 2009, 5, 1213–1221. 73. Jan, E.; Byrne, S. J.; Cuddihy, M.; Davies, A. M.; Volkov, Y.; Gun’ko, Y. K.; Kotov, N. A. High-content screening as a universal tool for fingerprinting of cytotoxicity of nanoparticles. ACS Nano 2008, 2, 928–938. 74. Cho, S. J.; Maysinger, D.; Jain, M.; Roder, B.; Hackbarth, S.; Winnik, F. M. Long-term exposure to CdTe quantum dots causes functional impairments in live cells. Langmuir 2007, 23, 1974–1980. 75. Emmanuel Chang, N. T.; William, W. Y.; Vicki, L. C. R. D. Evaluation of quantum dot cytotoxicity based on intracellular uptake13. Small 2006, 2, 1412–1417. 76. Derfus, A. M.; Chan, W. C. W.; Bhatia, S. N. Probing the cytotoxicity of semiconductor quantum dots. Nano Lett. 2004, 4, 11–18. 77. Yang, R. S.; Chang, L. W.; Wu, J. P.; Tsai, M. H.; Wang, H. J.; Kuo, Y. C.; Yeh, T. K.; Yang, C. S.; Lin, P. Persistent tissue kinetics and redistribution of nanoparticles, quantum dot 705, in mice: ICP-MS quantitative assessment. Environ. Health Perspect. 2007, 115, 1339–1343. 78. Liu, J.; Liu, Y.; Michalska, A. E.; Choo, K. H.; Klaassen, C. D. Distribution and retention of cadmium in metallothionein I and II null mice. Toxicol. Appl. Pharmacol. 1996, 136, 260– 268. 79. Ikeda, M.; Watanabe, T.; Zhang, Z. W.; Moon, C. S.; Shimbo, S. The integrity of the liver among people environmentally exposed to cadmium at various levels. Int. Arch. Occup. Environ. Health 1997, 69, 379–385. 80. Omarova, A.; Phillips, C. J. A meta-analysis of literature data relating to the relationships between cadmium intake and toxicity indicators in humans. Environ. Res. 2007, 103, 432–440.
P1: OTA/XYZ P2: ABC c04 JWBS037-Chen January 25, 2011
8:49
Printer Name: Yet to Come
REFERENCES
99
81. Yong, K. T.; Roy, I.; Ding, H.; Bergey, E. J.; Prasad, P. N. Biocompatible near-infrared quantum dots as ultrasensitive probes for long-term in vivo imaging applications. Small (Germany) 2009, 5, 1997–2004. 82. Lovric, J.; Cho, S. J.; Winnik, F. M.; Maysinger, D. Unmodified cadmium telluride quantum dots induce reactive oxygen species formation leading to multiple organelle damage and cell death. Chem. Biol. 2005, 12, 1227–1234. 83. Kirchner, C.; Liedl, T.; Kudera, S.; Pellegrino, T.; Munoz Javier, A.; Gaub, H. E.; Stolzle, S.; Fertig, N.; Parak, W. J. Cytotoxicity of colloidal CdSe and CdSe/ZnS nanoparticles. Nano Lett. 2005, 5, 331–338. 84. Zhang, Y.; He, J.; Wang, P.-N.; Chen, J.-Y.; Lu, Z.-J.; Lu, D.-R.; Guo, J.; Wang, C.-C.; Yang, W.-L. Time-dependent photoluminescence blue shift of the quantum dots in living cells: effect of oxidation by singlet oxygen. J. Am. Chem. Soc. 2006, 128, 13396–13401. 85. Chan, W. H.; Shiao, N. H.; Lu, P. Z. CdSe quantum dots induce apoptosis in human neuroblastoma cells via mitochondrial-dependent pathways and inhibition of survival signals. Toxicol. Lett. 2006, 167, 191–200. 86. Binil Itty Ipe, M. L.; Christof, M. N. On the generation of free radical species from quantum dots13. Small 2005, 1, 706–709. 87. Aldana, J.; Wang, Y. A.; Peng, X. Photochemical instability of CdSe nanocrystals coated by hydrophilic thiols. J. Am. Chem. Soc. 2001, 123, 8844–8850. 88. Smith, A. M.; Duan, H.; Rhyner, M. N.; Ruan, G.; Nie, S. A systematic examination of surface coatings on the optical and chemical properties of semiconductor quantum dots. Phys. Chem. Chem. Phys. 2006, 8, 3895–3903. 89. Dubertret, B.; Skourides, P.; Norris, D. J.; Noireaux, V.; Brivanlou, A. H.; Libchaber, A. In vivo imaging of quantum dots encapsulated in phospholipid micelles. Science 2002, 298, 1759–1762. 90. Maysinger, D.; Behrendt, M.; Lalancette-Hebert, M.; Kriz, J. Real-time imaging of astrocyte response to quantum dots: in vivo screening model system for biocompatibility of nanoparticles. Nano Lett. 2007, 7, 2513–2520. 91. Green, M.; Howman, E. Semiconductor quantum dots and free radical induced DNA nicking. Chem. Commun. 2005, 121–123. 92. Lovric, J.; Bazzi, H. S.; Cuie, Y.; Fortin, G. R.; Winnik, F. M.; Maysinger, D. Differences in subcellular distribution and toxicity of green and red emitting CdTe quantum dots. J. Mol. Med. 2005, 83, 377–385. 93. Liang, J.; He, Z.; Zhang, S.; Huang, S.; Ai, X.; Yang, H.; Han, H. Study on DNA damage induced by CdSe quantum dots using nucleic acid molecular “light switches” as probe. Talanta 2007, 71, 1675–1678. 94. Geys, J.; Coenegrachts, L.; Vercammen, J.; Engelborghs, Y.; Nemmar, A.; Nemery, B.; Hoet, P. H. M. In vitro study of the pulmonary translocation of nanoparticles: a preliminary study. Toxicol. Lett. 2006, 160, 218–226. 95. Gopee, N. V.; Roberts, D. W.; Webb, P.; Cozart, C. R.; Siitonen, P. H.; Warbritton, A. R.; Yu, W. W.; Colvin, V. L.; Walker, N. J.; Howard, P. C. Migration of intradermally injected quantum dots to sentinel organs in mice. Toxicol. Sci. 2007, 98, 249–257. 96. Chithrani, B. D.; Ghazani, A. A.; Chan, W. C. W. Determining the size and shape dependence of gold nanoparticle uptake into mammalian cells. Nano Lett. 2006, 6, 662–668. 97. De Jong, W. H.; Hagens, W. I.; Krystek, P.; Burger, M. C.; Sips, A. J.; Geertsma, R. E. Particle size-dependent organ distribution of gold nanoparticles after intravenous administration. Biomaterials 2008, 29, 1912–1919. 98. Murphy, C. J.; Gole, A. M.; Stone, J. W.; Sisco, P. N.; Alkilany, A. M.; Goldsmith, E. C.; Baxter, S. C. Gold nanoparticles in biology: beyond toxicity to cellular imaging. Acc. Chem. Res. 2008, 41, 1721–1730.
P1: OTA/XYZ P2: ABC c04 JWBS037-Chen January 25, 2011
100
8:49
Printer Name: Yet to Come
BIODISTRIBUTION AND PHARMACOKINETICS OF NANOPROBES
99. Ellen, E. C.; Judith, M.; Anand, G.; Catherine, J. M.; Michael, D. W. Gold nanoparticles are taken up by human cells but do not cause acute cytotoxicity. Small 2005, 1, 325–327. 100. Ofek, B.-I.; Ralph, M. A.; Valerie, E. F.; Darin, Y. F. Toxicity assessments of multisized gold and silver nanoparticles in zebrafish embryos. Small 2009, 5, 1897–1910. 101. Leonov, A. P.; Zheng, J.; Clogston, J. D.; Stern, S. T.; Patri, A. K.; Wei, A. Detoxification of gold nanorods by treatment with polystyrenesulfonate. ACS Nano 2008, 2, 2481–2488. 102. Tarantola, M.; Schneider, D.; Sunnick, E.; Adam, H.; Pierrat, S.; Rosman, C.; Breus, V.; Sonnichsen, C.; Basche, T.; Wegener, J.; Janshoff, A. Cytotoxicity of metal and semiconductor nanoparticles indicated by cellular micromotility. ACS Nano 2009, 3, 213–222. 103. Maria, T. H. K.; Wolfgang, B.; Helmut, E.; G¨unter, S. Cellular uptake and toxicity of Au55 clusters. Small 2005, 1, 841–844. 104. Manuela, S.-B. W. G. K.; Jens, L.; Stefanie, F.; Alexander, W.; Shinji, T.; G¨unter, S.; Wolfgang, B. Biodistribution of 1.4- and 18-nm gold particles in rats. Small 2008, 4, 2108–2111. 105. Lacerda, L.; Bianco, A.; Prato, M.; Kostarelos, K. Carbon nanotubes as nanomedicines: from toxicology to pharmacology. Adv. Drug Deliv. Rev. 2006, 58, 1460–1470. 106. Lam, C. W.; James, J. T.; McCluskey, R.; Hunter, R. L. Pulmonary toxicity of single-wall carbon nanotubes in mice 7 and 90 days after intratracheal instillation. Toxicol. Sci. 2004, 77, 126–134. 107. Lam, C. W.; James, J. T.; McCluskey, R.; Arepalli, S.; Hunter, R. L. A review of carbon nanotube toxicity and assessment of potential occupational and environmental health risks. Crit. Rev. Toxicol. 2006, 36, 189–217. 108. Schipper, M. L.; Nakayama-Ratchford, N.; Davis, C. R.; Kam, N. W. S.; Chu, P.; Liu, Z.; Sun, X.; Dai, H.; Gambhir, S. S. A pilot toxicology study of single-walled carbon nanotubes in a small sample of mice. Nat. Nano. 2008, 3, 216–221. 109. Chou, C. C.; Hsiao, H. Y.; Hong, Q. S.; Chen, C. H.; Peng, Y. W.; Chen, H. W.; Yang, P. C. Single-walled carbon nanotubes can induce pulmonary injury in mouse model. Nano Lett. 2008, 8, 437–445. 110. Warheit, D. B.; Laurence, B. R.; Reed, K. L.; Roach, D. H.; Reynolds, G. A.; Webb, T. R. Comparative pulmonary toxicity assessment of single-wall carbon nanotubes in rats. Toxicol. Sci. 2004, 77, 117–125. 111. Yu, Y.; Zhang, Q.; Mu, Q.; Zhang, B.; Yan, B. Exploring the immunotoxicity of carbon nanotubes. Nanoscale Res. Lett. 2008, 3, 271–277. 112. Al Faraj, A.; Cieslar, K.; Lacroix, G.; Gaillard, S.; Canet-Soulas, E.; Cre`ımillieux, Y. In vivo imaging of carbon nanotube biodistribution using magnetic resonance imaging. Nano Lett. 2009, 9, 1023–1027. 113. Helland, A.; Wick, P.; Koehler, A.; Schmid, K.; Som, C. Reviewing the environmental and human health knowledge base of carbon nanotubes. Environ. Health Perspect. 2007, 115, 1125–1131. 114. Shi Kam, N. W.; Jessop, T. C.; Wender, P. A.; Dai, H. Nanotube molecular transporters: internalization of carbon nanotube–protein conjugates into mammalian cells. J. Am. Chem. Soc. 2004, 126, 6850–6851. 115. Dumortier, H.; Lacotte, S.; Pastorin, G.; Marega, R.; Wu, W.; Bonifazi, D.; Briand, J.-P.; Prato, M.; Muller, S.; Bianco, A. Functionalized carbon nanotubes are non-cytotoxic and preserve the functionality of primary immune cells. Nano Lett. 2006, 6, 1522–1528. 116. Wu, P.; Chen, X.; Hu, N.; Tam, U. C.; Blixt, O.; Zettl, A.; Bertozzi, C. R. Biocompatible carbon nanotubes generated by functionalization with glycodendrimers. Angew. Chem. Int. Ed. 2008, 47, 5022–5025. 117. Yang, S.-T.; Wang, X.; Jia, G.; Gu, Y.; Wang, T.; Nie, H.; Ge, C.; Wang, H.; Liu, Y. Long-term accumulation and low toxicity of single-walled carbon nanotubes in intravenously exposed mice. Toxicol. Lett. 2008, 181, 182–189.
P1: OTA/XYZ P2: ABC c04 JWBS037-Chen January 25, 2011
8:49
Printer Name: Yet to Come
REFERENCES
101
118. Sheng-Tao Yang, K. A. F.; Jia-Hui, L.; Jing, W.; Hong-Fang, S.; Yuanfang, L.; Min, C.; Yipu, H.; Xin, W.; Haifang, W.; Ya-Ping, S. Covalently PEGylated carbon nanotubes with stealth character in vivo. Small 2008, 4, 940–944. 119. Liu, Z.; Davis, C.; Cai, W.; He, L.; Chen, X.; Dai, H. Circulation and long-term fate of functionalized, biocompatible single-walled carbon nanotubes in mice probed by Raman spectroscopy. Proc. Natl. Acad. Sci. U.S.A. 2008, 105, 1410–1415. 120. Poland, C. A.; Duffin, R.; Kinloch, I.; Maynard, A.; Wallace, W. A. H.; Seaton, A.; Stone, V.; Brown, S.; MacNee, W.; Donaldson, K. Carbon nanotubes introduced into the abdominal cavity of mice show asbestos-like pathogenicity in a pilot study. Nat. Nano 2008, 3, 423–428. 121. Tavill, A. S.; Bacon, B. R. Hemochromatosis: how much iron is too much? Hepatology 1986, 6, 142–145. 122. Malcolm, P.; Weir, J. F. G.; Peters, T. J. Haemosiderin and tissue damage. Cell Biochem. Funct. 1984, 2, 186–194. 123. Klausner, R. D.; Rouault, T. A.; Harford, J. B. Regulating the fate of mRNA: the control of cellular iron metabolism. Cell 1993, 72, 19–28. 124. Reimer, P.; Balzer, T. Ferucarbotran (Resovist): a new clinically approved RES-specific contrast agent for contrast-enhanced MRI of the liver: properties, clinical development, and applications. Eur. Radiol. 2003, 13, 1266–1276. 125. Gr¨uttner, C.; Teller, J. New types of silica-fortified magnetic nanoparticles as tools for molecular biology applications. J. Magn. Magn. Mater. 1999, 194, 8–15. 126. Berry, C. C.; Curtis, A. S. G. Functionalisation of magnetic nanoparticles for applications in biomedicine. J. Phys. D: Appl. Phys. 2003, 36, R198–R206. 127. Jain, T. K.; Reddy, M. K.; Morales, M. A.; Leslie-Pelecky, D. L.; Labhasetwar, V. Biodistribution, clearance, and biocompatibility of iron oxide magnetic nanoparticles in rats. Mol. Pharmacol. 2008, 5, 316–327. 128. Stuart, J. M.; Marie, R. M.; Maria, A. L.; Ricardo, A. F.; Micheal, N. E.; Dennis, R. F.; Scheetz, B.; Adeoye, Y. O. Phase I clinical evaluation of a new iron oxide MR contrast agent. J. Magn. Reson. Imag. 1994, 4, 301–307. 129. Weissleder, R. Liver MR imaging with iron oxides: toward consensus and clinical practice. Radiology 1994, 193, 593–595. 130. Nishimura, H.; Tanigawa, N.; Hiramatsu, M.; Tatsumi, Y.; Matsuki, M.; Narabayashi, I. Preoperative esophageal cancer staging: magnetic resonance imaging of lymph node with ferumoxtran10, an ultrasmall superparamagnetic iron oxide. J. Am. Coll. Surg. 2006, 202, 604–611. 131. Paciotti, G. F.; Myer, L.; Weinreich, D.; Goia, D.; Pavel, N.; McLaughlin, R. E.; Tamarkin, L. Colloidal gold: a novel nanoparticle vector for tumor directed drug delivery. Drug Deliv. 2004, 11, 169–183. 132. Oberdorster, G.; Oberdorster, E.; Oberdorster, J. Nanotoxicology: an emerging discipline evolving from studies of ultrafine particles. Environ. Health Perspect. 2005, 113, 823–839. 133. Maeda, H.; Wu, J.; Sawa, T.; Matsumura, Y.; Hori, K. Tumor vascular permeability and the EPR effect in macromolecular therapeutics: a review. J. Control. Release 2000, 65, 271– 284. 134. Maeda, H. The enhanced permeability and retention (EPR) effect in tumor vasculature: the key role of tumor-selective macromolecular drug targeting. Adv. Enzyme Regul. 2001, 41, 189– 207. 135. Ballou, B.; Lagerholm, B. C.; Ernst, L. A.; Bruchez, M. P.; Waggoner, A. S. Noninvasive imaging of quantum dots in mice. Bioconjug. Chem. 2004, 15, 79–86. 136. Soltesz, E. G.; Kim, S.; Laurence, R. G.; DeGrand, A. M.; Parungo, C. P.; Dor, D. M.; Cohn, L. H.; Bawendi, M. G.; Frangioni, J. V.; Mihaljevic, T. Intraoperative sentinel lymph node mapping of the lung using near-infrared fluorescent quantum dots. Ann. Thorac. Surg. 2005, 79, 269–277; discussion 269–277.
P1: OTA/XYZ P2: ABC c04 JWBS037-Chen January 25, 2011
102
8:49
Printer Name: Yet to Come
BIODISTRIBUTION AND PHARMACOKINETICS OF NANOPROBES
137. Fischer, H. C.; Pang, K. S.; Chan, W. C. W. Pharmacokinetics of nanoscale quantum dots: in vivo distribution, sequestration, and clearance in the rat. Adv. Funct. Mater. 2006, 16, 1299–1305. 138. Hama, Y.; Koyama, Y.; Urano, Y.; Choyke, P. L.; Kobayashi, H. Simultaneous two-color spectral fluorescence lymphangiography with near infrared quantum dots to map two lymphatic flows from the breast and the upper extremity. Breast Cancer Res. Treat. 2007, 103, 23–28. 139. Kobayashi, H.; Hama, Y.; Koyama, Y.; Barrett, T.; Regino, C. A. S.; Urano, Y.; Choyke, P. L. Simultaneous multicolor imaging of five different lymphatic basins using quantum dots. Nano Lett. 2007, 7, 1711–1716. 140. Zimmer, J. P.; Kim, S.-W.; Ohnishi, S.; Tanaka, E.; Frangioni, J. V.; Bawendi, M. G. Size series of small indium arsenide–zinc selenide core–shell nanocrystals and their application to in vivo imaging. J. Am. Chem. Soc. 2006, 128, 2526–2527. 141. Soo Choi, H.; Liu, W.; Misra, P.; Tanaka, E.; Zimmer, J. P.; Itty Ipe, B.; Bawendi, M. G.; Frangioni, J. V. Renal clearance of quantum dots. Nat. Biotechnol. 2007, 25, 1165–1170. 142. Chouly, C.; Pouliquen, D.; Lucet, I.; Jeune, J. J.; Jallet, P. Development of superparamagnetic nanoparticles for MRI: effect of particle size, charge and surface nature on biodistribution. J. Microencapsul. 1996, 13, 245–255. 143. Neuberger, T.; Sch¨opf, B.; Hofmann, H.; Hofmann, M.; von Rechenberg, B. Superparamagnetic nanoparticles for biomedical applications: possibilities and limitations of a new drug delivery system. J. Magn. Magn. Mater. 2005, 293, 483–496. 144. Jain, T. K.; Reddy, M. K.; Morales, M. A.; Leslie-Pelecky, D. L.; Labhasetwar, V. Biodistribution, clearance, and biocompatibility of iron oxide magnetic nanoparticles in rats. Mol. Pharmacol. 2008, 5, 316–327. 145. Briley-Saebo, K.; Bjornerud, A.; Grant, D.; Ahlstrom, H.; Berg, T.; Kindberg, G. M. Hepatic cellular distribution and degradation of iron oxide nanoparticles following single intravenous injection in rats: implications for magnetic resonance imaging. Cell Tissue Res. 2004, 316, 315–323. 146. Reddy, G. R.; Bhojani, M. S.; McConville, P.; Moody, J.; Moffat, B. A.; Hall, D. E.; Kim, G.; Koo, Y.-E. L.; Woolliscroft, M. J.; Sugai, J. V.; Johnson, T. D.; Philbert, M. A.; Kopelman, R.; Rehemtulla, A.; Ross, B. D. Vascular targeted nanoparticles for imaging and treatment of brain tumors. Clin. Cancer Res. 2006, 12, 6677–6686. 147. Montet, X.; Montet-Abou, K.; Reynolds, F.; Weissleder, R.; Josephson, L. Nanoparticle imaging of integrins on tumor cells. Neoplasia 2006, 8, 214–222. 148. Chen, L.-T.; Weiss, L. The role of the sinus wall in the passage of erythrocytes through the spleen. Blood 1973, 41, 529–537. 149. Haifang, W.; Jing, W.; Xiaoyong, D.; Hongfang, S.; Zujin, S.; Zhennan, G.; Yuanfang, L.; Yuliang, Z. Biodistribution of carbon single-wall carbon nanotubes in mice. J. Nanosci. Nanotechnol. 2004, 4, 1019–1024. 150. Deng, X.; Yang, S.; Nie, H.; Wang, H.; Liu, Y. A generally adoptable radiotracing method for tracking carbon nanotubes in animals. Nanotechnology 2008, 19, 075101. 151. McDevitt, M. R.; Chattopadhyay, D.; Jaggi, J. S.; Finn, R. D.; Zanzonico, P. B.; Villa, C.; Rey, D.; Mendenhall, J.; Batt, C. A.; Njardarson, J. T.; Scheinberg, D. A. PET imaging of soluble yttrium-86-labeled carbon nanotubes in mice. PLoS ONE 2007, 2, e907. 152. Elgrabli, D.; Floriani, M.; Abella-Gallart, S.; Meunier, L.; Gamez, C.; Delalain, P.; Rogerieux, F.; Boczkowski, J.; Lacroix, G. Biodistribution and clearance of instilled carbon nanotubes in rat lung. Particle Fibre Toxicol. 2008, 5, 20. 153. Hillyer, J. F.; Albrecht, R. M. Gastrointestinal persorption and tissue distribution of differently sized colloidal gold nanoparticles. J. Pharm. Sci. 2001, 90, 1927–1936. 154. Vijaya, K.; Kavita, K.; Sharanya, B.; Evan, J. B.; Stan W. C.; Genevieve M. F.; David J. R.; Meera, C.; Raghuraman K.; Kattesh V. K. Gum arabic as a phytochemical construct for
P1: OTA/XYZ P2: ABC c04 JWBS037-Chen January 25, 2011
8:49
Printer Name: Yet to Come
REFERENCES
155. 156. 157. 158.
159.
160.
161.
162.
163.
164. 165. 166. 167. 168. 169.
170.
171. 172.
103
the stabilization of gold nanoparticles: in vivo pharmacokinetics and X-ray-contrast-imaging studies. Small 2007, 3, 333–341. Decuzzi, P.; Pasqualini, R.; Arap, W.; Ferrari, M. Intravascular delivery of particulate systems: does geometry really matter? Pharm. Res. 2009, 26, 235–243. Champion, J.; Mitragotri, S. Shape induced inhibition of phagocytosis of polymer particles. Pharm. Res. 2009, 26, 244–249. Champion, J. A.; Mitragotri, S. Role of target geometry in phagocytosis. Proc. Natl. Acad. Sci. U.S.A. 2006, 103, 4930–4934. Lee, H. A.; Leavens, T. L.; Mason, S. E.; Monteiro-Riviere, N. A.; Riviere, J. E. Comparison of quantum dot biodistribution with a blood-flow-limited physiologically based pharmacokinetic model. Nano Lett. 2009, 9, 794–799. Niidome, T.; Yamagata, M.; Okamoto, Y.; Akiyama, Y.; Takahashi, H.; Kawano, T.; Katayama, Y.; Niidome, Y. PEG-modified gold nanorods with a stealth character for in vivo applications. J. Control. Release 2006, 114, 343–347. Singh, R.; Pantarotto, D.; Lacerda, L.; Pastorin, G.; Klumpp, C. d.; Prato, M.; Bianco, A.; Kostarelos, K. Tissue biodistribution and blood clearance rates of intravenously administered carbon nanotube radiotracers. Proc. Natl. Acad. Sci. U.S.A. 2006, 103, 3357– 3362. Wu, X.; Liu, H.; Liu, J.; Haley, K. N.; Treadway, J. A.; Larson, J. P.; Ge, N.; Peale, F.; Bruchez, M. P. Immunofluorescent labeling of cancer marker Her2 and other cellular targets with semiconductor quantum dots. Nat. Biotechnol. 2003, 21, 41–46. Xing, Y.; Smith, A. M.; Agrawal, A.; Ruan, G.; Nie, S. Molecular profiling of single cancer cells and clinical tissue specimens with semiconductor quantum dots. Int. J. Nanomed. 2006, 1, 473–481. Schipper, M. L.; Cheng, Z.; Lee, S.-W.; Bentolila, L. A.; Iyer, G.; Rao, J.; Chen, X.; Wu, A. M.; Weiss, S.; Gambhir, S. S. MicroPET-based biodistribution of quantum dots in living mice. J. Nucl. Med. 2007, 48, 1511–1518. Bruchez, M., Jr.; Moronne, M.; Gin, P.; Weiss, S.; Alivisatos, A. P. Semiconductor nanocrystals as fluorescent biological labels. Science (N.Y.) 1998, 281, 2013–2016. Chan, W. C. W.; Nie, S. Quantum dot bioconjugates for ultrasensitive nonisotopic detection. Science (N.Y.) 1998, 281, 2016–2018. Oliver, L.; Mnica L.-G.; Christine, S.; Jorg, A.; Horst, K.; Andreas R. B. Specific integrin labeling in living cells using functionalized nanocrystals. Small 2007, 3, 1560–1565. Pathak, S.; Davidson, M. C.; Silva, G. A. Characterization of the functional binding properties of antibody conjugated quantum dots. Nano Lett. 2007, 7, 1839–1845. Jayagopal, A.; Russ, P. K.; Haselton, F. R. Surface engineering of quantum dots for in vivo vascular imaging. Bioconjug. Chem. 2007, 18, 1424–1433. Wilhelm, C.; Gazeau, F.; Roger, J.; Pons, J. N.; Bacri, J. C. Interaction of anionic superparamagnetic nanoparticles with cells: kinetic analyses of membrane adsorption and subsequent internalization. Langmuir 2002, 18, 8148–8155. Papisov, M. I.; Bogdanov, A., Jr.; Schaffer, B.; Nossiff, N.; Shen, T.; Weissleder, R.; Brady, T. J. Colloidal magnetic resonance contrast agents: effect of particle surface on biodistribution. J. Magn. Magn. Mater. 1993, 122, 383–386. Jinxue, G.; Xiao, Z.; Qingnuan, L.; Wenxin, L. Biodistribution of functionalized multiwall carbon nanotubes in mice. Nucl. Med. Biol. 2007, 34, 579–583. Liu, Z.; Cai, W.; He, L.; Nakayama, N.; Chen, K.; Sun, X.; Chen, X.; Dai, H. In vivo biodistribution and highly efficient tumour targeting of carbon nanotubes in mice. Nat. Nano. 2007, 2, 47–52.
P1: OTA/XYZ P2: ABC c04 JWBS037-Chen January 25, 2011
104
8:49
Printer Name: Yet to Come
BIODISTRIBUTION AND PHARMACOKINETICS OF NANOPROBES
173. McDevitt, M. R.; Chattopadhyay, D.; Kappel, B. J.; Jaggi, J. S.; Schiffman, S. R.; Antczak, C.; Njardarson, J. T.; Brentjens, R.; Scheinberg, D. A. Tumor targeting with antibodyfunctionalized, radiolabeled carbon nanotubes. J. Nucl. Med. 2007, 48, 1180–1189. 174. Hirsch, L. R.; Stafford, R. J.; Bankson, J. A.; Sershen, S. R.; Rivera, B.; Price, R. E.; Hazle, J. D.; Halas, N. J.; West, J. L. Nanoshell-mediated near-infrared thermal therapy of tumors under magnetic resonance guidance. Proc. Natl. Acad. Sci. U.S.A. 2003, 100, 13549–13554. 175. Schwartz, J. A.; Shetty, A. M.; Price, R. E.; Stafford, R. J.; Wang, J. C.; Uthamanthil, R. K.; Pham, K.; McNichols, R. J.; Coleman, C. L.; Payne, J. D. Feasibility study of particle-assisted laser ablation of brain tumors in orthotopic canine model. Cancer Res. 2009, 69, 1659–1667. 176. McCarthy, J. R.; Jaffer, F. A.; Weissleder, R. A macrophage-targeted theranostic nanoparticle for biomedical applications. Small (Germany) 2006, 2, 983–987. 177. Jaffer, F. A.; Libby, P.; Weissleder, R. Optical and multimodality molecular imaging: insights into atherosclerosis. Arterioscler. Thromb. Vasc. Biol. 2009, 29, 1017–1024. 178. Bagalkot, V.; Zhang, L.; Levy-Nissenbaum, E.; Jon, S.; Kantoff, P. W.; Langer, R.; Farokhzad, O. C. Quantum dot-aptamer conjugates for synchronous cancer imaging, therapy, and sensing of drug delivery based on bi-fluorescence resonance energy transfer. Nano Lett. 2007, 7, 3065–3070. 179. Kester, M.; Heakal, Y.; Fox, T.; Sharma, A.; Robertson, G. P.; Morgan, T. T.; Altinoglu, E. I.; Tabakovic, A.; Parette, M. R.; Rouse, S.; Ruiz-Velasco, V.; Adair, J. H. Calcium phosphate nanocomposite particles for in vitro imaging and encapsulated chemotherapeutic drug delivery to cancer cells. Nano Lett. 2008, 8, 4116–4121. 180. Morgan, T. T.; Muddana, H. S.; Altinoglu, E. I.; Rouse, S. M.; Tabakovic, A.; Tabouillot, T.; Russin, T. J.; Shanmugavelandy, S. S.; Butler, P. J.; Eklund, P. C.; Yun, J. K.; Kester, M.; Adair, J. H. Encapsulation of organic molecules in calcium phosphate nanocomposite particles for intracellular imaging and drug delivery. Nano Lett. 2008, 8, 4108–4115. 181. Santimukul, S.; Charalambos, K.; Jan, G.; Perez, J. M. Drug/dye-loaded, multifunctional iron oxide nanoparticles for combined targeted cancer therapy and dual optical/magnetic resonance imaging. Small 2009, 5, 1862–1868. 182. Yezhelyev, M. V.; Qi, L.; O’Regan, R. M.; Nie, S.; Gao, X. Proton-sponge coated quantum dots for siRNA delivery and intracellular imaging. J. Am. Chem. Soc. 2008, 130, 9006–9012. 183. Chen, A. A.; Derfus, A. M.; Khetani, S. R.; Bhatia, S. N. Quantum dots to monitor RNAi delivery and improve gene silencing. Nucleic Acids Res. 2005, 33, e190.
P1: OTA/XYZ P2: ABC c05 JWBS037-Chen January 31, 2011
23:43
Printer Name: Yet to Come
PART II
NANOPARTICLES FOR SINGLE MODALITY MOLECULAR IMAGING
P1: OTA/XYZ P2: ABC c05 JWBS037-Chen January 31, 2011
23:43
Printer Name: Yet to Come
CHAPTER 5
Computed Tomography as a Tool for Anatomical and Molecular Imaging PINGYU LIU Palo Alto Unified School District, Palo Alto, California, USA
HU ZHOU Community Cancer Center of Roseburg, Roseburg, Oregon, USA
LEI XING Department of Radiation Oncology, Stanford University School of Medicine, Stanford, California, USA
5.1 INTRODUCTION X-ray and computed tomography (CT) imaging play a pivotal role in the diagnosis, staging, treatment planning, and image-guided intervention of various diseases. They are at the foundation of contemporary medical imaging and small animal imaging for biomedical research. It is estimated that 70–80% of all imaging procedures in medical applications entail the use of X-ray or CT imaging. The simplest implementation of X-ray imaging is the “plain film imaging,” and the next most common X-ray imaging approach is CT, which allows cross-sectional imaging of the body with exquisite depiction of anatomic detail. There are about 35,000 X-ray CT instruments installed worldwide for clinical applications. One of the primary advantages of X-ray imaging is the inherently simple basis of image contrast, which is the absorption of X-rays. For this reason, it has become the modality of choice in most radiology clinics. X-ray imaging can be used to image virtually every part of the body and is used for diagnosing orthopedic cases, cancer, heart disease, circulatory disease, and respiratory disease. In addition, it is also used as an aid for therapeutic interventions and as a means to follow the effects of treatment. During the last few years, innovations in X-ray detector technology have provided the capability to perform subsecond imaging of millimeter thin slices, which has opened up a host of new applications, such as blood vessel imaging, cardiac imaging, imaging of calcifications, large field imaging, and the ability to separately image the different vascular phases of medical diagnostic products. Finally, CT also plays an important role in molecular imaging of human diseases and animal model studies by providing unique and accurate spatial anatomy information. Nanoplatform-Based Molecular Imaging Edited by Xiaoyuan Chen C 2011 John Wiley & Sons, Inc. Copyright
107
P1: OTA/XYZ P2: ABC c05 JWBS037-Chen January 31, 2011
108
23:43
Printer Name: Yet to Come
COMPUTED TOMOGRAPHY AS A TOOL FOR ANATOMICAL AND MOLECULAR IMAGING
In this chapter, the physical and mathematical foundations of CT imaging are first reviewed. Recent advancements in CT technologies, such as multislice CT, cone-beam CT, respiration-gated CT, and four-dimensional (4D) CT, are highlighted. Novel strategies of scatter removal, noise suppression, and dose reduction are also discussed. In the second next of the chapter, we summarize the characteristics and performance of various contrast agents for enhanced CT imaging. The success of CT imaging techniques in molecular imaging is intimately dependent on the use of X-ray contrast agents to better differentiate soft tissues and to reveal the diseased regions. In reality, while contrast media find important applications in both micro-CT small animal imaging research and clinical settings, most commercially available CT agents possess blood clearance properties that are not suitable to the time scale of small animal CT protocols. The sensitivity of the agents is also problematic. Development of novel CT contrast agents represents a forefront of micro-CT and molecular imaging research. After a brief review of conventional iodine-based contrast media, we discuss some emerging agents for disease-specific applications. Clinically, the development of disease-specific contrast agents will allow us to realize the enormous promise of molecular biology research in routine practice.
5.2 PRINCIPLE OF COMPUTED TOMOGRAPHY Although the words of “computed tomography” (CT) could have much wider meaning, the terminology is accepted to specially refer to X-ray transmission computed tomography. In this technology the projections from the X-ray transmitted through the object under investigation are mathematically processed to construct a two-dimensional image. The word tomography originates from the Greek word tomos, which means “a section” or “a cutting.” Tomographic techniques are not restricted to transmitted X-rays. In positron emission tomography (PET) the information projections are from the emitted photons instead of the transmitted ones. In optical projection tomography (OPT) the source is visible light instead of X-rays. In traditional imaging technologies such as camera or X-ray radiography the recording media are used to store the image projected onto them. In tomography the image does not exist but is mathematically constructed from indirect information by computation based on the Radon transform theory. 5.2.1 Physics of X-rays X-rays are a form of electromagnetic radiation emitted when electrically charged particles release their potential or kinetic energies in the form of photons of energy from kilo electron volts (keV) to mega electron volts (MeV), or wavelength from picometers (10−9 meters) to nanometers (10−6 meters). In CT systems the X-ray is generated in an X-ray tube. The tube contains a thermal filament and a rotational anode target in a vacuum chamber. The target is set to a positive voltage relative to the filament, usually from 30 to 150 kV. The electrons emitted from the filament are accelerated by the electric field toward the target. When these electrons strike the target, they experience electrostatic forces from the electrons and nuclei in the target and are decelerated in a short distance. According to electrodynamics, a charged particle will emit electromagnetic radiation when it is accelerated or decelerated. The kinetic energies suddenly lost from the incident electrons are released in the form of electromagnetic
P1: OTA/XYZ P2: ABC c05 JWBS037-Chen January 31, 2011
23:43
Printer Name: Yet to Come
PRINCIPLE OF COMPUTED TOMOGRAPHY
109
waves called bremsstrahlung radiation. The electrons which lose most of their kinetic energies in the first collision emit high-energy photons, and those electrons which lose their energies gradually emit low-energy photons. Thus bremsstrahlung radiations form a continuous energy spectrum, whose maximum energy equals the acceleration potential times the electron charge. As in most of the X-ray tubes used for diagnostics, the acceleration potential is high enough so that electrons in the inner orbits of the target atoms may be kicked out by the incident electrons, or by the photons generated by the incident electrons, and the vacancies are filled up by higher orbit electrons. In this process a photon will be emitted whose energy equals the energy difference between the two orbits. Since this process has a resonance property, the photons emitted form the characteristic energy spikes on top of the bremsstrahlung spectrum. The energies of these spikes are specified by the target material, independent of the acceleration potential. The X-ray photons then pass through a window that separates the vacuum in the tube from the ambience, and form the X-ray beam. The window usually consists of a layer of millimeter-thick low-Z metal, which also plays a role in filtration, absorbing most of the low-energy photons. The spectrum from a kilovolt X-ray tube is specified by its acceleration potential, the target material, and the filtration. Figure 5.1 shows the output spectra from a tungsten target with acceleration potentials of 80 kV, 100 kV, and 120 kV with a 2.5-mm Al filter. The surface normal of the tungsten target is 10◦ with respect to the incident electron beam. The data of a spectrum specified in this way can be generated by Monte Carlo simulations or by empirical formulas [1]. When an X-ray beam passes through a material, the photons in the beam are either scattered or absorbed. Therefore the intensity of the outgoing beam is attenuated in comparison with the incident beam. If the incident photons are not monochromic, usually the lower-energy photons are attenuated more than the high-energy photons, so the transmitted 10
Photon Number (arbitrary)
9 8 7 6 5 4 3 2 1 0
0
0.02
0.04
0.06
0.08
0.1
0.12
0.14
Photon Energy (MeV) FIGURE 5.1 Energy spectrums from an X-ray tube with 80 kV (dotted), 100 kV (dashed), and 120 kV (solid) acceleration potentials, with 2.5 mm Al filtration. The spectrums are generated with program Spektr [2]. The continuous component is from bremsstrahlung radiation and the spikes are the tungsten characteristic peaks. Note that the maximum photon energy equals the acceleration voltage.
P1: OTA/XYZ P2: ABC c05 JWBS037-Chen January 31, 2011
110
23:43
Printer Name: Yet to Come
COMPUTED TOMOGRAPHY AS A TOOL FOR ANATOMICAL AND MOLECULAR IMAGING
beam will contain a higher fraction of high-energy photons than low-energy photons than the incident beam. This effect is called “beam hardening.” Since the X-ray emitted from the target contains a large portion of low-energy photons, and these photons do not make significant contributions to the image contrast, as we will discuss shortly, a filter of a few millimeters of aluminum is usually placed in front of the X-ray tube window to reduce the low-energy component. Also, beryllium filtration is used for low-energy and copper in high-energy systems. The filtration thus changes the shape of the output spectrum. A photon in the energy range between 1 and 200 kV may experience one of three types of interactions when it passes through a material: Rayleigh scattering, photoelectric absorption, and Compton scattering. In Rayleigh scattering, or coherent scattering, the photon is bounced by an electron or an atom without loss of its energy. It is this process that causes the sky to be blue. Rayleigh scattering mainly happens with low-energy photons. In X-ray radiography and CT it can cause image blurring. Photoelectric absorption happens when an electron is knocked out from an inner orbit of an atom. The energy of the photon is completely absorbed and causes ionization. In Compton scattering the photon energy is partially transferred to an electron. The outgoing photon changes its direction and has lower energy compared with the incident photon. The probabilities of all three interactions increase with electron density and are energy dependent. As the energy increases, the probabilities of Rayleigh scattering and the photoelectric absorption reduce. The probability of Compton scattering increases at low energy, then stays nearly constant over the energy range for X-ray imaging. When a monochromic X-ray beam passes through a material, the attenuation can be described by the Beer–Lambert law: I = I0 exp(−x)
(5.1)
where I and I 0 are the transmitted intensity and incident intensity, respectively, is the linear attenuation coefficient (in cm−1 ), which is energy dependent, and x is the thickness (in cm) of the body through which the beam penetrated. The Beer–Lambert law can easily be understood by looking at Figure 5.2. In this figure the X-ray beam of intensity I 0 is incident on a uniform material from the left side and transmits to the right side. The beam passes through the first block with some attenuation, and we denote the beam intensity outgoing from the block as I , and that from the second block as I , and so on. Because every block generates the same attenuation, I /I0 = I /I = · · · we have I/I 0 = (I /I 0 )n = exp[−n ln(I 0 /I )]. That is, the attenuation is an exponential function of the material thickness. The logarithm of the attenuation by a unit thickness of the material is the linear attenuation coefficient .
I0
I'
I"
...
I
FIGURE 5.2 Attenuation of X-ray through material. The material is made by n identical blocks.
P1: OTA/XYZ P2: ABC c05 JWBS037-Chen January 31, 2011
23:43
Printer Name: Yet to Come
PRINCIPLE OF COMPUTED TOMOGRAPHY
111
The linear attenuation coefficient is often expressed as the product of mass attenuation coefficient (/ ) and density : = (/ )
(5.2)
The significance of this expression comes from the fact that for all condensed matter the mass attenuation coefficient (/ ) is an intrinsic property of the material, depending only on its elemental compositions. Therefore water and ice have the same / , and so do diamond and graphite. Since the mass attenuation coefficient of all the elements have been tabulated over a wide range of photon energies (NIST), once the elemental compositions of a material are known, its (/ )mat can be calculated from the mass attenuation coefficients of the component elements (/ )elem and their mass weights welem (/ )mat =
welem (/ )elem
(5.3)
elem
Mass Attenuation Coefficients (cm²/g)
The mass attenuation coefficients of water, soft tissue, aluminum, copper, lead, and tungsten are shown in Figure 5.3. The mass attenuation coefficients of all the materials are strongly energy dependent in the energy range up to 100 kV. From Figure 5.3 it can be seen that the low-energy component of X-rays, especially at photon energies below 10 kV, all the materials have very large attenuation coefficients. Therefore the component of an X-ray beam in this energy range is not useful for imaging of a body because it will be entirely absorbed after a layer of a few millimeters. In other words, nearly all the materials are opaque. On the high-energy end of the spectrum, a photon of energy higher than 250–300 kV would have long penetration depth comparable with typical body size, making all the tissues nearly equally transparent so the X-ray would not create significant contrast. These phenomena dictate that the spectrum ranges of X-ray imaging systems
10
10
10
10
10
10
4
water soft tissue Al Cu Pb W
3
2
1
0
-1
10
0
10
1
10
2
10
3
Photon Energy (keV) FIGURE 5.3 Mass attenuation coefficients / of water, soft tissue, aluminum (Al), copper (Cu), lead (Pb), and tungsten (W). Note that mass attenuation coefficients of water and soft tissue are almost overlapped in this figure.
P1: OTA/XYZ P2: ABC c05 JWBS037-Chen January 31, 2011
112
23:43
Printer Name: Yet to Come
COMPUTED TOMOGRAPHY AS A TOOL FOR ANATOMICAL AND MOLECULAR IMAGING
must lie between 10 and 250 kV. The upper limit of the spectrum is set by the acceleration potential of the X-ray tube and the lower limit is set by filtration. If the photon energies in the incident X-ray beam expand to a spectrum S0 (E), after passing through an object of thickness x, the transmitted beam intensity I(x) is expressed by extending Eq. (5.1) as I (x) =
d E S(E, x) =
d E S0 (E) exp(−(E)x)
(5.4)
where S(E, x) is the transmitted beam spectrum at thickness x. If the material in the object is inhomogeneous, the mass attenuation coefficient and the density vary with the location, and Eq. (5.4) is further extended as I (x) =
d E S(E, x) =
S0 (E) exp −
x 0
(E, ) ()d d E
(5.5)
If a wide X-ray beam passes through an object, different parts of the beam pass through different parts of the body and will experience different attenuations. The spatial intensity variations then form an image projected to the record medium. This is how an X-ray radiographic image is formed. In cases where the spectrum is concentrated within a relatively narrow range of energy, the beam can be approximated to be monochromatic, and Eq. (5.5) is reduced to I (x) ≈ I0 exp −
x 0
() ()d
= I0 exp −
x
()d
(5.6)
0
or − ln
I ≈ I0
x
()d
(5.6a)
0
From Eq. (5.5) and (5.6a) it is seen that each pixel in a projection is an overlap of the linear attenuation coefficients of many layers of the object. If the object is divided into many small voxels, the logarithm of the projection pixel value would be proportional to the sum of the attenuations of the voxels along the path of the ray that projects to the pixel. If the X-ray is incident on the object from a different direction, in the new projection the logarithm of the pixel values would be proportional to the sum of another group of the voxel attenuations. Therefore if the X-ray beam can be arranged to pass through the object from many different directions, the three-dimensional distribution of the attenuations of the voxels could be resolved from the projections. In 1917 an Austrian mathematician, Johann Radon, proved that it is in principle possible to reconstruct a cross-sectional image of an unknown object using an infinite number of projections through the object [3], which set the theoretical foundation of computed tomography. Recently, Liu et al. [4] proposed a higher level image reconstruction theory, the P-transform, which unified X-ray and photoacoustic CT reconstruction theories into one theory. The Radon transform becomes a special case of the P-transform. Further description of the P-transform is beyond the scope of this chapter. Readers who are interested in the P-transform theory may refer to the original P-transfer theory [5].
P1: OTA/XYZ P2: ABC c05 JWBS037-Chen January 31, 2011
23:43
Printer Name: Yet to Come
PRINCIPLE OF COMPUTED TOMOGRAPHY
113
5.2.2 Mathematics of Computed Tomography
Radon Transformation
The Radon transform of a function f (x, y) can be expressed as
R(r, ␣) = R[ f (x, y)] =
f (x, y)␦(r − x cos ␣ − y sin ␣)d x d y
(5.7)
where r and ␣ are the polar coordinates of the projection, as shown in Figure 5.4a. The integration runs over the whole x-y plane. For a given value of ␣, a series of line integrals are computed along the lines r = x cos ␣ + y sin ␣, perpendicular to the vector r = (r, ␣), where each r corresponds to a line integral. Figure 5.4b demonstrates the result of the line integrals of the object in Figure 5.4a at angle ␣. Solving the function f (x, y) in Eq. (5.7) from known projections R(r, ␣) is called image reconstruction. There are many algorithms performing the function of image reconstruction.
Principle of Image Reconstruction mation takes the form
When applied to X-ray CT, the Radon transfor
− ln[I (r, ␣)/I0 ] = R(r, ␣) = R[(x, y)] =
(x, y)␦(r − x cos ␣ − y sin ␣)d x d y
(5.7a) where (x, y) is the linear attenuation coefficient at point (x, y), as in Eqs. (5.6) and (5.6a). Rotating the coordinate system (x, y) for an angle ␣ about its origin, so that x = x cos ␣ + y sin ␣ and y = y cos ␣ − x sin ␣ become the new coordinate system, Eq. (5.6) is then rewritten as (5.8) − ln[I (x , ␣)/I0 ] = R(x , ␣) = (x, y) dy
FIGURE 5.4 The Radon transformation of a cross section of an object. In this illustration f (x, y) is uniformly distributed inside the square in (a) so that in (b) the values of the Radon transformation are proportional to the lengths of the line segments across the object.
P1: OTA/XYZ P2: ABC c05 JWBS037-Chen January 31, 2011
114
23:43
Printer Name: Yet to Come
COMPUTED TOMOGRAPHY AS A TOOL FOR ANATOMICAL AND MOLECULAR IMAGING
Taking the Fourier transformation of Eq. (5.8) with respect to x , we have FR (k, ␣) =
∞
−∞
R(x , ␣)e−i2kx d x =
∞
−∞
(x, y)e−i2kx d x dy
(5.9)
The linear attenuation coefficient (x, y) can be solved by inverse Fourier transformation as
(x,y) =
∞
d␣ 0
−∞
FR (k, ␣)e
i2kx
|k|dk =
∞
d␣ 0
−∞
FR (k, ␣)ei2k(x cos ␣+y sin ␣) |k|dk (5.10)
Equation (5.10) is the back projection formula. With this equation the linear attenuation coefficient can be calculated from the projection data to form a cross-sectional image. A series of such cross-sectional images along a longitudinal axis are then stacked together to form a three-dimensional (3D) image. Therefore the essential step is the two-dimensional (2D) reconstruction of the cross-sectional image from the projected images. Although in principle Eq. (5.10) can be used to solve the problem of reconstruction, Eq. (5.7a) that relates (x, y) to the projected image I(r, ␣) is much simplified. This equation assumes the X-ray is monochromic or can be approximated to be monochromic, and the projection is free of noise. In modern CT systems more sophisticated mathematics are used in the reconstruction, which will be discussed in more detail as a specific topic.
5.3 EVOLUTION OF CT IMAGING TECHNOLOGY The basic components of a CT system includes an X-ray source, a detector or an array of detectors, a collimator that shapes the X-ray beam, a gantry carrying the source and detectors, a patient couch, and one or more computers that control the system, acquire the data, and reconstruct the image. In medical CT systems the computers are also responsible for administration, data archiving, data transferring, and image display.
5.3.1 The First to Fourth Generations of CT Scanners The first CT scanner was designed by Godfrey Hounsfield in the early 1970s by exactly realizing the Radon transformation shown in Figure 5.4 and Eq. (5.7). In the system an X-ray source and a single detector were set at the opposite sides of the object. The X-ray was confined to a narrow beam (pencil beam) aiming at the detector. The source–detector pair was moved together so that the object was scanned by parallel beams. The pair was then rotated a small angle and the procedure was repeated to scan the object at a different angle. This arrangement is shown in Figure 5.5a. The data acquisition ran through the angles over 360◦ . Hounsfield and A. M. Cormack shared the 1979 Nobel Prize in Medicine and Physiology for this invention. In a first generation CT scanner, the X-ray is confined to a narrow pencil beam, and the single detector senses only the primary photons. Because the scattered photons are going in different directions, they are ignored. The scattering reduces the intensity of the signal reaching the detector but this amount can be compensated by calibration.
P1: OTA/XYZ P2: ABC c05 JWBS037-Chen January 31, 2011
23:43
Printer Name: Yet to Come
EVOLUTION OF CT IMAGING TECHNOLOGY
(a)
(c)
115
(b)
(d)
FIGURE 5.5 Diagrams of four generations of CT: (a) first, (b) second, (c) third, and (d) fourth generations. Grey spheres represent X-ray sources.
The second generation of CT is shown in Figure 5.5b, in which the single detector is replaced by a linear array of detector elements. The X-ray was now confined in a fan beam so that all the detector elements can acquire data simultaneously. The source and the detector array translate to cover the subject being imaged, then rotate to the next angle. The data acquisition speed is multiplied because several detector elements work simultaneously, and the X-ray source is used more efficiently. But there is a price for these advantages: the noise due to scattering caused by using a wider X-ray beam. It is possible to add a grid in front of the detector array to reduce the scattering noise [6]. However, this solution has limitations because the grid not only adds extra weight to the detector but also takes away the detector area, thus reducing the efficiency. Because the wall between adjacent grid openings must be thick enough to create sufficient attenuation to the oblique photons, these walls will occupy detector area. In high spatial resolution detectors, the area taken by the grid wall becomes significant so that the benefit of noise reduction brought from the grid could be outweighed by the reduction of the element areas. In the third generation of CT, a longer array of detector elements is used and the X-ray from the single source is confined to a wider fan beam so that the whole width of the object can be covered. Therefore the translation is no longer necessary, and only the gantry that holds the source and detector array needs to rotate. The detectors are arranged on an arc, as shown in Figure 5.5c. Most modern clinical CT systems use this arrangement. The noise level of the third-generation CT scanner is higher than the second-generation scanner if all the components are the same, because of the wider X-ray beam angles. But
P1: OTA/XYZ P2: ABC c05 JWBS037-Chen January 31, 2011
116
23:43
Printer Name: Yet to Come
COMPUTED TOMOGRAPHY AS A TOOL FOR ANATOMICAL AND MOLECULAR IMAGING
the data acquisition speed is much faster, which brings tremendous benefit in medical applications, especially since it makes use of higher spatial resolution with smaller detector element area. Since the first CT scanner appeared in the 1970s, major efforts in CT engineering have been focused on increasing the data acquisition speeds and the spatial resolutions. This is discussed in the next sections. The fourth-generation CT has a set of detectors that are arranged in a ring over 360◦ around the object, and the source is mounted on a rotational gantry, as shown in Figure 5.5d. The detectors are on a fixed ring. Therefore the wiring can be greatly simplified compared to third-generation CT. The fixed relative positions of source and detectors in the third generation are no longer here. It is possible in principle to read out all the detector elements even if most of them are not irradiated, but such a procedure would slow down the acquisition speed and add work load to the reconstruction software. Therefore extra electronics and mechanical components are needed to synchronize the data acquisition—a complication that restricts the application of this generation. 5.3.2 The Fifth Generation of CT Scanner In CT history the fifth generation of CT is usually referred to as the electron beam CT, or the EBCT. The EBCT was the result of applying CT in a cardiology study, which requires a CT with higher temporal resolution. Research showed that in order to get a clear tomogram of the fastest moving part of a living heart, a complete data acquisition time should be no greater than 19.1 ms [7]. It’s obvious that there is no CT with mechanically rotating gantry that can reach this speed. An EBCT scanner can reach the goal by replacing the heavy X-ray tube with an electronically steered X-ray source. In an EBCT scanner, the X-ray source consists of a semicircular tungsten anode ring housed in a huge vacuum chamber with an electron gun, as shown in Figure 5.6. An electronically steered high-intensity electron beam generates a fast moving X-ray source along the tungsten target. Among various EBCT designs, GE’s eSpeed scanner is the fastest CT scanner in the world. Compared to other CT systems, EBCT technology has its own shortcomings, such as higher capital investment and lower signal-to-noise ratio (SNR). Currently, about 130 EBCT systems are in service worldwide. 5.3.3 CT Scanner with Nanotube X-ray Sources With the emergence of an X-ray source based on the field emission effect from the tip of a carbon nanotube [8], a new type of CT system was proposed by Alexei Ramotar in 2006 [9]. In this design the fixed detectors are arranged in a ring over 360◦ around the object as in the fourth generation, but the rotational single source is replaced by a fixed ring of many sources as shown in Figure 5.7. The sequence that turns the sources on and off is controlled by electronics. If the X-ray beam intensity of the nanotube source is sufficient, a scanner designed from this concept will completely omit the mechanical complication of the rotational gantry, and the data acquisition can be of very high speed and in more flexible patterns than just a rotational sequence. An example of possible applications of the technique is breast CT for mammograms. Compared with the current widely used breast-compression mammogram technique, breast CT offers patients more comfort and is highly sensitive in detecting early stage tumors [10]. There are restrictions to the application of traditional CT to mammograms, one of which is the consideration of radiation dose to the lungs and the heart, because of the geometry of
P1: OTA/XYZ P2: ABC c05 JWBS037-Chen January 31, 2011
23:43
Printer Name: Yet to Come
EVOLUTION OF CT IMAGING TECHNOLOGY
117
Focus, bend & steering coils
Electron gun
X-rays Shielding
Detector rings
Patient Electron beams Vacuum chamber
Tungsten anode rings
Fan x-rays
FIGURE 5.6 Side view (top) and cross view (bottom) of diagram of an EBCT scanner. (From Pingyu Liu’s 1997 proposal with author’s permission.)
the breast and the limited space between two breasts. Using the new technique the detector elements could be arranged on a spherical surface around the breast in layers parallel with breast base, for left and right breasts, respectively. The sources could be arranged on the same spherical surface with the beams collimated in the corresponding layers, with limited dose to the chest wall underneath.
FIGURE 5.7 Diagram of CT scanner with nanotube X-ray sources. The little dits represent the detectors and the black lines represent the X-ray sources. In this arrangement both the sources and detectors are fixed so that the rotation gantry can be completely omitted. The emissions from the sources are controlled by electronics so the data acquisition speed can be very high and the emission can be in flexible patterns other than a rotational sequence. In the figure one of the sources is collimated so that the X-rays are confined to certain detectors.
P1: OTA/XYZ P2: ABC c05 JWBS037-Chen January 31, 2011
118
23:43
Printer Name: Yet to Come
COMPUTED TOMOGRAPHY AS A TOOL FOR ANATOMICAL AND MOLECULAR IMAGING
5.3.4 Power and Data Transmission in a CT System Power and data transmission between the stationary components and the rotational gantry of a CT system can be realized in two possible ways. Conventionally, transmission is through cables wrapped on a rack. When the gantry rotates in one direction, the cables wind up on the rack. When the gantry rotates in another direction the cables unwind. Because the cable cannot be very long, the gantry needs to rotate alternately back and forth. This technique is used in early CT systems and most modern cone-beam CT systems, where a slow gantry speed can be tolerated. In most modern CT scanners a slip ring technique is used to transmit the power and data through contact brushes and slip rings. In some systems the data are transmitted through RF couplings, or fiberoptic rotary joints. In other CT systems a more flexible optical data coupling method is utilized. An important specification of a CT system is its gantry rotation time, which determines system’s temporal resolution. In a modern CT system, gantry rotation speed can reach 0.35 second per cycle. On the rim of a gantry 2 meters in diameter, the centrifugal acceleration can reach 24g. Therefore on the gantry the components must be carefully arranged to balance the momentum. 5.3.5 Helical CT If the patient couch moves simultaneously when the gantry is rotating, the patient will see the source moving along a helical path. This arrangement is called the helical or spiral CT. With slip rings the gantry can rotate continuously so the helix can extend over the length of the whole body. In third-generation CT with an array of detectors, the distance the couch moves during one gantry rotation period defines one slice thickness, which is the spatial resolution of the image along the axial direction. This technique is used in most modern medical CT systems, because it allows fast and high-resolution imaging. Fullbody CT imaging using such a system with 0.625-mm slice thickness can be completed in 1.5 minutes, and the image reconstruction can be completed on the fly. Fast imaging speed is important for medical applications. A shorter-time scan reduces the artifacts caused by patient motion. Many CT contrast agents stay in the body only a short period of time (a few minutes) and then are washed out by the urinary system. An imaging system must complete the data acquisition within that period. Most importantly, although all the CT systems are able to provide stationary anatomical information, a fast-imaging system can also provide dynamical and functional information. In cases where motion is periodic, as the respiration or heartbeat, the acquisition and reconstruction of time-resolved or 4D CT images can be accomplished with gating technology. During gated acquisition, the CT runs in scene mode, in which the couch moves step by step instead of continuously. In each step the gantry rotates several revolutions; each corresponds to a phase. In this mode the patient sees the source moving along circles. As the couch steps through the body of the patient, these circles stack to form a scan. The projections are then sorted according to the phases of the motion, and those belonging to a phase are picked up for reconstruction. When displayed, the reconstructed images are shown phase by phase as a 3D movie. Most helical CT systems can be operated in scene mode. 5.3.6 Multidetector CT The acquisition speed and spatial resolution of helical CT can be further improved by adding one or more rows of detectors to the array, so that during one gantry revolution more than
P1: OTA/XYZ P2: ABC c05 JWBS037-Chen January 31, 2011
23:43
Printer Name: Yet to Come
EVOLUTION OF CT IMAGING TECHNOLOGY
119
one slice of data can be collected at the same time. Multiple-slice CT, also referred to as multidetector CT (MDCT), makes a narrower slice thickness, more slices, and larger volume coverage simultaneously. MDCT scanners with 320 slices are now available in the market. There are many technical concerns in the design of MDCT systems. Because of the scattering effect, the MDCT data contains higher noise compared with single-slice CT. The increase in the number of slices is accomplished by reducing the row width of the detectors, and the reduction of the row width is limited by the detector sensitivity, which is proportional to the detector element area. Most manufacturers restrict the total detector width to 20 or 24 mm when the number of slices is 2, 4, 16, 20, and 24, and to 40 mm when the number of slices is 40 and 64. MDCT delivers the same X-ray dose to the object as single-slice CT for the same spatial resolution, but to get the same signal-to-noise ratio MDCT needs a higher dose because of the higher noise level. Especially, to take advantage of the high spatial resolution of MDCT, a significantly higher dose would be necessary to create a sufficient signal level from the smaller area of the detector elements. Many manufacturers have developed dose control mechanisms to automatically optimize the dose for each slice of a scan. The image reconstruction time of MDCT is expected to take longer than that of singleslice CT due to the larger amount of data. However, because of improved computer technology the reconstruction time does not significantly delay the total scan. The algorithm reconstructing on-the-fly can complete a slice reconstruction within half a second after acquisition. The fast speed and high spatial resolution brought by MDCT makes dynamic imaging of moving organs possible. An example is functional vascular imaging. In CT coronary angiography, cardiac data can be acquired continuously during 5 or 6 heartbeats. The dynamic images of the moving heart can be reconstructed corresponding to electrocardiographic gating signals. In polytrauma applications, the whole-body scan over 2 meters along the patient body can observe an upper extremity angiogram followed by pelvic and lower extremity angiography with one bolus of contrast. 5.3.7 Cone-Beam CT (CBCT) Replacement of a pencil beam and a single detector by a fan beam and a 1D detector array evolves CT from the first generation to the second generation, and extending the array size further evolves CT to the third generation [11, 12]. The array can also be extended in another direction to form a two-dimensional detector array. To irradiate the two-dimensional array the beam needs also to form a cone shape. In many cone-beam CT systems the detector areas are large enough that the beam can cover the whole region of interest in the body, so that during the acquisition of data the object does not have to move. These systems usually connect the stationary components and the gantry through cables, and rotate only one revolution for a scan. Typical scan time over 360◦ is about 1 minute, comparable to the acquisition time with helical CT. Compared to helical CT, cone-beam CT has a much simpler structure. The cable connection does not require high precision machining as the slip ring does; therefore the cost is significantly reduced. Also, cone-beam CT can be integrated into special systems such as a radiation treatment machine without interfering with the primary function of the system. The noise level of cone-beam CT is higher from 2D scattering. To obtain the same signal-to-noise level, cone-beam CT needs to deliver a several times higher dose to the object than helical CT. Also, a large-area 2D detector introduces another challenge—the
P1: OTA/XYZ P2: ABC c05 JWBS037-Chen January 31, 2011
120
23:43
Printer Name: Yet to Come
COMPUTED TOMOGRAPHY AS A TOOL FOR ANATOMICAL AND MOLECULAR IMAGING
data transmission time. A detector with 4000 × 3000 pixel size needs about 1 second to read out; therefore the frame rate cannot be faster unless the image pixels are combined in 2 × 2 binning or 4 × 4 binning. 5.3.8 Scattering Removal and Noise Reduction in CBCT In single-slice CT the X-ray beam and the detector span a narrow plane. Most of the scattered photons fall outside the plane. Only a small portion of the photons scattered in the plane are recorded. In cone-beam or multidetector CT the X-ray is collimated to a wide beam. The possibility of a scattered photon mixing with the primary photons is therefore much higher. The projection images from cone-beam or multidetector CT are therefore much noisier than those from single-slice CT, and the uncertainty in the reconstruction is correspondingly higher, which brings higher noise to the reconstructed image. Research in scattering removal techniques has shown that correction of the scattering could be carried out in several ways. If the CT systems allows object–detector distance variation, setting the distance larger helps to reduce the scattering effect, because the primary beam goes along the path defined by the desired geometry and the scattered photons move in large angles. The scattering effect can be measured by setting a small opaque shield between the source and the object, to block the primary beam from reaching the detector, so that the signal in the shielded area is all from the scattering from the object. A deconvolution algorithm in image processing based on a scattering model can be used to separate the primary beam and the scattered components in projections or sinographs. Monte Carlo simulations can fit the images as the summation of the primary beam and the scattered component. The detailed descriptions of these techniques can be found in the work by Zhu et al. [13] and the references quoted. A typical method to reduce the noise is to average a pixel in an image with its neighbors. This method also reduces the spatial resolution because it smoothes out the edges. In the work reported by Wang et al. [14], an algorithm was proposed that uses an anisotropic filter to reduce the noise. In this algorithm a pixel in a sinograph is replaced by the weighted average of the neighboring pixels. The weight is proportional to the exponential of difference between the pixel under consideration and its neighbor, so that the larger the difference the less the weight, to conserve the edges. After several iterations of such averaging the random noise can be significantly suppressed while the features remain. Numerical experiments using phantom images showed that faint features and a small spot originally flooded in noise in the original images can clearly be seen in the processed images. Advances in scattering removal and noise suppression techniques make it possible to use a lower dose to obtain sufficient image qualities in cone-beam CT and multidetector CT. The cost of using the techniques is a longer after-processing time.
5.4 SPECIAL CT SYSTEMS FOR DIFFERENT CLINICAL AND RESEARCH APPLICATIONS 5.4.1 Cardiac CT Since the advent of the first CT system for head imaging, scientists and engineers have been working on a more challenging task, cardiac CT. The challenge here is tremendous, and it stems from the restless, fast, and complex three-dimensional motion of the heart. To
P1: OTA/XYZ P2: ABC c05 JWBS037-Chen January 31, 2011
23:43
Printer Name: Yet to Come
SPECIAL CT SYSTEMS FOR DIFFERENT CLINICAL AND RESEARCH APPLICATIONS
121
FIGURE 5.8 Dynamic Spatial Reconstructor at the Mayo Clinic in Rochester, Minnesota (http://dpi.radiology.uiowa.edu/gallery/dsr.html).
get a clear CT image of a living heart requires significant advancement in both hardware and software of routine CT scanners. The most significant example of endeavor in hardware improvement for cardiac CT should perhaps be the Dynamic Spatial Reconstructor (DSR), which was installed at the Mayo Clinic in 1983. In CT history the DSR is known for its speed, weight, size, and complexity. The DSR was designed mainly for cardiac imaging. It has the ability to obtain up to 240 contiguous 0.9-mm thick sections in a time period as short as 1/60 second and to repeat this acquisition rate 60 times per second. The DSR consists of a gantry weighing approximately 17 tons with a length of 20.5 feet and a diameter of 15 feet. Fourteen X-ray guns reside in a hemicylindrical configuration. Figure 5.8 shows an artistic view of the DSR. Following the DSR, the next milestone in cardiac CT development was electron beam CT. All currently running EBCT systems are mainly for cardiac imaging. All main CT system manufacturers pushed out their fastest CT systems for cardiac applications, including a 64-slice detector with 0.33 second of rotation time. These newer CT scanners generate cardiac images with higher temporal and spatial resolution. Smarter data acquisition, preprocessing, and image reconstruction techniques have become increasingly available for cardiac CT images. A recent development of these techniques is the Matched Cardiac X-ray CT (MCCT). Figure 5.9 shows a diagram of the MCCT. Clinical studies have showed that reasonably good cardiac CT images can be obtained with the MCCT for most medical conditions with three heartbeat in routine CT scanners with a gantry rotation time of 1 second. 5.4.2 Micro-CT for Animal Imaging Experimental animals, especially small animals such as mice and rats, are increasingly being used to model human diseases [16] in the study of specific pathways for disease, potential therapies, and safety of new pharmaceuticals. In traditional techniques,
P1: OTA/XYZ P2: ABC c05 JWBS037-Chen January 31, 2011
122
23:43
Printer Name: Yet to Come
COMPUTED TOMOGRAPHY AS A TOOL FOR ANATOMICAL AND MOLECULAR IMAGING
FIGURE 5.9 Diagram of the MCCT (U.S. Patent 6233478, Apparatus and Method for Constructing Computed Tomography Image Slices of an Object Undergoing Cyclic Motion, May 15, 2001. Pingyu Liu) [15].
observation of disease development is based on the sacrifice and dissection of the studied animals. Such approaches are labor intensive and don’t permit longitudinal studies. The development of in vivo imaging technology in small animals revolutionarily altered the situation. The special challenge in animal imaging comes from the small size of the animals. The typical size of a mouse is a few centimeters, which is about 5% of human size (Fig. 5.10). The respiration period of a mouse is about 10% of that of human beings. A CT system designed for small animals, the micro-CT system, therefore requires both spatial and temporal resolutions an order of magnitude finer compared with a human system. A micro-CT scanner is not a simple scaling of a human CT scanner. Special considerations have to be taken into account from high resolution requirements and the associated radiation dose increase to the imaged animals. A micro-CT scanner can rotate either the object or the gantry. In rotate-object animal micro-CT systems, to avoid organ movement during rotation, the animal is set in the vertical position. This position is unnatural for animals and it induces biological effects. Most animal micro-CT systems rotate gantries, in which the animal is set in the more natural horizontal position. Cone-shaped X-ray beam, two-dimensional detector, and one-revolution gantry are most commonly used in animal CT systems [17], although slip-ring CT systems are also available in the market [18, 19]. The detector sensitivity is proportional to pixel size. The high spatial resolution of a micro-CT system requires much smaller pixel size than the detector on a human system. A typical pixel size of micro-CT detector is 50 m. The small area of the detector pixel requires higher flux density of the X-ray beam to obtain significant signal-to-noise ratio, which could result in a dose delivered to the experimental animal that is significant enough to alter its biology. In some systems if the accumulated imaging time is several hours the total dose delivered to the animal could reach the lethal level.
P1: OTA/XYZ P2: ABC c05 JWBS037-Chen January 31, 2011
23:43
Printer Name: Yet to Come
SPECIAL CT SYSTEMS FOR DIFFERENT CLINICAL AND RESEARCH APPLICATIONS
123
FIGURE 5.10 A micro-CT image of a mouse head, acquired using a GE micro-CT with 2 × 2 binning. The X-ray tube acceleration voltage is 70 kV and the tube current is 25 mA.
A special technical barrier of a two-dimensional X-ray detector is its read-out time, as described for cone-beam CT. The development of high-speed electronics is therefore urgently required. One of the factors influencing spatial resolution is the X-ray source focal spot size. The spot size of animal micro CT varies from 25 to 900 m. If the projection of the focal spot to the detector plane has a size comparable to or larger than the detector pixel size, a penumbral blurring of the image will occur. On the detector the penumbra width of a sharp-edged object is Penumbra width =
ODD × spot size SOD
where ODD is the object–detector distance and SOD is the source–object distance. On the other hand, for the object we can estimate its image size on the detector surface: SDD ODD × object size Image size = × object size = 1 + SOD SOD Where SDD = SOD + ODD is the distance from the X-ray source to the detector. Choosing smaller ODD and longer SOD is helpful in obtaining a sharper image through reduction of the penumbral blurring, if the image resolution is sufficient. Smaller source spot size can produce images with higher spatial resolution. However, it restricts the electron beam current because of the limit of the heat dissipation rate on the
P1: OTA/XYZ P2: ABC c05 JWBS037-Chen January 31, 2011
124
23:43
Printer Name: Yet to Come
COMPUTED TOMOGRAPHY AS A TOOL FOR ANATOMICAL AND MOLECULAR IMAGING
anode. A lower electron beam will generate lower X-ray fluence; therefore a longer imaging time would be required in order to create enough signal above the noise floor. As in the human system, short imaging time might be crucial in special cases as in respiration-gated imaging and when some contrast agents are applied whose wash-out time is within one or a few minutes. In practice, the selection of the spot size must be a compromise between the requirements of the X-ray fluence and the spatial resolution. 5.4.3 Dual-Modality Systems A CT system can provide detailed anatomic information for a body. Functional imaging systems such as positron emission tomography (PET) and single photon emission computed tomography (SPECT) are capable of providing dynamic metabolic information for the body, but not detailed information about the body’s anatomy. To associate the spatial distribution of the metabolic activities with the anatomic structure the subject can be scanned using CT and PET or SPECT separately, and then the two images can be combined through image registration. The subject must not move in -between the two scans. With a dual-modality system combining CT and PET or SPECT, such registration becomes simple. Also, dualmodality systems used together with molecular-specific agents play important roles in molecular imaging studies and applications, as discussed in more details in Section 5.5. In a PET–CT dual-modality system the two units are placed next to one another [20]. Figure 5.11 shows a Siemens Inveon microPET/CT imaging system with a bore size of 10 cm in diameter. Its main study objects are small animals. Note that in the picture a dedicated PET imaging system is in front of a CT imaging system. The animal couch moves through one modality to another. The system is capable of imaging a variety of radionuclides such as 18 F, 15 O, 13 N, and 11 C. The spatial resolution for PET imaging is 1.4 mm, and the X-ray CT system is on the order of 50 m and can achieve a resolution of 15 m depending on the size of the specimen . A PET scan is about 7–14 min, and a CT scan is 10–12 min. The two images are reconstructed separately. The user can fuse them or display them individually.
FIGURE 5.11 Siemens Inveon microPET/CT. Its bore size is 10 cm in diameter and can accommodate animals up to the size of a small rabbit.
P1: OTA/XYZ P2: ABC c05 JWBS037-Chen January 31, 2011
23:43
Printer Name: Yet to Come
SPECIAL CT SYSTEMS FOR DIFFERENT CLINICAL AND RESEARCH APPLICATIONS
125
SPECT uses radiation isotopes having half-life times of several hours to a few days and emitting ␥ -rays of about 100 keV, most commonly 99m Tc, 111 I, and 123 I. Glucose or drugs coded with these isotopes are injected into the patient and follow the blood flow to the target organ. The ␥ -rays are detected by a gamma camera, which consists of a lead or tungsten collimator and an array of photon multiplier tubes. Since a gamma camera can cover only a small solid angle, usually two to four such cameras are used in SPECT. During the CT scan, the SPECT system is turned off and shields are moved to the front of the highly sensitive gamma cameras, to avoid bombarding the photon multiplier tube with the high-intensity X-ray beam. When the SPECT system is on, the X-ray tube of the CT system is off to avoid possibly scattering photons into the gamma camera. 5.4.4 Dual Energy CT As shown in Eq. (5.5), the X-ray attenuation by a material depends on the density and the mass attenuation coefficient / of the material. In some applications only the geometry information is required, as in bone fracture diagnostics, in which attention is paid to increasing the contrast between different structures. In other applications, as in radiation treatment planning, both the densities and elemental components of tissues are important because the dose absorption is defined by the two. In regular conditions the density and elemental compositions are related and an empirical tissue model can be applied to find them from the CT numbers. A CT system with two energy spectra, on the other hand, can provide more direct information about the material properties [21]. In dual energy CT the same object is imaged twice with different X-ray tube acceleration potential settings, say, 80 and 140 kVp, respectively. The tube current of lower energy is usually adjusted higher than that of the higher energy so that the contrasts of the two images can be comparable. An algorithm proposed by Torikoshi et al. [22] modeled the linear attenuation coefficient of a CT voxel as = e [Z 4 F(E, Z ) + G(E, Z )]
(5.11)
where e Z 4 F(E, Z) and e G(E, Z) denote a photoelectric term and a scattering term, respectively, which can be derived from quantum mechanics; e is the electron density; Z is the atomic number; and E is the photon energy. Measuring the linear attenuation coefficients with two different energy X-rays, we obtain simultaneous equations with respect to the unknown variables of e and Z as follows: (E 1 ) = e [Z 4 F(E 1 , Z ) + G(E 1 , Z )] (E 2 ) = e [Z 4 F(E 2 , Z ) + G(E 2 , Z )]
(5.12)
Therefore if the X-ray beam is monoenergetic, Z4 =
(E 2 )G(E 1 , Z ) − (E 1 )G(E 2 , Z ) (E 1 )F(E 2 , Z ) − (E 2 )F(E 1 , Z )
(5.13)
and the effective value of Z can be solved numerically. If the X-ray beam energy expands to a spectrum, the equation becomes [23] [S1 (E)Z 4 F(E, Z ) + G(E, Z )]d E 1 =0 (5.14) − 2 [S2 (E)Z 4 F(E, Z ) + G(E, Z )]d E
P1: OTA/XYZ P2: ABC c05 JWBS037-Chen January 31, 2011
126
23:43
Printer Name: Yet to Come
COMPUTED TOMOGRAPHY AS A TOOL FOR ANATOMICAL AND MOLECULAR IMAGING
where S1 (E) and S2 (E) are the spectra of the two energy settings, respectively. Since all the functions in Eq. (5.14) are known, the ratio 1 /2 can be tabulated as the function of effective Z. Thus the Z image can be obtained from the CT images of two energy settings. Then by either form of Eq. (5.12) the electron density e can be mapped out. The capability of differentiating the effective Z of different tissues in a dual-energy CT image makes it possible to display clear soft tissue contrast. The challenges come from the facts that the effective Z solved from the Eq. (5.14) is highly sensitive to the uncertainty of 1 /2 , and the spectra of the two energy beams need to be known in high accuracies [40]. Dual-energy CT has been available in market, as introduced by Reference [41] for example. With these systems it is possible to “remove” out some structures, such as bone, from the reconstructed images and display the blood vessel only, which brings significant advantages in the clinical diagnostics.
5.5 CT CONTRAST MEDIA AND MOLECULAR CT The intrinsic contrasts of different types of tissues are often not sufficient for radiologists to make definitive diagnoses or to accurately determine the extent of diseased tissue or certain normal structures (e.g., blood vessels or prostate gland) in X-ray and CT imaging. Various contrast media have been developed over the years and used along with the X-ray or CT imaging. There are two types of X-ray contrast agents currently approved for human use: barium sulfate suspensions, which are used for GI tract imaging, and water-soluble aromatic iodinated contrast agents, which are used primarily for imaging of the blood vessels and/or urinary system. It is estimated that the annual uses of barium suspensions and iodinated media in the United States are 5 and 20 million, respectively. Despite the tremendous success of these iodinated contrast media, the search for better contrast agents has never ceased. The major problems with existing contrast media are that they have very low retention rate and are not tissue specific. In some rare cases, the iodinated media can be severely toxic, as manifested with cardiovascular, anaphylactic, and pain reactions. A novel contrast agent that circumvents these problems is thus highly desirable. In recent years, due to the advancement of nanotechnology, some new contrast agents have emerged and showed their promising future. The field of nanotechnology has experienced significant advance in synthesis, characterization, and novel applications for various nanoscale devices. One of the important areas of nanotechnology is the synthesis and characterization of polymer, metallic, or magnetic particles, which exhibit a variety of novel properties and functions with the possibility of molecular sensing and imaging. Interfacing nanotechnology, biotechnology, and medicine has opened new vistas for biomedical research and provided an unprecedented opportunity to achieve a fundamental understanding of biological processes and contribute to disease prevention, detection, and therapy. In the last decade, the magnetic nanoparticles for MRI T2 enhancement have been extensively studied and the results showed significant promise for clinical applications. In parallel, the potential of using metallic nanoparticles has been employed to augment the intrinsic contrasts of X-ray and CT imaging. While it is a known fact that heavy elements are efficient absorbers of X-rays, the use of metallic nanoparticles as contrast media has yet to be achieved. X-ray CT is among the most convenient and widely used imaging tools in hospitals today. However, in contrast to magnetic resonance imaging (MRI) and various nuclide imaging modalities, CT is generally not considered as a molecular imaging modality since targeted and molecularly specific contrast agents have not been fully developed. But this situation is changing. Popovtzer’s group synthesized gold nanorods (AuNRs) and conjugated them
P1: OTA/XYZ P2: ABC c05 JWBS037-Chen January 31, 2011
23:43
Printer Name: Yet to Come
CT CONTRAST MEDIA AND MOLECULAR CT
127
CT attenuation (HU)
HU (Hounsfield units)
240 200 160 120 80 40
La
ra l
O
ry
nx
ca
nc
er
+A La uN ca ry R nc nx +A er ca +A 9 nc u N O er R ra (w +A lc ith 9 an La ou c ry er tA nx (w uN ca ith R nc ou ) O er tA ra + uN lc Au an R N ) R ce +K r+ H Au R Fi N I-3 R br +K ob H la R st M I-3 G +A el ol u a d N n na om R +A no a+ 9 ro Au ds N in R +A wa 9 te rs ol ut io n W at er
0
FIGURE 5.12 CT attenuation (HU) of SCC head and neck cancer cells and positive and negative control samples. Bar graph with standard deviation of three samples: larynx and oral cancer cells that were targeted with A9 antibody-coated gold nanorods (AuNRs), larynx and oral cancer cells without gold nanorods, and larynx and oral cancer cells targeted with nanorods that are coated with nonmatching antibodies (KHRI-3); normal fibroblast and melanoma cells targeted with A9 antibodies, bare gold nanorods in water solution (2.5 mg/ mL), and water [21].
with UM-A9 antibodies, which home specifically to squamous cell carcinoma (SCC) head and neck cancer [24]. Two SCC human head and neck cancer cell lines (106 cells/mL) were used: oral cancer UM-SCC-1 and larynx cancer UM-SCC-5. Both cancerous cell lines were shown before to have a significant overexpression of the A9 antigen. CT imaging was performed on the SCC cells, which were targeted with the UM-A9 antibody-coated gold nanorods. When compared with the CT number of control samples, they found that the attenuation coefficient for the molecularly targeted cells is over 5 times higher than for identical but untargeted cancer cells or for normal cells, as shown in Figure 5.12. Kim et al. [25] also utilized gold nanoparticles (GNPs) as a contrast agent for X-ray CT imaging. They prepared uniform GNPs (similar to 30 nm in diameter) by general reduction of HAuCl4 by boiling with sodium citrate. The resulting GNPs were coated with polyethylene glycol (PEG) to impart antibiofouling properties, which extends their lifetime in the bloodstream. Measurement of the X-ray absorption coefficient in vitro revealed that the attenuation of PEG-coated GNPs is 5.7 times higher than that of the current iodinebased CT contrast agent, Ultravist. Furthermore, when injected intravenously into rats, the PEG-coated GNPs had a much longer blood circulation time (>4 h) than Ultravist (<10 min). Consequently, CT images of rats using PEG-coated GNPs showed a clear delineation of cardiac ventricles and main vessels. On the other hand, relatively high levels of GNPs accumulated in the spleen and liver, which contain phagocytic cells. Intravenous injection of PEG-coated GNPs into hepatoma-bearing rats resulted in a high contrast (similar to two-fold) between hepatoma and normal liver tissue on CT images. These results suggest that PEG-coated GNPs can be useful as a CT contrast agent for a blood pool and hepatoma imaging.
P1: OTA/XYZ P2: ABC c05 JWBS037-Chen January 31, 2011
128
23:43
Printer Name: Yet to Come
COMPUTED TOMOGRAPHY AS A TOOL FOR ANATOMICAL AND MOLECULAR IMAGING
FIGURE 5.13 An artistic view of a fullerene model of C60 with a heavy atom (center ball) inside the cage. (Reproduced from http://en.wikipedia.org/wiki/Endohedral fullerenes.)
Besides gold nanoparticles, other heavy metal nanoparticles have also found application in X-ray CT. Fullerenes are a new carbonic allotrope having a cage structure. Miyamoto et al. [26] investigated whether fullerenes containing one or two atoms of heavy metals could be used as X-ray contrast material. One or two atoms of dysprosium (Dy), erbium (Er), gadolinium (Gd), europium (Eu), and lutetium (Lu) were encapsulated into fullerene (C82 ), which was synthesized as a polyhydroxyl form (e.g., Gd@C82 (OH)n , n = 40, Gd— fullerenols). A model of a fullerene with one metal atom inside its cage is shown in Figure 5.13. It was found in their study that the concentrations and the CT numbers were relatively low for all solutions, and the correlation was not clear between the concentration and the CT number of the solutions. The group, however, was still optimistic about the approach and stated that “if nanotechnology progresses in the near future, it may prove to have a possibility as an X-ray contrast material.” Rabin et al. [27] developed long-circulated bismuth sulfide nanoparticles (BPNPs) as an X-ray CT imaging agent. Rabin’s group reported that they made use of a polymercoated Bi2 S3 nanoparticle preparation as an injectable CT imaging agent. This preparation demonstrates excellent stability at high concentrations (0.25 M Bi3+ ), high X-ray absorption (fivefold better than iodine), very long circulation times (<2 h) in vivo, and an efficacy/safety profile comparable to or better than iodinated imaging agents. They showed the utility of these polymer-coated Bi2 S3 nanoparticles for enhanced in vivo imaging of the vasculature, the liver, and lymph nodes in mice. CT values of different BPNP concentrations from their experiments are shown in Figure 5.14. A new nanoparticulated iodine contrast agent was prepared and tested in cardiovascular research with X-ray CT. A research team led by Fayad developed a new technique for early detection of high-risk plaque in coronary arteries using a contrast agent called N1177 containing iodinated nanoparticles [28]. High-risk plaque is characterized by its cellular and biologic structure. High-risk plaque rich in macrophages or cells can rupture, eventually causing a heart attack or stroke. Early identification of high-risk plaque in coronary arteries may be useful in preventing cardiac events but one major hurdle in detecting high-risk plaque
P1: OTA/XYZ P2: ABC c05 JWBS037-Chen January 31, 2011
23:43
Printer Name: Yet to Come
CT CONTRAST MEDIA AND MOLECULAR CT
5,000
129
2.36 M iodine (5,100)
3,000
0.55 M bismuth
CT value (HU)
4,000
2,000 PBS BPNP Iodine 1,000 Water (0)
0
Air (–1,000)
–1,000 0.001
0.01
0.1
1
Bismuth concentration (M)
FIGURE 5.14 X-ray absorption of BPNPs. Experimental CT opacity of bismuth solutions as a function of concentration. The horizontal dashed lines indicate the opacities of air, water, and 2.36 M (300 mg I/mL) iodine contrast agent, for comparison. Inset: X-ray fluoroscopy of tubes containing PBS (phosphate buffer saline), 0.5 M Bi2 S3 nanoparticle suspension (BPNPs), and 2.36 M iodine contrast agent (Iopromide). The error bars represent the standard deviation in the CT value of a set of 30 × 30 × 1 (900) boxels on a planar square aligned with the center of the sample receptacle [24].
is the lack of an imaging modality that allows physicians to see the composition of dangerous plaque. They tested the contrast agent N1177 for the detection of macrophages in an animal model with 64-slice CT. High-risk plaque in this animal model contained high levels of macrophages that are similar in size and content to human coronary plaques. Researchers compared the enhancement of macrophage-rich plaque after the injection of N1177 and a conventional CT contrast agent. The enhancement of the macrophage-rich plaque after the injection of N1177 was significantly higher and specific inside the vessel wall than after injecting the conventional CT contrast agent. Figure 5.15 shows axial views of the same atherosclerotic plaque (white arrowheads) in the aorta of a rabbit, obtained by CT before (a), during (b), and 2 h after the injection of N1177 (c) or a conventional contrast agent (d). Recently, Lee and Chen published a comprehensive review of dual-modality probes in molecular imaging [29]. This article introduced various emerging approaches enveloped by the term dual-modality probes with respect to their design strategies and applications for all medical imaging modalities, including PET, SPECT, MRI, CT, ultrasonography, optical CT, and photoacoustic CT. Besides extensive listing of current dual-modality imaging probes, they also discussed X-ray CT contrast agents containing bismuth, gold, and gadolinium. Alric and colleagues reported gadolinium chelate-coated gold nanoparticles as dual-imaging probes for CT–MRI [30]. They synthesized these particles by encapsulating gold nanoparticles within a multilayered DTPA shell, which is composed of gadolinium chelates bound to each other through disulfide bonds. The Au@DTDTPA-Gd nanoparticles showed dose-dependent X-ray absorption properties and relaxivities. The particles freely circulated in the blood pool without undesirable accumulation in the lung, liver, and spleen followed by either radiography or MRI. Given that surface modification of gold nanoparticles can be achieved by covalent grafting of targeting moieties, targeted Au@DTDTPA-Cid dual-imaging probes can be developed.
P1: OTA/XYZ P2: ABC c05 JWBS037-Chen January 31, 2011
130
23:43
Printer Name: Yet to Come
COMPUTED TOMOGRAPHY AS A TOOL FOR ANATOMICAL AND MOLECULAR IMAGING
(a)
(b)
(c)
(d)
FIGURE 5.15 Imaging of macrophages in atherosclerotic plaques. (A–D) Axial views of the same atherosclerotic plaque (white arrowheads) in the aorta of a rabbit, obtained by CT before (a), during (b), and 2 h after the injection of N1177 (c) or a conventional contrast agent (d). These images were acquired in two separate imaging sessions. Before the injection of the contrast agent, the atherosclerotic plaque could not be differentiated from the surrounding tissues, whereas a strong enhancement was detected in the atherosclerotic plaque after the injection of N1177 (but not the conventional contrast agent) [25].
It is interesting that the future molecular X-ray CT contrast agent should be a full-function entity, which combines targeting, imaging, and therapy functions in one nanoparticle. Correspondingly, the nanoparticle contains three parts in its structure: disease-specific targeting part, X-ray absorber, and disease-specific drug. A schematic diagram of the concept is shown in Figure 5.16. When treating a patient with a certain disease, first, a full-function contrast agent for the disease is selected. Then the patient is administered the agent. The disease-specific targeting part of the agent nanoparticle leads the whole nanoparticle to the diseased tissue. When most contrast agent particles are aggregated on the diseased tissue, an X-ray CT scan or radiography is taken. The images would yield
X-ray absorber
Disease-specific targeting part Disease-specific therapeutic agent
FIGURE 5.16 Schematic diagram of the future full-function molecular CT contrast agent.
P1: OTA/XYZ P2: ABC c05 JWBS037-Chen January 31, 2011
23:43
Printer Name: Yet to Come
MEDICAL APPLICATIONS OF CT
131
clinically valuable information about the size and severity of the disease. At the same time, the disease-specific therapy agent, which may be a drug or a radiation source, is activated to attack the diseased cells.
5.6 MEDICAL APPLICATIONS OF CT Modern X-ray CT is an excellent tool to show how powerful a joint effort of physics, mathematics, medicine, and engineering could be. Since X-ray CT is a noninvasive examining and measuring method, it finds a wide range of applications in medicine, bioengineering, pharmaceutics, heavy machinery, geology, and archeology. Here we just briefly discuss CT’s medical application. 5.6.1 Diagnosis Among all the applications, the primary function of CT is to diagnose abnormal structures in a body. As X-ray radiography, CT is a noninvasive diagnostic technique. With CT the cross sectional image can be clearly displayed without overlapping, enabling easy differentiation of the anatomic structures. CT images are digital, which makes their archiving and transferring much simpler than traditional X-ray films. The process of CT imaging is fast and safe. The spatial resolution can be adjusted for different applications A CT image is a map of X-ray attenuation. We have seen (Section 5.2.1) that the linear attenuation coefficient of a material is proportional to its density. A medical CT scan is carefully calibrated; its relation to the density is very stable and is sensitive to the change of electron densities. Therefore CT images show strong contrasts at material interfaces where the density abruptly changes, such as at the boundaries of body and ambience, bone and soft tissue, cranial bones and brain, and thorax muscles and lung. In a CT image these structures can easily be outlined, and abnormalities can be clearly located and associated to anatomic structures. The attenuation of a material in a CT image is displayed as its Hounsfield number. The Hounsfield numbers of all normal tissues are well known, so a change of tissue density or tissue material can easily be detected and quantified. Many diseases cause tissue and organ structural change. Shown in Figure 5.17 are tumors in the lung and sternum. The density of
FIGURE 5.17 Chest cross sections. (Left) A cross section of the right lung is entirely filled by a tumor. (Right) An abnormal mass is displayed in the sternum. In these images the letter I shown in the blue boxes indicates the views are from inferior, and the scale on the right-hand side of the images shows centimeters.
P1: OTA/XYZ P2: ABC c05 JWBS037-Chen January 31, 2011
132
23:43
Printer Name: Yet to Come
COMPUTED TOMOGRAPHY AS A TOOL FOR ANATOMICAL AND MOLECULAR IMAGING
FIGURE 5.18 (Left) Edema in right brain shown as volume with density close to water. (Right) A dense tumor in the left brain.
a normal lung is close to that of air. Therefore a solid lung tumor is clearly distinct from the normal lung tissue, as shown on the left of Figure 5.17. On the right, the tumor is revealed as an abnormal contour in the sternum region. On the left of Figure 5.18, a volume of edema can be observed as its density is lower than that of normal brain tissue. In the right a tumor is located because its density is higher than the normal gray matter. 5.6.2 Digitally Reconstructed Radiograph As a stack of 2D CT images is reconstructed from projections; a projection can also be reconstructed from the data contained in the stack of images. A digitally reconstructed radiograph (DRR) image looks like the image from an X-ray radiograph, but it can be created from different directions and any specified focal locations. A DRR is especially useful in radiation therapy simulation. In external X-ray beam radiation treatment the source comes from a point. It is essential to place the source at a number of directions so that the beams can focus at the tumor but avoid the organs at risk. Sometimes blocks made by heavy metals are placed in front of those organs to block the X-ray beam. A “beam eye’s view” DRR in which the focus location is set to the source point is constructed, from which the doctor can clearly see if the tumor is in the field of view and if the organs at risk can be blocked or are outside of the X-ray beam. By rotating the source around the body, DRRs from different directions can be investigated to select the best angles for radiation delivery; then the locations and the shapes of the blocks can be outlined on these DRRs. DRRs are also useful for guiding the patient setup before an actual treatment. An example of a DRR is shown in Figure 5.19. 5.6.3 Radiation Therapy In modern radiation therapy, a series of 2D images of the patient body are taken with CT, MRI, or ultrasound. The radiation dose can be delivered using an external beam or a
P1: OTA/XYZ P2: ABC c05 JWBS037-Chen January 31, 2011
23:43
Printer Name: Yet to Come
MEDICAL APPLICATIONS OF CT
133
FIGURE 5.19 CT image of a lung tumor shown in a radiation oncology treatment planning software: (top left) Transverse view, (bottom left) coronal view, (bottom right) sagittal view, and (top right) DRRs from different directions. The tumor is delineated by red loops in the views.
source placed inside the body. To ensure precise delivery of the prescribed radiation dose, a computer simulation is performed using the images of the patient for a specific type of treatment, and the sources are arranged in a way that optimizes the dose in the target and spares the surrounding tissues. In most cases CT images are used in the simulation. Thanks to the physics studies of radiation–tissue interaction over the past few decades, the dose deposited to the tissues along a path of the radiation beam can be predicted based on information provided by the CT images. Figure 5.19 shows a radiation treatment plan simulation, in which the target is irradiated by external X-ray beams from several directions. Mega voltage X-rays from a linear accelerator (linac) are commonly used in externalbeam radiation therapy because of their high penetration power. In a modern linac, an on-board kilovolt cone-beam CT device is usually installed for patient setup and to monitor the day-to-day anatomy change during a course of treatment [31–33].
5.6.4 CT Angiography With contrast agents injected into the bloodstream, detailed structures of artery and vein systems can clearly be displayed, as shown in Figure 5.20, which is crucial to detect vascular disorders such as malignant tumors. CT angiography requires short imaging time and has high spatial resolution. The multidetector helical CT device makes this application possible. The surface rendering in computer graphic technique makes the 3D blood vessel images appear more realistically. In some images the bone structures can be removed from the dataset to ease the evaluation of the vessels [21].
P1: OTA/XYZ P2: ABC c05 JWBS037-Chen January 31, 2011
134
23:43
Printer Name: Yet to Come
COMPUTED TOMOGRAPHY AS A TOOL FOR ANATOMICAL AND MOLECULAR IMAGING
FIGURE 5.20 Volume-rendered reconstruction of cranial vessels detected using iodine contrast agent [21]. The osseous structures are removed.
5.7 FUTURE PERSPECTIVES In this chapter various aspects of CT imaging and its applications are described. Recently, a CT scanner with flat-panel detector has become commercially available, which has the potential to push the spatial resolution of human CT imaging to a few hundred microns. Exciting advancements have also been made in image reconstruction methods. Thanks to the development of computer technologies (such as multicore PC and GPU) over the years, computationally expensive iterative reconstruction methods are entering into routine clinical image reconstruction. The major advantages of these iterative approaches include the ability of incorporating prior system knowledge, insensitivity to noise, lower radiation dose, and effectiveness in dealing with incomplete data. Additionally, some powerful signal processing techniques, such as compressed sensing [34–36] and total variation, are finding natural applications in CT image reconstruction with sparse or incomplete projection data. Indeed, with effective use of prior knowledge and the compressed sensing method, it is now possible to reconstruct a CT image with projection data that are deemed far from sufficient by the classical Nyquist–Shannon criterion. The approach also shows significant potential for CT reconstruction with noisy (or low-dose) and 4D CT image reconstruction. Phase contrast CT represents another interesting development and further enriches the armamentarium of tomographic imaging. This imaging technique reconstructs the refractive index distribution of a weakly absorbing object from a set of tomographic projection measurements. Early development of the imaging technique had relied on the use of a bright and coherent synchrotron radiation source, which greatly limits the application of the modality [37]. However, this has recently changed. By dividing a conventional X-ray beam into many line sources via an absorption grating, it has been demonstrated that phase contrast imaging is attainable [38, 39]. The coherence occurs in the direction perpendicular to the grating lines for each line source and the gradient of the phase distribution is measured
P1: OTA/XYZ P2: ABC c05 JWBS037-Chen January 31, 2011
23:43
Printer Name: Yet to Come
REFERENCES
135
with shear interferometry. A major advantage of the phase contrast CT is its superior capability in delineation of the boundaries of the involved structures. It is anticipated that a combination of phase contrast and absorption brain CT may yield soft tissue details comparable to MRI. Finally, there is a critical need for high temporospatial resolution molecular CT imaging. While much progress has been made toward this goal, CT imaging is still considered primarily an anatomical imaging modality because of the lack of effective disease-specific probes and, more fundamentally, the inherent low sensitivity of X-ray imaging. Currently, CT has been widely used in combination with other molecular imaging techniques, such as PET and optical imaging devices, to provide useful anatomical and biological information. With the recent advancements in nanoparticle-based contrast media, the role of CT imaging in biological research is being redefined.
ACKNOWLEDGMENT The authors thank Don Russell, Bob Bramall, and Jo Acord for their valuable input to the writing of this chapter. LX wishes to acknowledge research support from the National Cancer Institute (5R01 CA98523 and 1R01 CA133474), National Science Foundation (0854492), and Varian Medical Systems (Palo Alto, CA).
REFERENCES 1. Boone, J.M.; Fewell, T.R.; Jennings, R.J. Molybdenum, rhodium, and tungsten anode spectral models using interpolating polynomials with application to mammography. Med. Phys., 1997. 24(12), 1863–1874. 2. Siewerdsen, J.H.; et al. Spektr: a computational tool for X-ray spectral analysis and imaging system optimization. Med. Phys., 2004. 31(11), 3057–3067. ¨ 3. Radon, J. Uber die Bestimmung von Funktionen durch ihre Integralwerte l¨angs gewisser Mannigfaltigkeiten. Ber. Verh. Saechs. Akad. Wiss. Leipzig Math.-Nat. Kl. 1917, 69, 262–277. 4. Liu, P.; Wang, G.; Boyer, A. Unified imaging theory for X-ray and acoustic computerized tomography. In SPIE Conference. SPIE: Denver, CO; 2004. 5. Liu, P. The P-transform and photoacoustic image reconstruction Phy. Med. Biol., 1998. 43, 667–674. 6. Kwan, A.L.C.; Boonea, J.M.; Shah, N. Evaluation of X-ray scatter properties in a dedicated cone-beam breast CT scanner. Med. Phys., 2005, 32(9), 2967–2975. 7. Ritchie, C. J.; et al. Minimum scan speeds for suppression of motion artifacts in CT. Radiology 1992, (185), 37–42. 8. Yue, G. Z.; et al. Generation of continuous and pulsed diagnostic imaging X-ray radiation using a carbon-nanotube-based field-emission cathode. Appl. Phys. Lett. 2002, 81, 355–357. 9. Ramotar, A. General geometry computed tomography reconstruction. In Computer Science. University of Waterloo: Waterloo, Canada, 2006. 10. Doheny, K. Breast CT scans could be comfortable alternative to mammograms. Breast Cancer News 2008; Available at: http://www.aolhealth.com/condition-center/breastcancer/news/article/ a/breast-ct-scans-could-be-comfortable/n20080801141509990010. 11. Li, T.; et al. Four-dimensional cone-beam CT using an on-board imager. Med. Phys., 2006, 33, 3825–3833.
P1: OTA/XYZ P2: ABC c05 JWBS037-Chen January 31, 2011
136
23:43
Printer Name: Yet to Come
COMPUTED TOMOGRAPHY AS A TOOL FOR ANATOMICAL AND MOLECULAR IMAGING
12. Li, T.; Koong, A.; Xing, L. Enhanced 4D cone-beam computed tomography using an on-board imager. Med. Phys., 2007, 34, 3688–3695. 13. Zhu, L.; Wang, J.; Xing, L. Noise suppression in scatter correction for cone-beam CT. Med. Phys., 2009. 36(3), 741–752. 14. Wang, J., et al. Dose reduction for kilovotage cone-beam computed tomography in radiation therapy. Phys. Med. Bio, 2008, 53, 2897–2909. 15. Liu, P. Apparatus and method for constructing computed tomography image slices of an object undergoing cyclic motion, U.P.a.T. Office, Editor. 2001: United States of America. 16. Badea, C.T.; et al. In vivo small-animal imaging using micro-CT and digital subtraction angiography. Phys. Med. Biol. 2008, 53, R319–R350. 17. Holdsworth, D.W.; Thornton, M. Micro-CT in small animal and specimen imaging. Trends Biotechnol. 2002, 20, S34–S39. 18. Greschus, S. Potential applications of flat-panel volumetric CT in morphologic and functional small animal imaging. Neoplasia 2005, 7, 730–740. 19. Ford, N.L.; et al. Time-course characterization of the computed tomography contrast enhancement of an iodinated blood-pool contrast agent in mice using a volumetric flat-panel equipped computed tomography scanner. Invest. Radiol. 2006, 41, 384–390. 20. Xing, L. The value of PET/CT is being over-sold as a clinical tool in radiation oncology. For the proposition. Med. Phys. 2005, 32(6), 1457–1458. 21. Johnson, T. R. C., et al. Material differentiation by dual energy CT: initial experience. Eur. Radiol. 2007, 17, 1510–1517. 22. Torikoshi, M.; et al. Electron density measurement with dual-energy X-ray CT using synchrotron radiation. Phys. Med. Bio. 2003, 48, 673–685. 23. Bazalova, M.; et al. Tissue segmentation in Monte Carlo treatment planning: a simulation study using dual-energy CT images. Radiotherapy Oncol. 2008, 86, 93–98. 24. Popovtzer, R.; et al. Targeted gold nanoparticles enable molecular ct imaging of cancer. Nano Lett. 2008, 12(8), 4593–4596. 25. Kim, D.; et al. Antibiofouling polymer-coated gold nanoparticles as a contrast agent for in vivo X-ray computed tomography imaging. J. Am. Chem. Soc. 2007, 24(129), 7661–7665. 26. Miyamoto, A.; et al. Development of water-soluble metallofullerineces as X-ray contrast media. Eur. Radiol. 2006, 16, 1050–1053. 27. Rabin, O.; et al. An X-ray computed tomography imaging agent based on long-circulating bismuth sulphide nanoparticles. Nat. Mater. 2006, 5, 118–122. 28. Hyafil, F.; et al. Noninvasive detection of macrophages using a nanoparticulate contrast agent for computed tomography. Nat. Med. 2007, 13, 636–641. 29. Lee, S.; Chen, C. X. Dual-modality probes for in vivo molecular imaging. Mol. Imaging 2009, 8, 87–100. 30. Alric, C.; et al. Gadolinium chelate coated gold nanoparricles as contrast agents for both Xray computed tomography and magnetic resonance imaging. J. Am. Chem. Soc. 2008, 130, 5908–5915. 31. Xing, L.; et al. Overview of image-guided radiation therapy. Med. Dosim. 2006, 31(2), 91–112. 32. Wiersma, R.D.; Mao W.; Xing, L. Combined kV and MV imaging for real-time tracking of implanted fiducial markers. Med. Phys. 2008, 35(4), 1191–1198. 33. Lee, L.; Le, Q.; Xing, L. Retrospective IMRT dose reconstruction based on cone-beam computed tomography and the mlc positional log-file recorded during treatment. In Int. J. Radiat. Oncol. Biol. Phys. 2008, 634–644. 34. Sidky, E.Y.; Pan, X. Image reconstruction in circular cone-beam computed tomography by constrained, total-variation minimization. Phys. Med. Biol. 2008, 53(17), 4777–4807.
P1: OTA/XYZ P2: ABC c05 JWBS037-Chen January 31, 2011
23:43
Printer Name: Yet to Come
REFERENCES
137
35. Chen, G.T.; Kung, J.H.; Beaudette, K.P. Artifacts in computed tomography scanning of moving objects. Semin. Radiat. Oncol. 2004, 14(1), 19–26. 36. Zhu, L.; et al. Using total-variation regularization for segment-based dose optimization with field specific numbers of segments. Phys. Med. Biol. 2008, 53, 6653–6672. 37. McDonald, S. A.; et al. Advanced phase-contrast imaging using a grating interferometer. J. Synchrotron Radiat. 2009, 16 (Pt 4), 562–572. 38. Pfeiffer, F.; et al. Hard X-ray phase tomography with low-brilliance sources. Phys. Rev. Lett. 2007, 98(10), 108105. 39. Pfeiffer, F.; et al. Hard-X-ray dark-field imaging using a grating interferometer. Nat. Mater. 2008, 7(2), 134–137. 40. Bazalova, M.; Carrier, J.F.; Beautlien, L.; Verhaegen, F. Dual-energy CT-based material extraction for tissue segmentation in Monte Carlo dose calculations. Phys. Med. Biol. 2008, 53(9), 2439–2456. 41. http://www.siemens.com/innovation/en/news events/innovationnews/innovationnews-articles/ ct with dual energy simplifies diagnostics.htm.
P1: OTA/XYZ P2: ABC c06 JWBS037-Chen January 25, 2011
9:11
Printer Name: Yet to Come
CHAPTER 6
Carbon Nanotube X-Ray for Dynamic Micro-CT Imaging of Small Animal Models OTTO ZHOU Department of Physics and Astronomy, Curriculum in Applied Sciences and Engineering, and Lineberger Comprehensive Cancer Center, University of North Carolina at Chapel Hill, Chapel Hill, North Carolina, USA
GUOHUA CAO Department of Physics and Astronomy, University of North Carolina at Chapel Hill, Chapel Hill, North Carolina, USA
YUEH Z. LEE Department of Physics and Astronomy and Department of Radiology, University of North Carolina at Chapel Hill, Chapel Hill, North Carolina, USA
JIANPING LU Department of Physics and Astronomy, Curriculum in Applied Sciences and Engineering, University of North Carolina at Chapel Hill, Chapel Hill, North Carolina, USA
6.1 INTRODUCTION 6.1.1 Microcomputed Tomography Computed tomography (CT) pioneered by Hounsfield [1] and Cormack [2] is one of the most important breakthroughs for X-ray imaging. CT scanners are now widely used for diagnostic medical imaging [3]. Microcomputed tomography (micro-CT) has recently emerged as a powerful noninvasive imaging tool for preclinical cancer research [4–6]. In combination with careful selection of appropriate anesthesia, radiation energy, and contrast media, micro-CT has been applied successfully to image skeletal, heart, colorectal cancer, and lung cancer models [7–9]. Compared to clinical CT scanners, a much higher spatial resolution is needed to image small animals with research interest. A typical micro-CT scanner uses a fixed-anode microfocused X-ray source and a digital area X-ray detector and operates in the cone-beam geometry (Fig. 6.1). To collect all the projection images necessary for reconstruction, the source and detector pair rotates around a stationary and horizontally positioned mouse bed.
Nanoplatform-Based Molecular Imaging Edited by Xiaoyuan Chen C 2011 John Wiley & Sons, Inc. Copyright
139
P1: OTA/XYZ P2: ABC c06 JWBS037-Chen January 25, 2011
140
9:11
Printer Name: Yet to Come
CARBON NANOTUBE X-RAY FOR DYNAMIC MICRO-CT IMAGING OF SMALL ANIMAL MODELS
FIGURE 6.1 Schematic of a cone-beam CT scanner. In a typical micro-CT setup for small animal imaging, the source–detector pair rotates around the sample stage to collect a series of projection images from which 3D structure is reconstructed.
Some scanners are also constructed, typically in experimental settings, with a stationary source and detector pair and a vertically positioned rotating object stage. The spatial resolution of the micro-CT scanner depends primarily on the X-ray focal spot size, the resolution of the detector, and the scanner geometry [4]. The temporal resolution is determined by the detector readout speed and the X-ray exposure time, which is in turn determined by the fluency rate of the X-ray source. Current commercial micro-CT scanners have several limitations that are directly due to the performance of the conventional thermionic X-ray source technology. These include low temporal resolution, high imaging dose, lack of specificity, and long scanning time. These limitations are of particular importance for high-resolution longitudinal studies of live tumor models where motion-induced artifacts and accumulation of imaging dose are major concerns. The required imaging dose strongly depends on the imaging resolution, because the signal-to-noise ratio (SNR) is primarily determined by the amount of photons received per voxel. To maintain the same SNR, the smaller the voxel size, the higher the imaging dose required. The concerns of imaging dose ultimately limit the spatial resolution that can be obtained. A dose of 2–20 cGy is generally required for a micro-CT scan. An exposure of 6 Gy is considered as lethal for a small rodent [6]. Current commercial micro-CT scanners offer the capability of imaging objects ex vivo with high spatial resolution. Studies on lung and colon cancer models have shown that micro-CT is capable of sequentially tracking large tumors with high fidelity. Small tumors are still hardly distinguishable from the normal tissue, which is partially attributed to the motion-induced artifacts created by peristalsis, respiration, and cardiac motion. The motioninduced artifacts can, in principle, be minimized by prospective gating, where the image acquisition is synchronized with the physiological signals, as is commonly done in clinical CT scanners. However, performing such measurements on small animals is challenging because their physiological motions are about ten times faster than those of humans. A comparison of the size and physiological functions between human, rat, and mouse is summarized in Table 6.1. Motion-induced artifacts blur the micro-CT images, resulting in significantly deteriorated spatial resolution than the nominal values.
P1: OTA/XYZ P2: ABC c06 JWBS037-Chen January 25, 2011
9:11
Printer Name: Yet to Come
INTRODUCTION
141
TABLE 6.1 Comparison of Size and Physiological Functions Between Human, Rat, and Mouse Parameter
Human
Rat
Mouse
Weight (grams) Breaths/minute Heartbeats/minute
70,000 12 60
250 85 300
25 120 600
Several new designs have been considered for high-resolution CT imaging of live animals. One is to use a high-power rotating anode source instead of a fixed-anode microfocus X-ray tube [8]. The high flux of the rotating anode X-ray tube enables a short exposure time, but at the expense of large focal spot size. To achieve the high resolution, the object is placed as far away from the source and as close to the detector as the photon flux allows. Rotation of the large X-ray source requires complicated gantry design, similar to clinical scanners. Recently, micro-CT scanners with dual source–detector pairs have also been developed by commercial vendors to increase the imaging speed.
6.1.2 Conventional Thermionic X-ray Although there has been a tremendous amount of innovation and improvement in the X-ray detection technology and imaging algorithms over the years, the basic mechanism of generating X-ray radiation has remained the same since the inventions by Roentgen and Coolidge. A conventional X-ray tube comprises a metal filament or a film (cathode), which emits electrons when it is resistively heated to over 1000 ◦ C and a metal target (anode) that emits X-rays when bombarded by the accelerated electrons [10]. For a thermionic tube X-ray radiation is switched on and off by either changing the filament temperature, which is very slow, using a grid to control the electron beam current, which has the problem with space charge accumulation, pulsing the anode high voltage, or using an external mechanical shutter. Grid-controlled tubes are not commonly used for microfocused X-ray tubes, presumably due to issues with electron beam focusing. A high-speed mechanical shutter is used in some of the commercial micro-CT scanners for gated imaging [11] with limited temporal resolution. The spatial resolution of an X-ray source is determined by the size of the focal spot—the area on the X-ray anode that receives the electron beam. The intensity of the X-ray radiation is proportional to the electron beam current and the square of the acceleration voltage, and is limited primarily by heat dissipation of the anode [12]. The smaller the focal spot size the less power can be achieved. While a state-of-art high-power rotating anode medical X-ray tube can operate at ∼100 kW at an effective focal spot size of ∼1 × 1 mm, the power of a typical microfocus tube is limited to 10–100 W due to the small focal spot size. It has been shown [13] that the maximum power of a fixed-anode microfocus X-ray tube running at the continuous mode follows the relation of Pmax ≈ 1.4(X f,FWHM )0.88 , where Pmax is the maximum X-ray tube power in watts and X f,FWHM is the focal spot size in microns. Because of the limited output power, polychromatic radiation is used in essentially all clinical and preclinical imaging applications, although simulation and synchrotron studies have shown that the image quality can be significantly enhanced and the imaging dose reduced using monochromatic or quasi-monochromatic radiation.
P1: OTA/XYZ P2: ABC c06 JWBS037-Chen January 25, 2011
142
9:11
Printer Name: Yet to Come
CARBON NANOTUBE X-RAY FOR DYNAMIC MICRO-CT IMAGING OF SMALL ANIMAL MODELS
Essentially all current commercial X-ray tubes are single-pixel devices, where X-ray radiation is emitted from a single focal spot. For tomography imaging either the X-ray/detector pair or the object needs to be rotated to collect projection images from different viewing angles for CT reconstruction, which limits the scanning speed, complicates the mechanical design of the scanner, and restricts the scanner configuration. The concept of spatially distributed X-ray sources has been proposed and some have been developed, including the Dynamic Spatial Reconstructor (DSR) [14], the electron beam CT (EBCT) [15], and the scanning beam digital X-ray (SBDX) [16, 17]. The DSR system uses ∼30 individual X-ray tubes mounted on a gantry around a patient to collect the images. Although high temporal resolution has been demonstrated for cardiac imaging, the cost, size, and maintenance-related issues make the design prohibitive. Recently, a dual-source CT scanner was introduced, which uses two individual X-ray tubes that rotate simultaneously to reduce the imaging time. Both the EBCT scanner and the SBDX tube use an electromagnetic field to steer the electron beam to different spots on the X-ray target to produce a scanning X-ray beam, very much like the design of a cathode-ray-tube. Such X-ray tubes are in general large and have a limited viewing range due to the difficulties in steering the high-energy electron beam.
6.2 CARBON NANOTUBE FIELD EMISSION X-RAY 6.2.1 Electron Field Emission and Early Works on Field Emission X-ray Electron field emission is a quantum process, where under a sufficiently high external electrical field electrons can escape from the metal surface to the vacuum level by quantum tunneling [18]. Compared to thermionic emission, this is a preferred mechanism because no heating is required and the emission current can be controlled by the external field to give instantaneous response time. The basic physics of field emission is summarized by the Fowler–Nordheim equation [I = ␣V 2 exp(-b3/2 /V)], which states that the emission current (I) increases exponentially with increasing voltage (V) [18]. For a metal with a flat surface the threshold field is typically around 104 V/m, which is impractically high. All the electron field emitters rely on field enhancement () at the sharp tips/protrusions. One way to fabricate sharp field emitters is by a lithography process [19]. Such emitters have not been used in practical devices due to problems such as low emission current, poor stability, and high cost. X-ray tubes using field emission cathodes have been investigated in the past [20, 21]. In these early systems metal tips were used as the cathodes. Electrons were extracted by applying a pulsed high voltage between the target and cathode. X-ray radiation was generated when the field emitted electrons bombarded a target. The advantages of the field emission X-ray tubes compared to conventional thermionic X-ray tubes have been demonstrated in clinical studies in terms of resolution and exposure time [21]. However, the metal-tip emitters were inefficient. To obtain the high electrical field required to extract the electrons, a Max generator, which uses a series of discharging capacitors, was employed [22]. The X-ray tubes had a limited lifetime of 200–300 exposures and slow repetition rate. With a diode design, the acceleration voltage and the tube current could not be independently controlled. Field emission X-ray tubes using other types of emitters such as the Spindt tips and diamond emitter have also been investigated [23–25]. The highest electron current demonstrated in these X-ray tubes is only on the order of microamperes [25].
P1: OTA/XYZ P2: ABC c06 JWBS037-Chen January 25, 2011
9:11
Printer Name: Yet to Come
CARBON NANOTUBE FIELD EMISSION X-RAY
143
6.2.2 Carbon Nanotube Field Emitters and X-rays Carbon nanotubes (CNTs) are excellent electron field emitters [27]. Due to their atomically sharp tips and large aspect ratios (∼103 ), the CNTs have much larger field enhancement factors (), thus lower threshold fields are required for emission than with conventional emitters. They are stable at a high emission current: a stable emission current of >1 A has been observed from an individual SWNT [28] and a CNT bundle [26]; we have achieved a stable emission current density of over 1 A/cm2 from macroscopic cathodes (Fig. 6.2). This is attributed to their unique physical properties including high thermal and electrical conductivities, high temperature and chemical stability, and reasonable resistance to oxidation. These properties make them attractive as “cold” cathode materials for vacuum electronic device applications. A CNT-based field emission X-ray source has recently been demonstrated [29–32]. It utilizes the CNT field emitters as the “cold” cathode for X-ray generation. A typical CNT X-ray source comprised of a CNT cathode, a gate electrode, an electrostatic focusing unit, and a metal anode in a vacuum housing is illustrated in Figure 6.3. Field emitted electrons are extracted from the CNT cathode by applying a bias voltage on the gate electrode and are subsequently focused by the focusing lenses before reaching the anode.
Current (μA)
5 1μm
4 test 2
3 2 test 1
1
-12
2
(a)
2
0
2
6 4 Time (hrs)
8
10
-14
-16
1.0 1.5 2.0 2.5 3.0 -3
1/V (10 )
1
0
0
In(I/V )
test 3
Current Density (A/cm2)
3 6
0
4 2 Electric Field (V/μ μm)
6
(b)
(a)
Current Density (A/cm2)
3
2
1
0 0
50
100
150
Time (hrs)
(c)
FIGURE 6.2 (a) Stability of field emission current from a single CNT bundle measured at three different current levels [26]. (b) Current versus applied voltage curve and (c) high current stability from a 1-cm diameter CNT cathode. The high current data are taken from the Xintek website (www.xintek.com).
P1: OTA/XYZ P2: ABC c06 JWBS037-Chen January 25, 2011
144
9:11
Printer Name: Yet to Come
CARBON NANOTUBE X-RAY FOR DYNAMIC MICRO-CT IMAGING OF SMALL ANIMAL MODELS
Anode Electron trajectory Focusing III
Anode
Cathode assembly Focusing II Control circuit
Focusing I
CNT Cathode (a)
G
D
Ro S G
D
Ro SG
D
Ro S G
D
Ro SG
D
Ro S
(b)
FIGURE 6.3 (a) A schematic of a CNT X-ray source with a CNT cathode, a gate electrode, a focusing unit, and an anode. (b) A schematic of a multibeam field emission X-ray (MBFEX) source with five independently controllable X-ray “pixels.”
Compared to conventional thermionic X-ray tubes it has several intrinsic advantages that make it attractive for diagnostic medical imaging. Because of the nature of the field emission mechanism, the source has a fast response time and electronic programmability, which are highly desirable for gated imaging of moving objects [33]. The ability to switch the X-ray radiation on and off by readily turning on/off the electron beam enables efficient utilization of X-ray radiation and better thermal management of anode heat load. The use of a “cold” cathode rather than a “hot” filament makes it possible to manufacture field emission cathode arrays with individually controllable electron emitting pixels, which can generate X-ray radiation from different focal points on the X-ray anode. By sequentially switching on and off the individual pixels, a scanning X-ray beam can be generated to image an object from different viewing angles without mechanical motion [32, 34, 35]. Figure 6.3 shows the design of a CNT multibeam field emission X-ray (MBFEX) source that was first demonstrated in our lab [32]. The new source technology can potentially lead to a truly stationary and gantry-free CT with high scanning speed and high resolution. In addition, the MBFEX source makes multiplexing—simultaneous collection of multiple images at the same time—a reality. This is accomplished using either the binary multiplexing algorithm—activating a subset of the X-ray beams according to a preselected multiplexing algorithm—or the frequency division multiplexing algorithm [36, 37]. This potentially can lead to an orders of magnitude increase in imaging data collection speed, although the trade-off between the imaging speed and the signal-to-noise ratio needs to be investigated further. Besides our team at UNC, the CNT field emission X-ray technology is now being actively investigated by a number of academic groups and major companies in the world [30, 38–41]. Potential preclinical and clinical applications including micro-CT [42–44], stationary digital breast tomosynthesis [45], and tomosynthesis-based imaging guided radiation therapy are being developed. In the next below, we briefly introduce our work on developing CNT X-ray-based dynamic micro-CT for in vivo imaging of small animal models.
P1: OTA/XYZ P2: ABC c06 JWBS037-Chen January 25, 2011
9:11
Printer Name: Yet to Come
CARBON NANOTUBE FIELD EMISSION X-RAY
145
6.2.3 A Microfocus Field Emission X-ray Source for Dynamic Micro-CT To utilize the high temporal resolution and programming capability of the CNT X-ray we have developed a microfocus CNT field emission X-ray source using a modified Einzel type electrostatic lens [46]. As illustrated in Figure 6.4, the lens consists of three electrostatic focusing electrodes. Focusing electrodes 1 and 3 are made of planar metal diaphragms. Central focusing electrode 2 has the shape of a truncated cone. Electrode 1 is at the same potential as the gate electrode; electrodes 2 and 3 have independently controllable potentials. To obtain effective isotropic focal spot size, the CNT emission surface has an elliptical geometry. They are fabricated using a combined photolithography and electrophoretic deposition method developed in our lab [47]. An optical image of the finished CNT cathode
Anode
Cathode Structure Electron Beam
D d d = D tan θ
D=Long Axis d=Short Axis
Focus 2
Focus 1 Gate Cathode
Gate Mesh (a)
FSS (short axis) (µm)
450 400 350 300 250 200 150 100 50 0 0.0
0.5
1.0
1.5
2.0
2.5
Cathode Diameter (mm) (b)
FIGURE 6.4 (a) The electron beam optics modeling of CNT microfocus X-ray tube showing the details of the cathode, gate mesh, and focusing electrodes. The field emission cathode area is elliptical. The electron beam is uniformly focused on the anode. With a proper anode tilting the effective X-ray radiating focus spot is close to isotropic. (b) Simulated focus spot size as a function of the cathode size. Over a large range of cathode size an effective focusing power of 5 can be achieved
P1: OTA/XYZ P2: ABC c06 JWBS037-Chen January 25, 2011
146
9:11
Printer Name: Yet to Come
CARBON NANOTUBE X-RAY FOR DYNAMIC MICRO-CT IMAGING OF SMALL ANIMAL MODELS
CNT cathode
Mo/SiO2/Si
SiO2/Si (b)
(a)
Cathode and focusing feedthroughs
6 inches
mini ion pump
50-kV anode feedthrough
200-μm Be window
(c)
FIGURE 6.5 (a) A two-pixel CNT cathode on Si substrate. (b) Cross-sectional SEM image showing the arrangement of the CNTs on the cathode surface. (c) A compact CNT-based microfocus X-ray source developed at UNC.
and a cross-sectional SEM image of the CNT cathode are shown in Figure 6.5. The ratio between the elliptical long axis and the short axis is determined by the anode tilting angle. The dimensions of the focusing lenses were optimized by simulating the trajectory of the field emitted electrons. Figure 6.4 shows the relation between the cathode diameter and the focal diameter of the electron beam on the anode obtained from the simulation. The simulation results indicate that this design can give an optimum focusing factor of about 5. The effective focal spot size scales linearly with the linear dimension of the cathode. This means that for an isotropic effective focal spot size of 100 m an elliptical CNT cathode of 2.35 mm × 0.5 mm is needed at 12o anode tilting angle. Our first generation microfocus tube is housed inside a large vacuum chamber under dynamic pumping. The setup is convenient for laboratory experiments but has limited mobility, which restricts the geometry of the micro-CT scanner. Recently, we developed a compact microfocus X-ray tube (Fig. 6.5), which can readily be mounted on a rotating gantry and has similar performance as the one housed in the large vacuum chamber.
P1: OTA/XYZ P2: ABC c06 JWBS037-Chen January 25, 2011
9:11
Printer Name: Yet to Come
DYNAMIC MICRO-CT BASED ON THE CNT X-RAY TECHNOLOGY
147
Extensive studies have been carried out to evaluate the performance of this CNT-based field emission microfocus X-ray tube. The source can be operated at up to 55 kVp anode voltage, effective focal spot size down to 50 m, a tube current of 1.5 mA at 100-m focal spot size, and X-ray pulse width on the order of microseconds (although, in practice, the value is limited by the X-ray dose required for imaging).
6.3 DYNAMIC MICRO-CT BASED ON THE CNT X-RAY TECHNOLOGY 6.3.1 Design of the Micro-CT Scanner We have developed two prototype dynamic micro-CT scanners using the CNT microfocus X-ray technology at UNC. One uses the rotating object geometry and is called Cyclops; the other has a stationary and horizontally positioned mouse bed and a rotating source and detector pair and is called Charybdis. Below we briefly describe the designs of these two micro-CT scanners.
Cyclops System The Cyclops micro-CT scanner uses a stationary CNT X-ray source and stationary X-ray camera. The object is rotated to collect the projection images from different viewing angles, as illustrated in Figure 6.6. The object is suspended vertically in a custom built sample holder that is locked to a metal support attached to a computercontrolled rotation stage. The system is configured in the cone-beam geometry. The CNT microfocus X-ray source uses a 2.35-mm by 0.5-mm elliptical geometry cathode, which gives a ∼100-m focal spot size. It is capable of reliably delivering ∼1-mA peak anode current at up to 60-kV anode voltage. The camera is a high-speed CMOS flat-panel sensor with a CsI scintillator plate directly deposited on a photodiode array (Model C7940DK-02,
Camera
Mouse
CNT X-ray Source
Anode e-beam Focus 2
Rotation Trigger
Ph ys Trig i. ger
Physiological Monitoring System
Focus 1 Gate CNT
Camera Trigger
X-ray Trigger
Controlling Computer
FIGURE 6.6 A schematic of the Cyclops scanner. The CNT X-ray source placed inside a vacuum chamber and the camera are stationary. The mouse is placed vertically in a custom-built sample holder, which is rotated by a computer-controlled rotation stage. The respiration sensor is attached to the abdomen of the mouse from the top of the sample holder.
P1: OTA/XYZ P2: ABC c06 JWBS037-Chen January 25, 2011
148
9:11
Printer Name: Yet to Come
CARBON NANOTUBE X-RAY FOR DYNAMIC MICRO-CT IMAGING OF SMALL ANIMAL MODELS
Camera Trigger
0
1
1
0
1
0
1
0
1
tpw
Exposure Window Respiration Signal Physiological Trigger X-ray Trigger
2
1
3
tpw
Rotation Trigger
FIGURE 6.7 Illustrative timing diagram that the controlling computer used to gate the X-ray pulse firing and the camera frame acquisition to the respiration signal. The camera readout and integration regions are designated as 1 and 0, respectively. tpw is the temporal width of an X-ray pulse.
Hamamatsu). It has 2400 × 2400 pixels with a nominal pixel size of 50 m, giving an active area of 120 × 120 mm2 . Figure 6.7 is the timing diagram used to gate the X-ray pulse firing and the camera frame acquisition to the physiological signal from a live animal. The camera runs on a continuous, predefined frame rate. The camera frame rate, camera readout time, and the X-ray pulse width together determine the X-ray exposure window. The exposure window is defined as the time window between the end of the previous frame readout and the beginning of the next frame readout subtracted by the X-ray pulse width, tpw . This subtraction is to ensure the X-ray pulse would not overextend into the camera readout region. The physiological trigger is generated based on the respiration signal obtained from a BioVet physiological monitoring system (Spin Systems (QLD) Pty Ltd, Brisbane, Australia). The physiological trigger can be selected at any portion of the respiration cycle. The dynamic gating is achieved from the logic AND combination between the exposure window signal and the physiological trigger signal by an AND transistor-to-transistor logic (TTL) gate. The output of this AND gate is used as the trigger to generate the X-ray trigger signal that controls the X-ray pulse firing. An X-ray trigger pulse is enabled only by the rising edge of the TTL signal output from the logic AND gate. The physiological triggers from the respiratory gating of a free-breathing mouse may be nonperiodic, such as the three physiological triggers shown in Figure 6.7. The rising edges of the physiological triggers 1 and 3 come within the X-ray exposure window; thus they will each trigger an X-ray pulse. On the other hand, the physiological trigger 2 will not trigger an X-ray pulse because its rising edge comes outside the X-ray exposure window. In this gating scheme, the X-ray pulses (hence imaging sequences) are dynamically gated to the physiological signal of small animals (hence called dynamic scanner). It is worthwhile to point out that the X-ray firings are based on temporal coincidence between the X-ray exposure window and the physiological triggers. The total scan time for a dynamically gated scan critically depends on the rate of this temporal coincidence.
P1: OTA/XYZ P2: ABC c06 JWBS037-Chen January 25, 2011
9:11
Printer Name: Yet to Come
DYNAMIC MICRO-CT BASED ON THE CNT X-RAY TECHNOLOGY
149
The geometry parameters of the scanner are determined using a two-ball phantom following the procedure outlined by Noo et al. [48]. The source-to-object distance (SOD) is 135 mm. The source-to-detector distance (SDD) is 216 mm, which gives a magnification factor of 1.6. The cone angle is 14 ◦ and the field of view (FOV) of the scanner is 32 × 32 mm2 .
Charybdis Scanner The Charybdis system consists of a rotating source and detector pair and a stationary sample stage, as shown in Figure 6.8. In this design, the mouse is placed horizontally and stationary on the sample stage during scan. This configuration is preferred for in vivo measurement of the morphology and physiology of small animals. This configuration is made possible by the compact and portable CNT field emission microfocus X-ray tube recently developed in our lab [49]. The X-ray tube and the detector are mounted on two translational stages positioned on the opposite sides of a high-precision goniometer (Model 430, HUBER, Germany), which is driven by a digital stepping motor. This flexibility makes it relatively easy to optimize the scanner geometry, for example, for either high spatial resolution specimen imaging or in vivo animal imaging where a trade-off between the spatial resolution and the temporal resolution has to be considered. The X-ray detector is the same one that has been used in the Cyclops scanner. In both cases, the camera is configured to produce two-dimensional 1024 × 1024 pixels projection images at 50 × 50 m2 pixel resolution. This detector configuration together with the Charybdis scanner geometry gives an effective FOV of 39 × 39 mm2 and an effective digital sampling of 38.5 m at the object plane. The scanner is controlled by a homemade computer program written in LabVIEW. At the beginning of each CT scan, the program performs a series of initialization steps, which include homing the goniometer to its zero position. The program then instructs the X-ray tube, the detector, the goniometer, and a homemade dynamic gating electronic circuit (if gating is involved) to work sequentially so that projection images are acquired in the step-and-shoot mode. Multiple frame acquisition is also allowed for frame averaging. At the end of the scan, the LabVIEW program sends commands to rewind the goniometer to its zero position.
b
c
d
X-ray detector
Goniometer
Mouse holder
e HV power supplies
a
CNT x-ray tube (A)
(B)
FIGURE 6.8 (A) Picture of the Charybdis scanner using a compact CNT microfocus X-ray source. It mainly consists of (a) a compact CNT X-ray tube, (b) a flat-panel X-ray detector, (c) a mouse positioning stage, and (d) a goniometer. (B) CAD design of the scanner.
P1: OTA/XYZ P2: ABC c06 JWBS037-Chen January 25, 2011
150
9:11
Printer Name: Yet to Come
CARBON NANOTUBE X-RAY FOR DYNAMIC MICRO-CT IMAGING OF SMALL ANIMAL MODELS
6.3.2 Prospective Gated Micro-CT Imaging of Free-Breathing Mice
Respiratory Gated Micro-CT Imaging of Free-Breathing Mouse Respiratory motion is one of the major challenges in CT imaging of live mice. It has been the subject of numerous studies in recent years. It has been shown previously that the motions induced artifacts can be minimized by techniques including physically restricting the motion [7] and retrospective gating. In retrospective gating, projections from different phases of the respiratory cycles are taken but only the images belonging to a retrospectively selected phase are reconstructed. The method requires a significant increase in the imaging dose [50–52]. Prospective gating techniques, where the image acquisitions are synchronized with the physiological signals, have also been investigated. In most of these studies, animals are intubated and ventilator controlled [8, 53, 54]. Intubation requires dedicated equipment and special animal handling skills to avoid damage to the upper airways. The respiratory mechanics and lung function of the animal could be altered from paralysis. In addition, when animals are ventilator controlled, the pressure and volume settings can affect the morphology and mechanics, thereby making true measurements of function and structure difficult to obtain [55, 56]. Imaging of free-breathing animals is preferred for longitudinal studies and for disease models, where mechanical ventilation may introduce systematic errors into the physiologically relevant measurements. Imaging of free-breathing rodents using retrospective gating techniques has been reported in a few recent publications [50, 57]. It unavoidably requires redundant projections/exposures to minimize the missing view artifact if only one phase in the breathing cycle is to be imaged. Free-breathing rodents have also been imaged with prospective gating techniques using commercial scanners [57, 58]. However, the effective temporal resolution was only 170 ms due to limitations of the scanner. This is not sufficient to precisely define the peak inspiration phase in a mouse. In addition, each projection image had to be averaged from up to 10 frames in order to achieve a good image quality. We have recently demonstrated the potential of performing high-resolution prospective respiratory gated micro-CT imaging of anesthetized free-breathing mice using the Cyclops micro-CT scanner [59]. The imaging protocol is briefly summarized here. The scanning parameters for respiratory gated micro-CT of anesthetized free-breathing mice were 40 kVp, 0.7 mA anode current, 50-ms X-ray pulse width, 325 projections over a circular orbit of 195 ◦ with a step angle of 0.6 ◦ , and 50 bright and dark calibration images. The total scan time was typically 5–10 min depending on the mouse respiration rate. Reconstruction was done with the Feldkamp algorithm using the Cobra EXXIM software package (EXXIM Computing Corp., Livermore, CA). Data were reconstructed with Parker weighting as isotropic 512 × 512 × 512 arrays at a voxel size of 62 × 62 × 62 m3 . The system performance characteristics including the system modulation transfer function (MTF), contrast-to-noise ratio, and linearity of the system have been measured. The system modulation transfer function (MTF) measured using a 10-m diameter tungsten wire and the oversampling method proposed by Fujita et al. [60] is shown in Figure 6.9. The 10% system MTF is 6.2 lp/mm, which represents a resolution of 81 m. The measured CT imaging dose is ∼11 cGy. Respiratory-gated micro-CT imaging was performed on adult wild-type mice. The animals were anesthetized with 1–2% isoflurane delivered from a vaporizer. The animal breathing rates after anesthetization were typically in the range of 80–180 breaths/min. The anesthetized animals were placed over the pressure sensor in the mouse sample holder and secured with adhesive restraints. The animals were positioned such that the respiration
P1: OTA/XYZ P2: ABC c06 JWBS037-Chen January 25, 2011
9:11
Printer Name: Yet to Come
DYNAMIC MICRO-CT BASED ON THE CNT X-RAY TECHNOLOGY
151
1.0 MTF 10%
0.8
MTF
0.6
0.4
0.2
0.0 0
2
10 12 4 6 8 Spatial frequency (cycle per mm)
14
FIGURE 6.9 The 10% MTF of the scanner is 6.2 lp/mm, which represents a resolution of 81 m.
sensor was approximately in the position of the abdomen to achieve maximum sensor coupling. Experimental procedures carried out in this study had been approved by the University of North Carolina Institutional Animal Care and Use Committee. Each mouse was scanned with the standard scanning parameters and with respiration gated to the mouse’s peak inspiration, followed with another scan gated to its end expiration. Figure 6.10 shows the typical results from the respiratory-gated micro-CT of freebreathing mice. The small structures such as vessels and higher order branching airways can be identified in all those images. The lung in the peak inspiration images appears to have a lower CT number than that in the end expiration images. And airways in the peak inspiration images also appear larger in diameter than those in the end expiration images.
Cardiac Gated Imaging of Free-Breathing Mouse The small size (∼0.1 g and ∼7 mm in the long axis) and fast beating rate (∼600 beats/min) of the mouse heart make in vivo study of its cardiac metabolism, function, and morphology difficult. So far, micro-CT imaging of mouse heart has been demonstrated in two systems: a custom-made micro-CT scanner using a rotating mouse configuration [61] and the top-of-line micro-CT scanner from GE (Locus Ultra, General Electric). Prospective cardiac gated micro-CT imaging at 150-m spatial resolution and 10-ms temporal resolution has been reported using the first scanner. Retrospective cardiac gating at 300-m spatial resolution and 12-ms temporal resolution is reported using the second system [52]. Both systems use rotating-anode X-ray tubes to provide the high photon flux needed for the high temporal resolution (10 ms). In order to compensate the large focal spot size, the X-ray tube is placed as far away from the object and detector as the photon flux allows. The Charybdis system is the first micro-CT scanner that provides high spatial (<100 m) and temporal resolution (10–20 ms), stationary and horizontal mouse bed configuration, and prospective gating capability. The dynamical gating scheme is the same as the one used for Cyclops. The difference is the method for selecting physiological triggers. In the respiratory
P1: OTA/XYZ P2: ABC c06 JWBS037-Chen January 25, 2011
152
9:11
Printer Name: Yet to Come
CARBON NANOTUBE X-RAY FOR DYNAMIC MICRO-CT IMAGING OF SMALL ANIMAL MODELS
(a)
respiration signal
respiration signal
(a) Peak Inspiration 1400 1300 1200 1100 1000 900 800 700
(b)
4 mm 0 500 1000 (b) End Expiration
1500
2000
0
1500
2500
3000
4 mm
3500
(c)
(d)
1400 1300 1200 1100 1000 900 800 700 500
1000
2000
2500
3000
3500
4 mm
4 mm
time/ms
(A)
(B)
FIGURE 6.10 (A) Respiration signals of the mouse being imaged. Red bar indicates the 50-ms temporal windows for the peak inspiration/end respiration cycle where the CT images were acquired. (B) Reconstructed slice images show clear difference between peak inspiration (a) and (c), and end expiration (b) and (d) in the axial and coronal views, respectively.
gated case, physiological triggers were generated entirely based on the respiration signal itself, while for cardiac imaging, the physiological triggers were generated based on both respiration signal and cardiac signal. A physiological trigger is generated if and only if the first QRS complex of the ECG occurred within an acquisition window defined at a specific phase of the respiration cycle. An example for the physiological trigger selection is shown in Figure 6.11. Once the physiological triggers were successfully generated, they then are subject to the same timing diagram shown in Figure 6.7 to trigger the X-ray firings and imaging sequences. The imaging protocol was similar to that was used with the Cyclops system, except that ECG electrodes were taped to the footpads and iodinated contrast agent (Fenestra VC, Advanced Research Technologies Inc., Montreal, Canada) was infused via tail vein injection
Cardiac Signal
Respiration Signal
FIGURE 6.11 Physiological trigger selection in cardiac imaging. The red bars indicate the physiological triggers that were generated based on the temporal coincidence between the selected phase of the cardiac signal (the S wave of ECG in this case) and the acquisition window defined at the end expiration plateaus of the respiration signal.
P1: OTA/XYZ P2: ABC c06 JWBS037-Chen January 25, 2011
9:11
Printer Name: Yet to Come
CONCLUSION AND FUTURE DIRECTION
R
(a)
(b)
(c)
(d)
153
delay
Atrial Depolarization
T
P
Ventricular Repolarization
Q S
Ventricular Depolarization
(A)
(B)
FIGURE 6.12 (A) The physiological triggers were selected at 0-ms (systole) and 50-ms (diastole) delay after the first R wave of the ECG that occurred within the acquisition window. (B) The reconstructed slice images show clear differences between systole (a) and (c), and diastole (b) and (d) in the axial and coronal views, respectively.
(0.02 mL/gram of mouse) prior to imaging. For micro-CT imaging of mouse heart, 400 projections were acquired over a circular orbit of 199.5 ◦ with a stepping angle of 0.5 ◦ , and 50 bright and dark calibration images were collected. By running the detector at 1 frame/s (camera integration time = 500 ms), the scan time for a cardiac and respiratory gated mouse micro-CT was typically 15–30 min, depending on the mouse’s respiration and heart rates. Images were collected at 100-m spatial resolution and 20-ms temporal resolution. A representative result from respiratory and cardiac gated micro-CT of mouse heart is shown in Figure 6.12. As we can see, structures inside the heart are clearly delineated. The motion-induced artifacts are not present, indicating a proper gating.
6.4 CONCLUSION AND FUTURE DIRECTION The carbon nanotube field emission X-ray technology offers several intrinsic advantages over conventional thermionic X-ray tubes that are used in all current X-ray imaging systems. The programmability of the X-ray radiation, the high temporal resolution, the ability to produce a spatially distributed scanning X-ray beam, and the capability for multiplexing are highly attractive for X-ray imaging especially for X-ray tomography. After several years of intense research and development, the basic performance of the nanotube X-ray source including tube current, high voltage stability, focusing, lifetime, and reliability has reached the point where practical preclinical and clinical imaging applications can be expected. The novel X-ray source technology has also moved from a pure academic curiosity to industrial development and production (a MBFEX tube from XinRay Systems is illustrated in Figure 6.13; this particular model has 100 individually controllable X-ray pixels).
P1: OTA/XYZ P2: ABC c06 JWBS037-Chen January 25, 2011
154
9:11
Printer Name: Yet to Come
CARBON NANOTUBE X-RAY FOR DYNAMIC MICRO-CT IMAGING OF SMALL ANIMAL MODELS
(a)
3.2 cm diameter mouse holder
10 cm detector
Source array
Source array
10 cm detector (b)
FIGURE 6.13 (a) Schematic of a CNT MBFEX source with 100 individually controlled X-ray pixels produced by XinRay Systems (www.xinraysystems.com). (b) One of the proposed configurations for the stationary micro-CT scanner with two orthogonally positioned source arrays and two area detectors. (From Quan and LaLush [62].)
Next-generation imaging systems for image-guided radiation therapy, stationary digital breast tomosynthesis, and homeland security applications based on the nanotube X-ray technology are being actively developed by leading system vendors. In this chapter we summarized the ongoing research in our lab at UNC Chapel Hill on high-resolution dynamic micro-CT for in vivo imaging of small animal models. The research so far has focused on developing the field emission microfocus X-ray source and micro-CT scanner. Using a single-pixel CNT field emission X-ray source, we have demonstrated the capability to perform prospective respiratory and cardiac gated CT imaging of freebreathing anesthetized mice at 100-m spatial resolution and 20-ms temporal resolution. The configuration of the Charybdis system is very similar to the regular cone-beam CT scanner with a stationary mouse bed and rotating source and detector pair. With further development of the MBFEX technology and the imaging reconstruction algorithm, a gantryfree and stationary micro-CT scanner can potentially be constructed. Such a system would comprise a MBFEX source with several hundred microfocused X-ray beams, a panel of X-ray area detectors, and mouse bed, as illustrated in Figure 6.13. All the components are stationary. The project images will be collected by selectively activating the X-ray beams from different projection angles without mechanical motion of the object or the source. The feasibility of such a system was recently demonstrated in our lab using a prototype 20-beam microfocus MBFEX source [63]. There is also an ongoing development effort in the reconstruction algorithm [62]. Further advancement in the imaging speed is also possible using multiplexing, which allows the acquisition of many projection images at the same time [36, 37]. However, issues such as the noise propagation and cross-scattering effect remain to be studied. Advancement in the MBFEX technology will also make it possible to investigate some advanced X-ray imaging modalities currently only accessible at remote synchrotron facilities. One example is quasi-monochromatic CT. An X-ray source with a narrow energy bandwidth will improve contrast resolution and permit K-edge imaging and dual-energy
P1: OTA/XYZ P2: ABC c06 JWBS037-Chen January 25, 2011
9:11
Printer Name: Yet to Come
REFERENCES
155
imaging. Studies have also shown that a significant reduction of dose can be achieved using a quasi-monochromatic beam [64]. Although the advantages of monochromatic and quasi-monochromatic imaging have been convincingly demonstrated in synchrotron-based studies, it has so far not been adapted for preclinical and clinical imaging. The principal limitation is the prohibitive long imaging time if an X-ray tube is used [64–66]. We expect that by combining the MBFEX source with a multiplexing method, quasi-monochromatic micro-CT imaging can be performed in comparable scanning time as polychromatic microCT in the near future.
ACKNOWLEDGMENT A large number of current and former students and postdoctoral fellows in our group at UNC have worked on this interdisciplinary project over the last few years. In particular, we acknowledge contributions from Xiomara Calderon-Colon, Shabana Sultana, Rui Peng, Laurel Burk, Ramya Rajaram, Lei An, Yuan Cheng, Zejian Liu, and Jian Zhang. Without them this work would not have been possible. We also acknowledge contributions from our collaborators including Prof. David LaLush for advises on imaging reconstruction; Prof. Yue Xiong, Prof. Kay Lund, and Prof. Barbara Grubb for providing the mouse models; Prof. Eric Hoffman for valuable discussions; and Dr. Bo Gao and Dr. HuaiZhi Geng for assistance in cathode fabrication. The research on carbon nanotube field emission X-ray at UNC is generously supported by the Carolina Center for Cancer Nanotechnology Excellence (NCI U54CA119343), NIH-NIBIB (4R33EB004204-01), Xintek, Inc., and the UNC University Cancer Research Funds.
REFERENCES 1. Hounsfield, G. N. Computerized transverse axial scanning tomography: 1. Description of system. Br. J. Radiol. 1973, 46, 1016–1022. 2. Cormack, A. M. Reconstruction of densities from their projections with applications in radiological physics. Phys. Med. Biol. 1973, 18, 195–207. 3. Kalender, W. A. Computed Tomography. Publicis Corporate Publishing: Erlangen, 2005. 4. Paulus, M. J.; et al. High resolution X-ray computed tomography: an emerging tool for small animal cancer research. Neoplasia 2000, 2(1–2), 62–70. 5. Machin, K.; Webb, S. Cone-beam X-ray microtomography of small specimens. Phys. Med. Biol. 1994, 39, 1639–1657. 6. Ritman, E. L. Micro-computed tomography:current status and development. Annu, Rev. Biomed. Eng. 2004, 6, 185–208. 7. De Clerck, N. M.; et al. High-resolution X-ray microtomography for the detection of lung tumors in living mice. Neoplasia 2004, 6(4), 374. 8. Badea, C.; Hedlund, L. W.; Johnson, G. A. Micro-CT with respiratory and cardiac gating. Med. Phys. 2004, 31(12), 3325. 9. Pickhardt, P. J.; et al. Microcomputed tomography colonography for polyp detection in an in vivo mouse tumor model. PNAS 2005, 102(9), 3419–3422. 10. Bushong, S. C. Radiologic Science for Technologist. Mosby: St. Louis, MO, 1997.
P1: OTA/XYZ P2: ABC c06 JWBS037-Chen January 25, 2011
156
9:11
Printer Name: Yet to Come
CARBON NANOTUBE X-RAY FOR DYNAMIC MICRO-CT IMAGING OF SMALL ANIMAL MODELS
11. Namati, E. C. D.; Thiesse, J.; Hoffman, E. A.; de Ryk, J.; Ross, A.; McLennan, G. In vivo micro-CT lung imaging via a computer controlled intermittent iso-pressure breath hold (IIBH) technique. Phys. Med. Biol. 2006, 51, 6061. 12. Grider, D. E.; Wright, A.; Ausburn, P. K. Electron beam melting in microfocus X-ray tubes. J. Phys. D: Appl. Phys. 1986, 19, 2281–2292. 13. Flynn, M. J.; et al. Microfocus X-ray sources for 3-D microtomography. Nucl. Instrum. Meth. A 1994, 353, 312–315. 14. Robb, R. A.; et al. High-speed three-dimensional X-ray computed tomography: the Dynamic Spatial Reconstructor. Proc. IEEE 1983, 71(3), 308. 15. Lipton, M. J.; et al. Cardiac imaging with a high-speed Cine-CT Scanner: preliminary results. Radiology 1984, 152(3), 579–582. 16. Speidel, M. A.; et al. Scanning-beam digital X-ray (SBDX) technology for interventional and diagnostic cardiac angiography. Med. Phys. 2006, 33(8), 2714. 17. Schmidt, T. G.; et al. A prototype table-top inverse-geometry volumetic CT system. Med. Phys. 2006, 33(6), 1867. 18. Gomer, R. Field Emission and Field Ionization. Harvard University Press: Cambridge, MA, 1961. 19. Brodie, I.; Spindt, C. A. Vacuum microelectronics. Adv. Electronics Electron Phys. 1992, 83, 1–106. 20. Charbonnier, F. M.; Barbour, J. P.; Dyke, W. P. Resolution of field emission X-ray sources. Radiology 1974, 117, 165. 21. Hallenbeck, G. S. Clinical evaluation of the 350-KV chest radiography system. Radiology 1974, 117, 1–4. 22. Crooks, H. E.; Sangster, J.; Ardran, G. M. The fexitron 848 portabel X-ray apparatus. Radiology 1972, 456, 311. 23. Baptist, R. X-ray tube comprising an electron source with microtips and magnetic guiding means. U.S. Patent 6,259,765, 2001. 24. Whitlock, R. R.; et al. Transmission cathodes for X-ray production. U.S. Patent 6,333,968, 2001. 25. Rangstein, P.; et al. Field-emitting structures intended for a miniature X-ray source. Sensors and Actuators 2000, 82, 24–29. 26. Zhang, J.; et al. Efficient fabrication of carbon nanotube point electron sources by dielectrophoresis. Adv. Mater. 2004, 16(14), 1219–1222. 27. Zhu, W. Vacuum Microelectronics. Wiley: Hoboken, NJ, 2001. 28. Wei, Y.; et al. Stability of carbon nanotubes under electric field studied by scanning electron microscopy. Appl. Phys. Lett. 2001, 79(27), 4527–4529. 29. Zhou, O.; Lu, J. P.; New X-ray generating mechanism using electron field emission cathode. U.S. Patent 6,553,096, 2005. 30. Sugie, H.; et al. Carbon nanotubes as electron source in an X-ray tube. Appl. Phys. Lett. 2001, 78, 2578. 31. Yue, G. Z.; et al. Generation of continuous and pulsed diagnostic imaging X-ray radiation using a carbon-nanotube-based field-emission cathode. Appl. Phys. Lett. 2002, 81(2), 355. 32. Zhang, J.; et al. A stationary scanning X-ray source based on carbon nanotube field emitters. Appl. Phys. Lett. 2005, 86, 184104. 33. Cheng, Y.; et al. Dynamic X-ray radiography using a carbon nanotube field emission X-ray source. Rev. Sci. Instrum. 2004, 75(10), 3264. 34. Zhang, J.; et al. A multi-beam X-ray imaging system based on carbon nanotube field emitters. In Medical Imaging 2006, Proceedings of SPIE; Flynn, M. J., Hsieh J., Eds. SPIE: San Diego, CA, 2006, p. 614204.
P1: OTA/XYZ P2: ABC c06 JWBS037-Chen January 25, 2011
9:11
Printer Name: Yet to Come
REFERENCES
157
35. Zhou, O.; Lu, J. P.; Qiu, Q. Large-area individually addressable multi-beam X-ray system and method of forming same. U.S. Patent 6,876,724, 2005. 36. Zhang, J.; et al. Multiplexing radiography using a carbon nanotube based X-ray source. Appl. Phys. Lett. 2006, 89, 064106. 37. Lu, J. P.; Zhou, O.; Zhang, J. X-ray imaging systems and methods using temporal digital signal processing for reducing noise and for obtaining multiple images simultaneously. U.S. Patent 7,245,692, 2007. 38. Dunham, M. B.; et al. Solid-state CT system and method. U.S. Patent 6,385,292, 2002. 39. Heo, S. H.; Ihsan, A.; Cho, S. O. Transmission-type microfocus X-ray tube using carbon nanotube field emitters. App. Phys. Lett. 2007, 90(18), 183109. 40. Haga, A.; et al. A miniature X-ray tube. Appl. Phys. Lett. 2004, 84(12), 2208. 41. Yabushita, R.; Hata, K. Development of high spatial resolution X-ray radiography system equipped with multiwalled carbon nanotube field emission cathode. J. Vac. Sci. Technol. B. 2008, 26, 702. 42. Cao, G.; et al. Respiratory-gated micro-CT using a carbon nanotube based micro-focus field emission X-ray source. In SPIE Proceeding on Medical Imaging. 2008, 43. Zhou, O.; et al. Computed tomography scanning system and method using a field emission X-ray source. U.S. Patent 7,227,924, 2007. 44. Zhang, J.; et al. A nanotube-based field emission X-ray source for micro-computed tomography. Rev. Sci. Instrum. 2005, 76, 094301. 45. Yang, G.; et al. Stationary digital breast tomosynthesis system with a multi-beam field emission X-ray source array. In SPIE Proceedings on Medical Imaging. 2008, 46. Liu, Z.; et al. Carbon nanotube based microfocus field emission X-ray source for microcomputed tomography. Appl. Phys. Lett. 2006, 89, 103111. 47. Oh, S. J.; et al. Liquid-phase fabrication of patterned carbon nanotube field emission cathodes. Appl. Phys. Lett. 2004, 87(19), 3738. 48. Noo, F.; Mennessier, C.; White, T. A.; Roney, T. J. Analytic method based on identification of ellipse parameters for scanner calibration in cone-beam tomography. Phys. Med. Biol. 2000, 45, 3489. 49. Cao, G.; et al. A dynamic micro-CT scanner with a stationary mouse bed using a compact carbon nanotube field emission X-ray tube. In SPIE Proceedings on Medical Imaging, in press. 50. Hu, J.; Haworth, S. T.; Molthen, R. C.; Dawson, C. A. Dynamic small animal lung imaging via a postacquisition respiratory gating technique using micro-cone beam computed tomography. Acad. Radiol. 2004, 11, 961. 51. Ford, N. L.; Wheatley, A. R.; Holdsworth, D. W.; Drangova, M. Optimization of a retrospective technique for respiratory-gated high speed micro-CT of free-breathing rodents. Phys. Med. Biol. 2007(52), 5749. 52. Drangova, M.; Ford, N. L.; Detombe, S. A.; Wheatley, A. R.; Holdsworth, D. W. Fast retrospectively gated quantitative four-dimensional (4D) cardiac micro computed tomography imaging of free-breathing mice. Invest. Radiol. 2007, 42, 85. 53. Cavanaugh, D.; et al. In vivo respiratory-gated micro-CT imaging in small-animal oncology models. Mol. Imaging Biol. 2004, 3, 55. 54. Nahrendorf, M.; Badea, C.; Hedlund, L. W.; Figueiredo, J.-L.; Sosnovik, D. E.; Johnson, G. A.; Weissleder, R. High-resolution imaging of murine myocardial infarction with delayedenhancement cine micro-CT. Am J Physiol Heart Circ Physiol. 2007, 292, H3172. 55. Nilsson, R. Lung compliance and lung morphology following artificial ventilation in the premature and full-term rabbit neonate Scand. J. Respir. Dis. 1979, 60, 206.
P1: OTA/XYZ P2: ABC c06 JWBS037-Chen January 25, 2011
158
9:11
Printer Name: Yet to Come
CARBON NANOTUBE X-RAY FOR DYNAMIC MICRO-CT IMAGING OF SMALL ANIMAL MODELS
56. Nakamura, T.; Malloy, J.; McCaig, L.; Yao, L. J.; Joseph, M.; Lewis, J. ; Veldhuizen, R. Mechanical ventilation of isolated septic rat lungs: effects on surfactant and inflammatory cytokines. J. Appl. Physiol. 2001, 91, 811. 57. Ford, N. L.; Martin, E. L.; Lewis, J. F.; Veldhuizen, R. A. W.; Drangova, M.; Holdsworth, D. W. in vivo characterization of lung morphology and function in anesthetized free-breathing mice using micro-computed tomography. J. Appl. Physiol. 2007, 102, 2046. 58. Ford, N. L.; Nikolov, H. N.; Norley, C. J. D.; Thornton, M. M.; Foster, P. J.; Drangova, M.; Holdsworth, D. W. Prospective respiratory-gated micro-CT of free breathing rodents. Med. Phys. 2005 (32), 2888. 59. Guohua, C.; et al. Respiratory-gated micro-CT using a carbon nanotube based micro-focus field emission X-ray source, In Jiang, H.; Ehsan, S., Eds. SPIE: San Diego, CA. 2008, p. 691304. 60. Fujita, H. et al. Simple method for determining the modulation transfer function in digital radiography. IEEE Trans Med Imaging 1992, 11(1), 34–39. 61. Badea, C.; Hedlund, L. W.; Johnson, G. A. 4-D Micro-CT of the mouse heart. Mol. Imaging Biol. 2005, 4, 110. 62. Quan, E.; Lalush, D. S. A faster ordered-subset convex algorithm for iterative reconstruction in a rotation-free micro-CT system. Phys. Med. Biol. 2009, 54, 1061. 63. Peng, R.; et al. Stationary micro-CT scanner using a distributed multi-beam field emission X-ray source: a feasibility study. In SPIE Proceedings on Medical Imaging, in press. 64. Baldelli, P.; et al. A prototype of a quasi-monochromatic system for mammography applications. Phys. Med. Biol. 2005, 50, 2225–2240. 65. Mckinley, R. L.; et al. Simulation study of a quasi-monochromatic beam for X-ray computed mammotomography. Med. Phys 2004, 31(4), 800. 66. Marziani, M.; et al. Dual-energy tissue cancellation in mammography with quasi-monochromatic X-rays. Phys. Med. Biol. 2002, 47(2), 305.
P1: OTA/XYZ P2: ABC c07 JWBS037-Chen January 25, 2011
9:13
Printer Name: Yet to Come
CHAPTER 7
Quantum Dots for In Vivo Molecular Imaging YUN XING Department of Material Science and Engineering, University of Dayton, Dayton, Ohio, USA
During the last decade, semiconductor quantum dots (QDs) have sparked tremendous interest in the biomedical research field. The unique optical properties of QDs, such as superior brightness,photostability, large Stokes shift, and tunable emission spectra, made them appealing fluorescent probes in a wide variety of applications including cell labeling, single particle tracking inside living cells, and in vivo molecular imaging. The most commonly used QDs consist of an organic solvent-grown CdSe/ZnS core–shell structure made water soluble by either ligand exchange or polymer encapsulation. Surface funtionlization with targeting moieties such as antibody and peptide has allowed successful targeted imaging of tumors inside living animals. More recent developments include non-cadmium-based QDs, smaller QDs that emit in the near-infrared range as well as a QD–luciferase hybrid system (“self-illuminating quantum dots”), which permits more sensitive detection (higher signal-to-noise ratio) and a longer tissue penetration depth. In spite of the many successes, several critical issues remain to be solved for the clinical translation of QDs as a molecular imaging agent for the human body.
7.1 INTRODUCTION Semiconductor quantum dots (QDs) are tiny light-emitting nanoparticles that have captivated scientists and engineers over the past two decades owing to their fascinating optical and electronic properties, which are not available from either individual molecules or bulk solids. Recent research has stimulated considerable interest in developing these quantumconfined nanocrystals as optical probes for biomedical applications. Compared with organic dyes and fluorescent proteins, semiconductor QDs offer several unique advantages, such as size- and composition-tunable emission from visible to infrared wavelengths, large absorption coefficients across a wide spectral range, and very high levels of brightness and photostability [1]. The long-term photostability and superior brightness of QDs make them ideal candidates for live animal targeting and imaging [2–4]. Additionally, recent Nanoplatform-Based Molecular Imaging Edited by Xiaoyuan Chen C 2011 John Wiley & Sons, Inc. Copyright
159
P1: OTA/XYZ P2: ABC c07 JWBS037-Chen January 25, 2011
160
9:13
Printer Name: Yet to Come
QUANTUM DOTS FOR IN VIVO MOLECULAR IMAGING
studies have shown that QDs have two-photon excitation cross sections of magnitude larger than those of conventional fluorescent probes now in use, a further plus for deep tissue imaging [5]. In this chapter, we cover QD core synthesis, water solubilization, biological functionalization, and animal imaging applications and finish with a discussion on the issues/limitations and future perspectives.
7.2 QUANTUM DOT STRUCTURE QDs are somewhat spherical nanocrystals in the size range of 1–10 nm in diameter (Fig. 7.1a) [6, 7]. Semiconductor nanocrystals can also be produced with other shapes such as rods and tetrapods [8], but spherical QDs are often widely used for biological applications. These particles are generally made from hundreds to thousands of atoms of group II and VI elements (e.g., cadmium selenide (CdSe) and cadmium telluride (CdTe) or group III and V elements (e.g., indium phosphide (InP) and indium arsenide (InAs)).
(a)
(c)
(b)
(d)
FIGURE 7.1 Quantum dot structure and novel optical properties. (a) Structure of a multifunctional QD probe, showing the capping ligand TOPO, an encapsulating copolymer layer, tumor-targeting ligands (such as peptides, antibodies, or small molecule inhibitors), and polyethylene glycol (PEG) [18]. (b) Size-tunable emission spectra of QDs. This image shows ten distinguishable emission colors of ZnS-capped CdSe quantum dots excited with a near-UV lamp. From left to right (blue to red), the emission maxima are located at 443, 473, 481, 500, 518, 543, 565, 587, 610, and 655 nm [16]. (c) Excitation (dotted line) and fluorescence (solid line) spectra of fluorescein (top) and a typical water-soluble QD (bottom) [37]. (d) Superior photostability of QDs as compared to organic dyes [36].
P1: OTA/XYZ P2: ABC c07 JWBS037-Chen January 25, 2011
9:13
Printer Name: Yet to Come
NOVEL OPTICAL PROPERTIES
161
Recent advances have allowed the precise control of particle size, shape (dots, rods, or tetrapods) [9–11] and internal structure (core–shell, gradient alloy, or homogeneous alloy) [12, 13]. The novel optical property of quantum dots arises from the “quantum confinement” effect of the semiconductor materials. This refers to the size and composition dependence of the semiconductor bandgap energy. This effect is readily observed when one or more dimensions of a semiconductor are reduced to the nanometer regime. Absorption of a photon with energy above the semiconductor bandgap energy results in the creation of an electron–hole pair (or excitation). The absorption has an increased probability at higher energies (i.e., shorter wavelengths) and results in a broadband absorption spectrum, in marked contrast to standard fluorophores. For nanocrystals smaller than the Bohr excitation radius (a few nanometers), energy levels are quantized, with values directly related to the size of the nanoparticle. The radiative recombination of an excitation (characterized by a long lifetime, >10 ns [14]) leads to the emission of a photon in a narrow, symmetric energy band (Fig. 7.1c). This dependence of light emission on particle size allows the development of new fluorescence emitters with precisely tuned emission wavelengths (Fig. 7.1b). For example, the semiconductor cadmium selenide has a bulk bandgap of 1.7 eV (corresponding to 730-nm light emission). QDs of this material can be tuned to emit between 450 and 650 nm by changing the nanocrystal diameter from 2 to 7 nm. The composition of the material may also be used as a parameter to alter the bandgap of a semiconductor. A QD with a diameter of 5 nm can be tuned to emit between 610 and 800 nm by changing the composition of the alloy CdSex Te1 − x [15]. The classic and most commonly used QDs consist of a CdSe core and a shell layer made of ZnS or CdS. Fluorescent properties are determined by the core materials and the shell layer removes surface defects and prevents nonradiative decay (the disappearance of an excited species due to a radiationless transition), leading to a significant improvement in particle stability and fluorescence quantum yields. For biological imaging applications, these hydrophobic dots can be made water soluble by exchange with bifunctional ligands (mostly thiol and phosphine mono and multidentate ligands) or using amphiphilic polymers that contain both a hydrophobic segment or side chain (mostly hydrocarbons) and a hydrophilic segment or group (such as polyethylene glycol (PEG) or multiple carboxylate groups). Additionally, biomolecules such as antibodies and peptides can be attached to the QDs to achieve specific labeling and targeting (Fig. 7.1a).
7.3 NOVEL OPTICAL PROPERTIES As briefly noted above, QDs are made from inorganic semiconductors and have novel optical properties that can be used to optimize the signal-to-background ratio in fluorescence imaging. In comparison with organic dyes and fluorescent proteins, QDs have several advantages and unique applications. First, QDs have very large molar extinction coefficients on the order of 0.5–5 × 106 M-1 cm-1 [15], about 10–50 times larger than that (5–10 × 104 M-1 cm-1 ) of organic dyes. Therefore QDs are able to absorb 10–50 times more photons than organic dyes at the same excitation photon flux (i.e., the number of incident photons per unit area), leading to a significant improvement in probe brightness; this allows for brighter probes under photon-limited in vivo conditions (where light intensities are severely attenuated due to scattering and absorption). In theory, the lifetime-limited emission rates for single QDs are 5–10 times slower than those of single organic dyes, because of their longer excited
P1: OTA/XYZ P2: ABC c07 JWBS037-Chen January 25, 2011
162
9:13
Printer Name: Yet to Come
QUANTUM DOTS FOR IN VIVO MOLECULAR IMAGING
state lifetimes (20–50 ns vs. less than 10 ns). In practice, however, fluorescence imaging usually operates under absorption-limited conditions, in which the rate of absorption is the main limiting factor of fluorescence emission (versus the emission rate of the fluorephore). As a result, individual QDs have been found to be 10–20 times brighter than organic dyes [16]. Second, QDs are several thousand times more stable against photobleaching (the loss of fluorescence due to photoinduced chemical damages) than organic dyes (Fig. 7.1d) and are thus well suited for continuous tracking studies over a long period of time. In addition, the relatively longer excited state lifetimes of QDs can be used to separate the QD fluorescence from background fluorescence, in a technique known as time-domain imaging [17]; since QDs emit light slowly enough that most of the background autofluorescence emission is over by the time QD emission occurs. Third, the large Stokes shifts of QDs (measured by the distance between the excitation and emission peaks) can be used to further improve detection sensitivity. This factor becomes especially important for in vivo molecular imaging due to the high autofluorescence background often seen in complex biomedical specimens. The Stokes shifts of semiconductor QDs can be as large as 300–400 nm, depending on the wavelength of the excitation light. Organic dye signals with a small Stokes shift are often buried by strong tissue autofluorescence, whereas QD signals with a large Stokes shift are clearly recognizable above the background. A further advantage of QDs is that multicolor QD probes can be used to image and track multiple molecular targets simultaneously. This is a very desirable feature, because most complex human diseases such as cancer and atherosclerosis involve a large number of genes and proteins. The ability to track a panel of molecular markers at the same time will allow scientists to better understand, classify, and differentiate complex human diseases than using a single biomarker each time. Multiple parameter imaging, however, represents a significant challenge for magnetic resonance imaging, positron emission tomography, computed X-ray tomography, and related imaging modalities. By contrast, fluorescence optical imaging provides both signal intensity and wavelength information, and multiple wavelengths or colors can be resolved and imaged simultaneously (color imaging). Therefore different molecular or cellular targets can be tagged with different colors. In this regard, QD probes are particularly attractive, because their broad absorption profiles allow simultaneous excitation of multiple colors and their emission wavelengths can be continuously tuned by varying particle size and chemical composition. For organ and vascular imaging in which micrometer-sized particles could be used, optically encoded beads (polymer beads embedded with multicolor QDs at controlled ratios) could allow multiplexed molecular profiling in vivo at high sensitivities [18, 19]. The most common scheme of quantum dot production consists of four major steps: (1) synthesis of the QD core, most often CdSe, in a high-temperature organic solvent (aqueous-based QD synthesis methods have also emerged [20]; (2) growth of an inorganic shell (usually ZnS) on the core to protect the optical properties of the core; (3) water solubilization of QDs by ligand exchange or polymeric encapsulation; and (4) functionalization of the QD surface with biologically active molecules. Each of the four steps is discussed in the following section.
7.4 CORE SYNTHESIS The most effective and reproducible synthesis procedure for monodisperse quantum dot synthesis involves the addition of semiconductor precursors to a liquid coordinating solvent at high temperature, first described by Murray and co-workers [21]. In a typical synthesis
P1: OTA/XYZ P2: ABC c07 JWBS037-Chen January 25, 2011
9:13
Printer Name: Yet to Come
WATER SOLUBILIZATION
163
of CdSe quantum dots, room-temperature QD precursors, dimethylcadmium and elemental selenium dissolved in liquid trioctylphosphine (TOP), are swiftly injected into hot (290–350 ◦ C) trioctylphosphine oxide (TOPO), immediately initiating nucleation of QD crystals. The presence of coordinating solvents (i.e., TOPO or TOP) prevents the formation of bulk semiconductors by binding to the QD surfaces through their basic functional groups (phosphines, phosphine oxides). The alkyl chains from the coordinating ligands extend away from the QD surface, rendering them sterically stable as colloids and dispersible in many nonpolar solvents such as chloroform and hexane. By tuning the different parameters (precursors, coordinating ligands, etc.), QDs with diameters between 2 and 8 nm and emission spanning the entire visible spectrum can easily be synthesized with just one composition (CdSe). By also adjusting the composition (ZnS, CdS, CdSe, CdTe, PbS, PbSe, and their alloys), it is possible to generate QDs that emit all the way from 400 to 4000 nm [12, 22–24]. The quantum yield (ratio of the number of photons emitted to the number of photons absorbed, essentially the emission efficiency of a fluorophore) of the nanocrystal core synthesized as above is relatively low (less than 10%) due to the presence of gap surface states arising from surface nonstoichiometry, unsaturated bonds [21, 25]. Usually, a thin (a few atoms thick) shell of higher bandgap semiconductor material, such as ZnS or CdS, is epitaxially grown around the core to achieve better photostability and higher quantum yield. A higher bandgap material eliminates the surface defects. This layer improves quantum yield and protects the CdSe core against photo-oxidation (which is extremely important for minimizing cytotoxicity as well as for enhancing photostability). To produce a ZnS shell, a solution of dimethylzinc and hexamethyldisilathiane (in tri-noctylphosphine) is slowly added into the reaction solution of CdSe. Thickness of the shell is mediated by the amount of dimethylzinc and hexamethyldisilathiane injected in. By carefully designing the shell, quantum yields of QDs can be significantly improved (from 5% up to 90%) [25, 26]. The presence of a shell also improves QD photostability by several orders of magnitude relative to conventional fluorophores.
7.5 WATER SOLUBILIZATION As mentioned earlier, QDs made in organic solvents are coated with alkyl chains that render solubility only in nonpolar organic solvents. In order to make QDs useful for biological purposes, it is essential to modify the surface of QDs to make them water soluble. An ideal modification must meet the following criteria: (1) has good shelf stability, no flocculating over long term of storage; (2) efficiently coverts the insoluble QDs to water soluble; (3) maintains the brightness (i.e., quantum yield of the dots); and (4) keeps the overall size of the QDs small, best if below 10 nm. Unfortunately, none of the current coating strategies meet all four criteria, each with their advantages and shortcomings. Different QD solubilization strategies have been devised over the last decade: these strategies can be grouped into two major categories. The first uses “ligand exchange” and involves the substitution of the native TOP/TOPO with bifunctional ligands, each presenting a surface anchoring moiety to bind to the inorganic QD surface and an opposing hydrophilic end group that renders water compatibility. An array of thiol and phosphine mono and multidentate ligands have been tested and will be discussed in detail in the next section. The second method preserves the native TOP/TOPO on the QDs and uses variants of amphiphilic “diblock” and “triblock” copolymers and phospholipids to tightly interleave/interdigitate the alkylphosphine ligands through hydrophobic attraction, whereas the hydrophilic outer
P1: OTA/XYZ P2: ABC c07 JWBS037-Chen January 25, 2011
164
9:13
Printer Name: Yet to Come
QUANTUM DOTS FOR IN VIVO MOLECULAR IMAGING
block permits aqueous dispersion and further derivitization. Additionally, combinations of layers of different molecules conferring the required colloidal stability to QDs have also been exploited [27, 28]. In the ligand exchange scheme, TOPO-coated QDs are mixed with a solution containing a bifunctional ligand, which competes with TOPO for binding to a metal atom on the QD surface. With excess bifunctional ligands (e.g., mercapto acetic acid) in the solution, the thiol functional groups outcompete the phosphonic oxides (from the TOPO) for binding to the QDs and the QDs become hydrophilic. Various types of ligands have been used, including simple thiol-containing molecules such as mercaptoacetic acid [29], cysteine [30], dithiothreitol [31] and dihydrolipoic acid [32], oligomeric phosphine [22], dendrons [32], and peptides [33]. This method results in smaller QDs, generally not too much bigger than the organic core/shell structure. Drawbacks of the ligand exchange method are relatively low quantum yield, pH sensitivity, and poor stability in biological buffers or insufficient robustness for further chemical modification to introduce biorecognition molecules. In the second method, coordinating ligands on the QD surface are retained and used to interact with amphiphilic block copolymers [17, 34] in silica shells [35], phospholipid micelles [36], amphiphilic polysaccharides [37], or polyanhydrides [38]. The hydrophobic domains of these chemicals strongly interact with TOPO on the QD surface, whereas the hydrophilic groups face outwards and render the QDs water soluble. Note that the coordinating organic ligands (TOP or TOPO) are retained on the inner surface of QDs, a feature that is important for maintaining the optical properties of QDs and for shielding the core from the outside environment. This approach is more effective than ligand exchange at maintaining QD optical properties (higher quantum yield) and storage stability in aqueous buffer; the drawback is the increase in the overall size of the QDs. For example, phospholipids and block copolymer coatings tend to increase the diameter of CdSe–ZnS QDs from ∼4–8 nm before encapsulation to ∼20–30 nm, a size that although smaller than most mammalian cells can still limit intracellular mobility and may preclude distance-sensitive applications such as fluorescence resonance energy transfer [39]. In animal imaging applications, increase in size may present a hurdle by impeding the extravasations of QDs to the targeted sites and by causing rapid uptake by the reticuloendothelial system (RES) and thus decreasing the bioavailability of the nanoparticles.
7.6 BIOFUNCTIONIZATION To make QDs more useful for molecular imaging and other biological applications, QDs need to be conjugated to biological molecules without disturbing the biological function of these molecules. Several successful approaches have been used to link biological molecules to QDs, including nonspecific adsorption, electrostatic interaction, mercapto (–SH) exchange, and covalent linkage (Fig. 7.2) [40]. It has been reported that simple small molecules, such as oligonucleotides [41, 42] and various serum albumins [43], are readily adsorbed to the surface of water-soluble QDs. This adsorption is nonspecific and depends on ionic strength, pH, temperature, and the surface charge of the molecule. Mattousi and coworkers [34] presented a method of conjugating proteins to QD surfaces using electrostatic interactions. The protein of interest was engineered with a positively charged domain (polyhistidine), which in turn interacted electrostatically with the negatively charged surface of DHLA-capped QD. The protein–QD conjugates prepared in this way were very stable and the fluorescence quantum yield was even higher than that from the nonconjugated dots [32].
P1: OTA/XYZ P2: ABC c07 JWBS037-Chen January 25, 2011
9:13
Printer Name: Yet to Come
BIOFUNCTIONIZATION
165
FIGURE 7.2 Common methods used for QD biofunctionization, including direct linkage to the TOPO-coated QDs (biomolecules can be linked include thiolated DNA (ligand exchange) or peptides with adhesive domains electrostatic interaction, and covalent linking (see text for references).
P1: OTA/XYZ P2: ABC c07 JWBS037-Chen January 25, 2011
166
9:13
Printer Name: Yet to Come
QUANTUM DOTS FOR IN VIVO MOLECULAR IMAGING
Biological molecules containing thiol groups can be conjugated to the QD surface through a mercapto exchange process [44–48]. Unfortunately, the same problem of using thiol as an anchoring group on a ZnS surface occurs because the bond between Zn and thiol is not very strong and is dynamic. As a result, biomolecules can readily dissociate from the nanoparticle surface, causing QDs to precipitate from the solution. A more stable linkage is obtained by covalently linking biomolecules to the functional groups on the QD surface using crosslinker molecules [1, 29, 31, 34, 49]. Most water solubilization methods result in QDs covered with carboxylic acid, amino or thiol groups. Under these situations, it is easy to link QDs to biological molecules that also have these reaction groups. For example, the crosslinker 1-ethyl-3-(3-dimethylaminopropyl)carbodiimide (EDC) is commonly used to link –NH2 and –COOH groups, whereas 4-(N-maleimidomethyl)-cyclohexanecarboxylic acid N-hydroxysuccinimide ester (SMCC) can be used to crosslink –SH and –NH2 groups. Using the above methods, there have been numerous reports of conjugating QDs with various biological molecules, including biotin [35], oligonucleotides [50], peptides [32], and proteins, including avidin/streptavidin [34], albumin [51], adaptor proteins (e.g., protein A, protein G), and antibodies [34, 49]. In addition, the native functional groups (–COOH, –NH2 , or –SH) on a water-soluble QD surface can be further converted to other functional groups to allow more versatile conjugation of QDs to biomolecules (site-specific conjugation, molecules that are sensitive to EDC or SMCC modification). For instance, carboxylic acids on QDs have been converted to hydrazides, allowing attachments of biomolecules containing sugar groups [49]. Recently, Zhang et al. [53] presented another strategy for site-specific conjugation of biomolecules to the QD surface. This study utilized the specific and stable binding between HaloTag protein (HTP) and its ligand. QDs were first functionalized with HaloTag ligands, and the protein of interest (e.g., Renilla luciferase) was genetically fused to a HTP. When mixed together, QDs and Rluc8 can be immobilized on QDs through the HTP–HaloTag ligands linkage. In most cases, the biological functions of these molecules have been preserved during the conjugation process.
7.7 ANIMAL IMAGING APPLICATIONS OF QUANTUM DOTS Fluorescent proteins and small organic dyes have been used as a fluorescent contrast agent for living animal imaging. However, compared to other imaging modalities, such as PET and MRI, fluorescent imaging is still limited by the poor transmission of visible light through biological tissue. One way to attenuate this penetration problem is to use near-infrared (NIR, 650–900 nm) light, since hemoglobin and water, the major absorbers of visible and infrared light, respectively, have their lowest absorption coefficient in this region [53]. Few organic dyes, however, are available that emit brightly in this spectral region, and they suffer from the photobleaching problem. On the contrary, the novel optical properties of QDs allow the synthesis of bright and stable fluorescent labels that emit in the near-infrared spectrum by adjusting their size and composition. Because visible QDs are more synthetically advanced, most animal imaging studies implementing quantum dots have used CdSe/ZnS QDs that emit visible light and a few recent studies have started using near-infrared dots [54, 55]. Although still far from its mature stage, these studies have demonstrated the great performance and promise of QDs as fluorescent imaging agents in living animals, due to their superior ability to remain photostable and bright. In the following section, we first review the major successful applications of quantum dots in animal imaging and then discuss the current issues and limitations followed by future perspectives.
P1: OTA/XYZ P2: ABC c07 JWBS037-Chen January 25, 2011
9:13
Printer Name: Yet to Come
ANIMAL IMAGING APPLICATIONS OF QUANTUM DOTS
167
7.7.1 Nontargeted Animal Imaging QDs have been used for nontargeted imaging in various animal models [55–57]. Ballou and co-workers [57] injected PEG-coated QDs into the mouse bloodstream and investigated how the surface coating would affect their circulation lifetime. In contrast to small organic dyes, which are eliminated from the circulation within minutes after injection, PEG-coated QDs were found to stay in the blood circulation for an extended period of time (half-life more than 3 h). This long-circulating feature is due to the relatively large size of PEG-coated QDs, which falls within an intermediate size range: they are small enough and sufficiently hydrophilic to slow down opsonization and reticuloendothelial uptake, but are large enough to avoid renal clearance. Amazingly, these QDs maintained their fluorescence even after 4 months in vivo. In 2003, Larson et al. [5] intravenously injected green QDs (550 nm) in a living mouse and visualized them dynamically through the skin (in capillaries hundreds of micrometers deep) by using two-photon microscopy (Fig. 7.3a). In addition to the superior brightness and photostability, this study also found that QDs have two-photon excitation cross sections as high as 47,000 Goeppert-Mayer units, by far the largest of any label used in multiphoton microscopy. Two-photon excitation allows greater tissue penetration due to excitation within the NIR spectral range (e.g., 900 nm), but few fluorophores are bright
(a)
(c)
(b)
(d)
FIGURE 7.3 Animal imaging applications using quantum dots. (a) Two-photon fluorescence imaging of capillaries at the base of the dermis: (a) QD (1 M) intravenously delivered to the animal and (b) fluorescin (40 M) (adapted from Larson et al. [5]). (b) NIR QD fluorescence guided dissection of a lymph node in a pig; 400 pmol of NIR QDs were injected intradermally in the right groin [56]. (c) Targeted tumor imaging using quantum dots antibody conjugates [18]. (d) Tumor imaging using QD705 RGD bioconjugates; arrows indicate tumor sites [55].
P1: OTA/XYZ P2: ABC c07 JWBS037-Chen January 25, 2011
168
9:13
Printer Name: Yet to Come
QUANTUM DOTS FOR IN VIVO MOLECULAR IMAGING
enough for these purposes; QDs, with their large multiphoton excitation cross section, appear to be ideal probes for multiphoton microscopy imaging. Improved tissue penetration can also be achieved by tuning the QD emission to the NIR window [55, 56]. Kim et al. [56] prepared a core–shell nanostructure called type II QDs with fairly broad emission at 850 nm and a moderate quantum yield of ∼13%. In contrast to conventional QDs (type I), the shell materials in type II QDs have valence and conduction band energies both lower than those of the core materials. As a result, the electrons and holes are physically separated and the nanoparticles emit light at reduced energies (longer wavelengths). NIR QDs were injected intradermally in the left paw of a living mouse. Their results showed rapid uptake (5 min) of QDs into nearby lymph nodes and that they could be imaged virtually background-free (Fig. 7.4b). The same study also demonstrated the feasibility of image-guided surgery in a big animal model. Thus 400 pmol of NIR QDs injected intradermally permits sentinel lymph nodes 1 cm deep to be imaged easily in real time (and removed surgically) using an excitation fluorescence rate of only 5 mW/cm2 . QDs have also been used for cell tracking studies [17, 36, 57, 58]. QDs were delivered into live mammalian cells via three different mechanisms: nonspecific pinocytosis, microinjection, and peptide-induced transport (e.g., using the protein transduction domain of HIV-1 Tat peptide, Tat-PTD) [59]. A surprising finding was that two billion QDs could be delivered into the nucleus of a single cell, without compromising its viability, proliferation, or migration [35, 58, 59]. The ability to image single-cell migration and differentiation in real time is expected to be important to several research areas such as embryogenesis, cancer metastasis, stem-cell therapeutics, and lymphocyte immunology. These studies demonstrated the potential of using QDs to enable biologists to track cell (one polymer-coated QD is about one-thousandth of a mammalian cell), tissue, and organ developments over extended periods of time, a task not possible with small molecule organic dyes with fast photobleaching [60, 61]. 7.7.2 Targeted Animal Imaging The above studies demonstrated the capability of QDs for living animal imaging, but had only demonstrated image contrast at the tissue/organ level. The goal of molecular imaging is to generate image contrast due to the molecular difference in different tissues and organs. This requires a probe that has a targeting moiety to generate contrast only in locations specified by the targeting probe. Akerman et al. [49] were the first to explore the possibility of using QD–peptide conjugates to target tumor vasculatures in vivo. QDs coated with peptides targeting the lung vasculature, blood vessel and tumor cell, or lymphatic vessels were injected systematically into mouse. Although they were not able to image the QDs in a living animal, histological sections of different organs after 5 or 20 min of circulation revealed that QDs homed to tumor vessels guided by the peptides, but not to surrounding tissues, probably due to their larger size relative to organic dyes (which would stain surrounding tissues). Whole animal imaging of molecular-level detection was realized by Gao et al. [18] in 2004 using red fluorescent QDs conjugated to antibodies specific to prostate-specific membrane antigen (PSMA) on a human prostate cancer induced in a mouse. QD conjugates used in this study contained an amphiphilic triblock copolymer for in vivo protection, targeting ligands (anti-PSMA) for tumor antigen recognition, and multiple PEG molecules for improved biocompatibility and circulation. The QD-tagged PSMA antibodies recognized and bound at the tumor site and were clearly imaged in vivo (Fig. 7.3c). Control experiments with QD-PEG (no targeting ligand) or QD only confirmed
P1: OTA/XYZ P2: ABC c07 JWBS037-Chen January 25, 2011
9:13
Printer Name: Yet to Come
RESONANCE ENERGY TRANSFER BASED DETECTION AND IMAGING
169
the binding was specific and the accumulation of nontargeted QDs at tumor sites was marginal compared with the targeted ones. There are two possible mechanisms for the preferential accumulation of QDs at tumor sites: passive targeting due to the enhanced permeability and retention (EPR) effect and active targeting because of the antibody, PSMA, which is a cell surface marker for both prostate epithelial cells and neovascular endothelial cells. Because the QDs emit in the visible range, a spectral unmixing algorithm was used to separate QD signal from background autofluorescence. More recently, Cai and co-workers used near-infrared QDs for tumor imaging by targeting angiogenesis, the formation of new blood vessels from preexisting vasculature [55, 62, 63]. Amine-modified QD705 (emission maximum at 705 nm) was conjugated to ␣v 3 integrin-targeting cyclic RGD peptide and injected intravenously into living mice. Tumor fluorescence reached a maximum at 6 h postinjection with good contrast (Fig. 7.3d) Since angiogenesis is common to all tumors, this technique may aid cancer detection and management in general. It is worth noting that, in both studies, a significant portion of the injected QDs went to the RES system, including the liver, spleen, and lung [64].
7.8 RESONANCE ENERGY TRANSFER BASED DETECTION AND IMAGING 7.8.1 QD Fluorescence Resonance Energy Transfer Fluorescence resonance energy transfer (FRET) is a process in which energy is transferred from an excited donor to an acceptor via a resonant, near-field dipole–dipole interaction [65]. FRET is very sensitive to the distance between donor and acceptor and has been used to study biomolecule conformation, dynamics, and interactions. Several problems associated with FRET with organic fluorophores include fast photobleaching, and spectral overlap between the acceptor and donor. There are several examples of QDs used for FRET in biological systems [39, 44, 66, 67]. Medintz and co-workers engineered a histidine tag on maltose binding protein (MBP) that finally bound electrostatically to QDs, which served as the FRET donor [68] (Fig. 7.4A). With a quencher bound to the maltose-binding site, QD fluorescence was inhibited but could be recovered by adding maltose to displace the quencher. In 2005, the same group did an in-depth study using the MBP system by varying the QD size and the quantity of acceptor dye [66]. Increasing the amount of acceptor dye, or the degree of spectral overlap (by changing QD size), caused a substantial enhancement in energy transfer efficiency. This study demonstrated that QDs can be used as efficient energy donors in the FRET system and showed that by tuning their size, QDs can transfer energy to a number of organic dye molecules. From these studies, the unique advantages of QDs as FRET donors are apparent: size-tunable emission spectra can be used to improve spectral overlap with a particular acceptor dye, and the multivalency ability allows multiple acceptor dye molecules on a single QD donor and this will substantially improved FRET efficiency. QDs could even drive biosensors through a two-step FRET mechanism overcoming inherent donor–acceptor distance limitations, as demonstrated by another work by the same researchers. In this assembly, each 530-nm QD is surrounded by ∼10 MBPs (each monolabeled with Cy3). A second dye, -cyclodextrin-Cy3.5 (-CD-Cy3.5), specifically binds in the MBP central binding pocket. Excitation of the QD leads to FRET excitation of the MBPCy3, which, in turn FRET-excites the -CD-Cy3.5, overcoming the low direct QD-Cy3.5 FRET efficiency. Recently, QD-FRET has been successfully applied on proteolytic activity detection by the same group [40]. In this scenario, a rationally designed multifunctional
P1: OTA/XYZ P2: ABC c07 JWBS037-Chen January 25, 2011
170
9:13
Printer Name: Yet to Come
QUANTUM DOTS FOR IN VIVO MOLECULAR IMAGING
(a)
(b)
FIGURE 7.4 Resonance energy transfer based detection and imaging using QDs. (a) QD FRET. Formation of QD (560)-MBP--CD-QSY9 (maximum absorption ∼565 nm) results in quenching of QD emission. Added maltose displaces -CD-QSY9 from the sensor assembly, resulting in an increase in direct QD emission [68]. (b) QD BRET quantum dot (655) is covalently linked to a BRET donor, Rluc8. The bioluminescence energy of Luc8-catalyzed oxidation of coelenterazine is transferred to quantum dots in close proximity, resulting in QD emission (655 nm) [72].
peptide sequence was engineered between the donor (QD) and acceptor (organic dye). The peptide sequence consists of four functional domains including (1) polyhistidine sequence for self-assembly onto the DHLA capped QDs; (2) a helix-linker spacer to provide rigidity; (3) an exposed protease recognition/cleavage sequence; and (4) a C-terminal site-specific location (cysteine thiol) for dye attachment. When the donor–acceptor hybrid is integrated, FRET takes place and the emission spectrum is a composite of QD emission and the acceptor emission. Upon the addition of the active protease (caspase-1, collagenase, chymotrypsin, or thrombin), the acceptor dye is released from the donor (QD), and the observed fluorescent emission is purely QD emission. Since all the results so far are from experiments performed in buffers, and not in biological samples such as blood or serum, it remains to be determined whether these QD–FRET nanoassemblies can be applied for in vivo imaging of proteolytic activity. Paradoxically, the excellent QD donor properties (long fluorescence lifetime, broad absorption, and high extinction coefficient) may almost preclude their role as FRET acceptors for organic dyes [66]. However, this does not limit their use as acceptors in other applications such as bioluminescence resonance energy transfer as discussed below. 7.8.2 QD BRET Based Animal Imaging Bioluminescence resonance energy transfer (BRET) is a naturally occurring phenomenon whereby a light-emitting protein (the donor, e.g., R. reniformis luciferase) nonradiatively
P1: OTA/XYZ P2: ABC c07 JWBS037-Chen January 25, 2011
9:13
Printer Name: Yet to Come
RESONANCE ENERGY TRANSFER BASED DETECTION AND IMAGING
171
transfers energy to a fluorescent protein (the acceptor, e.g., GFP) in close proximity. BRET is analogous to FRET except that the energy comes from a chemical reaction catalyzed by the donor enzyme (e.g., R. reniformis luciferase-mediated oxidation of its substrate coelenterazine) rather than absorption of excitation photons. Compared to fluorescence imaging, bioluminescence has extremely high sensitivity for in vivo imaging purposes [69]. BRET has been used to study protein–protein interaction in living cells and animals [70, 71]. In a study by De and Gambhir, a hRluc8 (donor)–GFP2 (acceptor) BRET pair was utilized to study FKBP12 (fused with hRluc8) and FRB (fused with GFP2 ) interactions inside living mice [70]. BRET signal could be detected by using a cooled charge-coupled device (CCD) camera. Recently, work in the Rao laboratory has demonstrated the feasibility of using QDs as the acceptor in a bioluminescence resonance energy transfer system [72]. In this study, QDs were covalently conjugated by EDC coupling to the donor, Rluc8 protein, an eightmutation variant of the bioluminescence R. reniformis luciferase developed in the Gambhir lab [73]. The protein emits blue light with a peak at 480 nm upon the addition of the substrate, coelenterazine. If the QDs are in close proximity of the protein, they can be excited and emit at its emission maximum (Fig. 7.4B). The advantage of using bioluminescence versus fluorescence lies in the fact that no external excitation is needed. This “self-illuminating” feature allows cancer imaging in deeper tissue where light resources are limited. Since no excitation light is needed, the autofluorescence problem is automatically solved. In addition, since this coupling method is generic, any QDs with carboxyl groups (including those emitting in the NIR region, e.g., 705-, 800-nm QDs have been tested) can be used as a BRET accepter and therefore allows for multiplexed imaging. Compared with existing QDs, self-illuminating QD conjugates have greatly enhanced sensitivity in small animal imaging, with an in vivo signal-to-background ratio of >1000 for 5 pmol of conjugates subcutaneously injected. One critical issue in making these self-emitting QDs is the size of the nanoparticle, since like FRET, BRET is also a distance-sensitive process. Increase in the QD conjugate size results in greater distance between the protein and the fluorescent semiconductor core, and thereby decreases the energy transfer efficiency significantly. For instance, the authors have observed that increasing the protein–nanoparticle distance by only 2–3 nm causes the BRET ratio to drop from 1.29 to 0.37 [72]. By attaching targeting moieties such as tumor homing antibodies or peptides to the BRET assembly, it is possible to use BRET for targeted tumor imaging in living animals. More recently, QD-BRET was successfully applied to proteolytic activity detection in buffer with a slightly different coupling scheme [74]. In this approach, the luciferase–protease substrate recombinant protein was genetically modified with an additional intein segment. Carboxylated QDs were functionalized first with adipic dihydrazide because hydrazides are excellent nucleophiles to attack the thioester intermediate of inteins. The reaction proceeds rapidly upon mixing the two together and results in the cleavage of the intein and ligation of the C terminus of the recombinant protein to the QDs. Using this approach, the authors have successfully applied this method to synthesize a series of nanosensors for sensitive detection of MMP-2, MMP-7, and urokinase-type plasminogen activator (uPA). These prepared nanosensors can not only detect these proteases in complex biological media such as mouse serum and tumor lysates with a sensitivity of as low as 1 ng/mL, but they can also detect multiple proteases present in one sample. Considering that irregulation of proteolytic activity is an important hallmark of various diseases such as cancer progression and the development of athersclerotic plaques [75, 76], these types of QD-BRET probes have broad potential applications as in vivo molecular imaging agents.
P1: OTA/XYZ P2: ABC c07 JWBS037-Chen January 25, 2011
172
9:13
Printer Name: Yet to Come
QUANTUM DOTS FOR IN VIVO MOLECULAR IMAGING
7.9 RECENT DEVELOPMENTS 7.9.1 Reducing the Size One important issue with targeted in vivo imaging with nanoparticles is the opsonization and strong uptake by the RES system, which entraps the majority of the particles shortly after they enter the circulation system, thus preventing them from reaching the targted sites efficiently. RES uptake is influenced by many factors including particle size, shape, and surface chemistry. According to a study by the Frangioni laboratory [77], the hydrodynamic size (HD) of a nanoparticle has to be equal to or less than 5.5 nm in order to completely evade the RES organs (no accumulation in the liver, spleen, or lung) and be cleared by the renal system. However, the QDs are at least over 10 nm in HD. One approach to downsize QDs is through engineering the coating. Recently, Smith and Nie reported a new class of multifunctional multidentate polymer ligands, which not only minimized the HD of QDs but also preserved the colloidal stability and photobleaching/signal brightness [78]. A major finding is that a mixed composition of thiol (–SH) and amine (–NH2 ) groups grafted to a linear polymer chain can lead to a highly compact QD with long-term colloidal stability, strong resistance to photobleaching, and high fluorescence quantum yield. In contrast to the standing brush-like conformations of PEGylated dihydrolipoic acid ligands and monovalent thiols, these multidentate polymer ligands can wrap around the QD in a closed “loops-and-trains” conformation. Using this method, a new generation of bright and stable CdTe QDs with small hydrodynamic diameters between 5.6 and 9.7 nm, with fluorescence emission tunable from the visible (515 nm) to the near-infrared (720 nm), were prepared. In addition to CdTe nanocrystals, the same coating method is applicable to a broad range of core nanocrystals as well as core/shell nanostructures including CdS, ZnSe, CdSe/ZnS, and CdTe/CdS. The in vivo behavior and utility for imaging these nanoparticles remain to be evaluated.
7.9.2 Reducing/Eliminating Toxicity Cell culture studies [79] indicate that CdSe QDs are highly toxic to cultured cells under UV illumination for extended periods of time. This is not surprising because the energy of UV irradiation is close to that of a covalent chemical bond and dissolves the semiconductor particles in a process known as photolysis, releasing toxic cadmium ions into the culture medium. In the absence of UV irradiation, QDs with a stable polymer coating are likely to be much less toxic to cells and animals and this has been confirmed by a number of in vivo animal studies [17, 56]. Still, the perceived toxicity of cadmium has cast a doubtful future for cadmium-based QDs and intrigued the development of noncadmium QD alternatives. Among these alternatives are InAs QDs (will be discussed in Section 7.9.3), doped QDs, and carbon QDs. Doped QDs (d-dots) that do not contain toxic heavy metal ions have been actively studied by the Peng laboratory [80–82]. A d-dot often consists of a semiconductor nanocrystal core, such as ZnSe, doped with a transition metal ion such as Cu or Mn. In 2005, Peng and co-workers reported a method that yields doped ZnSe with high purity and tunable emission. The key feature of the related synthetic chemistry is decoupling the doping from nucleation and/or growth through nucleation-doping and growth-doping strategies [80]. Further optimization of the doping chemistry improved the PL quantum yield of the Mn-doped ZnSe d-dots (Mn:ZnSe d-dots) to the level of typical CdSe q-dots, as high as 60–70% PL QY measured against organic dyes [81]. Mn:ZnSe d-dots with a tunable
P1: OTA/XYZ P2: ABC c07 JWBS037-Chen January 25, 2011
9:13
Printer Name: Yet to Come
RECENT DEVELOPMENTS
173
photoluminescence peak position were made water soluble by coating with a monolayer of mercaptopropionic acid. The overall size of such d-dots/ligand complexes is only about 7–8 nm, implying an excellent permeability in biological issues. Carbon quantum dots are produced via laser ablation of a carbon target in the presence of water vapor with argon as carrier gas. The as-produced samples were dominated by nanoscale carbon particles in aggregates of various sizes and showed no detectable photoluminescence. Further acidic treatment and subsequent surface passivation resulted in bright luminescence emissions [83]. These dots are about 5 nm in diameter under TEM observation and have a continuous absorption spectra and can emit light from 450 to 700 nm depending on the excitation light. A follow-up study discovered the two-photon emissions and showed that they’re bright enough for cellular imaging [84]. Both the d-dots and carbon dots are promising nontoxic alternatives to the Cd-based QDs; however, more work is needed to make them comparable (optical properties, stability, and surface functionlization) to the CdSe QDs as in vivo molecular imaging probes. 7.9.3 Shifting to the Red: NIR QDs As mentioned earlier, the ideal QD emission for in vivo molecular imaging applications should be in the near-infrared (700–900 nm) [53]. This issue was noticed when QDs were first applied for in vivo molecular imaging and high-quality NIR QDs have been under active development since then. By far, there are two major caetgories of NIR QDs tested for in vivo imaging purposes: Cd-based (e.g., type II QDs have CdTe core [23, 55]) and non-Cd-based (InAs, InP, etc.) [85, 86]. The Bawendi laboratory at MIT and the Frangioni group at Harvard are the pioneers in developing NIR QDs and applying them for in vivo imaging [55, 85, 86]. Different formulations have been developed and tested in vivo. Type II QDs consisting of CdTe had been applied for sentinel lymph node mapping in 2004 [56]. Most recently, they reported the synthesis of a size series of (InAs)ZnSe (core)shell QDs that emit in the near-infrared and exhibit hydrodynamic size <10 nm [86] and demonstrated their utility in vivo by imaging multiple, sequential lymph nodes and showing extravasation from the vasculature in rat models. These QDs consist of an InAs core of <2 nm and a ZnSe shell and quantum yields were 6–9% in water. Coating DHLA-PEG renders water solubility and the total hydrodynamic size became 8.8 nm. When injected subcutaneously into the paw of a mouse or rat, the QDs rapidly migrated to the SLN, as previously reported for two other QD formulations. However, the new QDs did not get trapped completely in the SLN, but instead migrated further into the lymphatic system, imaging up to 5 lymph nodes sequentially, as well as the channels between nodes. A second new observation was the extravasations of the QDs from the vasculature when injected intravenously. And these were attributed to the unusual small size and the resistance to biofouling of the this new type of QDs. A more recent study showed that InAs-based QDs had longer circulating half-life with significant renal clearance. Xie et al. [87] reported the synthesis of small InAs /InP/ZnSe QDs (below 10 nm) that emit NIR light. After being administered to nude mice intravenously, the particle remained in the bloodstream for a prolonged time (110 min, circulation half-time vs. 6 min for its commercial peer) and showed evidence of improved renal clearance as compared to commercial NIR QDs both from wholebody imaging and ex vivo imaging of the different organs. Considering that InAs and InP are also more “tolerant” (compared to Cd-, Pd-, Hg-based QDs) in the body, these studies collectively have shown that InAs-based QDs are most likely the future for QD-based in vivo molecular imaging. As for Cd-based NIR dots, a recent study by Smith and co-workers [88]
P1: OTA/XYZ P2: ABC c07 JWBS037-Chen January 25, 2011
174
9:13
Printer Name: Yet to Come
QUANTUM DOTS FOR IN VIVO MOLECULAR IMAGING
demonstrated that the epitaxial deposition of a compressive shell (ZnS, ZnSe, ZnTe, CdS, or CdSe) onto a soft nanocrystalline core (CdTe) to form a lattice-mismatched quantum dot can dramatically change the conduction and valence band energies of both the core and the shell. In particular, standard type-I quantum-dot behavior is replaced by type-II behavior, which is characterized by spatial separation of electrons and holes, extended excited-state lifetimes, and giant spectral shifts. Moreover, the strain induced by the lattice mismatch can be used to tune the light emission—which displays narrow linewidths and high quantum yields—across the visible and near-infrared part of the spectrum (500–1050 nm). Owing to their near-infrared emission spectra, small sizes, and low cadmium content (only the CdTe core contains cadmium), this new class of QDs is also attractive for multicolor molecular, cellular, and in vivo imaging. 7.9.4 Getting More from One Particle: QDs in Multimodality Molecular Imaging The recognition that many biomedical imaging modalities provide complementary information has stimulated intense interest in multimodality imaging, using more than one modality to probe a sample of interest. For example, a marriage of MRI and optical techniques will provide anatomical/biodistribution info (MRI) and detailed information at the subcellular level (optical). QDs have been explored for multimodality molecular imaging by several groups around the world. Moulder et al. [89, 90] from The Netherlands has developed an MRI/optical dual-modality imaging system by encapsulating QDs with paramagnetic lipids such as high-density lipids (HDLs) and applied for the imaging of atherosclerotic plaques. In another example, a positron emission tomography/near-infrared fluorescence dual-modality imaging probe was made by conjugating NIR QDs with DOTA, which chelates 64 Cu, which is a radiotracer for PET imaging [91]. While multimodal probes can be made by simply conjugating probes of different functionality, a more elegant solution is to incorporate multiple functionalities in a single probe. Wang and co-workers reported the synthesis of core–shell quantum dots with high relaxivity and photoluminescence [92]. The dual-modality QDs have a composition of CdSe/Zn1-x Mnx S and can be made into watersoluble nanoparticles by capping the core/shell particles with amphiphilic polymer, and the QY values in water reached 21%. These materials also demonstrated high relaxivity with r1 values in the range of 11–18 mM-1 s-1 (at room temperature, 7 T). In vitro cell culture results showed that the QY and manganese concentration in the particles were sufficient to produce contrast for both modalities at relatively low concentrations of nanoparticles. The in vivo performance of these particles remains to be tested.
7.10 ISSUES AND PERSPECTIVES 7.10.1 Less RES Uptake and Longer Circulation Time The first issue is their relatively short circulation half-life, preventing long-term imaging or cell tracking studies; and this applies to all nanostructures injected systematically into a living body. Literature on the in vivo studies using QDs for imaging have revealed that their circulation half-time is influenced strongly by the surface chemistry and that they are cleared from the circulation primarily by phagocytosis of the nanoparticle by RES in the liver, spleen, and lymph nodes [17, 56]. Coating with PEG increases the circulation half-life, and
P1: OTA/XYZ P2: ABC c07 JWBS037-Chen January 25, 2011
9:13
Printer Name: Yet to Come
ISSUES AND PERSPECTIVES
175
attachment to targeting moieties (such as antibodies) reduces the dose needed to generate contrast between normal and tumor tissue [17]. Even with these strategies incorporated, the majority of nanoparticles still end up in the RES. A recent study using commercial nontargeted amino (PEG) QD705 for imaging glioma in rat suggested that maximal RES phagocytosis of QDs was reached between 3.4- and 8.5-nmol doses [93]. Significant tumor uptake of QDs was observed for doses higher than 8.5 nmole. Such a high dose might pose health hazards to living subjects. Thus there is an urgent need for engineering strategies to improve the circulation half-time of QDs, since the longer they can stay in the circulation system the better are the chances that they may be able to get to the tumor site. This might be achieved by minimizing the opsonization or other components of the RES. Intending to attack this problem, a few groups have started studying the interaction between blood components and nanoparticles [94, 95]. The rationale is that nanoparticles, once entered into the bloodstream, are immediately covered by plasma proteins; what the RES system “sees” and what defines the identity of the nanoparticle is largely the protein corona around the particle, not the core material. Elucidation of the profiles of adsorbed proteins on nanoparticles has the potential to facilitate engineering a surface chemistry that is less prone to opsonization and RES uptake. Another issue with nanoparticle-based molecular imaging is in vivo delivery and efficient extravasation to tumor sites. Using intravital microscopy, Smith et al. [96] discovered that QDs with tumor targeting arginine–glycine–aspartic acid (RGD) peptide were not able to extravasate in an SKOV-3 mouse ear tumor model; specific binding only occurred in the tumor neovasculature with aggregated QD conjugates. 7.10.2 Deeper Tissue Penetration Although the superior brightness and photostability of QDs made them attractive candidates for in vivo animal imaging, most of the current QDs still emit within the visible range. The ideal QDs for deep tissue imaging, that is, high-quality QDs with near-infrared-emitting properties, are not yet available. Recent developments include a promising water-based synthesis method that yields particles that emit from the visible to the NIR spectrum and are intrinsically water soluble, but the particles have yet to be tested in biological environments. Most materials (e.g., PdS, PdSe, CdHgTe, and CdSeTe) are either not bright enough or not stable enough for biomedical imaging applications. As such, there is an urgent need to develop bright and stable near-infrared-emitting QDs that are broadly tunable in the farred and infrared spectral regions [17]. In addition to high-quality NIR QDs, multiphoton fluorescence microscopy and novel illuminating mechanisms such as bioluminescence energy transfer can all be used to achieve deeper tissue penetration. 7.10.3 Perspectives Quantum dots as novel fluorescent probes have proved to be tremendously useful in many areas of biological and medical research, especially multiplexed tissue/cell labeling and live cell imaging as well as in vivo imaging. However, as an in vivo imaging agent, QDs are still at a very premature stage of development. Several issues (including toxicity, size, and trapping by the RES) remain to be solved before their potential can fully be exploited in this arena and applied to human subjects. Although the performances of visible QDs are greatly improved over conventional fluorophores, an ideal QD fluorophore should emit in the near-infrared/far-red region with high quantum yield and excellent stability. Apart from being an optical imaging agent, QD-based or multimodality molecular imaging have also
P1: OTA/XYZ P2: ABC c07 JWBS037-Chen January 25, 2011
176
9:13
Printer Name: Yet to Come
QUANTUM DOTS FOR IN VIVO MOLECULAR IMAGING
merged and will be the future of molecular imaging. In summary, quantum dot technology for in vivo cancer imaging is still an area of active research and will require the ongoing collaboration of chemists, biologists, and material scientists for its full potential to be released.
REFERENCES 1. Chan, W. C. W.; Maxwell, D. J.; Gao, X. H.; Bailey, R. E.; Han, M. Y.; Nie, S. M. Luminescent quantum dots for multiplexed biological detection and imaging. Curr. Opin. Biotechnol. 2002, 13, 40–46. 2. Medintz, I. L.; Mattoussi, H.; Clapp, A. R. Potential clinical applications of quantum dots. Int. J. Nanomed. 2008, 3(2), 151–167. 3. Bentolila, L. A.; Ebenstein, Y.; Weiss, S. Quantum dots for in vivo small-animal imaging J. Nucl. Med. 2009, 50, 493–496. 4. Li, Z. B.; Cai, W.; Chen, X. J. Semiconductor quantum dots for in vivo imaging. Nanosci. Nanotechnol. 2007, 7(8), 2567–2581. 5. Larson, D. R.; Zipfel, W. R.; Williams, R. M.; Clark, S. W.; Bruchez, M. P.; Wise, F. W.; Webb, W. W. Water-soluble quantum dots for multiphoton fluorescence imaging in vivo. Science 2003, 300, 1434–1436. 6. Alivisatos, A. P. Perspectives on the physical chemistry of semiconductor nanocrystals. J. Phys. Chem. 1996, 100, 13226–13239. 7. Sutherland, A. J. Quantum dots as luminescent probes in biological systems. Curr. Opin. Solid State Mater. Sci. 2002, 6, 365–370. 8. Peng, X. G. Mechanisms for the shape-control and shape evolution of colloidal semiconductor nanocrystals. Adv. Mater. 2003, 15, 459–463. 9. Manna, L.; Milliron, D. J.; Meisel, A.; Scher, E. C.; Alivisatos, A. P. Controlled growth of tetrapod-branched inorganic nanocrystals. Nat. Mater. 2003, 2, 382–385. 10. Milliron, D. J.; Hughes, S. M.; Cui, Y.; Manna, L.; Li, J.; Wang, L. W.; Alivisatos, A. P. Colloidal nanocrystal heterostructures with linear and branched topology. Nature 2004, 430, 190–195. 11. Dick, K. A.; Deppert, K.; Larsson, M. W.; Martensson, T.; Seifert, W.; Wallenberg, L. R.; Samuelson, L. Synthesis of branched “nanotrees” by controlled seeding of multiple branching events. Nat. Mater. 2004, 3, 380–384. 12. Yu, W. W.; Wang, Y. A.; Peng, X. G. Formation and stability of size-, shape-, and structurecontrolled CdTe nanocrystals: ligand effects on monomers and nanocrystals. Chem. Mater. 2003, 15, 4300–4308. 13. Hines, M. A.; Scholes, G. D. Colloidal PbS nanocrystals with size-tunable near-infrared emission: observation of post-synthesis self-narrowing of the particle size distribution. Adv. Mater. 2003, 15, 1844–1849. 14. Bailey, R. E.; Nie, S. Alloyed semiconductor quantum dots: tuning the optical properties without changing the particles. J. Am. Chem. Soc. 2003, 125, 7100–7106. 15. Efros, A. L.; Rosen, M. The electronics structure of semiconductor nanocrystals. Annu. Rev. Mater. Sci. 2000, 30, 475–521. 16. Smith, A. M.; Gao, X.; Nie, S. Quantum dot nanocrystals for in vivo molecular and cellular imaging. Photochem. Photobiol. 2004, 80, 377–385. 17. Leatherdale, C. A.; Woo, W. K.; Mikulec, F. V.; Bawendi, M. G. On the absorption cross section of CdSe nanocrystal quantum dots. J. Phys. Chem. B 2002, 106, 7619–7622.
P1: OTA/XYZ P2: ABC c07 JWBS037-Chen January 25, 2011
9:13
Printer Name: Yet to Come
REFERENCES
177
18. Gao, X.; Cui, Y.; Levenson, R. M.; Chung, L. W.; Nie, S. In vivo cancer targeting and imaging with semiconductor quantum dots. Nat. Biotechnol. 2004, 22, 969–976. 19. Dahan, M.; Laurence, T.; Pinaud, F.; Chemla, D. S.; Alivisatos, A. P.; Sauer, M.; Weiss, S. Time-gated biological imaging by use of colloidal quantum dots. Opt. Lett. 2001, 26, 825–827. 20. Chien, G. L.; Anselone, C. G.; Davis, R. F.; Van Winkle, D. M. Fluorescent vs. radioactive microsphere measurement of regional myocardial blood flow. Cardiovasc. Res. 1995, 30, 405–412. 21. Murray, C. B.; Norris, D. J.; Bawendi, M. G. Synthesis and characterization of nearly monodisperse CdE (E = S, Se, Te) semiconductor nanocrystals. J. Am. Chem. Soc. 1993, 115, 8706– 8715. 22. Han, M.; Gao, X.; Su, J. Z.; Nie, S. Quantum-dot-tagged microbeads for multiplexed optical coding of biomolecules. Nat. Biotechnol. 2001, 19, 631–635. 23. Deng, Z.; Lie, F. L.; Shen, S.; Ghosh, I.; Mansuripur, M.; Muscat, A. J. Water-based route to ligand-selective synthesis of ZnSe and Cd-doped ZnSe quantum dots with tunable ultraviolet A to blue photoluminescence. Langmuir 2009, 25(1), 434–442. 24. Kim, S.; Fisher, B.; Eisler, H. J.; Bawendi, M. G. Type-II quantum dots: CdTe/CdSe(core/shell) and CdSe/ZinTe(core/shell) heterostructures. J. Am. Chem. Soc. 2003, 125, 11466–11467. 25. Pietryga, J. M.; Schaller, R. D.; Werder, D.; Stewart, M. H.; Klimov, V. I.; Hollingsworth, J. A. Pushing the band gap envelope: mid-infrared emitting colloidal PbSe quantum dots. J. Am. Chem. Soc. 2004, 126, 11752–11753. 26. Zhong, X. H.; Feng, Y. Y.; Knoll, W.; Han, M. Y. Alloyed ZnxCd(1-x)S nanocrystals with highly narrow luminescence spectral width. J. Am. Chem. Soc. 2003, 125, 13559–13563. 27. Hines, M. A.; Guyot-Sionnest, P. Synthesis and characterization of strongly luminescing ZnSCapped CdSe nanocrystals. J. Phys. Chem. 1996, 100, 468–471. 28. Reiss, P.; Bleuse, J.; Pron, A. Highly luminescent CdSe/ZnSe core/shell nanocrystals of low size dispersion. Nano Lett. 2002, 2, 781–784. 29. Guo, W. H.; Li, J. J.; Wang, Y. A.; Peng, X. G. Luminescent CdSe/CdS core/shell nanocrystals in dendron boxes: superior chemical, photochemical and thermal stability. J. Am. Chem. Soc. 2003, 125 3901–3909. 30. Sukhanova, A.; Devy, J.; Venteo, L.; Kaplan, H.; Artemyev, M.; Oleinikov, V.; Klinov, D.; Pluot, M.; Cohen, J. H.; Nabiev, I. Biocompatible fluorescent nanocrystals for immunolabeling of membrane proteins and cells. Anal. Biochem. 2004, 324, 60–67. 31. Chan, W. C.; Nie, S. Quantum dot bioconjugates for ultrasensitive nonisotopic detection. Science 1998, 281, 2016–2018. 32. Soo, C. H.; Liu, W.; Misra, P.; Tanaka, E.; Zimmer, J. P.; Itty Ipe, B.; Bawendi, M. G.; Frangioni, J. V. Renal clearance of quantum dots. Nat. Biotechnol. 2007, 25, 1165–1170. 33. Pathak, S.; Choi, S. K.; Arnheim, N.; Thompson, M. E. Hydroxylated quantum dots as luminescent probes for in situ hybridization. J. Am. Chem. Soc. 2001, 123, 4103–4104. 34. Mattousi, H.; Mauro, J. M.; Goldman, E. R.; Anderson, A. P.; Sunder, V. C.; Mikulec, F. V.; Bawendi, M. G. Self-assembly of CdSe–ZnS quantum dots bioconjugates using an engineered recombinant protein. J. Am. Chem. Soc. 2000, 122, 12142–12150. 35. Pinaud, F.; King, D.; Moore, H. P.; Weiss, S. Bioactivation and cell targeting of semiconductor CdSe/ZnS nanocrystals with phytochelatin-related peptides. J. Am. Chem. Soc. 2004, 126, 6115–6123. 36. Wu, X. Y.; Liu, H. J.; Liu, J. Q.; Haley, K. N.; Treadway, J. A.; Larson, J. P.; Ge, N. F.; Peale, F.; Bruchez, M. P. Immunofluorescent labeling of cancer marker Her2 and other cellular targets with semiconductor quantum dots. Nat. Biotechnol. 2003, 21, 41–46. 37. Bruchez, M. J.; Moronne, M.; Gin, P.; Weiss, S.; Alivisatos, A. P. Semiconductor nanocrystals as fluorescent biological labels. Science 1998, 281, 2013–2016.
P1: OTA/XYZ P2: ABC c07 JWBS037-Chen January 25, 2011
178
9:13
Printer Name: Yet to Come
QUANTUM DOTS FOR IN VIVO MOLECULAR IMAGING
38. Dubertret, B.; Skourides, P.; Norris, D. J.; Noireaux, V.; Brivanlou, A. H.; Libchaber, A. In vivo imaging of quantum dots encapsulated in phospholipid micelles. Science 2002, 298, 1759–1762. 39. Pellegrino, T. M. L.; Kudera, S. Hydrophobic nanocrystals coated with an amphiphilic polymer shell: a general route to water soluble nanocrystals. Nano Lett. 2004, 4, 703–707. 40. Medintz, I. L.; Clapp, A. R.; Brunel, F. M.; Tiefenbrunn, T.; Uyeda, H. T.; Chang, E. L.; Deschamps, J. R.; Dawson, P. E.; Mattoussi, H. Proteolytic activity monitored by fluorescence resonance energy transfer through quantum-dot–peptide conjugates. Nat. Mater. 2006, 5, 581–589. 41. Alivisatos, P.; Gu, W.; Larabell, C. Quantum dots as fluorescent probes. Annu. Rev. Biomed. Eng. 2005, 7, 55–76. 42. Lakowicz, J. R.; Gryczynski, I.; Gryczynski, Z.; Nowaczyk, K.; Murphy, C. J. Time-resolved spectral observations of cadmium-enriched cadmium sulfide nanoparticles and the effects of DNA oligomer binding. Anal. Biochem. 2000, 280, 128–136. 43. Mahtab, R.; Harden, H. H.; Murphy, C. J. Temperature- and salt-dependent binding of long DNA to protein-sized quantum dots: thermodynamics of “inorganic protein”–DNA interactions. J. Am. Chem. Soc. 2000, 122, 14–17. 44. Hanaki, K.; Momo, A.; Oku, T.; Komoto, A.; Maenosono, S.; Yamaguchi, Y.; Yamamoto, K. Semiconductor quantum dot/albumin complex is a long-life and highly photostable endosome marker. Biochem. Biophys. Res. Commun. 2003, 302, 496–501. 45. Patolsky, F.; Gill, R.; Weizmann, Y.; Mokari, T.; Banin, U.; Willner, I. Lighting-up the dynamics of telomerization and DNA replication by CdSe–ZnS quantum dots. J. Am. Chem. Soc. 2003, 125, 13918–13919. 46. Mitchell, G. P.; Mirkin, C. A.; Letsinger, R. L. Programmed assembly of DNA functionalized quantum dots. J. Am. Chem. Soc. 1999, 121, 8122–8123. 47. Rosenthal, S. J.; Tomlinson, I.; Adkins, E. M.; Schroeter, S.; Adams, S.; Swafford, L.; McBride, J.; Wang, Y.; DeFelice, L. J.; Blakely, R. D. Targeting cell surface receptors with ligand conjugated nanocrystals. J. Am. Chem. Soc. 2002, 124, 4586–4594. 48. Winter, J. O.; Liu, T. Y.; Korgel, B. A.; Schmidt, C. E. Recognition molecule directed interfacing between semiconductor quantum dots and nerve cells. Adv. Mater. 2001, 13, 1673–1677. 49. Akerman, M. E.; Chan, W. C.; Laakkonen, P.; Bhatia, S. N.; Ruoslahti, E. Nanocrystal targeting in vivo. Proc. Natl. Acad. Sci. U.S.A. 2002, 99, 12617–12621. 50. Xing, Y.; Chaudry, Q.; Shen, C.; Kong, K. Y.; Zhau, H. E.; Chung, L. W.; Petros, J. A.; O’Regan, R. M.; Yezhelyev, M. V.; Simons, J. W.; Wang, M. D.; Nie, S. Bioconjugated quantum dots for multiplexed and quantitative immunohistochemistry. Nat. Protoc. 2007, 2, 1152–1165. 51. Xiao, Y.; Barker, P. E. Semiconductor nanocrystal probes for human metaphase chromosomes. Nucleic Acids Res. 2004, 32, e28. 52. Gao, X.; Chan, W. C.; Nie, S. Quantum dot nanocrystals for ultrasensitive biological labeling and multicolor optical encoding. J. Biomed. Opt. 2002, 7, 532–537. 53. Zhang, Y.; So, M. K.; Loening, A. M.; Yao, H.; Gambhir, S. S.; Rao, J. HaloTag protein-mediated site-specific conjugation of bioluminescent proteins to quantum dots. Angew. Chem. Int. Ed. Engl. 2006, 45, 4936–4940. 54. Weissleder, R. A clearer vision for in vivo imaging. Nat. Biotechnol. 2001, 19, 316–317. 55. Cai, W.; Shin, D. W.; Chen, K.; Gheysens, O.; Cao, Q.; Wang, S. X.; Gambhir, S. S.; Chen, X. Peptide-labeled near-infrared quantum dots for imaging tumor vasculature in living subjects. Nano Lett. 2006, 6, 669–676. 56. Kim, S.; Lim, Y. T.; Soltesz, E. G.; De Grand, A. M.; Lee, J.; Nakayama, A.; Parker, J. A.; Mihaljevic, T.; Laurence, R. G.; Dor, D. M.; Cohn, L. H.; Bawendi, M. G.; Frangioni, J. V. Nearinfrared fluorescent type II quantum dots for sentinel lymph node mapping. Nat. Biotechnol. 2004, 22, 93–97.
P1: OTA/XYZ P2: ABC c07 JWBS037-Chen January 25, 2011
9:13
Printer Name: Yet to Come
REFERENCES
179
57. Ballou, B.; Lagerholm, B. C.; Ernst, L. A.; Bruchez, M. P.; Waggoner, A. S. Noninvasive imaging of quantum dots in mice. Bioconjug. Chem. 2004, 15, 79–86. 58. Hoshino, A.; Hanaki, K.; Suzuki, K.; Yamamoto, K. Applications of T-lymphoma labeled with fluorescent quantum dots to cell tracing markers in mouse body. Biochem. Biophys. Res. Commun. 2004, 314, 46–53. 59. Lewin, M.; Carlesso, N.; Tung, C. H.; Tang, X. W.; Cory, D.; Scadden, D. T.; Weissleder, R. Tat peptide-derivatized magnetic nanoparticles allow in vivo tracking and recovery of progenitor cells. Nat. Biotechnol. 2000, 18, 410–414. 60. Schipper, M. L.; Cheng, Z.; Lee, S. W.; Bentolila, L. A.; Iyer, G.; Rao, J.; Chen, X.; Wu, A. M.; Weiss, S.; Gambhir, S. S. MicroPET-based biodistribution of quantum dots in living mice. J. Nucl. Med. 48(9), 1511–1518. 61. Schipper, M. L.; Iyer, G.; Koh, A. L.; Cheng, Z.; Ebenstein, Y.; Aharoni, A.; Keren, S.; Bentolila, L. A.; Li, J.; Rao, J.; Chen, X.; Banin, U.; Wu, A. M.; Sinclair, R.; Weiss, S.; Gambhir, S. S. Particle size, surface coating, and PEGylation influence the biodistribution of quantum dots in living mice. Small. 2009, 5(1), 126–134. 62. Cai, W.; Chen, X. Preparation of peptide-conjugated quantum dots for tumor vasculature-targeted imaging. Nat. Protoc. 2008, 3(1), 89–96. 63. Cai, W.; Chen, K.; Li, Z. B.; Gambhir, S. S.; Chen, X. Dual-function probe for PET and near-infrared fluorescence imaging of tumor vasculature. J. Nucl. Med. 2007, 48(11), 1862– 1870. 64. Chen, K.; Li, Z. B.; Wang, H; Cai, W.; Chen, X. Dual-modality optical and positron emission tomography imaging of vascular endothelial growth factor receptor on tumor vasculature using quantum dots. Eur. J. Nucl. Med. Mol. Imaging 2008, 35(12), 2235. 65. Van Der Meer, B. W.; Coker, G. III.; Chen, S.-Y. S. Resonance Energy Transfer: Theory and Data. Wiley: Hoboken, NJ, 1994. 66. Clapp, A. R.; Medintz, I. L.; Uyeda, H. T.; Fisher, B. R.; Goldman, E. R.; Bawendi, M. G.; Mattoussi, H. Quantum dot-based multiplexed fluorescence resonance energy transfer. J. Am. Chem. Soc. 2005, 127, 18212–18221. 67. Willard, D. M.; Carillo, L. L.; Jung, J.; Van Orden, A. CdSe–ZnS quantum dots as resonance energy transfer donors in a model protein–protein binding assay. Nano Lett. 2001, 1, 469–474. 68. Medintz, I. L.; Clapp, A. R.; Mattoussi, H.; Goldman, E. R.; Fisher, B.; Mauro, J. M. Selfassembled nanoscale biosensors based on quantum dot FRET donors. Nat. Mater. 2003, 2, 630–638. 69. Contag, C. H.; Bachmann, M. H. Advances in in vivo bioluminescence imaging of gene expression. Annu. Rev. Biomed. Eng. 2002, 4, 235–260. 70. De, A.; Gambhir, S. S. Noninvasive imaging of protein–protein interactions from live cells and living subjects using bioluminescence resonance energy transfer. FASEB J. 2005, 19, 2017–2019. 71. De, A.; Loening, A. M.; Gambhir, S. S. An improved bioluminescence resonance energy transfer strategy for imaging intracellular events in single cells and living subjects. Cancer Res. 2007, 67, 7175–7183. 72. So, M. K.; Xu, C.; Loening, A. M.; Gambhir, S. S.; Rao, J. Self-illuminating quantum dot conjugates for in vivo imaging. Nat. Biotechnol. 2006, 24, 339–343. 73. Loening, A. M.; Fenn, T. D.; Wu, A. M.; Gambhir, S. S. Consensus guided mutagenesis of Renilla luciferase yields enhanced stability and light output. Protein Eng. Des Sel. 2006, 19, 391–400. 74. Xia, Z.; Xing, Y.; So, M. K.; Koh, A. L.; Sinclair, R.; Rao, J. Multiplex detection of protease activity with quantum dot nanosensors prepared by intein-mediated specific bioconjugation, Anal. Chem. 2008, 80, 8649–8655. 75. Egeblad, M.; Werb, Z. New functions for the matrix metalloproteinases in cancer progression. Nat. Rev. Cancer 2002, 2(3), 161–174.
P1: OTA/XYZ P2: ABC c07 JWBS037-Chen January 25, 2011
180
9:13
Printer Name: Yet to Come
QUANTUM DOTS FOR IN VIVO MOLECULAR IMAGING
76. Malemud, C. J. Matrix metalloproteinases (MMPs) in health and disease: an overview. Front. Biosci. 2006, 11 (May 1), 1696–1701. 77. Choi, H. S.; Liu, W.; Misra, P.; Tanaka, E.; Zimmer, J. P.; Itty, I. B.; Bawendi, M. G.; Frangioni, J. V. Renal clearance of quantum dots, Nat. Biotechnol. 2007, 25, 1165–1170. 78. Smith, A. M.; Nie, S. Minimizing the hydrodynamic size of quantum dots with multifunctional, multidentate polymer ligands. J. Am. Chem. Soc. 2008, 130(34), 11278–11279. 79. Jackson, H.; Muhammad, O.; Daneshvar, H.; Nelms, J.; Popescu, A.; Vogelbaum, M. A.; Bruchez, M.; Toms, S. A. Quantum dots are phagocytized by macrophages and colocalize with experimental gliomas. Neurosurgery 2007, 50, 524–529. 80. Pradhan, N.; Goorskey, D.; Thessing, J.; Peng, X. An alternative of CdSe nanocrystal emitters: pure and tunable impurity emissions in ZnSe nanocrystals J. Am. Chem. Soc. 2005, 127(50), 17586–17587. 81. Pradhan, N.; Battaglia, D. M.; Liu, Y.; Peng, X. Efficient, stable, small, and water-soluble doped ZnSe nanocrystal emitters as non-cadmium biomedical labels. Nano Lett. 2007, 7, 312–317. 82. Pradhan, N.; Peng, X. Efficient and color-tunable Mn-doped ZnSe nanocrystal emitters: control of optical performance via greener synthetic chemistry. J. Am. Chem. Soc. 2007, 129(11), 3339–3347. 83. Sun, Y. P.; Zhou, B.; Lin, Y.; Wang, W.; Fernando, K. A.; Pathak, P.; Meziani, M. J.; Harruff, B. A.; Wang, X.; Wang, H.; Luo, P. G.; Yang, H.; Kose, M. E.; Chen, B.; Veca, L. M.; Xie, S. Y. Quantum-sized carbon dots for bright and colorful photoluminescence. J. Am. Chem. Soc. 2006, 128(24), 7756–7757. 84. Cao, L.; Wang, X.; Meziani, M. J.; Lu, F.; Wang, H.; Luo, P. G.; Lin, Y.; Harruff, B. A.; Veca, L. M.; Murray, D.; Xie, S. Y.; Sun, Y. P. Carbon dots for multi-photon bioimaging. J. Am. Chem. Soc. 2007, 129(37), 11318–11319. 85. Kim, S. W.; Zimmer, J. P.; Ohnishi, S.; Tracey, J. B. Frangioni, J. V.; Bawendi, M. G. Engineering InAsxP1-x/InP/ZnSe III–V alloyed core/shell quantum dots for the near-infrared. J. Am. Chem. Soc. 2005, 127(30), 10526–10532. 86. Zimmer, J. P.; Kim, S. W.; Ohnishi, S.; Tanaka, E.; Frangioni, J. V.; Bawendi, M. G. Size series of small indium arsenide–zinc selenide core–shell nanocrystals and their application to in vivo imaging. J. Am. Chem. Soc. 2006, 128, 2526–2527. 87. Xie, R.; Chen, K.; Chen, X.; Peng, X. InAs/InP/ZnSe core/shell/shell quantum dots as near-infrared emitters: bright, narrow-band, non-cadmium containing, and biocompatible InAs/InP/ZnSe QDs. Nano Res. 2008, 1(6), 457–464. 88. Smith, A. M.; Mohs, A. M.; Nie, S. Tuning the optical and electrical properties of colloidal nanocrystals by lattice strain. Nat. Nanotechnol. 2009, 4(1), 56–63. 89. Mulder, W. J.; Castermans, K.; van Beijnum, J. R.; Oude Egbrink, M. G.; Chin, P. T.; Fayad, Z. A.; L¨owik, C. W.; Kaijzel, E. L.; Que, I.; Storm, G.; Strijkers, G. J.; Griffioen, A. W.; Nicolay, K. Molecular imaging of tumor angiogenesis using alphav beta3 -integrin targeted multimodal quantum dots. Angiogenesis 2009, 12(1), 17–24. 90. Cormode, D. P.; Skajaa, T.; van Schooneveld, M. M.; Koole, R.; Jarzyna, P.; Lobatto, M. E.; Calcagno, C.; Barazza, A.; Gordon, R. E.; Zanzonico, P.; Fisher, E. A.; Fayad, Z. A.; Mulder, W. J. Nanocrystal core high-density lipoproteins: a multimodality contrast agent platform. Nano Lett. 2008, 8(11), 3715–3723. 91. Chen, K.; Li, Z. B.; Wang, H.; Cai, W. B.; Chen, X. Dual-modality optical and positron emission tomography imaging of vascular endothelial growth factor receptor on tumor vasculature using quantum dots. Eur. J. Nucl. Med. Mol. Imaging 2008, 35, 2235–2244. 92. Wang, S.; Jarrett, B. R.; Kauzlarich, S. M.; Louie, A. Y. Core/shell quantum dots with high relaxivity and photoluminescence for multimodality imaging. J. Am. Chem. Soc. 2007, 4; 129(13), 3848–3856.
P1: OTA/XYZ P2: ABC c07 JWBS037-Chen January 25, 2011
9:13
Printer Name: Yet to Come
REFERENCES
181
93. Jackson, H.,Muhammad, O.; Daneshvar, H.; Nelms, J.; Popescu, A.; Vogelbaum, M. A.; Bruchez, M.; Toms, A. Quantum dots are phagocytized by macrophages and colocalize with experimental gliomas. Neurosurgery 2007, 50, 524–529. 94. Cedervall, T.; Lynch, I.; Lindman, S.; Bergg˚ard, T.; Thulin, E.; Nilsson, H.; Dawson, K. A.; Linse, S. Understanding the nanoparticle–protein corona using methods to quantify exchange rates and affinities of proteins for nanoparticles. Proc. Natl. Acad. Sci. U.S.A. 2007, 104, 2050–2055. 95. Kim, H. R.; Andrieux, K.; Delomenie, C.; Chacun, H.; Appel, M.; Desma¨ele, D.; Taran, F.; Georgin, D.; Couvreur, P.; Taverna, M. Analysis of plasma protein adsorption onto PEGylated nanoparticles by complementary methods: 2-DE, CE and Protein Lab-on-chip system. Electrophoresis 2007, 28, 2252–2261. 96. Smith, B. R.; Cheng, Z.; De, A.; Koh, A. L.; Sinclair, R.; Gambhir, S. S. Real-time intravital imaging of RGD–quantum dot binding to luminal endothelium in mouse tumor neovasculature. Nano Lett., 2008, 8(9), 2599–2606.
P1: OTA/XYZ P2: ABC c08 JWBS037-Chen February 1, 2011
4:53
Printer Name: Yet to Come
CHAPTER 8
Biopolymer, Dendrimer, and Liposome Nanoplatforms for Optical Molecular Imaging DAVID PHAM, LING ZHANG, BO CHEN, and ELLA FUNG JONES Center for Molecular and Functional Imaging, Department of Radiology and Biomedical Imaging, University of California, San Francisco, California, USA
8.1 INTRODUCTION Recent advances in human genome mapping, molecular biology, and high-throughput screening have led to new discoveries of biomarkers and their corresponding binding ligands (peptides or small molecules) [1] for the development of targeted pharmaceuticals [2, 3]. However, the in vivo utility of small molecules as therapeutic or imaging agents often faces issues concerning their effective dose, stability, and rapid clearance. The appeal of macromolecular nanoconstructs as carriers is their multivalent and multifunctional attributes that allow high payload of drug molecules and active targeting groups to be incorporated to enhance therapeutic or imaging efficacy [4–6]. In addition, the ability to engineer the carrier surface with benign pharmacokinetic and/or biodistribution modifiers may provide a “stealth” environment to prevent uptake by macrophages and prolong in vivo blood half-life [7–9]. Biopolymers (60–250 nm), dendrimers (5–200 nm), and liposomes (20–250 nm) are common drug delivery systems. These nanoconstructs have long plasma half-life, and they can passively target aberrant tissues through the enhanced permeability and retention (EPR) effect [10]. These platforms set the precedent of effective drug delivery to diseased tissues while maintaining stability and therapeutic efficacy in circulation [11]. In a tumor environment, the defective leaky vessels and the lack of lymphatics favor the permeability and retention of nanoscale therapeutic and imaging agents [1, 12]. Normal vasculature with a uniform network has diameters ranging from 8 to 10 m and in between the vascular endothelial cells, there are gaps ranging from 5 to 10 nm. Depending on the tumor types, these gaps can be as large as 100–780 nm in a highly irregular network with vessel diameters ranging from 20 to 100 m [13–15]. Nanoscale pharmaceuticals take advantage of these physiological differences to differentiate pathological and normal tissues. However, with the rapid cell proliferation in tumorous regions, the blood vessels are
Nanoplatform-Based Molecular Imaging Edited by Xiaoyuan Chen C 2011 John Wiley & Sons, Inc. Copyright
183
P1: OTA/XYZ P2: ABC c08 JWBS037-Chen February 1, 2011
184
4:53
Printer Name: Yet to Come
BIOPOLYMER, DENDRIMER, AND LIPOSOME NANOPLATFORMS FOR OPTICAL
not always consistently leaky. In some regions, the vessels may be compressed or collapsed, and others may be leaky or even completely normal [16]. Nanoconstructs with differing chemical constitutions, size, shape, surface charge, and solubility may result in a wide range of pharmacokinetic and biodistribution profiles [11, 15, 17, 18]. It is therefore prudent to fully understand not only the in vivo properties of the nanoplatform itself, but also the intended disease target and clearance for the development of an effective therapeutic or imaging agent. In this chapter, we focus on the use of biopolymers, dendrimers, and liposomes as nanoplatforms for imaging. An overview of each type of platform and examples of their use in optical imaging are presented.
8.2 BIOPOLYMERS Biopolymers have been developed for medicinal use since the 1950s [19]. This class of nanoconstruct comprises three subtypes including (1) pseudosynthetic polymers, (2) synthetic polymers, and (3) natural polymers (see examples listed in Table 8.1) [9]. The application of biopolymers in therapeutics and imaging is limited by their consistent synthetic modification due to their polydispersity and entangling structure [37]. Nonetheless, biopolymers are well suited for use in controlled drug release in areas of cardiology, ophthalmology, endocrinology, orthopedics, and oncology [1]. Applications in optical imaging were not realized until the recent development of near-infrared (NIR) molecular beacons by Weissleder and co-workers [38]. 8.2.1 Optical Probe with a Pseudosynthetic Biopolymer Backbone Perhaps the most iconic application of biopolymers in optical imaging is the proteasesensing molecular beacons [38]. This class of optical nanoprobes consists of three components: the poly-(l-lysine) polymeric backbone, the self-quenching NIR cyanine dye, and the peptide substrate specific to protease known to overexpress in diseased tissues. These probes are assembled in a modular fashion. The PEGylated poly-l-lysine backbone is grafted with an optimal number of peptide substrates with an NIR fluorescent dye molecule conjugated at the N terminus. The resulting probe is fine-tuned with proteolytic cleavage sites for TABLE 8.1 Examples of Each Subtype of Biopolymer Biopolymers
Examples
Applications
Reference
Pseudosynthetic
Poly(amino acids), poly(glutamic acid) (PGA), poly(malic acid), and poly(aspartamides)
Anticancer Antiviral Tissue engineering
20, 21 22, 23 24, 25
Synthetic
N-(2-hydroxypropyl) methacrylamide (HPMA) copolymers, poly(vinylpyrrolidone) (PVP) and poly(ethyleneimine) (PEI)
Anticancer Multiple sclerosis Cholesterol control Drugs
26–29 30 31, 32 33
Natural
Dextran (␣-1,6 polyglucose), dextrin (␣-1,4 polyglucose), hyaluronic acid, and chitosans
Renal control HIV drug Antiviral
34 35 36
P1: OTA/XYZ P2: ABC c08 JWBS037-Chen February 1, 2011
4:53
Printer Name: Yet to Come
BIOPOLYMERS
185
efficient probe activation, while incorporating an optimal number of NIR dye molecules for effective self-quenching. In the absence of targeted protease, the self-quenched probe is optically silent, but upon exposure to targeted protease, the peptide substrate is cleaved and fragments containing dye molecules are released from the nanoprobe to give detectable fluorescent signal. Through phage display [39] and high-throughput screening, discoveries of potent peptide sequences have been identified for different diseases and cellular processes. The adaptation of various potent peptide sequences in the modular probe construct makes it an attractive approach to interrogate a wide variety of diseases such as cancers, rheumatoid arthritis, and arthrosclerosis [18]. 8.2.2 Nanoparticles Based on Synthetic Polymer Synthetic polymers have been used as protective coatings to improve the safety and stability of drugs. Gao and co-workers adapted the same strategy and created a triblock copolymer encapsulated quantum dot (QD) targeted optical imaging reporter [40]. The CdSe-ZnS QD is coated with a poly(butylacrylate)-b-poly(ethylacrylate)-b-poly(methylacrylic acid) block copolymer. The polymer-coated QD was then pegylated and functionalized with monoclonal antibody ligands that target prostate-specific membrane antigen (PSMA). When compared to the nontargeted PEGylated QD construct in mouse xenografts, the targeted probe showed a much stronger signal in the tumor. To fully take advantage of polyvalency [4], Cai and co-workers opted to use a PEG-coated QD conjugated with a smaller RGD peptide [41] and demonstrated a four-fold stronger signal-to-background ratio in in vivo imaging of integrin ␣v 3 -positive models than the corresponding nontargeted QD probe [42]. A novel reversibly photoswitchable polymeric NIR probe based on Nisopropylacrylamide (NIPAM) interspersed with styrene and divinylbenzene units was reported by Zhu and co-workers [43]. The polymer was conjugated with perylene diimide (PDI; ex ≈ 480 nm, em ≈ 535 nm) and spiropyran (SP). Spiropyran has two interchangeable conformations: the emissive opened merocyanine (mero form) (ex ≈ 588 nm, em ≈ 600–750 nm) and the nonemissive spiropyran (spiro form). The ring opening to merocyanine is induced by UV light, and the nonemissive spiro ring closure forms upon mild heating under visible light (Fig. 8.1). Although in water SP is not appreciably fluorescent in either form, the mero form is highly fluorescent in a hydrophobic environment. Upon UV irradiation, these particles produce a red emission by virtue of the fluorescence resonance energy transfer (FRET) effect from perylene dimide (PDI). The reverse effect is achieved through mild heating and exposure to visible light to give a green emission from the PDI molecules. A potential application of such a probe is to remove ambiguities in fluorescence microscopy stemming from tissue autofluorescence. 8.2.3 An NIR Theragnostic Agent Based on Natural Polymers Kwon and co-workers reported the derivatization of several glycol chitosan (GC) polymers (MW = 20 kDa, 100 kDa, and 250 kDa) with hydrophobic cholanic acid [44]. These nanoparticles were subsequently tagged with NIR dye Cy5.5 for in vivo pharmacokinetic and biodistribution studies, which showed greater tumor accumulation and longer blood circulation correlating with the higher molecular weight nanoparticles. This result led to the recent development of an optical theragnostic agent, an agent with both therapeutic and diagnostic attributes, based on the 250 kDa GC particle encapsulated with paclitaxel [45]. In mouse xenografts implanted with SCC7 tumor cells, preferential accumulation of the
blue excitation
186 uv
vis
uv
vis
blue excitation
red fluorescence
=
O
R
=
=
O
O
N
O
N
H
H
O_
Spiropyran (SP)
O
uv
O
O
O
+N
vis
O
R
NO2
NO2
O
N
O
O
perylene diimide (PDI)
Mero form
Spiro form
O
4:53
FIGURE 8.1 An illustration of the reversible dual-color fluorescence mechanism of the spiropyran polymeric nanoparticles. (Reproduced with permission from Zhu et al. [43]. Copyright © 2007 American Chemical Society.)
X
green fluorescence
P1: OTA/XYZ P2: ABC c08 JWBS037-Chen February 1, 2011 Printer Name: Yet to Come
P1: OTA/XYZ P2: ABC c08 JWBS037-Chen February 1, 2011
4:53
Printer Name: Yet to Come
DENDRIMERS
187
theragnostic agent in the tumor was observed by in vivo fluorescence imaging after 12 hours with signal persistence beyond 72 hours. Furthermore, the paclitaxel-loaded nanoparticle effectively inhibited tumor growth after the fifth administration on the 12th day.
8.3 DENDRIMERS Dendrimers are perfectly branched polymeric macromolecules that are synthetically constructed stepwise, layer-by-layer, emanating from a multifunctioned core or vice versa, known as divergent and convergent synthesis, respectively. Each additional layer, denoted as a generation, can drastically increase the number of the surface reactive functional groups. Unlike the traditional polymerization processes that result in polydispersed molecular weight, strict control over the synthesis of dendrimers results in monodispersed molecules, an important factor for reproducibility in pharmacological applications. Along with low polydispersity, dendrimers have well-defined surface functionalities for multivalency, and tunable parameters for optimization of pharmacokinetics. Dendrimers are typically 2–15 nm in diameter with a molecular weight ranging from 15 to 4000 kDa [46]. Pharmaceutically relevant dendrimers are based on polyamidoamines, polyamines, polyamides, poly(aryl ethers), polyesters, carbohydrates, and DNA [47]. Lower generation dendrimers are floppy disk-like structures, but beyond the fourth generation, they conform into the globular geometry similar to that of proteins [48]. The hydrodynamic size of a dendrimer depends on the compressibility of the chemical constituents, but the general trait for all dendrimers is that they become increasingly less compressible with increasing generations and steric hindrance. The shape and compressibility, along with biodegradability, are determining factors of the pharmacokinetics and feasibility of dendrimers in biological applications. Dendrimers of similar molecular weights with different geometrical shapes, like “bow-tie” [49] or “rigid rod” [50], and different sizes and cores [51] significantly alter the plasma half-life and biodistribution profiles [52]. PEGylated dendrimers prepared by conjugating poly(ethylene glycol) (PEG) to the dendrimer core have played an increasingly important role in biomedical research owing to their low toxicity, hemolytic properties, long circulation time, and high accumulation in tumor tissues [49]. 8.3.1 Optical Probes with a Dendrimer Backbone The earliest application of dendrimers in optical imaging was reported by McIntyre and co-workers on the development of fluorogenic peptidic dendrimer for the detection of matrix metalloproteinase-7 (MMP-7) activity in colon cancer [53]. Much like the selfquenched probe discussed earlier, this probe uses a fourth generation polyamidoamine (PAMAM) dendrimer as a backbone with fluorescein (ex = 493 nm; em = 517 nm) and tetramethylrhodamine (ex = 554 nm; em = 572 nm) fluorophores to promote energy transfer/self-quenching mechanisms. Although the use of this particular dye pair has limited tissue penetration for imaging, the synthetic versatility of dendrimer over a wide range of molecular weight for optimization of clearance, quenching, or proteolytic cleavage makes it an excellent design strategy for imaging. Besides probing for proteolytic activities, dendrimers have also been used as a carrier to characterize the tumor microenvironment. Almutairi and co-workers reported a pH-sensing NIR nanoprobe capable of reporting fluorescence lifetime under acidic conditions in vivo. The probe was synthesized by reacting the hydrazine on the NIR dye (cypate) with the
P1: OTA/XYZ P2: ABC c08 JWBS037-Chen February 1, 2011
188
4:53
Printer Name: Yet to Come
BIOPOLYMER, DENDRIMER, AND LIPOSOME NANOPLATFORMS FOR OPTICAL
ketone on the pegylated dendrimer to form acid-sensitive hydrazone linkers. At neutral pH, the NIR dye is “silent” where the fluorescence intensity is insensitive to the local environment. Under an acidic environment, hydrolysis of the hydrazone linkers releases the NIR dyes from the probe, thus increasing the intensity of the fluorescence signal [54]. 8.3.2 Multimodal Molecular Probes Radioscintigraphy and magnetic resonance (MR) are two imaging modalities with highest detection sensitivity and spatial resolutions respectively [55], but dendrimers provide a good platform for combining them into multimodal imaging probes. Dendrimerbased dual-modal probes combining optical imaging reporters with a MR contrast agent [56, 57] and a radionuclide [58] have been developed for in vivo lymphography. The MR/NIR optical probe was built on a sixth generation PAMAM (G6-PAMAM) dendrimer functionalized with a cyanine NIR dye (Cy5.5) and saturated with gadolinium 2-(4-isothiocyanatobenzyl)-6-methyl-diethylenetriaminepentaacetic acid (Bz-DTPA). The resulting dual-modal nanoprobe easily identified all sentinel lymph nodes, and the nodes can be resected under NIR-guided surgery. Similarly, the radionuclide/NIR probes were synthesized based on the same G6-PAMAM dendrimer. The dual-modal radionuclide/NIR probe was built with a modified Bz-DTPA chelater. With radioscintigraphy being much more sensitive than MR, only a single Bz-DTPA is chelated to 111 In and the rest of the ligands were left unlabeled. In addition to the 111 In/Cy5.5 probe, four other 111 In/NIR optical probes with different emission wavelengths were prepared. in vivo visualization was carried out on a mouse model, where the five probes were injected intracutaneously at five different sites and were visualized with a multiexcitation spectral fluorescence imaging station. Impressively, the report showed five lymphatic drainages simultaneously in different fluorescent hues that correlated with both in vivo and ex vivo radionuclide scintigraphic images. 8.3.3 Targeted Nanoprobes and Therapeutics The ease of conjugating different imaging reporters on the surface of dendrimers also pave the way to combine imaging with therapeutic properties in a single platform. Backer and co-workers synthesized a fifth generation PAMAM decorated with ∼110 decaborate (G5DB) moieties to be used as a boron neutron capture therapy agent for cancer treatment [59]. To this G5-DB nanotherapeutic construct was conjugated an NIR reporter (G5-DB-Cy5). Multiple G5-BD-Cy5 constructs were then linked to a vascular endothelial growth factor (VEGF). in vivo imaging of 4T1 mice bearing breast carcinoma showed selective uptake of VEGF-(G5-BD-Cy5)4 at the periphery of the tumor, whereas the nontargeted G5-BD-Cy5 construct show no accumulation in the tumor. Baker and co-workers [60, 61] and Brechbiel and co-workers [62] demonstrated that nanoprobes with targeting moieties, such as folic acid (FA), monoclonal antibodies for human epidermal growth factor receptor-2 (anti-HER2 mAb), and cyclic-(arginine-glycineaspartic acid) (RGD), can be incorporated with therapeutic drug molecules. Among these nanoprobes is the first dendritic therapeutic agent that not only demonstrated in vivo targeting of cancerous tissues, but also has increased therapeutic efficacy compared to the anticancer drug methotrexate (MTX) alone [60]. The Baker group’s dendritic prodrug was built on G5-PAMAM functionalized with five folic acid targeting moieties, five fluorescein isothiocyanate (FITC) reporters, and five MTX groups. In mice bearing human epithelial
P1: OTA/XYZ P2: ABC c08 JWBS037-Chen February 1, 2011
4:53
Printer Name: Yet to Come
LIPOSOMES
189
cancer cells, there is a 50% survival rate in mice treated with the dendritic prodrug after 67 days as opposed to no survival in the population treated with MTX. Furthermore, at the termination of the trial, the surviving mice showed no sign of toxicity in the heart or development of myopathy. The mice treated with the prodrug also show significant necrosis in the tumor.
8.4 LIPOSOMES Liposomes are spherical vesicles formed by an assembly of phospholipids containing hydrophilic head groups and hydrophobic aliphatic chains. They are usually prepared from biocompatible lipids that can be metabolized and cleared. In the past two decades, various forms of these lipid vesicles have been used to encapsulate drug molecules to improve in vivo delivery to disease targets [63]. Since the approval by the U.S. Food and Drug Administration (FDA), liposomes have become one of the most widely used drug delivery systems. However, an effective drug delivery using liposomes is not without challenges. The liposome’s size and surface charge are important elements that influence in vivo phagocytic events and biodistribution. Particles of appropriate size are phagocytized by cells in the reticuloendothelial system (RES): for example, Kupffer cells in the liver and macrophages in the spleen [64]. Additionally, the electrostatic charge on the liposome surface, whether it is neutral, positive, or negative, may also alter the overall pharmacokinetics, clearance, and uptake by disease targets [63]. Surface modification with hydrophilic polymers, such as PEG, is by far the most common strategy to improve the in vivo properties of liposomes. The introduction of PEG alters the surface hydration and steric factors, creating a “stealth” environment to bypass phagocytosis by the RES, and such strategy has been adapted for the development of a number of liposome-based imaging agents [65–68]. 8.4.1 Incorporation of Molecules in Liposomes Owing to the amphipathic nature of liposomes, a wide range of molecules may be incorporated into liposomes based on their interactions with the hydrophilic core, the hydrophobic shell, or the charge on the liposome surface. Additionally, desired ligands may be tethered onto the liposome surface through direct conjugation chemistry (Fig. 2). All these modes of molecular engineering have been explored for optimization of drug delivery [69], development of new imaging agents [70–72], and enhancement of imaging efficacy [73–79]. 8.4.2 Optical Imaging with Liposome-Based Nanoprobes Conjugation of organic dyes onto the liposomes has yielded liposome-based optical probes for both in vitro and in vivo imaging. Rhodamine [80, 81], fluorescein [81, 82], and Cy5.5 [83, 84] are among the most common fluorescent dye molecules being used. Besides organic dyes, CdSe/CdTe QD have also been incorporated for optical imaging [71]. Owing to their hydrophobic surface and potential in vivo toxicity [85, 86], the clinical application of QDs is often faced with skepticism. The encapsulation of QDs in liposomes therefore presents an attractive approach to improve QDs’ biocompatibility and toxicity [87].
P1: OTA/XYZ P2: ABC c08 JWBS037-Chen February 1, 2011
190
4:53
Printer Name: Yet to Come
BIOPOLYMER, DENDRIMER, AND LIPOSOME NANOPLATFORMS FOR OPTICAL
(b)
(a)
Hy m dro ole ph cu ilic les
c bi ho s op le dr lecu y H o m
ic at st n ro io ct ract e El nte I
co Dir nju ec ga t tio n
(c)
(d)
FIGURE 8.2 Incorporation of a variety of molecules in liposomes via (a) encapsulation into the hydrophilic core vesicle, (b) embedding hydrophobic molecules in the aliphatic shell, (c) electrostatic interactions, and (d) direct conjugation to the phospholipid.
Recently, it was also reported that liposomes consisting of hepatitis B surface antigen and fluorescently labeled polystyrene beads or plasmids were used for tissue and cellular targeting with high specificity [88]. Another method of liposome ligation was reported by Reulen et al. [89], which showed fluorescent proteins covalently coupled to cysteinefunctionalized phospholipids on the liposome surface for optical imaging of collagen.
8.4.3 Multimodal Functionality of the Liposome Platform The ease of incorporating various functional molecules in the core vesicle, the aliphatic shell, or the lipid surface makes the liposome an ideal platform for multimodal targeted imaging and therapy. As a delivery vehicle, the liposome is known for its ability to passively accumulate at tumor regions via the EPR effect. By conjugating active targeting ligands on the lipid surface, researchers have further improved liposomes’ delivery efficacy. Park and co-workers have shown that, by incorporating active targeting ligand such as antiHER2 (human epidermal growth factor receptor 2) or anti-EGFR (epidermal growth factor receptor) fragments, liposomes can not only accumulate at tumor sites but these ligands also facilitate active internalization by tumor cells [90, 91]. This combined passive and active targeting strategy has been further adapted with other disease binding ligands such as the RGD peptide for ␣v3-integrin [80], CREKA peptide for clotted plasma proteins [82], and GE11 peptide for EGFR [83]. The multimodal functionality is not only limited to targeting; multimodality imaging with optical and MR is often sought after using the liposome platform. For example, Mulder and co-workers reported the use of paramagnetic lipid to coat quantum dots for MR and optical dual imaging [66, 67, 80]. Alternatively, Kamaly et al. [68] synthesized a fluorescently labeled gadolinium lipid for cell labeling and tumor imaging. Another variation includes a dual-modal probe with fluorescence and magnetic properties constructed by
P1: OTA/XYZ P2: ABC c08 JWBS037-Chen February 1, 2011
4:53
Printer Name: Yet to Come
REFERENCES
191
linking the near-infrared fluorescent transferrin conjugate (Tf NIR ) on the surface of gadolinium encapsulated cationic liposome to image breast cancer [92].
8.5 PERSPECTIVE As shown in the examples presented here, biopolymers, dendrimers, and liposomes have proved to be versatile platforms for targeted delivery, controlled tissue distribution, and clearance. These nanoplatforms will continue to play a vital role in new agent development. With their characteristic multivalency and multifunctional attributes, highly sensitive and specific nanoagents will soon be applied in the clinic and the ability to simultaneously obtain in vivo disease information from different imaging modalities will soon be realized. Finally, the “stealth” environment provided by these platforms also opens new avenues to evaluate new materials that otherwise would be biologically incompatible (e.g., iron oxides and quantum dots). These platforms will allow a safe entry of many new agents for clinical evaluations in the future.
REFERENCES 1. Alexis, F.; Rhee, J.-W.; Richie, J. P.; Radovic-Moreno, A. F.; Langer, R.; Farokhzad, O. C. New frontiers in nanotechnology for cancer treatment. Urol. Oncol.-Semin. Ori. 2008, 26(1), 74–85. 2. Mankoff, D. A.; Link, J. M.; Linden, H. M.; Sundararajan, L.; Krohn, K. A. Tumor receptor imaging. J. Nucl. Med. 2008, 49(Suppl. 2), 149S–163. 3. Chanda, S. K.; Caldwell, J. S. Fulfilling the promise: drug discovery in the post-genomic era. Drug Discov. Today 2003, 8(4), 168–174. 4. Mammen, M.; Choi, S.-K.; Whitesides, G. M. Polyvalent Interactions in biological systems: implications for design and use of multivalent ligands and inhibitors. Angew. Chem., Int. Ed. Engl. 1998, 37(20), 2754–2794. 5. Lundquist, J. J.; Toone, E. J. The cluster glycoside effect. Chem. Rev. 2002, 102(2), 555–578. 6. Kitov, P. I.; Mulvey, G. L.; Griener, T. P.; Lipinski, T.; Solomon, D.; Paszkiewicz, E.; Jacobson, J. M.; Sadowska, J. M.; Suzuki, M.; Yamamura, K.-i.; Armstrong, G. D.; Bundle, D. R. In vivo supramolecular templating enhances the activity of multivalent ligands: a potential therapeutic against the Escherichia coli O157 AB5 toxins. Proc. Natl. Acad. Sci. U.S.A. 2008, 105(44), 16837–16842. 7. Lee, C. C.; Gillies, E. R.; Fox, M. E.; Guillaudeu, S. J.; Frechet, J. M. J.; Dy, E. E.; Szoka, F. C. A single dose of doxorubicin-functionalized bow-tie dendrimer cures mice bearing C-26 colon carcinomas. Proc. Natl. Acad. Sci. U.S.A. 2006, 103(45), 16649–16654. 8. Gabizon, A.; Shmeeda, H.; Barenholz, Y. Pharmacokinetics of pegylated liposomal doxorubicin: review of animal and human studies. Clin. Pharmacokinet. 2003, 42(5), 419–436. 9. Duncan, R., The dawning era of polymer therapeutics. Nat. Rev. Drug Discov. 2003, 2(5), 347–360. 10. Maeda, H. The enhanced permeability and retention (EPR) effect in tumor vasculature: the key role of tumor-selective macromolecular drug targeting. Adv. Enzyme Regul. 2001, 41(1), 189–207. 11. Khosravi-Darani, K.; Pardakhty, A.; Honarpisheh, H.; Rao, V. S. N. M.; Mozafari, M. R. The role of high-resolution imaging in the evaluation of nanosystems for bioactive encapsulation and targeted nanotherapy. Micron 2007, 38(8), 804–818.
P1: OTA/XYZ P2: ABC c08 JWBS037-Chen February 1, 2011
192
4:53
Printer Name: Yet to Come
BIOPOLYMER, DENDRIMER, AND LIPOSOME NANOPLATFORMS FOR OPTICAL
12. Haley, B.; Frenkel, E. Nanoparticles for drug delivery in cancer treatment. Urol. Oncol.-Semin. Ori. 2008, 26(1), 57–64. 13. Shubik, P. Vascularization of tumors: a review. J. Cancer Res. Clin. Oncol. 1982, 103(3), 211–226. 14. Hobbs, S. K.; Monsky, W. L.; Yuan, F.; Roberts, W. G.; Griffith, L.; Torchilin, V. P.; Jain, R. K. Regulation of transport pathways in tumor vessels: Role of tumor type and microenvironment. Proc. Natl. Acad. Sci. U.S.A. 1998, 95(8), 4607–4612. 15. Kim, K. Y. Nanotechnology platforms and physiological challenges for cancer therapeutics. Nanomed. Nanotechnol. Biol. Med. 2007, 3(2), 103–110. 16. Padera, T. P.; Stoll, B. R.; Tooredman, J. B.; Capen, D.; Tomaso, E. d.; Jain, R. K. Pathology: cancer cells compress intratumour vessels. Nature 2004, 427(6976), 695–695. 17. Caruthers, S. D.; Wickline, S. A.; Lanza, G. M. Nanotechnological applications in medicine. Curr. Opin. Biotechnol. 2007, 18(1), 26–30. 18. Kim, J.-H.; Park, K.; Nam, H. Y.; Lee, S.; Kim, K.; Kwon, I. C. Polymers for bioimaging. Prog. Polym. Sci. 2007, 32(8-9), 1031–1053. 19. Jatzkewitz, H. Peptamin (glycyl-l-leucyl-mescaline) bound to blood plasma expander (polyvinylpyrrolidone) as a new depot form of a biologically active primary amine(mescaline). Z. Naturforsch. 1955, 10, 27–31. 20. Li, C.; Yu, D.-F.; Newman, R. A.; Cabral, F.; Stephens, L. C.; Hunter, N.; Milas, L.; Wallace, S. Complete regression of well-established tumors using a novel water-soluble poly(l-glutamic acid)–paclitaxel conjugate. Cancer Res. 1998, 58(11), 2404–2409. 21. Boddy, A. V.; Plummer, E. R.; Todd, R.; Sludden, J.; Griffin, M.; Robson, L.; Cassidy, J.; Bissett, D.; Bernareggi, A.; Verrill, M. W.; Calvert, A. H. A Phase I and pharmacokinetic study of paclitaxel poliglumex (XYOTAX), investigating both 3-weekly and 2-weekly schedules. Clin. Cancer Res. 2005, 11(21), 7834–7840. 22. Gregoriadis, G.; McCormack, B.; Wang, Z.; Lifely, R. Polysialic acids: potential in drug delivery. FEBS Lett. 1993, 315(3), 271–276. 23. Sigal, G. B.; Mammen, M.; Dahmann, G.; Whitesides, G. M. Polyacrylamides bearing pendant ␣-sialoside groups strongly inhibit agglutination of erythrocytes by influenza virus: the strong inhibition reflects enhanced binding through cooperative polyvalent interactions. J. Am. Chem. Soc. 1996, 118(16), 3789–3800. 24. Langer, R.; Tirrell, D. A. Designing materials for biology and medicine. Nature 2004, 428(6982), 487–492. 25. Lutolf, M. P.; Hubbell, J. A. Synthetic biomaterials as instructive extracellular microenvironments for morphogenesis in tissue engineering. Nat Biotechnol. 2005, 23(1), 47–55. 26. Huang, P. S.; Oliff, A. Drug-targeting strategies in cancer therapy. Curr. Opin. Genet. Dev. 2001, 11(1), 104–110. 27. Vasey, P. A.; Kaye, S. B.; Morrison, R.; Twelves, C.; Wilson, P.; Duncan, R.; Thomson, A. H.; Murray, L. S.; Hilditch, T. E.; Murray, T.; Burtles, S.; Fraier, D.; Frigerio, E.; Cassidy, J. Phase I Clinical and pharmacokinetic study of PK1 [N-(2-hydroxypropyl)methacrylamide copolymer doxorubicin]: first member of a new class of chemotherapeutic agents—drug–polymer conjugates. Clin. Cancer Res. 1999, 5(1), 83–94. 28. Seymour, L. W.; Ferry, D. R.; Anderson, D.; Hesslewood, S.; Julyan, P. J.; Poyner, R.; Doran, J.; Young, A. M.; Burtles, S.; Kerr, D. J. Hepatic drug targeting: Phase I evaluation of polymer-bound doxorubicin. J. Clin. Oncol. 2002, 20(6), 1668–1676. 29. Gianasi, E.; Buckley, R. G.; Latigo, J.; Wasil, M.; Duncan, R. HPMA copolymer platinates containing dicarboxylato ligands. Preparation, characterisation and in vitro and in vivo evaluation. J. Drug Target. 2002, 10(7), 549–556. 30. Johnson, K. P.; Brooks, B. R.; Cohen, J. A.; Ford, C. C.; Goldstein, J.; Lisak, R. P.; Myers, L. W.; Panitch, H. S.; Rose, J. W.; Schiffer, R. B. Copolymer 1 reduces relapse rate and improves
P1: OTA/XYZ P2: ABC c08 JWBS037-Chen February 1, 2011
4:53
Printer Name: Yet to Come
REFERENCES
31. 32.
33. 34. 35.
36.
37. 38. 39.
40. 41. 42.
43.
44.
45. 46. 47. 48. 49.
193
disability in relapsing–remitting multiple sclerosis: results of a Phase III multicenter, doubleblind placebo-controlled trial. The Copolymer 1 Multiple Sclerosis Study Group. Neurology 1995, 45(7), 1268–1276. Mandeville, W. H.; Goldberg, D. I. The sequestration of bile acids, a nonabsorbed method for cholesterol reduction. A review. Curr. Pharm. Des. 1997, 3(1), 15–28. Slatopolsky, E. A.; Burke, S. K.; Dillon, M. A. RenaGel [reg], a nonabsorbed calcium- and aluminum-free phosphate binder, lowers serum phosphorus and parathyroid hormone. Kidney Int. 1999, 55(1), 299–307. Donaruma, L. G. Synthetic biologically active polymers. Prog. Polym. Sci. 1975, 4, 1–25. Mistry, C. D.; Gokal, R. The use of glucose polymer (icodextrin) in peritoneal dialysis: an overview. Perit. Dial. Int. 1994, 14(Suppl. 3), S158–S161. Shaunak, S.; Thornton, M.; John, S.; Teo, I.; Peers, E.; Mason, P.; Krausz, T.; Davies, D. S. Reduction of the viral load of HIV-1 after the intraperitoneal administration of dextrin 2-sulphate in patients with AIDS. AIDS 1998, 12(4), 399–409. Stafford, M. K.; Cain, D.; Rosenstein, I.; Fontaine, E. A.; McClure, M.; Flanagan, A. M.; Smith, J. R.; Taylor-Robinson, D.; Weber, J.; Kitchen, V. S. A placebo-controlled, doubleblind prospective study in healthy female volunteers of dextrin sulphate gel: a novel potential intravaginal virucide. J. Acquir. Immune Defic. Syndr. Hum. Retrovirol. 1997, 14(3), 213–218. Gillies, E. R.; Fr´echet, J. M. J. Dendrimers and dendritic polymers in drug delivery. Drug Discov. Today 2005, 10(1), 35–43. Funovics, M.; Weissleder, R.; Tung, C.-H. Protease sensors for bioimaging. Anal. Bioanal. Chem. 2003, 377(6), 956–963. Sidhu, S. S.; Lowman, H. B.; Cunningham, B. C.; Wells, J. A.; Jeremy Thorner, S. D. E. a. J. N. A. Phage display for selection of novel binding peptides. In Methods in Enzymology. Academic Press: San Diego, 2000; Vol. 328, pp. 333–358. Gao, X.; Cui, Y.; Levenson, R. M.; Chung, L. W. K.; Nie, S. In vivo cancer targeting and imaging with semiconductor quantum dots. Nat. Biotechnol. 2004, 22(8), 969–976. Cai, W.; Chen, X. Preparation of peptide-conjugated quantum dots for tumor vasculature-targeted imaging. Nat. Protocols 2008, 3(1), 89–96. Cai, W.; Shin, D.-W.; Chen, K.; Gheysens, O.; Cao, Q.; Wang, S. X.; Gambhir, S. S.; Chen, X. Peptide-labeled near-infrared quantum dots for imaging tumor vasculature in living subjects. Nano Lett. 2006, 6(4), 669–676. Zhu, L.; Wu, W.; Zhu, M.-Q.; Han, J. J.; Hurst, J. K.; Li, A. D. Q. Reversibly photoswitchable dual-color fluorescent nanoparticles as new tools for live-cell imaging. J. Am. Chem. Soc. 2007, 129(12), 3524–3526. Park, K.; Kim, J.-H.; Nam, Y. S.; Lee, S.; Nam, H. Y.; Kim, K.; Park, J. H.; Kim, I.-S.; Choi, K.; Kim, S. Y.; Kwon, I. C. Effect of polymer molecular weight on the tumor targeting characteristics of self-assembled glycol chitosan nanoparticles. J. Control. Release 2007, 122(3), 305–314. Kwon, I. C. Chitosan-based nanoparticles for cancer therapy; tumor specificity and enhanced therapeutic efficacy in tumor-bearing mice. J. Control. Release 2008, 132(3), e69–e70. Barrett, T.; Kobayashi, H.; Brechbiel, M.; Choyke, P. L. Macromolecular MRI contrast agents for imaging tumor angiogenesis. Eur. J. Radiol. 2006, 60(3), 353–366. Lee, C. C.; MacKay, J. A.; Frechet, J. M. J.; Szoka, F. C. Designing dendrimers for biological applications. Nat. Biotechnol. 2005, 23(12), 1517–1526. Farin, D.; Avnir, D. Surface fractality of dendrimers. Angew. Chem. Int. Ed. Engl. 1991, 30(10), 1379–1380. Gillies, E. R.; Frechet, J. M. J. Designing macromolecules for therapeutic applications: polyester dendrimer poly(ethylene oxide) “bow-tie” hybrids with tunable molecular weight and architecture. J. Am. Chem. Soc. 2002, 124(47), 14137–14146.
P1: OTA/XYZ P2: ABC c08 JWBS037-Chen February 1, 2011
194
4:53
Printer Name: Yet to Come
BIOPOLYMER, DENDRIMER, AND LIPOSOME NANOPLATFORMS FOR OPTICAL
50. Lee, C. C.; Yoshida, M.; Frechet, J. M. J.; Dy, E. E.; Szoka, F. C. In vitro and in vivo evaluation of hydrophilic dendronized linear polymers. Bioconjug. Chem. 2005, 16(3), 535–541. 51. Kobayashi, H.; Kawamoto, S.; Jo, S.-K.; Bryant, H. L.; Brechbiel, M. W.; Star, R. A. Macromolecular MRI contrast agents with small dendrimers: pharmacokinetic differences between sizes and cores. Bioconjug. Chem. 2003, 14(2), 388–394. 52. Gillies, E. R.; Dy, E.; Frechet, J. M. J.; Szoka, F. C. Biological evaluation of polyester dendrimer: poly(ethylene oxide) “bow-tie” hybrids with tunable molecular weight and architecture. Mol. Pharm. 2005, 2(2), 129–138. 53. McIntyre, J. O.; Fingleton, B.; Wells, K. S.; Piston, D. W.; Lynch, C. C.; Gautam, S.; Matrisian, L. M. Development of a novel fluorogenic proteolytic beacon for in vivo detection and imaging of tumour-associated matrix metalloproteinase-7 activity. Biochem. J. 2004, 377(3), 617–628. 54. Almutairi, A.; Guillaudeu, S. J.; Berezin, M. Y.; Achilefu, S.; Frechet, J. M. J. Biodegradable pH-sensing dendritic nanoprobes for near-infrared fluorescence lifetime and intensity imaging. J. Am. Chem. Soc. 2008, 130(2), 444–445. 55. Massoud, T. F.; Gambhir, S. S. Molecular imaging in living subjects: seeing fundamental biological processes in a new light. Genes Dev. 2003, 17(5), 545–580. 56. Talanov, V. S.; Regino, C. A. S.; Kobayashi, H.; Bernardo, M.; Choyke, P. L.; Brechbiel, M. W. Dendrimer-based nanoprobe for dual modality magnetic resonance and fluorescence imaging. Nano Lett. 2006, 6(7), 1459–1463. 57. Koyama, Y.; Talanov, V. S.; Bernardo, M.; Hama, Y.; Regino, C. A. S.; Brechbiel, M. W.; Choyke, P. L.; Kobayashi, H. A dendrimer-based nanosized contrast agent dual-labeled for magnetic resonance and optical fluorescence imaging to localize the sentinel lymph node in mice. J. Magn. Reson. Imaging 2007, 25(4), 866–871. 58. Kobayashi, H.; Koyama, Y.; Barrett, T.; Hama, Y.; Regino, C. A. S.; Shin, I. S.; Jang, B.; Le, N.; Paik, C. H.; Choyke, P. L.; Urano, Y. A multi-modal and multi-color dendrimer-based nano-sized contrast agent for near-infrared optical and radionuclide lymphatic imaging. Contrast Media Mol. Imaging 2007, 2(6), 279–280. 59. Backer, M. V.; Gaynutdinov, T. I.; Patel, V.; Bandyopadhyaya, A. K.; Thirumamagal, B. T. S.; Tjarks, W.; Barth, R. F.; Claffey, K.; Backer, J. M. Vascular endothelial growth factor selectively targets boronated dendrimers to tumor vasculature. Mol. Cancer Ther. 2005, 4(9), 1423–1429. 60. Kukowska-Latallo, J. F.; Candido, K. A.; Cao, Z.; Nigavekar, S. S.; Majoros, I. J.; Thomas, T. P.; Balogh, L. P.; Khan, M. K.; Baker, J. R., Jr. Nanoparticle targeting of anticancer drug improves therapeutic response in animal model of human epithelial cancer. Cancer Res. 2005, 65(12), 5317–5324. 61. Shukla, R.; Thomas, T. P.; Peters, J. L.; Desai, A. M.; Kukowska-Latallo, J.; Patri, A. K.; Kotlyar, A.; Baker, J. R. HER2 specific tumor targeting with dendrimer conjugated anti-HER2 mAb. Bioconjug. Chem. 2006, 17(5), 1109–1115. 62. Boswell, C. A.; Eck, P. K.; Regino, C. A. S.; Bernardo, M.; Wong, K. J.; Milenic, D. E.; Choyke, P. L.; Brechbiel, M. W. Synthesis, characterization, and biological evaluation of integrin ␣v 3 -targeted PAMAM dendrimers. Mol. Pharm. 2008, 5(4), 527–539. 63. Lian, T.; Ho, R. J. Y. Trends and developments in liposome drug delivery systems. J. Pharm. Sci. 2001, 90(6), 667–680. 64. Seltzer, S. E. The role of liposomes in diagnostic imaging. Radiology 1989, 171(1), 19–21. 65. Torchilin, V. P. Recent advances with liposomes as pharmaceutical carriers. Nat Rev Drug Discov 2005, 4(2), 145–160. 66. Mulder, W. J. M.; Strijkers, G. J.; Tilborg, G. A. F. v.; Griffioen, A. W.; Nicolay, K. Lipid-based nanoparticles for contrast-enhanced MRI and molecular imaging. NMR Biomed. 2006, 19(1), 142–164.
P1: OTA/XYZ P2: ABC c08 JWBS037-Chen February 1, 2011
4:53
Printer Name: Yet to Come
REFERENCES
195
67. Mulder, W. J. M.; Griffioen, A. W.; Strijkers, G. J.; Cormode, D. P.; Nicolay, K.; Fayad, Z. A. Magnetic and fluorescent nanoparticles for multimodality imaging. Nanomedicine 2007, 2(3), 307. 68. Kamaly, N.; Kalber, T.; Ahmad, A.; Oliver, M. H.; So, P.-W.; Herlihy, A. H.; Bell, J. D.; Jorgensen, M. R.; Miller, A. D. Bimodal paramagnetic and fluorescent liposomes for cellular and tumor magnetic resonance imaging. Bioconjug. Chem. 2008, 19(1), 118–129. 69. Drummond, D. C.; Meyer, O.; Hong, K.; Kirpotin, D. B.; Papahadjopoulos, D. Optimizing liposomes for delivery of chemotherapeutic agents to solid tumors. Pharmacol. Rev. 1999, 51(4), 691–744. 70. Keana, J. F. W.; Pou, S. Nitroxide-doped liposomes containing entrapped oxidant: an approach to the “reduction problem” of nitroxides as MRI contrast agents. Physiol. Chem. Phys. Med. NMR 1985, 17(22), 235–240. 71. Al-Jamal, W. T.; Kostarelos, K. Liposome nanoparticle hybrids for multimodal diagnostic and therapeutic applications. Nanomedicine 2007, 2, 85–98. 72. Navon, G.; Panigel, R.; Valensin, G. Liposomes containing paramagnetic macromolecules as MRI contrast agents. Magn. Reson. Med. 1986, 3(6), 876–880. 73. Najafi, A.; Amparo, E. C.; Johnson, R. F. J. Gadolinium labeled pharmaceuticals as potential MRI contrast agents for liver and biliary tract. J. Label. Compd. Radiopharm. 1987, 24(9), 1131–1141. 74. Devoisselle, J. M.; Vion-Dury, J.; Galons, J. P.; Confort-Gouny, S.; Coustaut, D.; Canioni, P.; Cozzone, P. J. Entrapment of gadolinium-DTPA in liposomes. Characterization of vesicles by P-31 NMR spectroscopy. Invest. Radiol. 1988, 23(10), 719–724. 75. Kabalka, G. W.; Buonocore, E.; Hubner, K.; Davis, M.; Huang, L. Gadolinium-labeled liposomes containing paramagnetic amphipathic agents: targeted MRI contrast agents for the liver. Magn. Reson. Med. 1988, 8(1), 89–95. 76. Unger, E. M.; Needleman, P.; Cullis, P. P.; Tilcock, C. P. Gadolinium-DTPA liposomes as a potential mri contrast agent: work in progress. Investig. Radiol. 1988, 23(12), 928–932. 77. Schwendener, R. A. Liposomes as carriers for paramagnetic gadolinium chelates as organ specific contrast agents for magnetic resonance imaging (MRI). J. Liposome Res. 1994, 4(2), 837–855. 78. Unger, E. C.; Winokur, T.; MacDougall, P.; Rosenblum, J.; Clair, M.; Gatenby, R.; Tilcock, C. Hepatic metastases: liposomal Gd-DTPA-enhanced MR imaging. Radiology 1989, 171(1), 81–85. 79. Unger, E. C.; MacDougall, P.; Cullis, P.; Tilcock, C. Liposomal Gd-DTPA: Effect of encapsulation on enhancement of hepatoma model by MRI. Magn. Reson. Imaging 1989, 7(4), 417–423. 80. Mulder, W. J. M.; Strijkers, G. J.; Habets, J. W.; Bleeker, E. J. W.; van der Schaft, D. W. J.; Storm, G.; Koning, G. A.; Griffioen, A. W.; Nicolay, K. MR molecular imaging and fluorescence microscopy for identification of activated tumor endothelium using a bimodal lipidic nanoparticle. FASEB J. 2005, 19(14), 2008–2010. 81. Kunisawa, J.; Masuda, T.; Katayama, K.; Yoshikawa, T.; Tsutsumi, Y.; Akashi, M.; Mayumi, T.; Nakagawa, S. Fusogenic liposome delivers encapsulated nanoparticles for cytosolic controlled gene release. J. Control. Release 2005, 105(3), 344–353. 82. Simberg, D.; Duza, T.; Park, J. H.; Essler, M.; Pilch, J.; Zhang, L.; Derfus, A. M.; Yang, M.; Hoffman, R. M.; Bhatia, S.; Sailor, M. J.; Ruoslahti, E. Biomimetic amplification of nanoparticle homing to tumors. Proc. Natl. Acad. Sci. U.S.A. 2007, 104(3), 932–936. 83. Song, S.; Liu, D.; Peng, J.; Sun, Y.; Li, Z.; Gu, J.-R.; Xu, Y. Peptide ligand-mediated liposome distribution and targeting to EGFR expressing tumor in vivo. Int. J. Pharm. 2008, 363(1-2), 155–161. 84. Saad, M.; Garbuzenko, O. B.; Ber, E.; Chandna, P.; Khandare, J. J.; Pozharov, V. P.; Minko, T. Receptor targeted polymers, dendrimers, liposomes: Which nanocarrier is the most efficient for tumor-specific treatment and imaging? J. Control. Release 2008, 130(2), 107–114.
P1: OTA/XYZ P2: ABC c08 JWBS037-Chen February 1, 2011
196
4:53
Printer Name: Yet to Come
BIOPOLYMER, DENDRIMER, AND LIPOSOME NANOPLATFORMS FOR OPTICAL
85. Dubertret, B.; Skourides, P.; Norris, D. J.; Noireaux, V.; Brivanlou, A. H.; Libchaber, A. In vivo imaging of quantum dots encapsulated in phospholipid micelles. Science 2002, 298(5599), 1759–1762. 86. Chen, C.-S.; Yao, J.; Durst, R. Liposome encapsulation of fluorescent nanoparticles: quantum dots and silica nanoparticles. J. Nanopart. Res. 2006, 8(6), 1033–1038. 87. Dudu, V.; Ramcharan, M.; Gilchrist, M. L.; Holland, E. C.; Vazquez, M. Liposome delivery of quantum dots to the cytosol of live cells. J. Nanosci. Nanotechnol. 2008, 8, 2293–2300. 88. Jung, J.; Matsuzaki, T.; Tatematsu, K.; Okajima, T.; Tanizawa, K.; Kuroda, S. I. Bio-nanocapsule conjugated with liposomes for in vivo pinpoint delivery of various materials. J. Control. Release 2008, 126(3), 255–264. 89. Reulen, S. W. A.; Brusselaars, W. W. T.; Langereis, S.; Mulder, W. J. M.; Breurken, M.; Merkx, M. Protein–liposome conjugates using cysteine-lipids and native chemical ligation. Bioconjug. Chem. 2007, 18(2), 590–596. 90. Kirpotin, D. B.; Park, J. W.; Hong, K.; Shao, Y.; Shalaby, R.; Colbern, G.; Benz, C. C.; Papahadjopoulos, D. Targeting of liposomes to solid tumors: the case of sterically stabilized anti-Her2 immunoliposomes. J. Liposome Res. 1997, 7(4), 391–417. 91. Mamot, C.; Drummond, D. C.; Noble, C. O.; Kallab, V.; Guo, Z.; Hong, K.; Kirpotin, D. B.; Park, J. W. Epidermal growth factor receptor-targeted immunoliposomes significantly enhance the efficacy of multiple anticancer drugs in vivo. Cancer Res. 2005, 65(24), 11631–11638. 92. Shan, L.; Wang, S. S., R.; Bhujwalla, Z. M.; Wang, P. C. Dual probe with fluorescent and magnetic properties for imaging solid tumor xenografts. Mol. Imaging 2007, 6(2), 85–95.
P1: OTA/XYZ P2: ABC c09 JWBS037-Chen August 29, 2010
13:16
Printer Name: Yet to Come
CHAPTER 9
Nanoplatforms for Raman Molecular Imaging in Biological Systems ZHUANG LIU Institute of Functional Nano & Soft Materials, Soochow University, Suzhou, Jiangsu, China
9.1 INTRODUCTION Traditional optical imaging of biological samples mostly relies on bioluminance and fluorescence imaging. The latter, in particular, has been widely utilized to study biological systems in vitro and in vivo. Despite the powerfulness of this technique, a few factors still limit its applications [1]. Rapid photobleaching is a problem for most organic fluorescent dyes, preventing long-term imaging and monitoring of fluorescently labeled substances [2]. The autofluorescence background of biological tissues also restricts the sensitivity of fluorescence imaging [3]. Moreover, fluorescent molecules normally have wide emission peaks with full-width at half-maximum (FWHM) as large as 100 nm, which leads to spectral overlay between different dyes and limits the imaging multiplicity. This is especially problematic for in vivo imaging, in which the ideal excitation and emission wavelengths are in a narrow tissue transparency window in the near-infrared (NIR) region (700–900 nm) [4, 5]. Therefore new imaging techniques as well as novel contrast agents are desired to circumvent those problems. As a photon scattering phenomenon, Raman spectroscopy has distinctive features from fluorescence emission. The Raman scattering of molecules exhibits narrow spectral fingerprints with FWHM within a few nanometers and thus allows a dramatically increased imaging multiplicity compared with fluorescent imaging [6]. Tissue autofluorescence is no longer a problem because the sharp peaks in Raman spectra can easily be distinguished from the fluorescence background, imparting high imaging and detection sensitivity. In addition, Raman scattering signals are generally more robust than fluorescence with much less photobleaching. Therefore Raman imaging is a promising technique for the next generation of biomedical imaging [7–9]. Despite the above advantages in Raman imaging, the photon efficiency in Raman scattering is very low. Traditional Raman spectroscopy measurement of small molecules has very low detection sensitivity due to poor Raman scattering signals. In the past decade, the development of nanotechnology has provided a number of breakthroughs to Raman Nanoplatform-Based Molecular Imaging Edited by Xiaoyuan Chen C 2011 John Wiley & Sons, Inc. Copyright
197
P1: OTA/XYZ P2: ABC c09 JWBS037-Chen August 29, 2010
198
13:16
Printer Name: Yet to Come
NANOPLATFORMS FOR RAMAN MOLECULAR IMAGING IN BIOLOGICAL SYSTEMS
imaging by enhancing Raman scattering intensities through various mechanisms including resonance Raman scattering and SERS, boosting applications of this imaging modality in biomedical research. In this chapter, resonance Raman scattering imaging with single walled carbon nanotubes (SWNTs) in biological systems is reviewed. Recent advances of SERS Raman imaging based on nanoplatforms are summarized. Nanotechnology, in combination with modern optical techniques in Raman spectroscopy, is likely to bring novel opportunities to the realm of molecular imaging.
9.2 RAMAN SCATTERING For light scattering by particles much smaller than the wavelength, a photon is incident on a molecule and interacts with the electric dipole of the molecule, inducing an excitation to a virtual state, followed by nearly coincident deexcitation by a photon emission [10]. If the final state of the molecule is at the same vibration level as the initial state, the scattering photon will have exactly the same energy as the incident photon and this scattering is Rayleigh scattering. In Raman scattering, in contrast, the final state has a different vibration level from the initial state, causing a shift of wavelength in the scattered photon. For most molecules, Rayleigh scattering normally overwhelms Raman scattering. In Stokes Raman scattering, molecules in the ground state (the dominant population at room temperature) will be excited to the virtual state and relax back to the excited vibrational state with higher energy. Therefore the scattered light will show a red shift, which is called the Stokes Raman shift (Fig. 9.1). On the other hand, in rare cases if molecules are in the excited vibrational state before photon excitation and relax to the ground state after photon scattering, the scattered photon will have a blue shift with a wavelength lower than the excitation photon. This anti-Stokes Raman scattering has much lower intensity than Stokes Raman scattering but no fluorescence background. When measuring samples with high fluorescence emissions, anti-Stokes Raman scattering may be used because of the higher signal-to-noise ratio despite its weak scattering intensity. Raman scattering of small molecules usually has very low photon efficiency because the majority of incident photons are elastically scattered by Rayleigh scattering. Only a very small fraction of scattered light (approximately one out of a million photons) is Raman scattering [10]. Therefore the spontaneous Raman scattering of molecules is normally too weak to be used in biological detection and imaging, unless enhanced through certain mechanisms including resonance Raman scattering and SERS, as is discussed in the following two sections.
9.3 RESONANCE RAMAN SCATTERING OF SWNTs FOR IN VITRO AND IN VIVO MOLECULAR IMAGING 9.3.1 Resonance Raman Scattering of SWNTs In resonance Raman scattering, the wavelength of the incoming photon is adjusted to match the electronic transition energy of the molecule. Instead of being excited to a virtual energy state, the molecule could undergo a real electronic transition after excitation by an incident photon [11]. When the laser beam has a frequency close to an electronic
P1: OTA/XYZ P2: ABC c09 JWBS037-Chen August 29, 2010
13:16
Printer Name: Yet to Come
RESONANCE RAMAN SCATTERING OF SWNTs FOR IN VITRO AND IN VIVO MOLECULAR IMAGING
199
Virtual energy level
Energy
ve
vs
ve
vs
ve
vs
Excited vibrational state Ground state
ve = vs Rayleigh scattering
ve > vs
Stokes Raman scattering
ve < vs Anti-Stokes Raman scattering
FIGURE 9.1 Rayleigh and Raman light scattering. In Rayleigh scattering, the scattered photon has the same energy as the excitation light. In Raman scattering, the final state of the molecule has a different vibrational energy level from the initial state, affording a frequency shift of the scattered light, which is red-shifted for Stokes Raman scattering and blue-shifted for anti-Stokes Raman scattering.
transition (in resonance), the Raman scattering of vibration models associated with that particular electronic transition will be greatly enhanced by many orders of magnitude. The largely increased Raman scattering intensity allows much higher detection sensitivity in biological systems. Since different molecules have different electronic transitions, the resonance Raman spectroscopy technique became more popular after the advent of tunable lasers in the 1970s. SWNTs are one-dimensional quantum wires whose density of states (DOS) is characterized by the so-called van Hove singularities (VHSs) (Fig. 9.2a) [13, 14]. The sharp VHS defines narrow energy ranges where the DOS intensity is very high, allowing a single nanotube to behave like a molecule with well-defined electronic energy levels at each VHS. The electronic structures of SWNTs are associated with their chiralities and diameters, with electronic transition energies Eii (E11 and E22 are in the near-infrared to visible regions) varying for nanotubes with different chiral indices (n, m) (Fig. 9.2b). For a single SWNT, if the visible or NIR laser frequency is close enough to the E11 or E22 energy, an electronic transition will be induced, followed by the strongly enhanced resonance Raman scattering. Most of as-prepared SWNT samples have various populations of nanotubes with different (n, m). Therefore only a certain fraction of SWNTs in a bulk or solution sample will be in resonance with the laser. Nonresonance Raman signals from other populations of nanotubes are relatively weak. SWNTs prepared by different methods have different chirality and diameter distributions. To obtain the highest resonance effect in Raman spectroscopy, the laser wavelength could be chosen to match the electronic transition energy of the most populated nanotubes at certain chiral indices (n, m). For example, SWNTs prepared by the high-pressure carbon monoxide (Hipco) method exhibit strong Raman scattering signals under a 785-nm laser excitation because many nanotubes including (10, 5), (9, 7), and (11, 3) ones in a Hipco SWNT sample are in or close to the resonance of the 785-nm laser [15]. SWNTs have multiple Raman peaks. The radial breathing model (RBM) and tangential G-band are two important and usually the strongest features in the SWNT Raman spectrum
P1: OTA/XYZ P2: ABC c09 JWBS037-Chen August 29, 2010
200
13:16
Printer Name: Yet to Come
NANOPLATFORMS FOR RAMAN MOLECULAR IMAGING IN BIOLOGICAL SYSTEMS
FIGURE 9.2 Electronic structure of SWNTs. (a–c) Theoretically calculated density of electronic state for an armchair (10, 10) SWNT (a), chiral (11, 9) SWNT (b), and zigzag (22, 0) SWNT (c). (d) Electronic transition energies Eii for all the (n, m) SWNTs with diameters from 0.4 and 3.0 nm using a simple first-neighbor tight binding model [12]. (Reproduced with permission from Jorio et al. [13]. Copyright © 2003 New Journal of Physics.)
[13]. The RBM peak of SWNTs occurring at wave numbers between 100 cm−1 and 300 cm−1 corresponds to nanotube “breathing,” the vibration of carbon atoms in the radial direction (Fig. 9.3a). The vibration energy and thus the RBM peak position is determined by the diameter of nanotubes. The G-band peak at around 1580 cm−1 is due to tangential vibration of carbon atoms in the graphitic plane (Fig. 9.3b). Depending on the vibrational direction, the SWNT G-band has two subbands, G+ and G− peaks, which correspond to
FIGURE 9.3 Schematic drawings of atom vibrations in RMB model and G-band model of SWNTs. Depending on the vibration direction, the SWNT G-band has two subbands—G+ band vibration in parallel with the tube axis and G− band vibration perpendicular to the tube axis.
P1: OTA/XYZ P2: ABC c09 JWBS037-Chen August 29, 2010
13:16
Printer Name: Yet to Come
RESONANCE RAMAN SCATTERING OF SWNTs FOR IN VITRO AND IN VIVO MOLECULAR IMAGING
201
vibrations along the tube axis direction and the circumferential direction, respectively. For semiconducting nanotubes, the G+ peak at ∼1590 cm−1 is normally stronger than the G− peak at ∼1570 cm−1 . Unlike RBM peaks, the SWNT G-band peak is less affected by the nanotube diameter. Both RBM and G-band SWNT Raman peaks have been used in Raman imaging studies in biological systems.
9.3.2 Raman Imaging with SWNTs in Biological Systems In recent years, carbon nanotubes have been widely explored for applications in biomedicine, including drug delivery, biological sensing, and molecular imaging (see Chapter 20 of this book). Studies have shown that functionalized SWNTs taken up by living systems escape from cells by exocytosis [16], and from mice mainly via biliary and possibly via renal excretions [17], without exhibiting obvious toxicity [18] in both cases. Carbon nanotubes are able to enter cells via the endocytosis mechanism, shuttling various biological molecules including proteins, DNA, RNA, and drugs into cells. In vivo cancer treatment in an animal model has been realized with a nanotube drug carrier [19]. Ultrasensitive ex vivo protein microarrays have been developed utilizing the resonance Raman scattering of SWNTs in combination with SERS, achieving a protein detection limit down to 1 fM [20]. Moreover, the intrinsic optical properties of SWNTs including resonance Raman scattering and photoluminance in the NIR region have also been used in biological imaging in vitro and in vivo. Raman imaging of SWNTs in live cells was first reported by Heller et al. [21] in 2005 [21]. Hipco SWNTs were functionalized by DNA and used to incubate with 3T3 fibroblast and myoblast stem cells. A 785-nm laser was used as the excitation source for Raman spectroscopic mapping. The RMB peak intensity at each mapping pixel was integrated and color-scaled to create a Raman image. Raman imaging of those cells showed high SWNT Raman signals inside cells, a direct proof of nanotube cellular entry without using additional fluorescent labels. Interestingly, they observed ultrahigh photostability of SWNTs, many orders of magnitudes better than organic fluorescent dyes and NIR quantum dots. The high photostability is due to the inert chemical structure of SWNTs and allows long-term tracking and imaging of SWNTs in biological systems. This was further shown by Liu et al. [17] in a later work, in which ex vivo Raman spectroscopy measurement was used to track SWNT biodistribution in mice over 3 months. Besides in vitro imaging of cells, in vivo Raman imaging of SWNTs in live animals has been achieved by the Gambhir group at Stanford [22, 23]. Hipco SWNTs functionalized by phospholipid–polyethylene glycol (PL-PEG) were conjugated with a targeting Arg–Gly–Asp (RGD) peptide to recognize integrin ␣v 3 receptors overexpressed on tumor vasculature and U87MG human glioblastoma cancer cells [24]. U87MG tumor bearing mice were intravenously injected with SWNT-RGD conjugate before Raman imaging. Whole mouse Raman imaging was conducted on a commercial Renishaw Raman microscope with small modifications. Raman images of tumors on live mice revealed time-course information of SWNT levels in tumors. As expected and consistent to a earlier study using position emission tomography (PET) to image radiolabeled SWNTs in mice [24], high SWNT Raman signals were observed on tumors of mice injected with SWNT-RGD but not from those injected with plain SWNTs, owing to the specific binding between the RGD peptide conjugated on SWNTs and the integrin ␣v 3 protein expressed on cancer cells
P1: OTA/XYZ P2: ABC c09 JWBS037-Chen August 29, 2010
202
13:16
Printer Name: Yet to Come
NANOPLATFORMS FOR RAMAN MOLECULAR IMAGING IN BIOLOGICAL SYSTEMS
(a)
(b)
(c)
FIGURE 9.4 In vivo Raman imaging with SWNTs. (a) A digital photo of tumor bearing mouse used for Raman imaging. The tumor area in the black box was scanned by a Raman spectroscopic microscope. (b) Panel of tumor images from mouse receiving SWNT-RGD at various time points postinjection starting from left to right with 2 h, 8 h, 24 h, 48 h, and 72 h. (c) Panel of tumor images from mouse receiving plain SWNTs without RGD conjugation at various time points postinjection starting from left to right with 2 h, 8 h, 24 h, 48 h, and 72 h. Strong SWNT G-band signals were observed on tumors from mice injected with SWNT-RGD but not from those injected with plain nanotubes. (Reproduced with permission from Zavaleta et al. [22]. Copyright © 2008 American Chemical Society.)
and the vasculature (Fig. 9.4). This is the first proof-of-principle study achieving in vivo molecular imaging using Raman spectroscopy.
9.3.3 Multiplexed Raman Imaging with SWNTs The frequency of carbon–carbon bond vibration in the SWNT G-band model is dependent on the atomic mass. Changing of carbon isotope from normal C12 to C13 changes the position of the SWNT Raman G-band peak, lowering the Raman shift by ∼62 cm−1 . It has also been found that the SWNT G-band frequency shifts as the variation of C13 isotope composition ratios in nanotubes. Utilizing the isotopically modified SWNTs, multiplexed multicolor Raman imaging has been realized by Liu et al. [25] in Dai group at Stanford. Pure C12, pure C13, and mixed C12/C13 SWNTs synthesized by the chemical vapor deposition (CVD) method exhibiting Raman G-band peaks at positions of 1590 cm−1 , 1528 cm−1 , and 1544 cm−1 , respectively, are used as three different Raman “colors” for multiplexed imaging (Fig. 9.5a,b). Those nanotubes were functionalized and conjugated with three different targeting ligands including RGD peptide, anti-EGFR/Her1, and
P1: OTA/XYZ P2: ABC c09 JWBS037-Chen August 29, 2010
13:16
Printer Name: Yet to Come
RESONANCE RAMAN SCATTERING OF SWNTs FOR IN VITRO AND IN VIVO MOLECULAR IMAGING
203
(a)
(b)
(c)
FIGURE 9.5 Multicolor Raman imaging with isotopically modified SWNTs. (a) Schematic SWNTs with three different isotope compositions (C13-SWNT, C12/C13-SWNT, C12-SWNT) conjugated with different targeting ligands. (b) Solution phase Raman spectra of the three SWNT conjugates under a 785-nm laser excitation. Different G-band peak positions were observed. (c) A deconvoluted confocal Raman spectroscopic image of a mixture of three cell lines with different receptor expressions incubated with the three-color SWNT mixture. Scale bar = 100 m. (Reproduced with permission from Liu et al. [25]. Copyright © 2008 American Chemical Society.)
anti-Her2 antibodies to selectively recognize specific receptors on cancer cells. After staining by such a three-color SWNT mixture, cells were imaged by a confocal Raman microscope to obtain spectroscopic maps. The recorded spectrum was deconvoluted into three individual spectra of different SWNT “colors” by curve fitting, obtaining relative intensities of three “colors” at each imaging pixel. Three cancer cell lines with different receptor expressions were well differentiated by their labeled SWNT Raman “colors” in multicolor Raman images (Fig. 9.5c), demonstrating the ability of multiplexed Raman imaging by SWNTs with different isotope compositions, for probing and imaging several biological species simultaneously. By varying C12/C13 ratios in the nanotube synthesis, more SWNT Raman “colors” can be obtained. It is estimated that at least five SWNT Raman “colors” can be imaged and distinguished by this method. Besides the isotopically controlled G-band model shift, the RMB model of SWNTs also shows varied peak positions for nanotubes with different diameters. More Raman colors may be obtained if diameter-separated SWNTs are used [26, 27]. Compared with SERS nanoparticles to be discussed in the next section, Raman signals of SWNTs are weaker than some SERS nanoparticles on a per particle basis [23].
P1: OTA/XYZ P2: ABC c09 JWBS037-Chen August 29, 2010
204
13:16
Printer Name: Yet to Come
NANOPLATFORMS FOR RAMAN MOLECULAR IMAGING IN BIOLOGICAL SYSTEMS
However, a SERS nanoparticle contains many Raman active molecules absorbed on the surface of a heavy noble metal particle, while a SWNT is a single molecule with much less weight. Therefore if these two are compared at the same weight concentration, the Raman intensity of SWNTs could be close to that of SERS nanoparticles. Taken together, SWNTs are promising Raman tags for multiplexed Raman molecular imaging in vitro and potentially in vivo.
9.4 NANOPLATFORMS FOR SERS RAMAN IMAGING 9.4.1 Surface Enhanced Raman Scattering The development of SERS has generated large impacts in the areas of nanotechnology, single-molecule spectroscopy, as well as detection and analysis in biomedical systems. It has been proved that Raman intensities can be enormously enhanced by as much as 1014 –1015 -fold when Raman active molecules are adsorbed on a rough noble metal surface or nanoparticles [6, 28]. The huge enhancement factor allows ultrasensitive Raman spectroscopic measurement of molecules down to a single-molecule level, useful in biological detection and imaging. The exact mechanisms of the enhancement effect in SERS still require further explorations. Currently, two theories—the long-range electromagnetic effect [29] and the shortrange chemical effect—have been proposed to explain this unique phenomenon. These two mechanisms, although substantially different from each other, may simultaneously contribute to the SERS effect. In the electromagnetic mechanism, the surface plasmon of the metal is excited by incident photons, enhancing the electromagnetic field at the surface. If the laser frequency is in resonance with the surface plasmon, the field enhancement effect will be the highest. Silver and gold are thus the most popularly used metals for SERS because their plasmon resonance frequencies fall in the visible to NIR range. In addition, the plasmon oscillation must be perpendicular to the surface in order for scattering to occur. Therefore a rough metal surface or metal nanoparticles are required for the SERS effect [29, 30]. Although the electromagnetic effect applies for many molecules located close to the metal nanoparticle surface, it cannot fully explain the large magnitude of enhancement observed in many experiments. As reported by Xu et al. [31], the maximal enhancement of the electromagnetic field is about 1011 , which only occurs for a strongly coupled structure such as dimer configurations or sharp protrusions. Another effect, the chemical effect, also contributes to the SERS effect and bridges the gap to the observed maximal SERS enhancement of the order 1014 . In the chemical effect [32–34], the Raman active molecule is bound to a certain active site on the metal surface, allowing a charge transfer between each other [35]. The electronic transition energy between the highest occupied molecular orbital (HOMO) and lowest unoccupied molecular orbital (LUMO) in most molecules is usually higher than the energy of visible and NIR light. The charge transfer could lower the transition energy to be close to the incident laser frequency, inducing an enhancement effect similar to that of resonance Raman scattering. In contrast to the electromagnetic effect, which is effective in a long range (10–100 nm), the chemical effect requires direct chemical bonds between molecules and surface active sites (e.g., terraces and steps) [34], and thus only occurs in a very short range on the atomic scale.
P1: OTA/XYZ P2: ABC c09 JWBS037-Chen August 29, 2010
13:16
Printer Name: Yet to Come
NANOPLATFORMS FOR SERS RAMAN IMAGING
205
9.4.2 Fabrication of SERS Nanoprobes Numerous SERS nanoprobes have been developed in the past decade for biological detection and imaging. A typical SERS nanoprobe is composed of four major components (Fig. 9.6a). A metal nanoparticle core is needed to provide the SERS effect. Raman active dye molecules are absorbed on the metal particle to contribute Raman signatures. A surface coating is necessary to afford nanoprobes stability and biocompatibility. Targeting ligands (e.g., antibodies) are conjugated on those nanoprobes to allow selective binding of the nanoparticle to a specific target. Silver and gold are the most commonly used metals for SERS nanoparticles because their plasmon peaks are located in the visible to NIR range. Although silver often offers a higher SERS enhancement factor than gold, gold nanomaterials are preferred for SERS nanoplatforms used in biological imaging especially for in vivo imaging, owing to their plasmon peaks at higher wavelengths, close to or in the NIR window that is ideal for in vivo optical imaging. Besides nanoparticles with single metal composition, Au/Ag alloy nanoparticles [36] and core–shell nanoparticles [37] have also been developed for SERS Raman enhancement. The shape of nanomaterials is not limited to spherical particles. Many nanomaterials with interesting shapes such as gold nanorods [38] and nanocages [39] have also been utilized as SERS platforms. Those nanomaterials have the resonance plasmon in the NIR region and may be particularly useful for in vivo detection and imaging. In Raman active vibration models, the dipole moment of a molecule has to be kept constant during vibration. A number of small molecules, usually aromatic ones, are used as Raman active dyes in SERS nanoplatforms. Pyridine is the first Raman active molecule involved in SERS observation [40]. Many thiol containing aromatic molecules, such as thiophenol, naphthalenethiol, and 4-mercaptobenzoic acid, are widely employed as the Raman dyes on SERS nanoparticles [41–43]. Various fluorescent dyes such as rhodamine derivatives and Cy dyes are also useful Raman reporter molecules in SERS systems [6, 9, 36].
(a)
(b)
(c)
Silica bead
Au or Ag nanoparticles Coating shell
Raman active dyes Targeting ligand
FIGURE 9.6 Schematic drawings of a few commonly used SERS nanoprobes for biological detection and imaging. (a) A SERS nanoprobe based on a single metal nanoparticle. A coating layer envelops the metal nanoparticle labeled with Raman dyes. Targeting ligands such as antibodies are then linked on the SERS nanoparticle. (b) A SERS bead with many small metal nanoparticles set on a big silica bead. (c) A SERS nano- or microprobe with a nucleus formed by aggregated small metal nanoparticles. Raman dyes (green) are deposited on metal nanoparticles while targeting ligands are conjugated on the surface of the outer layer in (b) and (c).
P1: OTA/XYZ P2: ABC c09 JWBS037-Chen August 29, 2010
206
13:16
Printer Name: Yet to Come
NANOPLATFORMS FOR RAMAN MOLECULAR IMAGING IN BIOLOGICAL SYSTEMS
Because every molecule has its own unique Raman fingerprint, multiple targets can be identified in a signal measurement using SERS nanoparticles labeled by different Raman active dyes, allowing a high degree of multiplicity in imaging and detection [44]. A coating layer is required on SERS nanoparticles to protect Raman active dyes from falling off, prevent SERS enhancement of Raman signals of nontarget molecules nonspecifically adsorbed on the metal core, and impart nanoparticles biocompatibility. The first generation SERS Raman nanoprobes were synthesized by coattaching of antibodies and Raman dyes on Au nanoparticles without an additional protection layer [41]. Although functioning, those SERS Raman probes are less stable and have substantial nonspecific absorption signals because they are not well shielded from the external environment. A silica or glass layer coating is the most commonly used strategy in the fabrication of SERS nanoprobes [45–48]. Raman dyes are absorbed on metal nanoparticles first. A layer of silica is then introduced to encapsulate the nanoparticle. Undesired binding of interfering Raman active molecules on the metal surface are avoided by the silica coating, which also improves the stability of SERS Raman nanoprobes. However, bare silica-coated nanoparticles still tend to nonspecifically bind to proteins and cell surface due to limited hydrophilicity of the silica surface [49, 50]. An Additional hydrophilic coating such as PEG polymer may be helpful to further improve the performance of this type of SERS nanoprobes. Recently, the Nie group has developed a simple strategy to acquire stable and biocompatible SERS nanoprobes for in vivo Raman imaging [9]. In their work, green isothiocyanate was absorbed on the surface of a gold nanoparticle, which was further protected by thiolated PEG. Addition of PEG-SH did not replace Raman dye molecules absorbed on the gold surface and improved the stability and biocompatibility of nanoparticles. Nonspecific binding of nanoparticles on cells was also minimized. Interestingly, this method applies for other nonsulfur-containing positively charged Raman active molecules, which are absorbed on the gold surface due to an electrostatic force and delocalized -electron interaction. To create “smart” SERS nanoprobes with molecular recognition ability, targeting ligands are conjugated to SERS nanoparticles. Various antibodies have been linked to SERS nanoprobes to detect or image different molecular species in biological systems [9, 45–48, 51, 52]. Oligonucleotides can also be used to functionalize SERS nanoparticles to make gene probes with high detection sensitivities [36, 44]. Besides SERS nanoplatforms with single nanoparticle cores, SERS beads with multiple metal nanoparticles in one bead have also been developed for Raman detection and imaging. In one type of SERS beads, small gold or silver nanoparticles labeled with Raman dyes are anchored on the surface of a big silica bead from 100 nm to a few micrometers (Fig. 9.6b) [45, 53]. The bead is further coated with a layer of silica to acquire better stability. Another type of SERS beads have Raman dye labeled small Ag or Au nanoparticles aggregated in the nucleus, and an outer lay of polymer or silica coating to envelop the beads (Fig. 9.6c) [51, 52]. Composite organic–inorganic nanoparticles (COINs) developed by Intel Corp. fall into this category [54, 55]. The SERS enhancement effect of such a coupled nanostructure in the “hot” bead core with metal nanoparticle aggregates could be higher than that of a single individual metal nanoparticle. However, the near micrometer sizes of some SERS beads are not ideal for certain applications such as in vivo molecular imaging. 9.4.3 Biological Detection and Diagnosis by SERS Detection of biological species by SERS spectroscopy has been widely explored in recent years. The Vo-Dinh group has carried out many pioneer studies in this area [56–59]. In an
P1: OTA/XYZ P2: ABC c09 JWBS037-Chen August 29, 2010
13:16
Printer Name: Yet to Come
NANOPLATFORMS FOR SERS RAMAN IMAGING
207
earlier work, they developed a SERS gene probe to selectively detect the DNA of human immunodeficiency virus (HIV) [56]. A primer sequence labeled with a Raman dye was used to amplify the HIV gene by polymerase chain reaction (PCR), affording amplified labeled strands, which were then captured by complementary oligonucleotides anchored on a substrate. Silver was evaporated on the substrate to enhance Raman signals of the dye labeled gene probe through SERS, achieving high sensitivity in the detection of the HIV gene. However, the silver evaporation step used in this protocol is not compatible for a conventional diagnostic lab. Silver colloid deposition was thus employed in a later study to replace the silver evaporation [59]. Similar strategies have been used for SERS detection of various cancer genes on substrates (Fig. 9.7a) [60–63]. Another type of nanoparticle based SERS detection method has been reported by the Mirkin group [36, 44, 64]. In those strategies, Raman dye labeled targeting gold nanoparticles were used to recognize DNA and RNA through hybridization of complementary sequences, or proteins by antibody–antigen binding (Fig. 9.7b). Silver deposition was then
Laser
(a) 1.
(Target DNA)
2.
(Gene probe)
Ag Deposition
(Raman dye)
(Target DNA)
(b)
Laser
(c)
(SERS probe)
(Target molecule) Raman Tag
(d)
1. aHSA Primer antibody
SWNT-GaM-IgG
Laser aHSA
HSA
HSA
Annealing at 400 oC in H2
Au/6armPEGCOOH
2. SWNT-GaM-IgG Secondary antibody
FIGURE 9.7 Schematic drawings of different SERS detection strategies in biological assays. (a) Raman labeled gene probes for DNA detection. Ag deposition is applied to enhance Raman signals by SERS [56, 60–63]. (b) Gold nanoparticle based SERS detection of DNA or proteins [36, 44, 64]. The scattering of Raman dyes on gold nanoparticles is further enhanced by deposition of silver nanoparticles. (c) A conventional strategy of protein sandwich assay using SERS nanoprobes [41, 52, 54, 65, 66]. (d) Protein microarray assay using SWNTs as the Raman tag [20]. Ultrasensitive detection is achieved by combining the resonance enhancement effect and the SERS enhancement effect. (Reproduced with permission from Cao et al. [44]. Copyright © 2002 Science Magazine, 2008 Nature Publish Group.)
P1: OTA/XYZ P2: ABC c09 JWBS037-Chen August 29, 2010
208
13:16
Printer Name: Yet to Come
NANOPLATFORMS FOR RAMAN MOLECULAR IMAGING IN BIOLOGICAL SYSTEMS
introduced on those gold nanoparticles captured either on a substrate or a glass bead, to obtain dramatically enhanced SERS effects for Raman detection. Multiplexed detection of six DNA targets with high sensitivity (20 femtomolar) was realized by using six different Raman dyes with easily distinguishable sharp spectroscopic fingerprints [44]. Many other groups have also reported protein or gene detection with SERS nanoprobes (Fig. 9.7c) [41, 52, 54, 65, 66]. Ni et al. [41] used antibody conjugated SERS nanoparticles to replace fluorescence dyes in a sandwich immunoassay and achieved a detection limit at the nanomolar level. In another work by Sun et al. [65], thiolated oligonucleotide functionalized gold nanoparticles were labeled with various nonfluorescent Raman dyes for multiplexed DNA detection. Up to eight different DNA targets were detected simultaneously through SERS Raman spectroscopy with a detection limit at about 10−7 M. COIN nanoprobes have also been used in the protein sandwich assay, showing a high sensitivity of 1 pg/mL in the detection of interleukin-2 protein [54]. In a recent study by Dai and co-workers, the intense resonance Raman scattering of SWNTs was combined with SERS, extending the detection limit of traditional fluorescence assays from approximately 1 pM [67] to the femtomolar level [20]. In this work, the sandwich protein assay was performed on a gold coated substrate with antibody conjugated SWNTs as Raman probes (Fig. 9.7d). After annealing at 400 ◦ C in H2 , the gold film aggregated and formed small particle islands, which afforded a nearly 100-fold increase in SWNT Raman scattering intensity, boosting the detection sensitivity by two orders of magnitude. Multiplexed protein microarray was also realized using isotopically modified SWNTs with shifted G-band peaks. 9.4.4 Molecular Imaging with SERS Nanoprobes SERS nanoprobes conjugated with targeting ligands can be used to selectively recognize specific receptor molecules. In vitro Raman imaging is thus able to reveal the molecular information of targeted cells. Lee and co-workers used antibody conjugated glass coated SERS beads to recognize HER2 and CD10 proteins expressed on cancer cells and showed high targeting specificity [45]. In a later work by the same group, simultaneous multiplexed Raman imaging of two proteins on cells and tissue slices was performed, also using antibody conjugated SERS beads [53]. Another Korean group used Au/Ag core/shell nanoparticles as the SERS platform with rhodamine 6G Raman dye placed between the core/shell structure [37]. Those nanoparticles were further coated with a PEG layer and conjugated with antiPLC␥ 1 antibody for cell labeling and Raman imaging. Instead of using additional Raman dyes, Kneipp et al. [68, 69] were able to image unlabeled gold nanoparticles inside cells by Raman spectroscopy and detect chemical changes in the intracellular environment of these nanostructures. Biological molecules inside cells were adsorbed on the gold nanoparticle surface after endocytosis of particles, affording strong SERS enhanced Raman signatures of those molecules to be detected by Raman spectroscopy. In vivo Raman detection and imaging with SERS nanoprobes have been relatively less explored. To date, there have been two groups reporting on Raman imaging in small animal models using SERS nanoparticles as contrast agents. Nie and co-workers [9] coated Raman dye labeled gold nanoparticles with PEG. and an infrared dye diethylthiatricarbocyanine (DTTC) was used as the Raman reporter. Those nanoparticles were further conjugated with a single chain variable fragment (ScFv) antibody to recognize the epidermal growth factor receptor (EGFR), an important cancer biomarker overexpressed on various types of cancer cells. The obtained SERS nanoprobes exhibited strong Raman signals and high
P1: OTA/XYZ P2: ABC c09 JWBS037-Chen August 29, 2010
13:16
Printer Name: Yet to Come
NANOPLATFORMS FOR SERS RAMAN IMAGING
(a)
209
(b)
(c)
FIGURE 9.8 In vivo tumor targeting and Raman spectroscopic detection by using SERS nanoprobes [9] (a, b) Raman spectra obtained from the tumor and the liver locations 5 hours after injection of targeted (a) and nontargeted (b) SERS nanoparticles. High tumor uptake of targeted nanoparticles was evidenced by the strong Raman signals from the tumor. (c) Photographs showing a laser beam focusing on the tumor site or on the anatomical location of liver. (Reproduced with permission from Qian et al. [9]. Copyright © 2008 Nature Publishing Group.)
targeting specificity in vitro on EGFR positive neck squamous carcinoma Tu686 cells. After intravenous injection of SERS nanoprobes into mice bearing Tu686 xenograft tumors, Raman spectra were recorded at the tumor site and the anatomical location of the liver (Fig. 9.8). Mice injected with targeted SERS nanoparticles showed strong Raman signatures of DTTC reporter molecules at the tumor site, indicating high tumor uptake of nanoprobes owing to the specific ScFv antibody–EGFR binding (Fig. 9.8a). Nontargeted nanoparticle injected mice, in contrast, showed low Raman signals in the tumor (Fig. 9.8b). Although nanoparticles exhibited high uptake in the reticuloendothelial system (RES) including liver and spleen, the observed Raman signals from the liver were low due to the limit of tissue penetration depth of light in optical imaging (Fig. 9.8a,b). This work is the first success of using SERS nanoparticles for in vivo detection of cancer biomarkers with potential high sensitivity. In another work, Gambhir and co-workers have used glass coated gold nanoparticles as the SERS platform for in vivo Raman imaging in mice [23]. Several types of SERS nanoparticles labeled by different Raman reporters with distinctive spectral fingerprints were subcutaneously injected into mice for imaging (Fig. 9.9). Raman spectroscopic scanning was carried out on a live mouse, revealing different Raman “colors” after spectral
P1: OTA/XYZ P2: ABC c09 JWBS037-Chen August 29, 2010
210
13:16
Printer Name: Yet to Come
NANOPLATFORMS FOR RAMAN MOLECULAR IMAGING IN BIOLOGICAL SYSTEMS
3000
(a)
(d)
Counts
2000
1000
S421 14 60 15 60 16 60 17 60
96 0 10 60 11 60 12 60 13 60
86 0
76 0
0
Raman Shift cm-1 3000
Counts
(b)
2000
1000
S440 60
60
17
60
16
60
15
14
60 12 60 13 60
60
11
0
10
96
0 76
86 0
0
Raman Shift cm-1
(c)
5000
Counts
4000 3000 2000
Equal Mix
1000
0 11 60 12 60 13 60 14 60 15 60 16 60 17 60
0
10 6
96
86 0
76 0
0
Raman Shift cm-1
(e)
(f)
Head
L
L
Tail
FIGURE 9.9 In vivo Raman imaging with SERS nanoparticles [23]. (a, b) Raman spectra of two types of SERS nanoparticles (a, S421; b, S440) with different Raman labels. (c) Raman spectrum of a mixture of S421 and S440 nanoparticles. (d) A multicolor Raman image of a mouse subcutaneously injected with S421 nanoparticles, S440 nanoparticles, and a mixture of the two nanoparticles. Artificial red and green colors were assigned to Raman signals from S421 and S440 nanoparticles, respectively. (e) A whole-body Raman image of a mouse intravenously injected with SERS nanoparticles. (f) A Raman image of liver area showing high uptake of nanoparticles in the liver. (Reproduced with permission from Keren et al. [23]. Copyright © 2008 National Academy Society.)
P1: OTA/XYZ P2: ABC c09 JWBS037-Chen August 29, 2010
13:16
Printer Name: Yet to Come
CONCLUSION AND PERSPECTIVES
211
unmixing. In addition to the proof-of-principle imaging of SERS probes under the skin, they intravenously injected a mouse with SERS nanoparticles and performed whole-mouse imaging, observing strong Raman signals from the liver owing to the RES uptake of nanoparticles. This is the first study showing the feasibility of the whole-animal Raman imaging in a multiplexed manner. However, targeted multiplexed molecular imaging in vivo with SERS nanoprobes requires further efforts.
9.5 CONCLUSION AND PERSPECTIVES Raman imaging has various advantages over fluorescence imaging. The sharp spectral signatures impart to Raman imaging a higher imaging multiplicity compared with fluorescence imaging. The benefit is the most obvious in the case of in vivo imaging, when excitations and emissions in the narrow NIR window are necessary to allow the maximal tissue penetration of light. The narrow peak features in a Raman spectrum also enable us to easily distinguish Raman peaks and subtract the autofluorescence background from a recorded spectrum. Removal of interfering background allows higher detection and imaging sensitivity compared with the fluorescent technique, whose detection limit is largely affected by the autofluorescence from biological species. Moreover, Raman tags generally have higher photostability than fluorescent molecules, good for long-term imaging and tracking in biological systems. In this chapter, the basic physics of Raman light scattering, resonance Raman scattering, and surface enhanced Raman scattering are briefly introduced. Applications of resonance Raman scattering of carbon nanotubes in biological imaging are summarized. Pioneer works using SWNTs for in vivo Raman molecular imaging and in vitro multiplexed imaging are described. In a related vein, various SERS nanoplatforms and their applications are reviewed. Commonly used strategies of fabrication of SERS nanoprobes, gene and protein assays with SERS, and biological imaging with SERS nanoparticles are discussed. Despite the rapid progress and great potential in Raman imaging, a few challenges still hamper future applications of Raman nanoplatforms. Current Raman imaging suffers from slow scanning speed. Point-by-point scanning is the most commonly used imaging methodology with a spectrum at each mapping point recorded by a spectrometer. This is a relatively slow process limited by both instruments and the nanoplatforms. Regarding instruments, new Raman microscopes with better designs and the capability of fast mapping are on the market now. However, Raman mapping of a large area with high spatial resolution is still a time-consuming process that often requires tens of minutes or up to a few hours. On the other hand, the absolute intensities of Raman scattering of Raman nanoplatforms including SWNTs and SERS nanoparticles are still much weaker than the fluorescence of high quantum yield fluorescent dyes. Brighter Raman nanoplatforms are desired for future imaging applications. Compared with SERS nanoparticles, SWNTs have a weaker Raman signal per particle. However, currently used SWNTs in Raman detection and imaging are mixtures of nanotubes with different chiralities. Therefore only a small percentage of SWNTs are in resonance with the laser source. Separated SWNTs with a single chirality composition are expected to exhibit a much stronger resonance effect when a proper laser wavelength is applied to match the E11 or E22 energy of this particular type of SWNTs. In addition, significant SERS enhancement of SWNT Raman scattering has been reported earlier on the substrate for
P1: OTA/XYZ P2: ABC c09 JWBS037-Chen August 29, 2010
212
13:16
Printer Name: Yet to Come
NANOPLATFORMS FOR RAMAN MOLECULAR IMAGING IN BIOLOGICAL SYSTEMS
biological detection. Efforts may be put to develop SERS enhanced SWNT nanoplatforms for Raman imaging to achieve stronger signals and faster mapping speeds. In the aggregated noble metal nanoparticle system, the interparticle distance is very small, affording a much more localized region with ultrahigh enhancement of the electromagnetic field [31]. SERS nanoplatforms such as COINs [54] with a core of aggregated metal nanoparticles (Fig. 9.6c) are expected to show stronger SERS enhancement effect than single metal nanoparticles. However, the aggregation of metal nanoparticles is not perfectly controlled, raising potential batch-to-batch variation problems of those SERS platforms. For in vivo Raman imaging, the safety issue of nanoplatforms is a concern. Like other nanomaterials, carbon nanotubes and SERS nanoparticles all tend to accumulate in RES organs after intravenous administration into animals. Better surface coating is able to reduce RES uptake. Although PEGylated SWNTs appear to be safe in mice at tested doses, much more work is required to fully understand the potential side effects of those inorganic nanomaterials to a living subject. Besides conventional Raman spectroscopy, coherent anti-Stokes Raman scattering spectroscopy (CARS), a nonlinear optical process with three laser beams involved, has also been widely used in biological imaging [8, 70–72]. CARS spectroscopy is free of fluorescence background and yields higher total signals than Raman spectroscopy owing to the coherent addition of emissions from molecules in the laser beam focused region. Therefore CARS imaging has a much faster scanning speed and enables real-time imaging of biological systems with chemical selectivity in the absence of additional labels. However, though without interference from the fluorescence, CARS spectroscopy contains an inherent nonresonant background, which limits the detection of CARS signals from molecules at low concentrations. Studies have shown that gold nanoparticles are able to enhance CARS signals of their surrounding molecules, allowing indirect detection of gold nanoparticles in biological samples [73, 74]. Combination of CARS spectroscopy and nanotechnology may offer new opportunities in biological imaging. In summary, Raman imaging with nanoplatforms has been intensively investigated in the past decade. Various problems including relatively weak Raman signals, slow imaging process, and concerns of in vivo toxicity of inorganic nanoparticles remain to be resolved. Nevertheless, Raman imaging is a promising technique for future highly sensitive multiplexed molecular imaging in vitro and in vivo.
REFERENCES 1. Wagnieres, G. A.; Star, W. M.; Wilson, B. C. In vivo fluorescence spectroscopy and imaging for oncological applications. Photochem. Photobiol. 1998, 68(5), 603–632. 2. Song, L. L.; Hennink, E. J.; Young, I. T.; Tanke, H. J. Photobleaching kinetics of fluorescein in quantitative fluorescence microscopy. Biophys. J. 1995, 68(6), 2588–2600. 3. Aubin, J. E. Autofluorescence of viable cultured mammalian-cells. J. Histochem. Cytochem. 1979, 27(1), 36–43. 4. Frangioni, J. V. In vivo near-infrared fluorescence imaging. Curr. Opin. Chem. Biol. 2003, 7(5), 626–634. 5. Chen, X. Y.; Conti, P. S.; Moats, R. A. In vivo near-infrared fluorescence imaging of integrin a,alpha(v)beta(3) in brain tumor xenografts. Cancer Res. 2004, 64(21), 8009–8014.
P1: OTA/XYZ P2: ABC c09 JWBS037-Chen August 29, 2010
13:16
Printer Name: Yet to Come
REFERENCES
213
6. Nie, S. M.; Emery, S. R. Probing single molecules and single nanoparticles by surface-enhanced Raman scattering. Science 1997, 275(5303), 1102–1106. 7. Hanlon, E. B.; Manoharan, R.; Koo, T. W.; Shafer, K. E.; Motz, J. T.; Fitzmaurice, M.; Kramer, J. R.; Itzkan, I.; Dasari, R. R.; Feld, M. S. Prospects for in vivo Raman spectroscopy. Phys. Med. Biol. 2000, 45(2), R1–R59. 8. Baena, J. R.; Lendl, B. Raman spectroscopy in chemical bioanalysis. Curr. Opin. Chem. Biol. 2004, 8(5), 534–539. 9. Qian, X. M.; Peng, X. H.; Ansari, D. O.; Yin-Goen, Q.; Chen, G. Z.; Shin, D. M.; Yang, L.; Young, A. N.; Wang, M. D.; Nie, S. M. In vivo tumor targeting and spectroscopic detection with surface-enhanced Raman nanoparticle tags. Nat. Biotechnol. 2008, 26(1), 83–90. 10. Bohren, C. F.; Huffman, D. R. Absorption and scattering of light by small particles. Wiley-VCH 1983, x, x–x. 11. Bobovich, Y. S.; Bortkevich, A. V. Resonance stimulated Raman light-scattering (review article). Kvantovaya Elektronika 1977, 4(3), 485–512. 12. Saito, R.; Dresselhaus, G.; Dresselhaus, M. S. Physical properties of carbon nanotubes. World Scientific Publishing Company 1998, x, x–x. 13. Jorio, A.; Pimenta, M. A.; Souza, A. G.; Saito, R.; Dresselhaus, G.; Dresselhaus, M. S. Characterizing carbon nanotube samples with resonance Raman scattering. New J. Phys. 2003, 5, 139.1–139.17. 14. Wildoer, J. W. G.; Venema, L. C.; Rinzler, A. G.; Smalley, R. E.; Dekker, C. Electronic structure of atomically resolved carbon nanotubes. Nature 1998, 391(6662), 59–62. 15. Liu, Z.; Tabakman, S.; Welsher, K.; Dai, H. Carbon nanotubes in biology and medicine: in vitro and in vivo detection, imaging and drug delivery. Nano Res. 2009, 2(2), 85–120. 16. Jin, H.; Heller, D. A.; Strano, M. S. Single-particle tracking of endocytosis and exocytosis of single-walled carbon nanotubes in NIH-3T3 cells. Nano Lett. 2008, 8(6), 1577–1585. 17. Liu, Z.; Davis, C.; Cai, W.; He, L.; Chen, X.; Dai, H. Circulation and long-term fate of functionalized, biocompatible single-walled carbon nanotubes in mice probed by Raman spectroscopy. Proc. Natl. Acad. Sci. U.S.A. 2008, 105(5), 1410–1415. 18. Schipper, M. L.; Nakayama-Ratchford, N.; Davis, C. R.; Kam, N. W. S.; Chu, P.; Liu, Z.; Sun, X.; Dai, H.; Gambhir, S. S. A pilot toxicology study of single-walled carbon nanotubes in a small sample of mice. Nat. Nanotechnol. 2008, 3(4), 216–221. 19. Liu, Z.; Chen, K.; Davis, C.; Sherlock, S.; Cao, Q.; Chen, X.; Dai, H. Drug delivery with carbon nanotubes for in vivo cancer treatment. Cancer Res. 2008, 68, 6652–6660. 20. Chen, Z.; Tabakman, S. M.; Goodwin, A. P.; Kattah, M. G.; Daranciang, D.; Wang, X.; Zhang, G.; Li, X.; Liu, Z.; Utz, P. J.; Jiang, K.; Fan, S.; Dai, H. Protein microarrays with carbon nanotubes as multicolor Raman labels. Nat. Biotechnol. 2008, 26, 1285–1292. 21. Heller, D. A.; Baik, S.; Eurell, T. E.; Strano, M. S. Single-walled carbon nanotube spectroscopy in live cells: towards long-term labels and optical sensors. Adv. Mater. 2005, 17(23), 2793– 2799. 22. Zavaleta, C.; de la Zerda, A.; Liu, Z.; Keren, S.; Cheng, Z.; Schipper, M.; Chen, X.; Dai, H.; Gambhir, S. S. Noninvasive Raman spectroscopy in living mice for evaluation of tumor targeting with carbon nanotubes. Nano Lett. 2008, 8(9), 2800–2805. 23. Keren, S.; Zavaleta, C.; Cheng, Z.; de la Zerda, A.; Gheysens, O.; Gambhir, S. S. Noninvasive molecular imaging of small living subjects using Raman spectroscopy. Proc. Nat. Acad. Sci. U.S.A. 2008, 105(15), 5844–5849. 24. Liu, Z.; Cai, W. B.; He, L. N.; Nakayama, N.; Chen, K.; Sun, X. M.; Chen, X. Y.; Dai, H. J. In vivo biodistribution and highly efficient tumour targeting of carbon nanotubes in mice. Nat. Nanotechnol. 2007, 2(1), 47–52.
P1: OTA/XYZ P2: ABC c09 JWBS037-Chen August 29, 2010
214
13:16
Printer Name: Yet to Come
NANOPLATFORMS FOR RAMAN MOLECULAR IMAGING IN BIOLOGICAL SYSTEMS
25. Liu, Z.; Li, X.; Tabakman, S. M.; Jiang, K.; Fan, S.; Dai, H. Multiplexed multi-color Raman imaging of live cells with isotopically modified single walled carbon nanotubes. J. Am. Chem. Soc. 2008, 130, 13540–13541. 26. Tu, X.; Zheng, M. A DNA-based approach to the carbon nanotube sorting problem. Nano Res. 2008, 1(3), 185–194. 27. Zheng, M.; Jagota, A.; Semke, E. D.; Diner, B. A.; Mclean, R. S.; Lustig, S. R.; Richardson, R. E.; Tassi, N. G. DNA-assisted dispersion and separation of carbon nanotubes. Nat. Mater. 2003, 2(5), 338–342. 28. Xu, H. X.; Bjerneld, E. J.; Kall, M.; Borjesson, L. Spectroscopy of single hemoglobin molecules by surface enhanced Raman scattering. Phys. Rev. Lett. 1999, 83(21), 4357–4360. 29. Moskovits, M. Surface-roughness and enhanced intensity of Raman-scattering by molecules adsorbed on metals. J. Chem. Phys. 1978, 69(9), 4159–4161. 30. Kerker, M.; Wang, D. S.; Chew, H. Surface enhanced Raman-scattering (SERS) by molecules adsorbed at spherical-particles. Appl. Opt. 1980, 19(24), 4159–4174. 31. Xu, H. X.; Aizpurua, J.; Kall, M.; Apell, P. Electromagnetic contributions to single-molecule sensitivity in surface-enhanced Raman scattering. Phys. Rev. E 2000, 62(3), 4318–4324. 32. Otto, A. What is observed in single molecule SERS, and why? J. Raman Spectrosc. 2002, 33(8), 593–598. 33. Liang, E. J.; Kiefer, W. Chemical effect of SERS with near-infrared excitation. J. Raman Spectrosc. 1996, 27(12), 879–885. 34. Doering, W. E.; Nie, S. M. Single-molecule and single-nanoparticle SERS: examining the roles of surface active sites and chemical enhancement. J. Phys. Chem. B 2002, 106(2), 311–317. 35. Lombardi, J. R.; Birke, R. L.; Lu, T. H.; Xu, J. Charge-transfer theory of surface enhanced Raman-spectroscopy—Herzberg–Teller contributions. J. Chem. Phys. 1986, 84(8), 4174–4180. 36. Jin, R. C.; Cao, Y. C.; Thaxton, C. S.; Mirkin, C. A. Glass-bead-based parallel detection of DNA using composite Raman labels. Small 2006, 2(3), 375–380. 37. Lee, S.; Kim, S.; Choo, J.; Shin, S. Y.; Lee, Y. H.; Choi, H. Y.; Ha, S. H.; Kang, K. H.; Oh, C. H. Biological imaging of HEK293 cells expressing PLC gamma 1 using surface-enhanced Raman microscopy. Anal. Chem. 2007, 79(3), 916–922. 38. Nikoobakht, B.; Wang, J. P.; El-Sayed, M. A. Surface-enhanced Raman scattering of molecules adsorbed on gold nanorods: off-surface plasmon resonance condition. Chem. Phys. Lett. 2002, 366(1-2), 17–23. 39. Au, L.; Chen, Y.; Zhou, F.; Camargo, P. H. C.; Lim, B.; Li, Z.-Y.; Ginger, D. S.; Xia, Y. Synthesis and optical properties of cubic gold nanoframes. Nano Res. 2008, 1, 441–449. 40. Fleischm, M; Hendra, P. J.; Mcquilla, A. J. Raman-spectra of pyridine adsorbed at a silver electrode. Chem. Phys. Lett. 1974, 26(2), 163–166. 41. Ni, J.; Lipert, R. J.; Dawson, G. B.; Porter, M. D. Immunoassay readout method using extrinsic Raman labels adsorbed on immunogold colloids. Anal. Chem. 1999, 71(21), 4903–4908. 42. Viets, C.; Hill, W. Laser power effects in SERS spectroscopy at thin metal films. J. Phys. Chem. B 2001, 105(27), 6330–6336. 43. Michota, A.; Bukowska, J. Surface-enhanced Raman scattering (SERS) of 4-mercaptobenzoic acid on silver and gold substrates. J. Raman Spectrosc. 2003, 34(1), 21–25. 44. Cao, Y. W. C.; Jin, R. C.; Mirkin, C. A. Nanoparticles with Raman spectroscopic fingerprints for DNA and RNA detection. Science 2002, 297(5586), 1536–1540. 45. Kim, J. H.; Kim, J. S.; Choi, H.; Lee, S. M.; Jun, B. H.; Yu, K. N.; Kuk, E.; Kim, Y. K.; Jeong, D. H.; Cho, M. H.; Lee, Y. S. Nanoparticle probes with surface enhanced Raman spectroscopic tags for cellular cancer targeting. Anal. Chem. 2006, 78(19), 6967–6973.
P1: OTA/XYZ P2: ABC c09 JWBS037-Chen August 29, 2010
13:16
Printer Name: Yet to Come
REFERENCES
215
46. Gong, J. L.; Liang, Y.; Huang, Y.; Chen, J. W.; Jiang, J. H.; Shen, G. L.; Yu, R. Q. Ag/SiO2 core–shell nanoparticle-based surface-enhanced Raman probes for immunoassay of cancer marker using silica-coated magnetic nanoparticles as separation tools. Biosensors & Bioelectronics 2007, 22(7), 1501–1507. 47. Liang, Y.; Gong, J. L.; Huang, Y.; Zheng, Y.; Jiang, J. H.; Shen, G. L.; Yu, R. Q. Biocompatible core–shell nanoparticle-based surface-enhanced Raman scattering probes for detection of DNA related to HIV gene using silica-coated magnetic nanoparticles as separation tools. Talanta 2007, 72(2), 443–449. 48. Mulvaney, S. P.; Musick, M. D.; Keating, C. D.; Natan, M. J. Glass-coated, analyte-tagged nanoparticles: a new tagging system based on detection with surface-enhanced Raman scattering. Langmuir 2003, 19(11), 4784–4790. 49. Jones, B. M.; Wagner, J. C.; Edwards, J. H. Absorption of serum-proteins by inorganic dusts. Br. J. Industrial Med. 1972, 29(3), 287-288. 50. Kaufmann, S.; Tanaka, M. Cell adhesion onto highly curved surfaces: one-step immobilization of human erythrocyte membranes on silica beads. Chemphyschem 2003, 4(7), 699–704. 51. McCabe, A. F.; Eliasson, C.; Prasath, R. A.; Hernandez-Santana, A.; Stevenson, L.; Apple, I.; Cormack, P. A. G.; Graham, D.; Smith, W. E.; Corish, P.; Lipscomb, S. J.; Holland, E. R.; Prince, P. D. SERRS labelled beads for multiplex detection. Faraday Discuss. 2006, 132, 303–308. 52. Huang, P. J.; Chau, L. K.; Yang, T. S.; Tay, L. L.; Lin, T. T. Nanoaggregate-embedded beads as novel Raman labels for biodetection. Adv. Functional Mater. 2009, 19(2), 242–248. 53. Yu, K. N.; Lee, S. M.; Han, J. Y.; Park, H.; Woo, M. A.; Noh, M. S.; Hwang, S. K.; Kwon, J. T.; Jin, H.; Kim, Y. K.; Hergenrother, P. J.; Jeong, D. H.; Lee, Y. S.; Cho, M. H. Multiplex targeting, tracking, and imaging of apoptosis by fluorescent surface enhanced Raman spectroscopic dots. Bioconjug. Chem. 2007, 18(4), 1155–1162. 54. Su, X.; Zhang, J.; Sun, L.; Koo, T. W.; Chan, S.; Sundararajan, N.; Yamakawa, M.; Berlin, A. A. Composite organic–inorganic nanoparticles (COINs) with chemically encoded optical signatures. Nano Lett. 2005, 5(1), 49–54. 55. Sun, L.; Sung, K. B.; Dentinger, C.; Lutz, B.; Nguyen, L.; Zhang, J. W.; Qin, H. Y.; Yamakawa, M.; Cao, M. Q.; Lu, Y.; Chmura, A. J.; Zhu, J.; Su, X.; Berlin, A. A.; Chan, S.; Knudsen, B. Composite organic–inorganic nanoparticles as Raman labels for tissue analysis. Nano Lett. 2007, 7(2), 351–356. 56. Isola, N. R.; Stokes, D. L.; Vo-Dinh, T. Surface enhanced Raman gene probe for HIV detection. Anal. Chem. 1998, 70(7), 1352–1356. 57. Vo-Dinh, T.; Yan, F.; Wabuyele, M. B. Surface-enhanced Raman scattering for medical diagnostics and biological imaging. J. Raman Spectrosc. 2005, 36(6-7), 640–647. 58. Yan, F.; Vo-Dinh, T. Surface-enhanced Raman scattering detection of chemical and biological agents using a portable Raman integrated tunable sensor. Sensors and Actuators B—Chemical 2007, 121(1), 61–66. 59. Culha, M.; Stokes, D.; Allain, L. R.; Vo-Dinh, T. Surface-enhanced Raman scattering substrate based on a self-assembled monolayer for use in gene diagnostics. Anal. Chem. 2003, 75(22), 6196–6201. 60. Allain, L. R.; Vo-Dinh, T. Surface-enhanced Raman scattering detection of the breast cancer susceptibility gene BRCA1 using a silver-coated microarray platform. Anal. Chim. Acta 2002, 469(1), 149–154. 61. Vo-Dinh, T.; Allain, L. R.; Stokes, D. L. Cancer gene detection using surface-enhanced Raman scattering (SERS). J. Raman Spectrosc. 2002, 33(7), 511–516. 62. Pal, A.; Isola, N. R.; Alarie, J. P.; Stokes, D. L.; Vo-Dinh, T. Synthesis and characterization of SERS gene probe for BRCA-1 (breast cancer). Faraday Discuss. 2006, 132, 293–301.
P1: OTA/XYZ P2: ABC c09 JWBS037-Chen August 29, 2010
216
13:16
Printer Name: Yet to Come
NANOPLATFORMS FOR RAMAN MOLECULAR IMAGING IN BIOLOGICAL SYSTEMS
63. Volkan, M.; Stokes, D. L.; Vo-Dinh, T. A sol-gel derived AgCl photochromic coating on glass for SERS chemical sensor application. Sensors and Actuators B—Chemical 2005, 106(2), 660–667. 64. Cao, Y. C.; Jin, R. C.; Nam, J. M.; Thaxton, C. S.; Mirkin, C. A. Raman dye-labeled nanoparticle probes for proteins. J. Am. Chem. Soc. 2003, 125(48), 14676–14677. 65. Sun, L.; Yu, C. X.; Irudayaraj, J. Surface-enhanced Raman scattering based nonfluorescent probe for multiplex DNA detection. Anal. Chem. 2007, 79(11), 3981–3988. 66. Fritzsche, W.; Taton, T. A. Metal nanoparticles as labels for heterogeneous, chip-based DNA detection. Nanotechnology 2003, 14(12), R63–R73. 67. Espina, V.; Woodhouse, E. C.; Wulfkuhle, J.; Asmussen, H. D.; Petricoin, E. F., 3rd; Liotta, L. A. Protein microarray detection strategies: focus on direct detection technologies. J. Immunol. Methods 2004, 290(1-2), 121–33. 68. Kneipp, K.; Haka, A. S.; Kneipp, H.; Badizadegan, K.; Yoshizawa, N.; Boone, C.; Shafer-Peltier, K. E.; Motz, J. T.; Dasari, R. R.; Feld, M. S. Surface-enhanced Raman spectroscopy in single living cells using gold nanoparticles. Appl. Spectrosc. 2002, 56(2), 150–154. 69. Kneipp, J.; Kneipp, H.; McLaughlin, M.; Brown, D.; Kneipp, K. In vivo molecular probing of cellular compartments with gold nanoparticles and nanoaggregates. Nano Lett. 2006, 6(10), 2225–2231. 70. Tolles, W. M.; Nibler, J. W.; Mcdonald, J. R.; Harvey, A. B. Review of theory and application of coherent anti-stokes Raman-spectroscopy (CARS). Appl. Spectrosc. 1977, 31(4), 253–271. 71. Zumbusch, A.; Holtom, G. R.; Xie, X. S. Three-dimensional vibrational imaging by coherent anti-Stokes Raman scattering. Phys. Rev. Lett. 1999, 82(20), 4142–4145. 72. Xie, X. S.; Yu, J.; Yang, W. Y. Perspective—living cells as test tubes. Science 2006, 312(5771), 228–230. 73. Ichimura, T.; Hayazawa, N.; Hashimoto, M.; Inouye, Y.; Kawata, S. Local enhancement of coherent anti-Stokes Raman scattering by isolated gold nanoparticles. J. Raman Spectrosc. 2003, 34(9), 651–654. 74. Tong, L.; Lu, Y.; Lee, R. J.; Cheng, J. X. Imaging receptor-mediated endocytosis with a polymeric nanoparticle-based coherent anti-stokes raman scattering probe. J. Phys. Chem. B 2007, 111(33), 9980–9985.
P1: OTA/XYZ P2: ABC c10 JWBS037-Chen January 25, 2011
9:31
Printer Name: Yet to Come
CHAPTER 10
Single-Walled Carbon Nanotube Near-Infrared Fluorescent Sensors for Biological Systems JINGQING ZHANG and MICHAEL S. STRANO Department of Chemical Engineering, Massachusetts Institute of Technology, Cambridge, Massachusetts, USA
10.1 INTRODUCTION Semiconducting single-walled carbon nanotubes (SWNTs) are a unique class of materials that fluoresce in the near-infrared (NIR) range, typically from 820 to above 1600 nm in wavelength, and are unique in that they are one of the very few photostable species that do so. The optical properties of SWNTs have been thoroughly investigated both theoretically and experimentally, and a complete review of their photophysics is beyond the scope of this chapter (readers are referred to Refs. 1–3). Rather, we focus on summarizing the work from our laboratory and others in using near-infrared fluorescence from SWNTs to detect a broad range of biomolecules. This chapter first summarizes the inherent properties of carbon that motivate this work, then describes the development of sensor assays of various types, and examines future efforts and applications of these platforms to solve biological and biomedical problems.
10.2 FAVORABLE PROPERTIES FOR SENSOR APPLICATIONS Semiconducting single-walled carbon nanotubes have several advantages for sensor applications in the life sciences: (1) completely photostable emission, with no blinking or bleaching, even at high fluence [4]; (2) near-infrared emission where many biological tissues are transparent and many others do not autofluoresce [4, 5–8]; (3) sensitivities that enable even a single-molecule docking to the SWNT sidewall to be detected [3, 9–11]; (4) if encapsulated or functionalized appropriately, they are noncytotoxic and demonstrate favorable pharmacological clearance and biodistribution in vivo [12]; (5) they lack surface states and the requirement of surface passivation to emit, meaning that target molecules can Nanoplatform-Based Molecular Imaging Edited by Xiaoyuan Chen C 2011 John Wiley & Sons, Inc. Copyright
217
P1: OTA/XYZ P2: ABC c10 JWBS037-Chen January 25, 2011
218
9:31
Printer Name: Yet to Come
SINGLE-WALLED CARBON NANOTUBE NEAR-INFRARED FLUORESCENT SENSORS
access the SWNT surface reversibly in many cases where it can be detected [1, 13]; (6) there are large classes of proteins and enzymes [11, 13], DNA oligonucleotides [14–16], peptides [17, 18], and polymers [19] that can encapsulate or wrap SWNTs to enable selective interaction [19] and detection of target molecules, a subfield that is just beginning.
10.3 TISSUE IMPLANTABLE SENSORS Any fluorescent sensor designed for long-term and real-time monitoring in vivo has two principal figures of merit (FOM) that need to be considered, quantum yield and photostability. The quantum yield determines the maximum emitted signal from a sensor for a given excitation intensity [20–23]. However, photobleaching is a property of all organic [7] and most nanoparticle fluorophores [6] investigated to date except for SWNTs [24]. Fluorophores that undergo photochemistry necessarily have limited viability with respect to sensor lifetime (Table 10.1) [25]. The photostability of the fluorescent probe dictates the lifetime of the optical device. For example, for electrochemical glucose sensors, a maximum baseline drift of 20% is tolerated before recalibration is needed. If this is used as the upper limit on the working lifetime ( ), one can express it in terms of the photobleaching rate constant k [25]: =−
ln (0.8) k
(10.1)
Table 10.1 uses photobleaching rate constants obtained from the literature to estimate lifetimes of some common organic fluorophores and quantum dots [25]. These rate constants are highly dependent on the environment and the experimental setup. However, even highly optimized quantum dots still suffer from prominent photobleaching. SWNTs are the only optical probe investigated to date that possesses no photobleaching threshold [24, 26]. In order to have continuous, real-time monitoring, a nonphotobleaching sensor is necessary. In addition, for any in vivo sensor, tissue transparency is another necessary prerequisite. The absorption and scattering of the emitted light within the tissue is strongly wavelength dependent (Fig. 10.1) [4, 25] and limits the fluorescent probe that can be used for any in vivo detection purpose. SWNT sensors have advantages for sensing in tissues as the fluorescence is of longer wavelength where these materials have greater transparency. In many fluorometric assays, autofluorescent background is a key limitation. Near-infrared TABLE 10.1 Comparison of Photobleaching Tendency of Common Organic and Nanoparticle Fluorophores Fluorophore IR-Dye 78-CA [6] Cy5 [6] Indocyanine Green [7] Type II NIR QD [6] SWNT [24] a
Photobleaching Rate Constant (h−1 )
Fluence (mW/cm2 )
Nominal Sensor Lifetimea
250.92 20.52 0.0412 0.0827 0
600 600 28 600 1.0 × 106
3.2 s 39.1 s 5.4 h 2.7 h ∞
Estimated optical sensor device lifetime calculated from Eq. (10.1).
Source: Barone et al. [25].
P1: OTA/XYZ P2: ABC c10 JWBS037-Chen January 25, 2011
9:31
Printer Name: Yet to Come
TISSUE IMPLANTABLE SENSORS
µ (oxy)
100 Cy5 Absorption coefficient
219
µ (deoxy) N-IR Qdots
Average
10
1
Indocyanine Green SWNT
0.1 600
700
800
900
1000
1100
Wavelength (nm)
FIGURE 10.1 Absorption coefficient of human whole blood (the dominant absorber for in vivo applications) for oxygenated, deoxygenated, and the average value. Potential fluorophores are compared. (Reproduced with permission from Barone and co-workers [14, 24, 25].)
fluorophores avoid this background that is typically strong at visible wavelengths but not in the near-infrared where SWNTs fluoresce. Consider an optical sensor implanted a distance d into a tissue sample with absorption coefficient and quantum yield . A simple one-dimensional, absorption and fluorescence model can be used to estimate the signal intensity of different fluorophores and the lifetime of the device [25]: Is = I0 e−2d−k
(10.2)
where I 0 = incident excitation intensity I s = fluorescent optical signal outside the body k = pseudo-first-order photobleaching rate constant = aggregate excitation exposure lifetime For the maximum emitted signal from a sensor implanted a distance d into tissue, the fluorophore properties should be such that the figure of merit (FOM) = e−2d is maximized. Table 10.2 reports this value for some common organic and nanoparticle fluorophores. This calculation clearly indicates that visible fluorophores are strongly attenuated in blood samples. Figure 10.2 shows a wavelength window (near-infrared wavelength) at which SWNTs naturally fluoresce, whereas human tissues strongly absorb and autofluoresce at visible wavelengths. 10.3.1 Single-Walled Carbon Nanotubes as Fluorometric Glucose Sensors The first example of a SWNT-based fluorescence sensor was by Barone et al. [13] and introduced the concept of modulating the fluorescence of SWNTs in response to a target analyte such as glucose.
P1: OTA/XYZ P2: ABC c10 JWBS037-Chen January 25, 2011
220
9:31
Printer Name: Yet to Come
SINGLE-WALLED CARBON NANOTUBE NEAR-INFRARED FLUORESCENT SENSORS
TABLE 10.2 Comparison of Commonly Utilized Visible and Near-Infrared Organic and Nanoparticle Fluorescent Probes Based on Quantum Yield (Q.Y.) and Absorbance in Human Whole Blood () [4, 25] Standards Cy5 Fluorescein Rhodamine 6G Rhodamine B Indocyanine Green Indocyanine Green Type II NIR QD SWNT a Based
Q. Y. (%)
Conditions for Q. Y. Excitation (cm−1 ) (cm−1 ) Measurement (nm) Oxy Deoxy
27 95 95 31 0.266
PBS 0.1 M NaOH, 22 ◦ C Water Water Water (0.15 g/L)
620 496 488 514 820
2 150 200 110 0.96
60 120 105 190 0.77
3.20 × 10−28 5.23 × 10−118 3.30 × 10−133 1.60 × 10−131 4.72 × 10−4
1.14
Blood (0.08 g/L)
830
1.01
0.7788
1.91 × 10−3
13 0.1
PBS PBS
840 1042
1.05 0.889
0.778 0.12
2.09 × 10−2 3.65 × 10−4
FOMa
on a 1-cm implanted sensor device, an average is utilized.
Source: Adapted from Saito et al. [3].
Because of the large need, with over 194 million diabetics worldwide [27], and the tangible benefits of frequent blood glucose measurements [28], a significant effort has been made toward the development and production of a real-time, continuous glucose sensor capable of in vivo operation. Current glucose sensors operate electrochemically and are the standard, partly because they are already commercially available [29] and have FDA approval for use in humans. The majority of these devices utilize transdermal electrodes that require frequent replacement every few days. Devices in development utilize telemetry, and require some form of surgical implantation [30]. If instead, sensors are based on optical signal transmission, they can either be minimally invasive, subcutaneously implanted fluorescence-based devices or involve completely noninvasive spectroscopic detection
1
8
0.8
6
0.6
4
0.4
2
0.2
Blood Abs.
0 400
Water Abs. 650
900
1150
Normalized fluorescence
Absorbance (cm-1)
10
0
1400
Wavelength (nm)
FIGURE 10.2 Nanotubes fluoresce at near-infrared wavelengths where tissue does not absorb. (Adapted from Wray and et al. [4].)
P1: OTA/XYZ P2: ABC c10 JWBS037-Chen January 25, 2011
9:31
Printer Name: Yet to Come
DNA SENSORS
221
techniques. The former are typically based on either FRET between fluorophores [5, 31–33], or some type of competitive binding [34, 35], while the latter currently lack analyte specificity and dynamic range [36], as detailed in the reviews of Moschou, Wickramasinghe, and Wilson [29, 37, 38]. In addition to fluorescence-based techniques, several photonics applications have also been proposed for glucose sensing [39, 40]. However, the high-resolution gratings and notch filters necessary for Raman-based schemes are expensive [39, 40]. A fluorescencebased device would require lower resolution and would be relatively inexpensive due to the affordability of current instrumentation and its amenability to microfabrication techniques [41]. Additionally, a small fluorescence based device could readily be implanted beneath the skin without requiring surgical procedures. In an effort to produce a glucose sensor based on SWNT fluorescence, the Strano lab has explored noncovalent methods of modulating the nanotube emission. In this work, new types of noncovalent functionalization using electron-withdrawing molecules were shown to provide sites for transferring electrons in and out of the nanotube. More specifically, they have shown [13] that it is possible to link enzyme activity or protein binding to electron withdrawal from or donation to a nanotube through a variety of surface chemistries [42, 43]. In this first glucose-sensing work, an electroactive species such as potassium ferricyanide, K3 Fe(CN)6 , can irreversibly adsorb at the surface of the nanotube and shift the Fermi level into the valence bands, or quench the emission after photoexcitation. This quenched state can be partially reversed by hydrogen peroxide (H2 O2 ), resulting in a fluorescence recovery. As glucose oxidase catalyzes the reaction of -d-glucose to the d-glucono-1,5lactone with a H2 O2 coproduct, it is possible to couple the carbon nanotube fluorescence to the glucose concentration [13]. Because the ferricyanide adsorption is irreversible, the sensing solution can be loaded into a sealed dialysis capillary where glucose is free to diffuse across the capillary boundary while the nanotube sensing complex is retained (Fig. 10.3). The nanotube fluorescence from such a capillary is easily imaged in the NIR through a human epidermal tissue sample. The fluorescence of one nanotube is shown to respond to glucose after an 80-s transient, even in a whole blood specimen maintained at 37 ◦ C and pH 7.4. The calculated detection limit is approximately 2.2 molecules/nm of nanotube length. The results demonstrate new opportunities for nanoparticle optical sensors that operate in strongly absorbing media of relevance to medicine or biology. In addition, Barone et al. [25] have compared a direct concentration measuring sensor with flux-based sensors and concluded that the former is more stable to membrane biofouling.
10.4 DNA SENSORS Fluorescence detection of specific DNA sequences through hybridization with a complementary DNA probe has many applications in biological and medical sciences. Jeng et al. [14] have demonstrated the optical detection of DNA hybridization on the surface of single-walled carbon nanotubes (SWNTs) in solution through a SWNT bandgap fluorescence modulation. In this work, they first developed a novel method to suspend SWNTs using DNA and remove free DNA through a two-step dialysis method. The addition of complementary DNA (cDNA) to this DNA–SWNT suspension caused a 2-meV hypsochromic shift in the (6,5) SWNT fluorescence. In contrast, the addition of a noncomplementary DNA (nDNA) does not introduce any significant shift. The normalized energy shift can be
P1: OTA/XYZ P2: ABC c10 JWBS037-Chen January 25, 2011
222
9:31
Printer Name: Yet to Come
SINGLE-WALLED CARBON NANOTUBE NEAR-INFRARED FLUORESCENT SENSORS
(a)
HO HO
O OH OH OH
Enzyme/ SWNT solution
Porous dialysis capillary 13 kDa MW cutoff 105 5
nIR image of filled capillary beneath tissue sample
4.5
200 µm
3.5 3 2.5 2 1.5 1
Epidermal Tissue Sample
0.5
(b)
0
1 cm
1.10
(i)
Relative Intensity
(c) 4.2 mM
0.80
2.4 mM
0.50
1.4 mM 0.20 0
10
Time (min)
Relative Intensity
4
20
0.80 0.60 0.40 0.20 0.00 0
5
10
Glucose concentration (mM)
FIGURE 10.3 (a) A 200 m × 1 cm, 13-kDa microdialysis capillary, shown to scale on a human finger, is loaded with the nanotube solution allowing glucose to diffuse through the membrane with the containment of the sensing medium. Placing the capillary beneath a human epidermal tissue sample (above) and the nanotube fluorescence can be clearly mapped from the capillary, seen in the two-dimensional (2D) profile. (b) After ferricyanide surface reaction at 37 ◦ C and pH buffered at 7.4 (i) the fluorescence response to the addition of 1.4 mM, 2.4 mM, and 4.2 mM glucose is scaled by the difference between minimum and maximum intensities. (c) The response function relates the normalized intensity to the local glucose concentration in the range of blood glucose detection with a type I absorption isotherm. (Adapted from Barone’s work [1, 3, 13].)
P1: OTA/XYZ P2: ABC c10 JWBS037-Chen January 25, 2011
9:31
Printer Name: Yet to Come
DNA SENSORS
223
fitted to a normalized exciton binding energy, which is a function of the surface coverage of DNA on the SWNT, yielding an estimated initial fractional coverage of 0.25, and a final coverage of 0.5 due to DNA hybridization. As the exposed area on the SWNT increases, the local dielectric environment of the SWNT changes, resulting in a decrease in the SWNT emission energy [43]. The hybridization process on the nanotube is found to be significantly slower than hybridization of the same sequence free in solution. Therefore the kinetics and thermodynamics of this hybridization mechanism have been studied to elucidate the reasons for this discrepancy in hybridization rate. The kinetic rates for hybridization were fitted using a two-step mechanism in which the DNA first adsorbed on the nanotube in a fast step, followed by a much slower hybridization step. The rates were also measured at different temperatures so the energetics could be determined. The conclusion of this study was that the cause for the slow hybridization was energetic. Hybridization on the SWNT was compared to transition state thermodynamics where the transition state was represented by the stage when DNA was adsorbed to the SWNT, but had not yet hybridized. Both the original DNA and cDNA strands are adsorbed to the SWNT before hybridization, resulting in a reduced initial free energy. Therefore the effective activation energy is increased between the unhybridized and transition states for DNA–SWNT (Fig. 10.4). As a follow-up work, Jeng et al. [45] have shown for the first time that the detection of a single nucleotide polymorphism (SNP) is indeed possible using SWNT fluorescence. It is known that genomic mutations could lead to genetic disorders [46] and predisposition to diseases [47, 48], such as cancer. The most common mutation investigated so far is the single nucleotide polymorphism (SNP) [47–49], a type of variation where one base can be occupied by either of two different bases [46, 47, 50]. In this work, a SNP sensor was synthesized by adsorbing a probe DNA on the SWNT surface via the two-step dialysis described earlier. The presence of the SNP DNA strand can be detected through an emission energy increase in the peak maximum of the (6,5) nanotube fluorescence, and distinguished from the response caused by completely complementary DNA, as shown in Figure 10.5. Kinetics experiments suggested a two-step model including a fast adsorption step, followed by a slower hybridization step [51]. In addition, the slower
2.5 Normalized Intensity
ΔE (meV)
2 1.5 1 0.5 0 -0.5
0
500 1000 cDNA or nDNA (nM) (a)
1500
1.05 1 0.95 0.9 0.85 987
992 997 Wavelength (nm) (b)
FIGURE 10.4 (a) Addition of complementary DNA (cDNA) causes an increase in energy of the steady state (6,5) fluorescence peak while there is negligible energy change with noncomplementary DNA (nDNA). The solid line is a fit of the dielectric model to the cDNA energy shifts. (b) Sample spectra of the fluorescence peak blue shift with cDNA addition. (Reproduced with permission from Jeng et al. [14].)
P1: OTA/XYZ P2: ABC c10 JWBS037-Chen January 25, 2011
224
9:31
Printer Name: Yet to Come
SINGLE-WALLED CARBON NANOTUBE NEAR-INFRARED FLUORESCENT SENSORS
(a)
(b)
FIGURE 10.5 The presence of a single nucleotide polymorphism (SNP) in the complement results in incomplete hybridization on the nanotube surface and a smaller surface coverage. (a) Transient detection of the energy shifts for cDNA and SNP at 37 ◦ C were fitted with a two-step model of adsorption followed by hybridization of the cDNA or SNP. While the equilibrium constants are the same, the partial hybridization of SNP comes to steady state slower than cDNA hybridization. (b) A diagram showing the different nanotube coverage that can result from complementary DNA (cDNA), and cDNA containing a SNP.
path to reach steady state of the SNP strand is consistent with a disruption in hybridization at the mismatched nucleobase pair in the center of the DNA sequence. The diagram in Figure 10.5b illustrates how full hybridization of the complement with the probe DNA that is already on the nanotube would result in denser DNA packing and more expulsion of water than the partial hybridization that would occur between the SNP and probe DNA. Our observations, that the cDNA + probe DNA on the SWNT result in a higher fluorescence energy than the SNP + probe DNA on the SWNT is consistent with work by Ohno and others [52, 53]. Although the (6,5) nanotube can be used to detect SNP for this particular sequence, this method cannot yet be applied to any sequence generically. One reason for this discrepancy is the sequence-dependent secondary structures that can form. More work is needed to develop a generic method to detect any DNA sequence using SWNT fluorescence.
10.5 SENSORS FOR CELLULAR IMAGING As mentioned earlier, a fluorescence-based SWNT senor has the capability for real-time, long-term cellular and tissue-based monitoring. This is made possible because fluorescence
P1: OTA/XYZ P2: ABC c10 JWBS037-Chen January 25, 2011
9:31
Printer Name: Yet to Come
SINGLE-MOLECULE SENSORS
225
does not blink or photobleach and lies within the NIR wavelength window where water or media do not absorb, scatter, or autofluoresce. Heller et al. [15] used the fluorescence of SWNTs to study the B–Z transition of DNA. This work identified the transition of DNA secondary structure from an analogous B to Z conformation, which modulates the dielectric environment of the SWNT around which it is adsorbed. The SWNT bandgap fluorescence undergoes a red shift when an encapsulating 30-nucleotide oligomer is exposed to cations that bind to the nucleobases and screen the charged backbone of DNA. The transition is thermodynamically identical for DNA on and off the nanotube, except that the propagation length of the former is shorter by fivesixths. The magnitude of the energy shift is described by using an effective medium model and the DNA geometry on the nanotube sidewall. The detection of the B–Z change was demonstrated in whole blood, tissue, and from within living mammalian cells. Each type of SWNT has two distinct optical modes for sensing, intensity modulation, and emission wavelength shift. Heller et al. [16] showed that a pair of single-walled nanotubes provides at least four modes that can be used to uniquely fingerprint agents by the degree to which they alter both the emission band intensity and wavelength. They validated this identification method in vitro by demonstrating the detection of six genotoxic analytes, including chemotherapeutic drugs and reactive oxygen species, which are spectroscopically differentiated into four distinct classes, and also by demonstrating single molecule sensitivity in detecting hydrogen peroxide. Finally, the SWNT sensors detected and identified these analytes in real time within live 3T3 cells, demonstrating multiplexed optical detection from a nanoscale biosensor and the first label-free tool to optically discriminate between genotoxins (Fig. 10.6). A major challenge in the synthesis of nanotube or nanowire sensors is to impart selective analyte binding through means other than covalent linkages, which compromise electronic and optical properties. Kim et al. [19] synthesized a 3,4-diaminophenylfunctionalized dextran (DAP-dex) wrapping for single-walled carbon nanotubes (SWNTs) that imparts rapid and selective fluorescence detection of nitric oxide (NO), which is a gaseous free radical playing an important role as an intracellular and intercellular messenger for signaling. The near-infrared (NIR) fluorescence of SWNTDAP-dex is immediately and directly bleached by NO, but not by other reactive nitrogen and oxygen species. It is found that the fluorescence bleaching of SWNTDAP-dex is reversible and caused by electron transfer from the top of the valence band of the SWNT to the lowest unoccupied molecular orbital (LUMO) of the NO radical. The resulting optical sensor is able to spatiotemporally detect NO produced by stimulating inducible NO synthase (iNOS) in Raw 264.7 macrophage cells, as shown in Figure 10.7. They have also demonstrated the potential of the optical sensor for in vivo detection of NO in a tissue model. The ability of SWNTDAP-dex optical sensor to provide temporal and spatial distribution of NO at nanomolar ranges in a living organism demonstrates its potential for a variety of biological applications.
10.6 SINGLE-MOLECULE SENSORS As mentioned earlier, SWNTs are very sensitive to perturbation of their electronic structure. This type of sensitivity can go down to the single-molecule level. Cognet et al. [10] observed single-molecule adsorption and desorption events by studying single step quenching and brightening of fluorescence of sodium dodecylbenzenesulfonate wrapped SWNT. Jin et al. [11] have developed a type 1 collagen film, similar to those used as 3D cell scaffolds for tissue engineering, embedded with SWNT sensors capable of reporting
P1: OTA/XYZ P2: ABC c10 JWBS037-Chen January 25, 2011
226
9:31
Printer Name: Yet to Come
SINGLE-WALLED CARBON NANOTUBE NEAR-INFRARED FLUORESCENT SENSORS O
A
OH NH2
O RO P O O–
O O P –O O
O
O
O O P O O– N
N O
NH
N O NH H2N
O
O RO P O O OR O–
N
N
O O P –O O
O
O O P O O– N
O
O RO P O O–
O OR
N O
Cl
NH
O
N O
NH
N H2 N
O
NH
O
N O
NH
O
1015
•OH
1050
980
Wavelength (nm)
Alkylating agent
1.0
E
O O P O O–
N O
NH
F
0.9
H2O 2
1.00
0.96
(6,5)
0.94
0.7
0.92
(7,5)
O
100
200
300
O O P O O– N N NH
X O
O RO P O O–
O OR N O
NH
G
1.0 0.8
1O
H
2
0.6 0.4
10
20
30
O
O O P –O O
NH
O
O O P O – O O
O N
NH
N H2N
NH
O
O OR
O N NH O
100
200
300
0.998 0.997
1.00
Time (min)
0.999 0.998 0.997
Time (min)
0.99
200
300
400
0.0
0.5
0.0
0.5
1.0
1.5
2.0
2.5
1.0
1.5
2.0
2.5
1.000
0.98
0.996
0.996
0.6
400
Norm. Energy
0.999
I
0.8
1.001
1.000 Norm. Energy
1.000
•OH
1.0
0.4 0
40
1.001
100
O
N O
O
0.2 0
400
1.001
0
NH
O
2
N H2N
O
0.90 0
N
1050
0.98
0.8
O NH
H2N
Wavelength (nm)
0.6
Norm. Energy (E/Emax )
O
Norm. Energy
Norm. Intensity (I/Imax)
1.1
1015
N
O
O
•OH O RO P O O–
D
NH
O
1
C
(7,5) 2
N
Norm. Energy
Relative Int. 980
H2O 2
OR
O
H2 O2
Norm. Energy
Relative Int.
Alk. B agent
1O
O O P O O– N
O
N
Alkylating agent
(6,5)
O O P O O–
O
N
0.999 0.998 0.997 0.996
0
10
Time (min)
20
30
40
Time (min)
0
100
200
300
Time (min)
400
Time (min)
All Data with Multiplexed PC1 vs PC2
•OH
Alkylating agent 0.5
Principal Component 2
1O 2
0.1 –0.6
H2O2
-0.1
0.4
0.9
1.4
J
Alkylating Agent Conc Peroxide Conc Singlet Oxygen Conc Hydroxyl Radical Conc Alkylating Agent Trans Peroxide Trans
–0.3 Singlet Oxygen Trans Hydroxyl Radical Trans Multiplexed Trans
-0.7 Principal Component 1
FIGURE 10.6 Multimodal detection of four reaction pathways. (A) Scheme of interactions on the DNA–SWNT complex: alkylating agent reaction with guanine, hydrogen peroxide (H2 O2 ) adsorption on the nanotube sidewall, singlet oxygen (1 O2 ) reaction with DNA, and hydroxyl radical (•OH) damage to DNA. (B) DNA–SWNT photoluminescence spectra before (blue) and after (green): introducing mechlorethamine, (C) hydrogen peroxide, (D) singlet oxygen, and (E) hydroxyl radicals. (F) Transient responses of photoluminescence intensity (top) and energy (bottom) of the (6,5) nanotube (black) and (7,5) nanotube (red) upon introducing mechlorethamine (blue border), (G) hydrogen peroxide (magenta border), (H) singlet oxygen (orange border), and (I) hydroxyl radicals (green border). (J) Plot of first two principal component scores of transient (closed points) and concentration-dependent (open points) detection data (from Supplementary Figure S1). Black crosses represent simultaneous singlet oxygen and hydroxyl radical generation (from Supplementary Figure S2b). Area-minimized ovals encompass all data sets taken for each analyte, including those not shown. Arrows denote direction of increasing concentration or time. (Reproduced with permission from Heller et al. [15].)
P1: OTA/XYZ P2: ABC c10 JWBS037-Chen January 25, 2011
9:31
Printer Name: Yet to Come
SINGLE-MOLECULE SENSORS
227
FIGURE 10.7 In vitro NO detection using SWNTDAP-dex sensor. (a) NIR fluorescence images of Raw 264.7 cells incorporating SWNTDAP-dex before and after NO addition. After the cells were first incubated with SWNTDAP-dex (1 g/mL) for 12 h, and washed with PBS to remove SWNTs in the growth medium, an extra solution of NO (5 M, PBS) was added into the cell media at 330 s. A control experiment was also conducted without NO addition. Fluorescence was monitored in real time with 658-nm laser excitation (35 mW) and 63× objective. (b) Real-time tracking of NIR fluorescence response inside Raw 264.7 cells with an addition of NO solution (5 and 0.5 M) at around 330 s. Fluorescence bleaching of SWNTDAP-dex was observed. (c) Averaged NIR fluorescence response to NO produced by iNOS in Raw 264.7 cells stimulated with LPS (20 ng/mL) and IFN-␥ (20 U/mL). The NIR fluorescence images were taken 0 and 12 h after stimulation. As a control experiment, the NIR fluorescence image was taken 12 h after incubation of Raw 264.7 cells without stimulation. (d) Averaged fluorescence intensity from each cell region responding to NO produced by iNOS in Raw 264.7 cells stimulated with LPS and IFN-␥ . After the cells (blue control, without any treatment; red, stimulated with LPS and IFN-␥ ; green, pretreated with NMA (2 mM) for 1 h, and then stimulated with LPS and IFN-␥ ) were monitored at a 2-h interval 6 h after stimulation. (Reproduced with permission Kim et al. [19].)
single-molecule adsorption and desorption. Hidden Markov modeling (HMM) is used to link single-molecule adsorption events to rate constants, as they have demonstrated for H2 O2 , H+ , and Fe(CN6 ) [1, 11]. Of the three kinds of reactant molecules studied, H2 O2 showed the highest quenching equilibrium constant of 1.59 at 20 M while H+ was so insensitive that a similar equilibrium constant was achieved at a concentration as high as 0.1 M (pH 1). These results were self-consistent as reverse (dequenching) rate constants (600 s−1 for H2 O2 , 1130 s−1 for H+ , and 400 s−1 for Fe(CN6 )3− ) were concentration independent while the forward (quenching) rate constants varied with concentration. These developments provide the groundwork for SWNTs to function as single-molecule stochastic biosensors.
P1: OTA/XYZ P2: ABC c10 JWBS037-Chen January 25, 2011
228
9:31
Printer Name: Yet to Come
SINGLE-WALLED CARBON NANOTUBE NEAR-INFRARED FLUORESCENT SENSORS
In another work by Jin et al. [54], human carcinoma epidermal A431 cells were grown on the porous collagen–SWNT film (Fig. 10.8a), and relatively stable Reactive Oxygen Species (ROS) generated from the cells like H2 O2 can diffuse into the film. The definition of ROS is broad. Species including singlet oxygen, superoxide anion radical, hydroxyl radical, hydrogen peroxide, and nitric oxide are all considered to be ROS. However, most of these species are highly unstable with a lifetime between a nanosecond and a microsecond in solution [55, 56]. For instance, the lifetime of singlet oxygen is calculated to be 4 s in vivo [55]; the lifetimes of superoxide radical anion and hydroxyl radical are 1 s and 1 ns, respectively [56]. As a result, only the relatively stable species can diffuse into the film and act as quenchers to the nanotubes over hours of experiment. The SWNTs embedded in the film serve as single-molecule sensors, reporting the fluorescence quenching reactions [57] with the H2 O2 (Fig. 10.8a,c) from EGF–EGFR binding [58]. The existence of the cells causes stepwise quenching and dequenching reactions to occur (top green trace,
FIGURE 10.8 (a) A431 cell was cultured on the collagen–SWNT film where the SWNT sensors were excited by a 658-nm excitation laser at 1 mW of the sample through an Alpha Plan-Apo 100x/1.46 oil emersion objective. The fluorescence emission from the SWNT was collected by the InGaAs nIR camera at 1 frame per second. The quenching activity of each sensor from hydrogen peroxide was binned into 16 categories represented by 16 different color bars with red the highest and black the lowest number of quenching transitions. Scale bars represent 10 microns. (b) Representitive confocal image for A431 cells with EGFR (red) labeled with rabbit polyclonal to EGFR and Alexa Fluor 568 donkey anti-rabbit IgG. The nuclei (blue) is labeled with 4 ,6-diamidino-2-phenylindole (DAPI). (c) Example trace of one SWNT sensor that shows stepwise fluorescence quenching from hydrogen peroxide adsorption, compared to the stable fluorescence trace of SWNT sensor in the absence of a cell. The quenching is reversible. (d) Hydrogen peroxide molecules are detected from the real-time quenching of SWNT sensors for both live and fixed A431 cells with the quenching rate (#/sensor/s) documented in real time. (Adapted from Jin et al. [54].)
P1: OTA/XYZ P2: ABC c10 JWBS037-Chen January 25, 2011
9:31
Printer Name: Yet to Come
REFERENCES
229
Fig. 10.8c), compared to the control experiment where there is no cell (bottom green trace, Fig. 10.8c). As has been detailed in our previous study [57], HMM can be used to recover the actual steps (red traces, Fig. 10.8c) from the data (green traces, Fig. 10.8c). The reversible quenching of collagen-wrapped SWNT by H2 O2 is the most predominant among other relative stable ROS as has been discussed previously. In order to image EGFR, A431 cells were immunostained using rabbit polyclonal to EGFR as the primary antibody and Alexa Fluor 568 donkey anti-rabbit IgG as the secondary antibody. As can be seen in the representative confocal image (Fig. 10.8b), A431 cells express a large amount of EGFR, which is considered to be generating H2 O2 upon biding to EGF [58]. The quenching rate was calculated in real time for EGF stimulation (500 ng/mL EGF was added at t = 0) on live/fixed A431 cells (Fig. 10.8d). Compared to the control where there is no EGF stimulation, the quenching rate of both live and fixed A431 cells has been increased from EGF stimulation. As can be seen from Figure 10.8d, the behaviors of single A431 cells after EGF stimulation are similar: the quenching rate increased rapidly right after stimulation. However, the time point of maximal response ranges from 600 to 1800 s after stimulation. There is no significant difference between live and fixed A431 cells. Compared to the ensemble measurement on thousands of cells, this platform allows the real-time measurement on a single-cell level for the first time.
ACKNOWLEDGMENTS We thank Paul W. Barone, Daniel A. Heller, Esther S. Jeng, Hong Jin, and Jong-Ho Kim for providing valuable input into this chapter.
REFERENCES 1. O’Connell, M. J.; Bachilo, S. M.; Huffman, C. B.; Moore, V. C.; Strano, M. S.; Haroz, E. H.; Rialon, K. L.; Boul, P. J.; Noon, W. H.; Kittrell, C.; Ma, J. P.; Hauge, R. H.; Weisman, R. B.; Smalley, R. E. Band gap fluorescence from individual single-walled carbon nanotubes. Science 2002, 297(5581), 593–596. 2. Bachilo, S. M.; Strano, M. S.; Kittrell, C.; Hauge, R. H.; Smalley, R. E.; Weisman, R. B. Structure-assigned optical spectra of single-walled carbon nanotubes. Science 2002, 298(5602), 2361–2366. 3. Saito, R.; Dresselhaus, M. S.; Dresselhaus, G. Physical Properties of Carbon Nanotubes. Imperial College Press: London, 1998. 4. Wray, S.; Cope, M.; Delpy, D.; Wyatt, J.; Reynolds, E. Characterization of the near infrared absorption spectra of cytochrome aa3 and haemoglobin for the non-invasive monitoring of cerebral oxygenation. Biochim. Biophys. Acta 1988, 933, 184–192. 5. McCartney, L. J.; Pickup, J. C.; Rolinski, O. J.; Birch, D. J. S. Near-infrared fluorescence lifetime assay for serum glucose based on allophycocyanin-labeled concanavalin A. Anal. Biochem. 2001, 292(2), 216–221. 6. Kim, S.; Lim, Y.; et al. Near-infrared fluorescent type II quantum dots for sentinel lymph node mapping. Nat. Biotechnol. 2004, 22(1), 93–97. 7. Saxena, V.; Sadoqi, M.; Shao, J. Degradation kinetics of indocyanine green in aqueous solution. J. Pharm. Sci. 2003, 92(10), 2090–2097.
P1: OTA/XYZ P2: ABC c10 JWBS037-Chen January 25, 2011
230
9:31
Printer Name: Yet to Come
SINGLE-WALLED CARBON NANOTUBE NEAR-INFRARED FLUORESCENT SENSORS
8. Klonis, N.; Quazi, N. H.; Deady, L. W.; Hughes, A. B.; Tilley, L. Characterization of a series of far-red-absorbing thiobarbituric acid oxonol derivatives as fluorescent probes for biological applications. Anal. Biochem. 2003, 317(1), 47–58. 9. Durkop, T.; Getty, S. A.; Cobas, E.; Fuhrer, M. S. Extraordinary mobility in semiconducting carbon nanotubes. Nano Lett. 2004, 4(1), 35–39. 10. Cognet, L.; Tsyboulski, D. A.; Rocha, J. D. R.; Doyle, C. D.; Tour, J. M.; Weisman, R. B. Stepwise quenching of exciton fluorescence in carbon nanotubes by single-molecule reactions. Science 2007, 316(5830), 1465. 11. Jin, H.; Heller, D. A.; Kim, J.-H.; Strano, M. M. Stochastic analysis of stepwise fluorescence quenching reactions on single-walled carbon nanotubes: single molecule sensors. Nano Lett. 2009, 8(12), 4299–4304. 12. Liu, Z.; Chen, K.; Davis, C.; Sherlock, S.; Cao, Q.; Chen, X.; Dai, H. Drug delivery with carbon nanotubes for in vivo cancer treatment. Cancer Res. 2008, 68(16), 6652. 13. Barone, P. W.; Baik, S.; Heller, D. A.; Strano, M. S. Near-infrared optical sensors based on single-walled carbon nanotubes. Nat. Mater. 2005, 4(1), 86–U16. 14. Jeng, E. S.; Moll, A. E.; Roy, A. C.; Gastala, J. B.; Strano, M. S. Detection of DNA hybridization using the near-infrared band-gap fluorescence of single-walled carbon nanotubes. Nano Lett. 2006, 6(3), 371–375. 15. Heller, D. A.; Jeng, E. S.; Yeung, T. K.; Martinez, B. M.; Moll, A. E.; Gastala, J. B.; Strano, M. S. Optical detection of DNA conformational polymorphism on single-walled carbon nanotubes. Science 2006, 311, 508–511. 16. Heller, D. A.; Jin, H.; Martinez, B. M.; Patel, D.; Miller, B. M.; Yeung, T. K.; Jena, P. V.; Hartner, C.; Ha, T.; Silverman, S. K. Multimodal optical sensing and analyte specificity using single-walled carbon nanotubes. Nat. Nanotechnol. 2008, 4(2), 114–120. 17. Pender, M. J.; Sowards, L. A.; Hartgerink, J. D.; Stone, M. O.; Naik, R. R. Peptide-mediated formation of single-wall carbon nanotube composites. Nano Lett. 2006, 6(1), 40–44. 18. Witus, L. S.; Rocha, J. D. R.; Yuwono, V. M.; Paramonov, S. E.; Weisman, R. B.; Hartgerink, J. D. Peptides that non-covalently functionalize single-walled carbon nanotubes to give controlled solubility characteristics. J. Mater. Chem. 2007, 17(19), 1909–1915. 19. Kim, J. H.; Heller, D. A.; Jin, H.; Barone, P. W.; Song, C.; Zhang, J.; Trudel, L. J.; Wogan, G. N.; Tannenbaum, S. R.; Strano, M. S. The rational design of nitric oxide selectivity in single-walled carbon nanotube near-infrared fluorescence sensors for biological detection. Nat. Chem. 2009, 1(6), 473–481. 20. Mujumdar, R.; Ernst, L.; Mujumdar, S.; Lewis, C.; Waggoner, A. Cyanine dye labeling reagents: sulfoindocyanine succinimidyl esters. Bioconjug. Chem. 1993, 4, 105–111. 21. Benson, R.; Kues, H. Fluorescence properties of indocyanine green as related to angiography. Phys. Med. Biol. 1978, 23(1), 159–163. 22. Lakowicz, J. Principles of Fluorescence Spectroscopy, 2nd ed. Academic/Plenum Publishers: New York, 1999. 23. Magde, D.; Rojas, G.; Seybold, P. Solvent dependence of the fluorescence lifetimes of xanthene dyes. Photochem. Photobiol. 1999, 70, 737. 24. Heller, D. A.; Baik, S.; Eurell, T. E.; Strano, M. S. Single-walled carbon nanotube spectroscopy in live cells: towards long-term labels and optical sensors. Adv. Mater. 2005, 17, 2793–2798. 25. Barone, P.; Parker, R.; Strano, M. S. In-vivo fluorescence detection of glucose using a single walled carbon nanotube optical sensor: design, fluorophore properties, advantages and disadvantages. Anal. Chem. 2005, 77, 7556–7562. 26. Cherukuri, P.; Bachilo, S. M.; Litovsky, S. H.; Weisman, R. B. Near-infrared fluorescence microscopy of single-walled carbon nanotubes in phagocytic cells. J. Am. Chem. Soc. 2004, 126, 15638–15639.
P1: OTA/XYZ P2: ABC c10 JWBS037-Chen January 25, 2011
9:31
Printer Name: Yet to Come
REFERENCES
231
27. Diabetes Atlas (http://www.worlddiabetesfoundation.org/). 28. Shamoon, H.; Duffy, H.; Fleischer, N.; Engel, S.; Saenger, P.; et al. The effect of intensive treatment of diabetes on the development and progression of long-term complications in insulindependent diabetes-mellitus. N Engl. J. Med. 1993, 329(14), 977–986. 29. Wilson, G. S.; Gifford, R. Biosensors for real-time in vivo measurements. Biosensors & Bioelectronics 2005, 20(12), 2388–2403. 30. Heller, A. Implanted electrochemical glucose sensors for the management of diabetes. Annu. Rev. Biomed. Eng. 1999, 1, 153–175. 31. Pickup, J. C.; Hussain, F.; Evans, N. D.; Rolinski, O. J.; Birch, D. J. S. Fluorescence-based glucose sensors. Biosensors & Bioelectronics 2005, 20(12), 2555–2565. 32. Cote, G. L. Noninvasive and minimally-invasive optical monitoring technologies. J. Nutr. 2001, 131(5), 1596S–1604S. 33. Rolinski, O. J.; Birch, D. J. S.; McCartney, L. J.; Pickup, J. C. A time-resolved near-infrared fluorescence assay for glucose: opportunities for trans-dermal sensing. J. Photochem. Photobiol. B—Biol. 2000, 54(1), 26–34. 34. Ballerstadt, R.; Polak, A.; Beuhler, A.; Frye, J. In vitro long-term performance study of a nearinfrared fluorescence affinity sensor for glucose monitoring. Biosensors & Bioelectronics 2004, 19(8), 905–914. 35. Ballerstadt, R.; Schultz, J. S. A fluorescence affinity hollow fiber sensor for continuous transdermal glucose monitoring. Anal. Chem. 2000, 72(17), 4185–4192. 36. Gough, D. A.; Armour, J. C.; Baker, D. A. Advances and prospects in glucose assay technology. Diabetologia 1997, 40, S102–S107. 37. Moschou, E. A.; Sharma, B. V.; Deo, S. K.; Daunert, S. Fluorescence glucose detection: advances toward the ideal in vivo biosensor. J. Fluorescence 2004, 14(5), 535–547. 38. Wickramasinghe, Y.; Yang, Y.; Spencer, S. A. Current problems and potential techniques in in vivo glucose monitoring. J. Fluorescence 2004, 14(5), 513–520. 39. Berger, A. J.; Itzkan, I.; Feld, M. S. Feasibility of measuring blood glucose concentration by near-infrared Raman spectroscopy. Spectrochim. Acta Part A—Mol. Biomol. Spectrosc. 1997, 53(2), 287–292. 40. Wang, S. Y.; Hasty, C. E.; Watson, P. A.; Wicksted, J. P.; Stith, R. D.; March, W. F. Analysis of metabolites in aqueous-solutions by using laser Raman-spectroscopy. Appl. Opt. 1993, 32(6), 925–929. 41. Kostov, Y.; Rao, G. Low-cost optical instrumentation for biomedical measurements. Rev. Sci. Instrum. 2000, 71(12), 4361–4374. 42. Zaric, S.; Ostojic, G. N.; Engroff, A. D.; Kono, J.; Shaver, J.; Moore, V. C.; Strano, M. S.; Hauge, R. H.; Smalley, R. E.; Wei, X. Interband magneto-optics in single-walled carbon nanotubes. Superlattices and Microstructures 2003, 34(3-6), 413–418. 43. Strano, M. S.; Moore, V. C.; Miller, M. K.; Allen, M. J.; Haroz, E. H.; Kittrell, C.; Hauge, R. H.; Smalley, R. E. The role of surfactant adsorption during ultrasonication in the dispersion of single-walled carbon nanotubes. J. Nanosci. Nanotechnol. 2003, 3(1-2), 81–86. 44. Barone, P. W.; Heller, D. A.; Baik, S.; Strano, M. S. Solution phase near infrared sensors based on single walled carbon nanotubes. Nat. Mater. 2005, 4, 86–92. 45. Jeng, E. S.; Nelson, J. D.; et al. Detection of a single nucleotide polymorphism using single-walled carbon-nanotube near-infrared fluorescence 13. Small 2010, 6(1), 40–43. 46. Schork, N. J.; Fallin, D.; Lanchbury, J. S. Single nucleotide polymorphisms and the future of genetic epidemiology. Clin. Genet. 2000, 58(4), 250–264. 47. Wang, D. G.; Fan, J. B.; Siao, C. J.; Berno, A.; Young, P.; Sapolsky, R.; Ghandour, G.; Perkins, N.; Winchester, E.; Spencer, J.; Kruglyak, L.; Stein, L.; Hsie, L.; Topaloglou, T.; Hubbell, E.; Robinson, E.; Mittmann, M.; Morris, M. S.; Shen, N. P.; Kilburn, D.; Rioux, J.; Nusbaum,
P1: OTA/XYZ P2: ABC c10 JWBS037-Chen January 25, 2011
232
48. 49. 50. 51.
52.
53. 54.
55. 56. 57.
58.
9:31
Printer Name: Yet to Come
SINGLE-WALLED CARBON NANOTUBE NEAR-INFRARED FLUORESCENT SENSORS
C.; Rozen, S.; Hudson, T. J.; Lipshutz, R.; Chee, M.; Lander, E. S. Large-scale identification, mapping, and genotyping of single-nucleotide polymorphisms in the human genome. Science 1998, 280(5366), 1077–1082. Altshuler, D.; Brooks, L. D.; Chakravarti, A.; Collins, F. S.; Daly, M. J.; Donnelly, P. A haplotype map of the human genome. Nature 2005, 437(7063), 1299–1320. Twyman, R. M. SNP discovery and typing technologies for pharmacogenomics. Curr. Top. Med. Chem. 2004, 4(13), 1423–1431. McCarthy, J. J.; Hilfiker, R. The use of single-nucleotide polymorphism maps in pharmacogenomics. Nat. Biotechnol. 2000, 18(5), 505–508. Jeng, E. S.; Barone, P. W.; Nelson, J. D.; Strano, M. S. Hybridization kinetics and thermodynamics of DNA adsorbed to individually dispersed single-walled carbon nanotubes. Small 2007, 3, 1602–1609. Miyauchi, Y.; Saito, R.; Sato, K.; Ohno, Y.; Iwasaki, S.; Mizutani, T.; Jiang, J.; Maruyama, S. Dependence of exciton transition energy of single-walled carbon nanotubes on surrounding dielectric materials. Chem. Phy. Lett. 2007, 442(4-6), 394–399. Ando, T. Theory of electronic states and transport in carbon nanotubes. J. Phys. Soc. Japan 2005, 74(3), 777–817. Jin, H.; Heller, D. A.; Kalbacova, M.; Kim, J.-H.; Zhang, J.; Boghossian, A. A.; Maheshri, N.; Strano, M. S. Detection of single molecule H2 O2 signaling from epidermal growth factor receptor using fluorescent single-walled carbon nanotubes. Nat. Nanotechnol. 2010, 5(4), 302–309. Foote, C. S. In Free Radicals in Biology; Pryor, W. A., Ed. Academic Press: New York, 1976; pp. 85–133. Ziyatdinova, G. K.; Gil’metdinova, D. M.; Budnikov, G. K. Reactions of superoxide anion radical with antioxidants and their use in voltammetry. J. Anal. Chem. 2005, 60(1), 49–52. Jin, H.; Heller, D. A.; Kim, J.-H.; Strano, M. S. Stochastic analysis of stepwise fluorescence quenching reactions on single-walled carbon nanotubes: single molecule sensors. Nano Lett. 2008, 8(12), 4299–4304. DeYulia, G. J.; Carcamo, J. M.; Borquez-Ojeda, O.; Shelton, C. C.; Golde, D. W. Hydrogen peroxide generated extracellularly by receptor–ligand interaction facilitates cell signaling. Proc. Nat. Acad. Sci. U.S.A. 2005, 102(14), 5044–5049.
P1: OTA/XYZ P2: ABC c11 JWBS037-Chen February 2, 2011
21:47
Printer Name: Yet to Come
CHAPTER 11
Microparticle- and Nanoparticle-Based Contrast-Enhanced Ultrasound Imaging ¨ NIRUPAMA DESHPANDE and JURGEN K. WILLMANN Department of Radiology and Molecular Imaging Program at Stanford, Stanford University School of Medicine, Stanford, California, USA
11.1 INTRODUCTION TO ULTRASOUND IMAGING IN MEDICINE Ultrasound refers to sound waves of 20,000 or more vibrations per second, which are far above the frequency heard by the human ear. Ultrasound imaging in medicine is based on the ability of these sound waves to reflect back from tissues in the human body, allowing an anatomical picture to be drawn on a screen of what is inside the body. Thus in technical terms ultrasound imaging is based on the reception, analysis, and display of acoustic signals produced by reflection or backscatter of sound (echo) above the audible frequency range for humans (20 Hz to 20 kHz). Ultrasound is primarily a tomographic modality, meaning that it generates an image that is typically a cross section of the tissue volume under investigation. It is also a soft-tissue modality, given that current ultrasound methodology does not provide useful images of or through bone or bodies of gas, such as found in the lung and bowel. In the past decade ultrasound imaging has gained immense popularity in medical diagnostic and therapy due to at least four characteristics: (1) it is a noninvasive, real-time imaging modality, (2) it does not utilize ionizing irradiation for image acquisition, (3) it is a quantitative imaging modality, and (4) it has a relatively low cost compared with other imaging modalities. 11.1.1 Historical Review Karl Theodore Dussik of Austria published the first paper on medical ultrasound in 1942, attempting to locate the cerebral ventricles in the brain by measuring the transmission of ultrasound beams through the skull. The first experimental investigations were performed with A-mode (amplitude modulation) on the cathode-ray tube display made up of a vacuum tube containing a source of electrons and a fluorescent screen used to create an image. On the cathode-ray tube display one axis represents the time required for the return of the echo and the other corresponds to the strength of the echo: thus providing a very simple display Nanoplatform-Based Molecular Imaging Edited by Xiaoyuan Chen C 2011 John Wiley & Sons, Inc. Copyright
233
P1: OTA/XYZ P2: ABC c11 JWBS037-Chen February 2, 2011
234
21:47
Printer Name: Yet to Come
MICROPARTICLE- AND NANOPARTICLE-BASED CONTRAST-ENHANCED ULTRASOUND IMAGING
plotted as a series of peaks, the height of which represents the depth of the echoing structure from the transducer. In the early 1950s the first B-mode (brightness modulation) images were obtained. In B-mode, the position of a spot on the cathode-ray tube display corresponds to the time elapsed as well as the position of the echogenic surface and the brightness of the spot corresponds to the strength of the echo. Movement of the ultrasound transducer over tissue produces a sweep of the ultrasound beam and a tomographic scan of a cross section of the body. After 1960, two-dimensional (2D) compound procedures were developed with ultrasound information generated from multiple angles being combined in one 2D image. This technique led to a boom in obstetric and gynecologic ultrasound worldwide. In the 1970s B-mode gray scale imaging became available, wherein the returning sound pulses have different shades of gray depending on their intensities, with solid areas appearing white and liquid areas appearing black. With progress of computer technology, ultrasonic imaging became better and faster and a wider range of clinical applications evolved over the following years. Real-time B-mode imaging, high-resolution scanners with digital beam forming, high transducer frequencies, multichannel focus, and broadband transducer technology have developed. Color Doppler and Duplex became available and sensitivity for low flow has continuously improved [1]. In Doppler ultrasonography, shifts in frequency between emitted ultrasonic waves and their echoes are used to measure the velocities of moving objects. Based on the principle of the Doppler effect, this technique is frequently used to examine cardiovascular blood flow but cannot be used to assess tissue microvascular blood flow. Duplex ultrasonography combines the standard real-time B-mode display with pulsed Doppler signals, allowing analysis of frequency shifts in an ultrasonographic signal, reflecting motion within a tissue (e.g., blood flow). This led to the invention of the power Doppler technique that is sensitive to low blood flow, enabling detailed visualization of vascular blood structure. The flow information is based on the amplitude of echoes received from moving cells and not on frequency shifts. Today, machines with advanced ultrasound system performance are equipped with real-time compound imaging. Furthermore, with the advent of contrast agents that enhance ultrasound signal, many techniques such as tissue harmonic imaging, contrast harmonic imaging, pulse inversion imaging, and 3D and 4D ultrasound with panoramic view became available. Harmonic imaging and pulse inversion imaging utilize the oscillation of gas-filled contrast microbubbles that are circulating in the blood flow, leading to scattering of the ultrasound wave. When an acoustic wave encounters a microbubble, it alternately compresses the microbubble on the positive pressure and expands it on the negative pressure. On the positive portion of the wave, the microbubbles are compressed in a different fashion from the way they expand in the negative portion. This results in an asymmetric, nonlinear bubble oscillation. Instead of producing a sinusoidal echo with a clean frequency spectrum, it produces an echo with asymmetric top and bottom. It is this asymmetry that produces harmonics and can be utilized to enhance the signals from the microbubble contrast agents. Ultrasonic waves of very low sound intensity lead to a linear (resonant) oscillation of the microbubbles. Greater sound amplitudes cause nonlinear (asymmetrical) oscillation of the microbubbles, during which other frequencies besides the output frequency are also emitted during the signal response. These other frequencies are the so-called harmonic frequencies, or overtones and undertones. The frequency components that are harmonic overtones are multiples of (i.e., double, triple, quadruple, etc.) the fundamental frequency used and/or its half, third, quarter, and so on in the case of undertone frequency component. The latter are also called subharmonic frequencies. Among the harmonic overtone frequencies, the second harmonic overtone frequency is of special importance in contrast-agent-specific
P1: OTA/XYZ P2: ABC c11 JWBS037-Chen February 2, 2011
21:47
Printer Name: Yet to Come
INTRODUCTION TO ULTRASOUND IMAGING IN MEDICINE
235
sonography, due to its high amplitude. The basic frequency received from tissue and the microbubble contrast agent is suppressed by special filtering techniques during signal processing and only the contrast-agent-specific parts of the second harmonic frequency are used to generate the image. This procedure results in a clear reduction of the tissue signal portion, resulting in an improvement of the signal-to-noise ratio. In pulse inversion imaging technology, two successive ultrasonic pulses are emitted, whereby the phase of the second pulse is inverted with the phase of the first. From the received (backscattered) signals the two sequential waves (positive and negative waves) are subtracted from each other, whereby mathematically, the results cancel each other with linear signals. On the other hand, the nonlinear signals of the microbubbles contrast agent will not be cancelled out and can be visualized [2]. Finally three-dimensional (3D) ultrasound utilizes several 2D images acquired by moving the ultrasound transducer across the body surface, thereby resulting in a volumetric image. Four-dimensional (4D) ultrasound takes 3D ultrasound images and adds the element of time to the progress so that a moving 3D image is generated [3].
11.1.2 Ultrasound Equipment A basic ultrasound imaging system has the following components: (1) the ultrasound transducer is the main part of the ultrasound imaging system from which the ultrasonic waves are transmitted and received. The waves are generated by electrically stimulating a piezoelectric (pressure electricity) crystal, which produces elastic vibrations that are transmitted to the material being investigated. (2) The central processing unit (CPU) is the computational unit of the ultrasound machine that contains a microprocessor, memory, and amplifiers. It also contains the electrical power supplies for itself and the ultrasound transducer. The CPU does the calculations and data processing and forms the image on the monitor. The CPU also stores the processed data and/or images. (3) The transducer pulse control controls the changes in amplitude, frequency, and duration of the ultrasound pulses emitted from the ultrasound transducer. (4) Keyboard cursors are used for data input and measurement of images. (5) Finally, the monitor displays the images from the ultrasound data processed by the CPU [3].
11.1.3 Safety and Side Effects of Ultrasound Imaging There have been no substantiated side or ill effects of ultrasound imaging documented in studies in both humans and animals. Nevertheless, since ultrasound produces energy, there have been some concerns about the safety of ultrasound due to the development of heat or thermal effect. The thermal effect of ultrasound is caused by absorption of the ultrasound beam energy. As the ultrasound waves are absorbed, their energies are converted into heat. Energy of an ultrasound wave is described by pressure, temperature, density, and particle displacement. An ultrasound wave propagating through a medium loses its energy as a function of distance via reflection, absorption, and scattering. This lost energy is converted into heat. The thermal effect is highest in tissue with a high absorption coefficient, particularly in bone, and is low where there is little absorption such as fluid. The temperature rise is also dependent on the thermal characteristics of the tissue (conduction of heat and perfusion), the ultrasound intensity, and the length of examination time [4].
P1: OTA/XYZ P2: ABC c11 JWBS037-Chen February 2, 2011
236
21:47
Printer Name: Yet to Come
MICROPARTICLE- AND NANOPARTICLE-BASED CONTRAST-ENHANCED ULTRASOUND IMAGING
11.1.4 General Diagnostic Indications of Ultrasound Imaging Ultrasound imaging in medicine is the most widespread imaging technique worldwide; it is easy to use, and the equipment is small and portable and therefore suitable for a doctor’s private practice use and ambulance, field, or emergency room applications. One of the main advantages of ultrasound imaging is that it is relatively inexpensive compared to other imaging modality such as magnetic resonance imaging (MRI), computed tomography (CT), or positron emission tomography (PET) imaging (Fig. 11.1) [5]. Diagnostic ultrasonography today is used in many medical fields as the main advantage is that certain structures can be observed without using radiation and it is relatively inexpensive with easy portability. Ultrasound produces real-time images of soft tissue and can capture structure of internal organs in the abdomen, pelvis, and thoracic cavity. For example, in obstetrics and gynecology, ultrasound is the first-line imaging modality for assessment of the female pelvis, for characterization of adnexal and uterine lesions, for diagnosing pregnancy, and monitoring fetal development. In cardiovascular imaging, ultrasound is used, for example, to evaluate the presence of blood flow by Doppler ultrasound through the major arteries and veins, to reveal stenosed or occluded arteries or veins, and to identify aneurysms and thrombus formations. Other imaging indications for ultrasound include assessment of size, shape, location, and diseases of solid organs in the abdomen including diseases of the liver, gallbladder, spleen, pancreas, kidneys, and pancreas. In addition, ultrasound is the first-line imaging modality for evaluation of the thyroid. In musculoskeletal radiology, ultrasound is used for assessment of muscles and ligaments and allows for accurate diagnosis of ligamentous tears in the shoulder. There is also an established use of ultrasound as a rapid imaging tool for diagnosis in the emergency room including the fast diagnosis of free intraperitoneal fluid from solid organ injuries [6].
FIGURE 11.1 Summary of advantages and disadvantages of some of the most common imaging modalities that are used for preclinical molecular imaging research. All of the imaging modalities have the potential for clinical translation. Compared with other imaging modalities, ultrasound is widely available in the imaging community throughout the world, is portable, and is relatively cheap compared with other imaging modalities. Due to its advantages, there is an increasing interest in evaluating ultrasound as a molecular imaging tool. (Reproduced with permission from Willmann and van Bruggen [5].)
P1: OTA/XYZ P2: ABC c11 JWBS037-Chen February 2, 2011
21:47
Printer Name: Yet to Come
CONTRAST-ENHANCED ULTRASOUND IMAGING
237
11.2 CONTRAST-ENHANCED ULTRASOUND IMAGING Contrast-enhanced ultrasound imaging can be defined as the application of ultrasound contrast agents to traditional medical ultrasonic imaging. Ultrasound waves travel through tissue or media under the influence of sound pressure and because molecules of tissue are bound elastically to one another, the excess pressure results in a wave propagating through the tissue. Thus acoustic impedance is important in the determination of acoustic transmission and reflection at the boundary of two materials having different acoustic impedances. The acoustic impedance (Z) of a material is defined as the product of its density (p) and acoustic velocity (V), given by the equation, Z = p × V. Ultrasound contrast agents are designed on the principle of acoustic impedance. Contrast agents usually consist of a gaseous core enclosed by an outer shell for stability. These agents act as echo enhancers when circulating in blood flow due to the high difference in acoustic impedance at the interface between gas and blood. The enhanced echo intensity is proportional to the change in acoustic impedance as the sound beam crosses from the blood to the gas. The contrast agent being gas filled is much more compressible than surrounding aqueous biological fluid; therefore ultrasound pressure waves in the megahertz range cause compression and expansion of contrast agents resulting in effective ultrasound backscattering [7]. These scattered ultrasound waves can then be detected by the ultrasound transducer. The contrast agents are administered intravenously into the systemic circulation and remain within the vasculature due to their size of several microns (1–4 m) with no leakage and hence have access to the microvascular level. Several types of contrast-enhanced agents have been developed; the properties and application of each of these contrasts are discussed in the Section 11.3. Contrast-enhanced ultrasound can be classified into two general approaches: (1) contrastenhanced ultrasound imaging using nontargeted contrast agents, which are already being used in clinical practice, and (2) contrast-enhanced ultrasound imaging using molecularly targeted ultrasound contrast agents that are currently still restricted to preclinical research.
11.2.1 Nontargeted Contrast-Enhanced Ultrasound Imaging Nontargeted contrast-enhanced ultrasound imaging makes use of contrast agents, such as microbubbles, consisting of a gaseous core stabilized by a biocompatible outer shell and injected intravenously into the systemic circulation in a small bolus. Microbubbles once injected remain in circulation for a period of time, during which ultrasound waves are directed at the area of interest. When microbubbles in the blood flow past the area, which sees an ultrasound wave, they reflect a unique echo that stands in stark contrast to the surrounding tissue due to the orders of magnitude mismatch between microbubble and tissue echogenicity. The ultrasound system converts the strong echogenicity into a contrastenhanced image of the area of interest. In this way, the bloodstream’s echo is enhanced, thus allowing distinguishing blood from surrounding tissues. There are several applications of nontargeted contrast agents, which are listed below. 1. Edge Delineation in the Heart. Accurate detection of differences in tissue structures is crucial in ultrasound imaging of the heart, where a thinning, thickening, or irregularity in the heart wall indicates a serious heart condition that requires either monitoring or treatment. Microbubbles enhance the contrast at the interface between the heart
P1: OTA/XYZ P2: ABC c11 JWBS037-Chen February 2, 2011
238
21:47
Printer Name: Yet to Come
MICROPARTICLE- AND NANOPARTICLE-BASED CONTRAST-ENHANCED ULTRASOUND IMAGING
tissue and the circulating blood and allow better delineation of morphological cardiac pathologies. 2. Assessment of Blood Volume and Perfusion. Contrast-enhanced ultrasound holds the promise for evaluating the degree of blood perfusion and blood volume in an organ or area of interest. The body consists of 90% water and is therefore acoustically relatively homogeneous. Blood and surrounding tissues have similar echogenicities; thus it is difficult to clearly discern the degree of blood flow through an organ (perfusion), or the interface between the tissue and blood using traditional ultrasound [8]. Color Doppler ultrasound that measures velocity of every voxel (3D equivalent of pixel) in a given region is conventionally used for detection of blood flow. The use of ultrasound contrast agents enhances Doppler flow signals by providing more and better acoustic scatterering. This results in improved detection of blood flow from vessels by providing high-resolution real-time assessment of flow without color Doppler artifacts. 3. Diagnosing Liver Diseases. Ultrasound is a widely used modality for imaging liver pathologies. By using nondestructive low-acoustic-power ultrasound with perfluorocarbon or sulfur hexafluoride-filled microbubbles, contrast-enhanced ultrasound has allowed improved detection and characterization of focal liver lesions. By analyzing the enhancement patterns of different focal liver lesions following intravenous administration of contrast agents, focal liver lesions can often be characterized by ultrasound without the need for further follow-up procedures such as biopsy. For example, malignant liver lesions such as hepatocellular carcinoma (HCC) are usually hypoechoic compared to normal liver parenchyma on delayed phase images (acquired several minutes after contrast administration), whereas benign lesions such as hemangiomas remain iso- or hyperechoic without washout of contrast on delayed phase images [9]. 11.2.2 Targeted Contrast-Enhanced Ultrasound Imaging (Molecular Ultrasound Imaging) The latest ultrasound approach combines ultrasound imaging technology with actively targeted ultrasound contrast agents; that is, agents to which one or more biologically active molecules have been attached such that the agents localize in a specific area of interest. Detection of bound targeted contrast agents can then show the area of interest expressing that particular molecule, which can be indicative of a certain disease state, or identification of a particular cell type. Using targeted agents, clinical and experimental roles of ultrasound can even further be expanded into applications including noninvasive detection of pathology using disease-associated molecular signatures, detection of gene expression, investigation of drug localization, and delineation of molecular mechanisms of disease. Targeted contrastenhanced ultrasound imaging is being developed for a variety of medical applications and is detailed in subsequent parts of this chapter. 11.2.3 Disadvantages of Contrast-Enhanced Ultrasound Imaging The introduction of contrast agents can provide nuclei for the generation of acoustic cavitation [10]. Cavitation is any activity of highly compressible transient or stable bubbles of gas or vapor, generated by ultrasonic power in the propagation medium. One of the side
P1: OTA/XYZ P2: ABC c11 JWBS037-Chen February 2, 2011
21:47
Printer Name: Yet to Come
ULTRASOUND CONTRAST AGENTS
239
effects of ultrasound energy is the production of heat and increasing heat is associated with the risk of cavitation. However, the threshold for cavitation is high and does not occur at current levels of diagnostic ultrasound. Cavitation can be described as inertial or noninertial. Inertial cavitation, formerly called “transient” cavitation, occurs when the acoustic pressure amplitude is sufficiently high and above a threshold level. Under this condition the microbubble contrast agents will first grow in volume and then implode violently. If the core of a microbubble is gas with high specific heat ratio (constant pressure/constant volume), high temperature may result during implosions and highly reactive free radicals may be generated. Thus inertial cavitation has the most potential to damage tissue and cause local microvasculature ruptures and hemolysis [11]. Ultrasound produces more heat as the frequency increases. Therefore the ultrasonic frequency must be carefully monitored. Microbubbles burst at low ultrasound frequencies and at high mechanical indices (MIs—defined as the measure of the acoustic power output of the ultrasound imaging system). Increasing MI increases image quality, but there are trade-offs with microbubble destruction and thus inertial cavitation. Noninertial cavitation is the process in which microbubbles in a liquid are forced to oscillate asymmetrically in the presence of an acoustic field, when the intensity of the acoustic field is insufficient to cause total bubble collapse. Asymmetric contraction of an inertial cavitation bubble generates a microjet-like small stream that has a localized but unfavorable mechanical effect, such as membrane damage of a cell [12]. Targeted contrast agents to date have been restricted to preclinical trials due to the binding chemistry used so far for targeting microbubbles. Introduction of currently used targeting ligands can be immunogenic in human subjects as the binding ligands such as monoclonal antibodies are often derived from animal culture and the often used coupling strategy with biotin–streptavidin can cause allergic and immunogenic reactions [13]. However, humanized and chimeric antibodies or other small peptides can be used as an alternative source of ligands to functionalize microbubbles for clinical translation, such as synthetically manufactured targeting peptides that perform the same function, but without being immunogenic [14]. In addition, couple strategies other than biotin–streptavidin interactions are currently explored to further help moving targeted ultrasound contrast agents into the clinic [15, 16].
11.3 ULTRASOUND CONTRAST AGENTS 11.3.1 Background Blood is two to three orders of magnitude less echogenic than tissue due to the relatively small acoustic impedance difference between red blood cells and plasma. This low bloodto-issue signal ratio hinders the detection of small blood vessels. Ultrasound contrast agents, typically between 0.1 and 5 m in diameter, increase the echo amplitude from the blood pool and allow the detection of vessels even as small as capillaries. The earliest introduction of contrast agents was by Gramiak and Shah [17], when they injected agitated saline into the ascending aorta and cardiac chambers during echocardiography to opacify the left heart chamber. Strong echoes were produced due to the acoustic mismatch between the air bubbles and surrounding blood. Today the contrast agents are a long way away from being produced by simple agitation, which causes the air bubbles to quickly dissipate and dissolve in the blood. The basic principle of contrast agents is using gas bubbles,
P1: OTA/XYZ P2: ABC c11 JWBS037-Chen February 2, 2011
240
21:47
Printer Name: Yet to Come
MICROPARTICLE- AND NANOPARTICLE-BASED CONTRAST-ENHANCED ULTRASOUND IMAGING
which are encapsulated with a shell for stabilization. The ideal qualities in a contrast agent are high echogenicity, low attenuation, low blood solubility, low diffusivity, and lack of biological side effect. Basic principles for fabrication of various types of contrast agents are summarized in later sections. The commercial development of contrast agents began in the 1980 with a great effort on stabilization of the small microbubbles. Now, ultrasound contrast media sales have been growing at a compound rate of over 30% per year since 2004, with sales of $77.1 million in 2006 [18]. However, FDA approved indications are still limited. In addition, currently there is no FDA approved targeted contrast agent available. 11.3.2 Types of Ultrasound Contrast Agents Ultrasound contrast agents are biocolloids (colloidal particles made from biocompatible materials) and can be divided into two main classes: (1) microbubble based contrast agents, and (2) non-microbubble based contrast agents. Microbubbles are gas–liquid emulsions whereas non microbubble based contrast agents include liquid–liquid emulsions (nanodrops), solid nanoparticles, and liposomes [19]. Contrast agents measured for performance are based on factors such as minimal toxicity in vivo, ease of administration, and contrast to tissue ratio (CTR), which is the ratio between the acoustic backscatter coming from the contrast agent and tissue. The degree of acoustic backscatter depends on the intrinsic properties of the contrast agent, such as the compressibility of the contrast agent, the density difference between the contrast agent and surrounding tissue, and nonlinear effects (see Section 11.1.1) and resonance. The latest ultrasound technique is the development of targeted ultrasound contrast agents. Targeted ultrasound contrast agents are specially modified such that they recognize specific molecular markers and localize in a specific area of interest. Targeted contrast agents provide a powerful tool to evaluate and treat tissue in disease states in whole animal models and patients. The idea that contrast agents can be coupled to ligands for specific proteins and imaged to report molecular targets was developed in the late 1990s, therefore giving birth to the principle of molecular ultrasound. Targeted contrast agents can be generated by either passive or active targeting. Passive targeting is a nonspecific accumulation of contrast agents at the target site. The desired mechanism of passive targeting can be specified by manipulating the properties of the contrast agent, such as size, chemical properties such as type of encapsulated gas, electrical charge of the shell, the route of injection, and physiological processes of the immune system. Lindner et al. [20] were able to show during inflammatory processes that activated leukocytes engulf and absorb microbubbles, which are surrounded with an albumin or a lipid shell. The adhesion of lipid or albuminencapsulated microbubbles to activated leukocytes is due to binding to 2-integrin, Mac 1, and C3, all of which are molecular markers of inflammatory process. Fischer Scientific (Hanover Park, IL) [21] manufactured lipid encapsulated microbubbles with negative and neutral surface charges. In a subsequent ultrasonic investigation in the myocardium of dogs, a myocardial contrast could only be achieved with the negatively charged microbubbles 10 min after intravenous injection. On the other hand, microbubbles that have no charge did not show any contrast enhancement in this case. Oussoren et al. [22] examined the influence of the size of liposomes with respect to their transport from the injection site into the lymphatic vessels and their enhancement in regional lymph nodes (popliteal lymph nodes and iliac lymph nodes) in rats during the period of 52 h after subcutaneous injection. Accordingly, they observed that after 52 h only 26% of the liposomes with a size of 40 nm
P1: OTA/XYZ P2: ABC c11 JWBS037-Chen February 2, 2011
21:47
Printer Name: Yet to Come
ULTRASOUND CONTRAST AGENTS
241
were present; however, there were still 95% of the liposomes with a size of 400 nm in the area of the injection site. In contrast to this, the difference in enhancement was not as clear in the regional lymph nodes, which indicates that larger liposomes become more enhanced in the lymph node than do smaller ones. Thus these experiments suggest that liposomes can be passively targeted to an area of interest based on size and route of injection. Active targeting requires modification of the contrast agent shell to allow selective binding to cellular epitopes or other receptors of interest. The attachment of targeting ligands can either be integrated into the shell during the production process or after manufacturing of the contrast agent. The ligands can be attached through covalent or noncovalent procedures [11, 13, 23]. Furthermore, the attachment of ligands can either be through direct coupling to shell or indirect coupling wherein a linker is used to connect the contrast agent shell and the ligand. This is necessary in cases where there is no direct coupling between contrast agent shell and ligand possible or when the ligand does not extend far enough from the encasement into the outside environment after direct coupling and thus is not in a position to connect to the molecular target structures. In this case, the linker works as a spacer to the contrast agent surface. These ligands can be attached to the contrast agent shell using several binding approaches such as carbodiimide, maleimide, or biotin–streptavidin coupling [11]. Biotin–streptavidin is the most popular coupling strategy because biotin’s affinity for streptavidin is very strong and it is easy to label the ligands with biotin. Currently, in preclinical studies these ligands are often monoclonal antibodies produced from animal cell cultures that bind specifically to receptors and molecules expressed by the target cell type. Peptides, which have the ability to recognize cell epitopes, can also be directly incorporated in the microbubble shell or be conjugated onto the microbubble shell [14, 15]. Visualization and quantification of various molecular molecules by ultrasound contrast agents has been shown for several pathological processes such as angiogenesis, inflammation, or thrombus formation, which will be explained in more detailed in Section 11.3.3.
Microbubbles Microbubbles are a gas–liquid emulsion with sizes ranging from 0.1 to 5 m in diameter. The microbubble signal is dependent on the compressibility and density of the gas core, the viscosity and density of the surrounding medium, the frequency and power of ultrasound applied, and the bubble size. A microbubble is an ideal ultrasound contrast agent due to its excellent ultrasound response. A microbubble undergoes oscillation due to its gas core, which allows it to shrink and expand with the passage of an acoustic wave, resulting in the generation of strong acoustic signals that greatly exceed conventional ultrasound backscatter produced by reflection in acoustic impedance. The natural reaction time of a microbubble to a rapid pressure variation is on the order of microseconds; thus they resonate at frequencies typically used in ultrasound imaging, and resonance can be exploited to generate a very strong echo. Due to this high degree of backscatter in comparison to plasma and blood cells, an ultrasound system is capable of detecting the signature from a single microbubble in an in vitro setting. Klibanov et al. [24] imaged very dilute microbubble dispersion in plastic bags as well as attached microbubbles on a Petri dish. On imaging the plastic bags and Petri dishes, distinct white foci were observed. Concentration at the site of imaging was consistent with signals from individual microbubbles as determined by a Coulter counter. These results imply that the ultrasound system can resolve backscatter signals from individual microbubbles both in solution as well as attached to a surface if the surrounding background signal is kept to near zero.
P1: OTA/XYZ P2: ABC c11 JWBS037-Chen February 2, 2011
242
21:47
Printer Name: Yet to Come
MICROPARTICLE- AND NANOPARTICLE-BASED CONTRAST-ENHANCED ULTRASOUND IMAGING
Types of Microbubbles Microbubble particles have a protective shell that prevents gas loss and particle fusion. There are a variety of microbubble contrast agents depending on their shell makeup, gas core makeup, and whether or not they are targeted. Microbubble shell material determines how easily the microbubble is taken up by the immune system. A more hydrophilic material tends to be taken up more easily, which reduces the microbubble residence time in the circulation. This reduces the time available for contrast imaging. The shell material also affects microbubble mechanical elasticity. The more elastic the material, the more acoustic energy it can withstand before bursting [25]. Currently, microbubble shells are composed of albumin, galactose, lipid, or polymers [8]. Ultrasound contrast agents can also be classified according to the rigidity of their shell [26]. Soft-shell microbubbles either have no coating or are covered by a thin monolayer of surfactant molecules such as palmitic acid or phospholipids and are very sensitive to pressure changes. Hard-shell microbubbles have a more rigid shell made of polymers or denatured albumin, which can increase their stability to a certain degree. Introduction of a grafted arm of polyethylene glycol (PEG) polymer onto the microbubble shell provides additional stearic protection, prevents microbubble aggregation, and makes the microbubble more nonreactive. It temporarily “hides” the microbubble from immune system uptake, increasing the amount of circulation time and hence imaging time (Fig. 11.2) [11]. The microbubble gas core is the most important part of the ultrasound contrast microbubble because it determines the echogenicity. Gas cores can be composed of air or high molecular weight or heavy gases like perfluorocarbon, sulfur hexafluoride, or nitrogen [8]. Heavy gases are less water soluble so they are less likely to leak out from the microbubble to impair echogenicity and thus prolonging their effective life in circulation [25]. Various methods have been employed to fabricate microbubbles for ultrasound imaging. The most popular method is dispersion of a gas through an aqueous phase by mechanical agitation of the gas–liquid interface. Dispersion is performed with the aid of a sonicator (sonication) or by high-shear mixing (amalgamation). These techniques rely on stochastic events that produce a polydispersed size distribution, generally ranging between submicrometer to tens of micrometers in diameter. Size fractionation techniques can be employed, which are based on buoyancy. Several methods have been described to encapsulate gas in a polymer shell. Dispersion and ionic gelation have been used to create alignate-shelled microbubbles. Organic solvents have been used to dissolve and disperse the polymer, which is then resuspended to form hollow polymer capsules. Polymerization at the air–liquid interface during agitation of an acidic medium was also used. Each of these methods has produced microbubbles with enhanced stability [19]. Lastly, there is a category of targeted microbubbles which retain the same general features as untargeted microbubbles, but they are outfitted with ligands that bind specific receptors expressed by cell types of interest, such as inflamed cells or angiogenic endothelial cells. This is explained in more detail in Section 11.3.3. Once in the blood flow, bound targeted contrast agents need to be differentiated from freefloating agents. This can be achieved if the targeted agent has a long lifetime at the target site, while nonadherent agents have a short circulating half-life. Another approach to differentiate bound from nonbound microbubbles is through the destruction and replenishment method described by Chomas [27] and Pollard [28]. After intravenous injection of the targeted contrast agent, ultrasound imaging is paused for several minutes to allow retention of contrast agents, following which a series of image frames are captured to obtain a signal from the tissue from adherent as well as freely circulating contrast agents. A single or a series of high-power destructive pulses are then given to destroy all contrast agents within
P1: OTA/XYZ P2: ABC c11 JWBS037-Chen February 2, 2011
21:47
Printer Name: Yet to Come
ULTRASOUND CONTRAST AGENTS
Targeted shell Streptavidin Antibody Blotin
Shell
Microbubble contrast agent
Hydrophobic interactions
Peptide
243
Lipid or
PEG Arm
Disulfide bonds
Protein e.g. Albumin
Gas Gas core Perfluorocarbon Sulphur hexafluoride Nitrogen Air
Steric stabilizer Shell Biomarkers Blood vessel wall
Endothelial cell
Ultrasound wave Transducer
FIGURE 11.2 Schematic representation of a microbubble in a blood vessel. A microbubble consists of a gas core enclosed by a shell. Various types of gas and shell compositions can be used. The gas core can consist of perfluorocarbon, sulfur hexafluoride, nitrogen, or air. The shell can be made of various substances such as phospholipid molecules held together with hydrophobic interactions or a protein like albumin with disulfide bonds to hold the molecules together. The shell can be modified to create a targeted microbubble by incorporating peptides that recognize specific ligands of interest or antibodies linked either directly onto the shell or indirectly through streptavidin–biotin moieties. Adding a polyethylene glycol (PEG) arm onto the shell provides stability, prevents microbubbles from aggregation, and helps prevent uptake by the immune system.
the beam elevation. After a brief period to allow freely circulating microbubbles to replenish the tissue of interest, a series of image frames are again captured. The signals before and after destruction are averaged and digitally subtracted from the initial predestruction frames to derive the signal attributable on adherent contrast agent only (Fig. 11.3).
Biodistribution of Microbubbles Biodistribution studies revealed that microbubbles have low circulation residence times as they are rapidly removed by the reticuloendothelial system (RES), which is part of the immune system, consisting of the phagocytic cells located in reticular connective tissue, primarily monocytes and macrophages in the liver (called Kupffer cells), spleen, and bone marrow. Walday et al. [30] studied the organ distribution of nontargeted air-filled iodine-125-labeled albumin microspheres (Albunex) in rats and pigs up to 90 min after intravenous administration. Three minutes after administration in rats, about 60% of the radiolabeled albumin microspheres were recovered in the liver, 9% of them were recovered in the spleen, and 5% of them were recovered in the lungs, with only small amounts in other organs such as the brain, heart, and kidneys. In contrast, in pigs, more than 90% of the albumin microspheres were recovered in the lungs, most likely because of the presence of intravascular macrophages in the porcine lungs. Biodistribution
P1: OTA/XYZ P2: ABC c11 JWBS037-Chen February 2, 2011
Printer Name: Yet to Come
MICROPARTICLE- AND NANOPARTICLE-BASED CONTRAST-ENHANCED ULTRASOUND IMAGING
(a)
(b)
(c) 40 35 30
Video Intensity
244
21:47
25 20 15 10 5 0 0
50
100
150 200 Frames
250
300
350
P1: OTA/XYZ P2: ABC c11 JWBS037-Chen February 2, 2011
21:47
Printer Name: Yet to Come
ULTRASOUND CONTRAST AGENTS
245
on air-filled iodine-123-labeled human serum albumin microspheres (Quantison) was studied in humans for 58 h by using radionuclide imaging by Perkins et al. [31]. One hour after intravenous administration, the greatest amount of radiolabeled microsphere uptake was in the liver (41.8%), followed by the spleen (11%) and the lungs (3.4%). Small amounts of radiotracer were also visible in bone and heart tissue in that study. Willmann et al. [32] performed in vivo analysis in living mice for lipid shell perfluorocabon-filled microbubbles targeted to VEGFR2 via anti-VEGFR2 antibodies by using dynamic micropositron emission tomography (micro-PET). Anti-VEGFR2 antibodies were radiolabeled by conjugating radiofluorination agent N-succinimidyl-4-[18 F]fluorobenzoate to the antibodies as a tracer for in vivo assessment of targeted microbubble biodistribution using micro-PET imaging [32]. In that study it was shown that 50% of targeted microbubbles cleared from the blood pool after ∼3.5 min and ∼95% were cleared from the blood circulation after 30 min. These findings also agree with results from repeated in vivo targeted contrast-enhanced ultrasound imaging, which have shown that the US imaging signal in blood vessels returns to almost background levels several minutes after administration of targeted microbubbles [33, 34]. Palmowski et al. [35] studied the biodistribution of hard-shelled microbubbles. The microbubbles consisted of cyanoacrylate and were targeted to VEGFR2 using monoclonal antibodies and streptavidin–biotin interactions. In that study it was observed that 1 min after injection of these microbubbles, 90% of the microbubbles were cleared from the blood and pooled in the lungs, liver, and spleen. VEGFR2 targeted microbubbles also accumulated significantly within the tumor vasculature. A biphased distribution of the microbubbles was observed in the blood. After injection of microbubbles, an initial increase was first seen in the blood, followed by a high concentration in the lungs, which then decreased during the ← FIGURE 11.3 (a) Schematic representation of targeted microbubbles (blue) attached to ligands in tumor vasculature of a solid tumor after intravenous administration. Tumor vasculature consists of vessels with neoplastic endothelial cells (orange). Here, both the tumor cells (gray) and the endothelial cells express an angiogenesis marker such as vascular endothelial growth factor receptor type 2 (VEGFR2). The microbubbles remain predominantly in the vasculature due to their size (several micrometers) and thus adhere only to the endothelial cell and not tumor cells. Not all microbubbles adhere; some are floating freely in the circulation system. After a high-power destructive pulse, adherent microbubbles are destroyed and freely circulating microbubbles replenish from outside the imaging plane after several seconds. (Reproduced with permission from Willmann et al. [29].) (b) This graph represents the method to calculate the signal from targeted microbubbles attached to vascular markers on angiogenic tumor vessels using the destruction–replenishment method. After intravenous administration, microbubbles float into the tumor and start binding to molecular markers to which they are targeted such as VEGFR2. In this example, a 4-min interval is allowed to pass for microbubbles to attach to the vascular target. The measured ultrasound signal consists of the background signal from the tissue, and signal from still freely floating microbubbles and microbubbles attached to the vascular targets. All microbubbles within the beam elevation are then destroyed by an external destructive pulse. The difference in imaging signal before and several seconds after destruction (to allow freely floating microbubbles to recirculate into the imaging plane) corresponds to the imaging signal from microbubbles attached to the vascular markers. (Reproduced with permission from Willmann et al. [29].) (c) Quantitative representation of molecular ultrasound imaging experiment using the destruction–replenishment method. The measured video intensity from the ultrasound image is plotted against the number of imaging frames. The difference between the predestruction and postdestruction imaging signal is automatically calculated and corresponds to the imaging signal from attached microbubbles.
P1: OTA/XYZ P2: ABC c11 JWBS037-Chen February 2, 2011
246
21:47
Printer Name: Yet to Come
MICROPARTICLE- AND NANOPARTICLE-BASED CONTRAST-ENHANCED ULTRASOUND IMAGING
subsequent 2 h. The decreasing microbubble concentration in the lungs was accompanied by a decent recirculation of microbubbles in the blood. These studies show that, after the first pass, most nontargeted and targeted microbubbles are cleared from the blood circulation and accumulate in the RES. However, due to variable characteristics such as size or shell thickness within the population of injected microbubbles, the microbubbles may contribute to varying levels of stability.
Non-Microbubble Based Contrast Agents Non-microbubble based contrast agents are submicron and nanosized contrast agents that are attracting considerable interest as contrast agents for molecular ultrasound imaging. Their small size makes them also suited to targeting cells outside the capillary vasculature, such as cancer cells [36, 37]. Nonmicrobubble based contrast agents can be defined as particles ranging in size between 10 and 1000 nm. Their unique size range promotes long circulation half-lives and high tissue extravasation rates (the process of exuding or passing out of a vessel into surrounding tissues). Larger particulates greater than 1 m in diameter are quickly recognized and cleared by the body’s reticuloendothelial system (RES) after intravenous injection [38]. Individual nanoparticles when circulating in the bloodstream cannot be detected with ultrasound due to their poor echogenicity. This is due to the inability of nanoparticles to oscillate significantly in an ultrasound field, as their interior core, being mostly liquid, is incompressible. But as the nanoparticles accumulate and aggregate on a surface, the coating on the surface changes the acoustic impedance difference between blood and the surface, making the surface acoustically more reflective and detectable with ultrasound. At diagnostic frequency, nanoparticles are less echogenic than microbubbles but are more suitable for high-frequency imaging approaches used for intravascular imaging, providing near microscopic resolution [38]. Nanoparticles of particular relevance according to their submicron size are perfluorocarbon emulsion (PFC) nanodroplets, polylactide acid (PLA) nanobubbles, solid nanoparticles, and echogenic liposomes (Fig. 11.4) that are discussed in the following sections. Perfluorocarbon Emulsion Nanodroplets Perfluorocarbon emulsion (PFC) nanodroplets are liquid–liquid emulsions consisting of a PFC liquid core encapsulated by a phospholipid monolayer with typical diameters of 200–400 nm. The liquid core, which is 98% by volume of the nanodroplets, can have any one of the numbers of PFCs—namely, perflurooctylbromide (PFOB), perfluoro-15-crown-5 ether (CE), and perfluorodichlorooctane, with PFOB the most commonly used. PFC nanodroplets are stabilized for use in vivo with surfactants, the most common being phospholipids, which restrict the particles to coalesce with one another. PFC nanodroplets are formed by mechanical diminution of the fluorocarbon liquid into an aqueous phase. The dispersed phase consisting of liquid fluorocarbon is immiscible with the continuous aqueous phase and is highly hydrophobic; thus PFC nanodroplets exhibit robust stability against handling, pressure, heat, and shear. The size of these particles can be controlled by extrusion. Individual PFC nanodroplets are less sensitive contrast agents due to poor echogencity of its perfluorocarbon core. However, when collective deposition occurs on the surfaces of tissues or a cell in a layering effect, these particles create a local acoustic impedance mismatch that produces a strong ultrasound signal without a concomitant increase in the background level [39]. Thus PFC nanodroplets are less liable to nonspecific signal enhancement as larger binding events take place for a detectable signal. The nongaseous nature of PFC nanodroplets has advantages for ultrasound imaging as their liquid composition
P1: OTA/XYZ P2: ABC c11 JWBS037-Chen February 2, 2011
21:47
Printer Name: Yet to Come
ULTRASOUND CONTRAST AGENTS
Phospholipid shell
247
Polylactic acid shell
Liquid core: Perfluoro octylbromidePerfluoro-15-crown-5-ether PLA nanobubble 40-200 nm Perfluorodichlorooctane
Gas core
PFC nanoparticle 200–400 nm Air pocket Solid particle Aqueous core Phospholipid bilayer
Gas entrapped in pockets Solid nanoparticles 100 nm
Liposome 20 nm to 10 µm
FIGURE 11.4 Schematic representation of various kinds of non-microbubble based contrast agents. Perfluorocarbon emulsion nanodroplets (PFC nanodroplets) are liquid–liquid emulsions, polylactide acid (PLA) nanobubbles are gas–liquid emulsions, liposomes consists of a phospholipid bilayer enclosing an air pocket, and solid nanoparticles are solid amorphous substances that have gas or air entrapped in their pores or fissures.
makes the nanodroplets resistant to pressure or mechanical stress, they are not easily deformed, and they do not cavitate during the imaging process [40]. Another advantage of PFC nanodroplets includes the fact that, based on the properties of perfluorocarbon in their liquid cores, they can be vaporized into gas bubbles visible with standard ultrasound imaging [41]. When low boiling point perfluorocarbons such as dodecaperfluoropentane are used in nanodroplets, they spontaneously vaporize into gas pockets caused by the temperature increase due to a body temperature of ∼37 ◦ C in a living subject. The challenge for these low boiling point nanodroplets is for enough contrast agents to have accumulated at the desired target site before spontaneous vaporization occurs. In order to decrease spontaneous vaporization, high-boiling point nanodroplets using liquid perfluorocarbons such as perflurohexyl bromide have been synthesized to minimize spontaneous vaporization at physiological body temperature. However, with these types of high boiling point nanodroplets, higher energies have to be applied to transfer them into visible gas pockets. To overcome this trade-off, Amirriazi et al. showed in vitro that nanodroplets filled with high boiling point perfluorcarbons can be modified by adding iron oxide nanoparticles as nucleation sites inside the nanodroplets, thereby substantially lowering the boiling point of nanodroplets (high enough to minimize spontaneous vaporization and low enough to decrease the required energy for vaporization) [42]. Future studies are warranted to assess in vivo targeting efficiency of these contrast agents. biodistribution of perfluorocarbon emulsion nanodroplets The circulatory halflife period of PFC nanoparticles is in excess of 1 h and the tissue half-life residency is many
P1: OTA/XYZ P2: ABC c11 JWBS037-Chen February 2, 2011
248
21:47
Printer Name: Yet to Come
MICROPARTICLE- AND NANOPARTICLE-BASED CONTRAST-ENHANCED ULTRASOUND IMAGING
days. Studies done to visualize PFC nanoparticles, which were used as a substitute to create artificial blood in the vascular system of rats using MRI, showed retention of PFC for several days. Images obtained at intervals of approximately 2 h, 2 days, 2 weeks, and 2 months after infusion of PFC nanoparticles showed identification of PFC in the heart, lung, liver, spleen, and large vessels both in vivo and postmortem [43]. Furthermore, it was shown that the retention of PFC nanoparticles in the circulation was prolonged more than 300% when coated with lecithin [44]. This prolonged retention in the circulation resulted in a marked decrease in the amount of perfluorochemicals retained in the liver and spleen.
Polylactic Acid (PLA) Nanobubbles Nanobubbles are gas–liquid emulsions enclosed with a biodegradable polymer, polylactic acid (PLA), ranging from 40 to 200 nm in size. Use of targeted nanobubbles was tested by Liu et al. [45]. The surface of PLA nanobubbles was conjugated to an anti-Her2 antibody (i.e., Herceptin) and the nanobubbles were tested in vitro for specific binding to breast cancer cells that overexpress Her2 receptors. In high-resolution ultrasound B-mode images, the average gray scale of the Her2-positive cells was consistently and significantly higher after nanobubble treatment than negative controls in vitro. Rapoport et al. [46] describe a novel method of combining ultrasound imaging of cancer with ultrasound enhanced drug delivery treatment. Nanobubbles filled with the chemotherapy drug doxorubicin were injected into mice. The bubbles accumulated in the tumors, where they combined to form larger “microbubbles.” When exposed to ultrasound, the bubbles generated echoes, which made it possible to image the tumor. Furthermore, a destructive ultrasound wave destroying the bubbles caused release of the drug that suspended tumor growth [46]. At the time of this writing, biodistribution studies of nanobubbles have not yet been published. Solid Nanoparticles Solid nanosized particles, which include amorphous solid particles like silica or iron oxide particles that contain gas pockets in their pores and fissures, have been discovered to produce detectable backscatter for ultrasound imaging. One of the advantages of these solid nanosized particles is the high tissue extravasation rate due to their unique size range and therefore their potential use to target molecular markers outside the vasculature. In addition, these solid nanoparticles remain stable during insonication; thus they can also generate good tissue enhancement, owing to the acoustic-impedance mismatch between solids and soft tissue compared to that between liquids and soft tissue. Liu et al. [47] tested silica nanospheres, 100 nm in diameter, in a saline suspension and intravenously injected in mice for imaging the liver with a transducer frequency of 30 MHz. It was shown that the mean acoustic intensity increases in the liver following particle administration. Also, these silica nanospheres were tested in vitro on tissue phantoms consisting of particles dispersed in agarose gel. Intensity was seen to increase with increasing particle size and concentration. Targeted solid nanoparticles are yet to be designed and could be the next step since nanoparticles can be utilized effectively as targeted contrast agents because they can be efficiently surface modified with biorecognition molecules given their relatively large surface area. Surface modification could result in nanoparticles escaping the RES, leading to prolonged half-lives in circulation and an increased number of passes through the vasculature to achieve satisfactory targeting. In addition, tumor vasculature exhibits a permeability and retention effect that results in exaggerated extravasation and retention of particles that are smaller than the pore size of tumor endothelia [48], making these solid nanoparticles very
P1: OTA/XYZ P2: ABC c11 JWBS037-Chen February 2, 2011
21:47
Printer Name: Yet to Come
ULTRASOUND CONTRAST AGENTS
249
useful for cancer imaging. It was shown that superparamagnetic iron oxide nanoparticles, another solid nanoparticle agent, provided unexpected enhancement of the delineation of ex vivo rat brain tumors [49] and central nervous system lesions [50] in sonograms. These studies suggest that the potential for using nanoparticles as molecularly targeted ultrasound imaging agents should be explored. Toxicity and biocompatibility issues will also need to be considered if nanoparticles are to be considered for clinical imaging applications.
Liposomes The discovery of liposomes or lipid vesicles emerged from self-forming enclosed lipid bilayers with an aqueous core upon hydration. Liposomes can be unilamellar or multilamellar, and their size can be controlled by sonication and extrusion (forcing the liposomes through microscale pores) to range from ∼20 nm to >10 m in diameter. A semipermeable, hydrophobic membrane separates the interior and exterior compartments of a liposome. Echogenic liposomes contain air pockets within the lipid bilayer to generate acoustic reflectivity. The mechanism of echo contrast appears to be the backscatter from entrapped pockets of air within the liposomes that form during rehydration of freeze-dried liposomes. The size, charge, and surface properties of liposomes can easily be changed by adding ingredients to the lipid mixture before liposome preparation and/or by varying the preparation methods. The ability of liposomes to easily entrap different substances into both the aqueous phase and the liposome membrane compartments makes them suitable for manipulations to generate targeted contrast agents. Echogenic liposomes can be prepared by incorporating gases such as air, argon, or nitrogen into the liposome or by inducing the formation of gas bubbles directly inside the liposome as a result of a chemical reaction [51]. Liposomes composed of phospholipids are not echogenic, but when made with phosphatidylcholine (PC), phosphatidylethanolamine (PE), phosphatidylglycerol (PG), and cholesterol [52] they are highly acoustically reflective. These liposomes improved intravascular contrast by approximately 300% relative to blood and 150% relative to agitated saline when injected in swine [53]. Antibodies can be conjugated on these lipid shells without hindering the echogenecity. biodistribution of liposomes Phospholipid liposomes introduced into the circulation system are very rapidly sequestered by the cells of the reticuloendothelial system (RES) [54]. The clearance half-time of the liposome is usually within 30 min. A large number of liposomes need to accumulate at the target tissue to render it acoustically reflective. This can be achieved through multiple passes of the blood containing the liposomes over the target tissue. A prolonged blood circulation of the liposomes was achieved with the addition of a polyethylene glycol (PEG) coating, which efficiently minimizes their removal by macrophages of the reticuloendothelial system [55]. Marik et al. [56] conducted PET imaging and biodistribution studies of liposomes. To track liposomes, radioactive [18 F]fluorodipalmitin ([18 F]FDP) was incorporated into the lipid molecule of the phospholipid bilayer of the liposome. As a control, free [18 F]FDP was also injected in rats. Maximum intensity projection images obtained from 90-min continuous bed motion scans were used to illustrate the full body distribution of free and liposome-encapsulated [18 F]FDP. Freely injected [18 F]FDP showed an initial concentration of 3% ID/cc and was cleared from the blood within a few minutes. The greatest concentration of free [18 F]FDP was observed in the liver, 5.5% ID/cc and in the spleen at a concentration of 4.2% ID/cc. Liposomal [18 F]FDP remained in the blood circulation at near constant levels for at least 90 min, with a peak concentration near 2.5% ID/cc. In addition, liposomal [18 F]FDP quickly reached a
P1: OTA/XYZ P2: ABC c11 JWBS037-Chen February 2, 2011
250
21:47
Printer Name: Yet to Come
MICROPARTICLE- AND NANOPARTICLE-BASED CONTRAST-ENHANCED ULTRASOUND IMAGING
steady state in the liver, spleen, kidney, and lungs, which was maintained throughout the scan, reflecting the activity within the vasculature of these organs.
11.3.3 Applications of Molecular Ultrasound Imaging Molecular ultrasound imaging with the advent of targeted contrast agents has opened a variety of diagnostic and therapeutic potentials. Using targeted agents, the clinical and experimental roles of ultrasound are further explored, into applications including noninvasive detection of pathology using disease-associated molecular signatures, detection of gene expression, investigation of drug localization, and delineation of molecular mechanisms of disease. Furthermore, the ability to differentiate between different disease-indicative molecular signatures could allow early assessment of pathology and expedite the design of customized treatments. There are many potential targets available for study by targeted contrast imaging. These molecular signatures can be used to localize ultrasound contrast agents through the use of complementary receptor ligands attached to the contrast agent shell such that the ligand–receptor interaction tethers the agent to the cell of interest. Studies published on targeted ultrasound contrast agents are briefly summarized below and in Tables 11.1 and 11.2.
Molecular Ultrasound Imaging of Tumor Angiogenesis Angiogenesis, the process of new blood vessel formation, plays an important role in tumor growth and metastasis and diverse disease processes such as atherosclerotic plaque growth and adaptation to chronic ischemic disease. Novel molecular imaging strategies that allow direct visualization and quantification of expression levels of key molecular markers of these diseases would be ideal tools for detection as well as monitoring therapeutic treatment in patients. To illustrate this concept, many molecular markers such as integrins or vascular endothelial growth factor receptor type 2 (VEGFR2) have been shown to be upregulated on angiogenic and metastatic endothelial cells in an actively growing tumor vasculature specifically (Fig. 11.5). The introduction of microbubbles targeted to key molecular markers of tumor angiogenesis is now being explored as a novel molecular imaging strategy with ultrasound [72, 73]. A number of studies have addressed assessment of tumor angiogenesis with molecular ultrasound in preclinical studies. Integrins, which are extracellular matrix molecules, have been extensively evaluated for targeting of imaging agents, drugs, and particles to the tumor endothelium. Integrin ␣v 3 , in particular, has received a lot of attention as it is highly expressed on activated endothelium and almost absent on normal vessels, making it very useful for detection of tumor formation. Various strategies were used to generate integrin ␣v 3 targeted microbubbles, namely, the use of streptavidin–biotin coupling chemistry along with monoclonal antibodies for integrin ␣v 3 , and the use of a cyclic RRL (arginine–arginine–leucine) peptide and echistatin as ligands on the microbubble shell that bind integrin ␣v 3 . Echistatin is a viper venom disintegrin containing an RGD (arginine–glycine–aspartic acid) peptide that binds integrin receptors. Ellegala et al. [57] imaged malignant gliomas in athymic rats by intracerebral implantation of U87MG human glioma cells. An increase in the acoustic reflectivity of malignant gliomas was demonstrated with the use of echistatin incorporated microbubbles targeted to ␣v 3 . Weller et al. [58] used the RRL peptide incorporated microbubbles to show a significant accumulation of targeted microbubbles within subcutaneously implanted, human prostate carcinoma xenografts in mice. Although the RRL peptide is known to bind
Human KDR/ VEGFR2 Endoglin
Angiogenesis
Inflammation Inflammation Inflammation Inflammation Inflammation Inflammation Arteriosclerosis Transplant rejection Transplant rejection Thrombus Thrombus Lymph node
Angiogenesis
Angiogenesis VEGFR2, Integrin alphavbeta3, Endoglin Leukocytes P-selectin P-selectin P-selectin P-selectin MadCAM-1 VCAM-1 Leukocyte ICAM-1 GP IIb/IIIa GP IIb/IIIa L-selectin
Integrin ␣v 3 VEGFR2 VEGFR2 VEGFR2 Human KDR/ VEGFR2 VEGFR2
Angiogenesis Angiogenesis Angiogenesis Angiogenesis Angiogenesis
mAb RGD peptide RGD peptide MECA-79 ligand
Microbubble Microbubbles Microbubbles Microbubbles Microbubbles Microbubbles Microbubbles Microbubbles Microbubbles Microbubbles Microbubbles Microbubbles
Microbubbles
mAb
mAb Sulfo-Le-AAA peptide mAb mAb mAb mAb
Microbubble
Microbubbles
Microbubble
Microbubble Microbubble Microbubble Microbubble Microbubbles
Microbubble Microbubble
Contrast Agent
Heterodimeric KDR-targeted peptide mAb
Echistatin peptide Tumor endothelial cell (target was not identified) Knottin peptides mAb mAb mAb Heterodimeric KDR-targeted peptide mAb
Ligand
Animal Model
Tissue necrosis factor-treated cremaster muscle of mice Mouse model for postischemic injury in kidney Rat model for postischemic injury in the myocardium Mouse model for postischemic injury Mouse model for postischemic injury Mouse model for inflammatory bowel disease Mouse model for arteriosclerotic plaque in ApoE deficient mouse Rat cardiac transplantation model Rat cardiac transplantation model Mouse cremaster muscle model for arteriolar and venular clots Thrombi formation in dogs Lymph nodes of mice and dogs
Subcutaneous tumor model in mouse for human pancreatic cancer and orthotopic pancreatic cancer model in mice Subcutaneous tumor model in mouse for human breast, ovarian and pancreatic cancer
Subcutaneous tumor model in mouse for human pancreatic cancer and orthotopic pancreatic cancer model in mice Subcutaneous human colon cancer xenografts in mice
Subcutaneous tumor model in mouse for human ovarian cancer Subcutaneous tumor model in mouse for angiosarcoma and glioma Subcutaneous tumor model in mouse for human melanoma Subcutaneous tumor model in mouse for murine breast cancer Orthotopic rat breast cancer model in rats
Subcutaneous tumor model in rat for human glioma Subcutaneous tumor model in mouse for human prostrate carcinoma
66 67 68 33 76 69 70
20 63 64 65
62
34
15
34, 35
14 59 60 61 16
57 58
References
21:47
Angiogenesis
Integrin ␣v 3 Integrin ␣v 3
Molecular Target
Angiogenesis Angiogenesis
Disease
TABLE 11.1 Summary of Published Studies on the Use of Targeted Microbubbles for Ultrasound Imaging of Angiogenesis, Inflammation, Thrombus Formation, and Lymph Nodes
P1: OTA/XYZ P2: ABC c11 JWBS037-Chen February 2, 2011 Printer Name: Yet to Come
251
P1: OTA/XYZ P2: ABC c11 JWBS037-Chen February 2, 2011
252
21:47
Printer Name: Yet to Come
MICROPARTICLE- AND NANOPARTICLE-BASED CONTRAST-ENHANCED ULTRASOUND IMAGING
TABLE 11.2 Summary of Published Studies on the Use of Targeted Nanoparticles for Ultrasound Imaging of Inflammation and Thrombus Formation Molecular Target
Disease
Ligand
Contrast Agent
Atherosclerosis ICAM-1, Fibrinogen
mAb
Liposome
Thrombus
ICAM-1
mAb
Liposome
Thrombus
VCAM-1
mAb
Liposome
Thrombus
Fibrin
mAb
PFC nanoparticles
Thrombus
Stretch induced
Tissue factor mAb
PFC nanoparticles
Animal Model
References
Yucatan miniswine model for atherosclerosis Yucatan miniswine for different stages of atheroma Yucatan miniswine for different stages of atheroma Thrombi formation in the carotid artery of pigs Pig model for balloon stretch induced injury to carotid artery
53
71
71
38
72
tumor endothelium, the molecular target is still not identified. Recent studies conducted by Willmann et al. [14] showed the use of a novel peptide—knottin—that binds to integrin ␣v 3 with high affinity. Knottins are small, compact peptides (20–60 amino acids) that consist of a core of at least three disulfide bonds that are interwoven into a “knotted” conformation. The in vivo molecular ultrasound imaging signal seen with knottin peptide when conjugated to the shell of contrast microbubbles was similar or higher when compared with microbubbles
(a)
(b)
FIGURE 11.5 (a) Transverse B-mode ultrasound image of a subcutaneous human colon adenocarcinoma xenograft tumor (arrows) in a nude mouse. (b) Targeted contrast-enhanced ultrasound image of tumor angiogenesis in same xenograft tumor (arrows) 4 min after intravenous injection of microbubbles targeted to VEGFR2; imaging signal from contrast microbubbles bound to VEGFR2 is shown as green overlay on B-mode image. Note: Molecular ultrasound imaging signal was measured using the destruction replenishment approach described in Figure 11.3.
P1: OTA/XYZ P2: ABC c11 JWBS037-Chen February 2, 2011
21:47
Printer Name: Yet to Come
ULTRASOUND CONTRAST AGENTS
253
conjugated to a monoclonal antibody for integrin ␣v 3 . Knottins have great potential for in vivo applications due to their resistance to proteolysis and their high thermal stability and are thought to be nonimmunogenic. Finally, knottins may be a promising platform for designing novel contrast agents for molecular ultrasound imaging [14]. Another angiogenic marker, VEGFR2, has also been studied as a target for molecular ultrasound imaging using high-frequency ultrasound. Overexpression of VEGFR2 has been associated with tumor progression and poor prognosis in several tumors including colorectal, gastric, and pancreatic carcinomas; angiosarcoma; breast, prostate, and lung cancers; malignant gliomas; and melanomas. Several studies reported enhanced contrast images of tumor with the use of VEGFR2 targeted microbubbles [29, 32, 34, 35, 59–62]. Willmann et al. [29] used targeted microbubbles by coupling monoclonal antibodies as the binding ligand against VEGFR2 via streptavidin–biotin coupling chemistry; as a control, microbubbles were also conjugated to isotype matched antibody along with unlabeled microbubbles. The ability of these microbubbles to adhere to the target was first confirmed in cell culture experiments. In vivo ultrasound images were then acquired with a 40-MHz linear transducer in a mouse subcutaneous tumor model for angiosarcoma and malignant glioma. Accumulation of VEGFR2 targeted microbubbles was reported within subcutaneously implanted tumors compared to unspecific control microbubbles, which was further confirmed by immunofluorescence analysis of VEGFR2 expression in both tumor types [29]. Rychak et al. [60] reported similar results on subdermal tumors derived from human melanoma cells in mice. Studies performed by Lyshchik et al. [61] compared expression of VEGFR2 by targeted microbubbles on highly invasive metastatic and nonmetastatic murine models of breast cancer cells demonstrating that targeted ultrasound can be used to characterize angiogenic activity corresponding to the degree of malignancy. They observed a higher accumulation of VEGFR2 targeted microbubbles within the more aggressive tumors. Korpanty and colleagues [34] investigated the use of targeted microbubbles to follow vascular response of therapy in addition to detection of tumor angiogenesis by molecular ultrasound. VEGFR2 targeted microbubble accumulation was assessed to quantify vascular effects of two different antitumor therapies—namely, with antivascular endothelial growth factor (VEGF) monoclonal antibodies and/or gemcitabine. The model systems used were subcutaneous as well as orthotopic pancreatic cancer tumors in mice. They detected decreasing marker densities after tumor- suppressive therapy, which correlated with the observed effects of treatment. In addition, they performed a multimarker imaging, assessing VEGFR2 followed by Endoglin targeted microbubbles in the same tumor after a long interval between scans to ensure passive clearance of previously injected microbubbles. Endoglin is a cell membrane glycoprotein that is involved in vascular development and remodeling and is overexpressed on tumor-associated vascular endothelium. Targeted microbubbles showed significant enhancement of tumor vasculature when compared with untargeted or control IgG targeted microbubbles that correlated with ex vivo expression analysis. In the same light Palmowski and colleagues [35] demonstrated an upregulation of VEGFR2 and of integrin ␣v 3 during the growth of untreated tumors, and a downregulation of both markers after antiangiogenic therapy using microbubbles conjugated to VEGFR2 antibodies and cyclic RGD peptides. Recently, Pysz et al. [15] evaluated a novel, clinically-translatable microbubble targeted to human kinase insert domain receptor (KDR; the human protein analogous VEGFR-2) for monitoring antiangiogenic therapy in subcutaneous human colon cancer xenografts in mice. In mice receiving antiangiogenic therapy in vivo molecular ultrasound imaging using novel KDR-targeted microbubbles significantly
P1: OTA/XYZ P2: ABC c11 JWBS037-Chen February 2, 2011
254
21:47
Printer Name: Yet to Come
MICROPARTICLE- AND NANOPARTICLE-BASED CONTRAST-ENHANCED ULTRASOUND IMAGING
decreased as early as 24 hours after initiation of antiangiogenic therapy. In contrast, no difference in molecular ultrasound KDR-targeted imaging signal was observed in nontreated mice. Furthermore, KDR-associated molecular ultrasound signal was observed prior to any changes in tumor size; thus, demonstrating the advantage of early assessment of antiangiogenic therapy using molecular ultrasound imaging prior to overt morphological-anatomical changes become visible in tumors. Another aspect observed by Willmann et al. [59] was that multiple markers conjugated on microbubbles could significantly enhance the imaging signals, which could be a useful tool for detection of tumors early enough for effective therapy. Toward this goal, it was demonstrated that dual-targeted microbubbles carrying antibodies for VEGFR2 and integrin ␣V accumulate to higher intensities compared to single targeted microbubbles in tumor angiogenesis of human ovarian cancer in mice xenografts. These results may be of great significance for early detection of cancer when tumors are too small to cause detectable morphologic changes but large enough to induce tumor angiogenesis (Fig. 11.6). Molecular ultrasound imaging also allows noninvasive mapping of expression levels of angiogenic markers in tumor angiogenesis. In three different tumor types (breast, ovarian, and pancreatic cancer) and using noninvasive molecular ultrasound imaging, Deshpande et al. [62] showed varying expression levels of the three angiogenic markers integrin ␣v 3 , endoglin, and VEGFR2 during tumor growth. In this study, Using ex vivo western blotting as reference standard, the study confirmed that molecular ultrasound imaging allows longitudinal noninvasive assessment of the temporal tumor angiogenic molecular marker expression levels in vivo [62]. The results of this study provided further insights into the biology of tumor angiogenesis and may help in defining promising imaging targets for
Endothelial cell
Endothelial cell
Endothelial cell
Endothelial cell
FIGURE 11.6 Single and dual targeted microbubbles and interaction with tumor vessel endothelial cells. Gas-filled microbubbles (MBs) conjugated to ligands bind molecular markers of angiogenesis (e.g., VEGFR2) expressed on tumor vessel endothelial cells. Dual targeted MBs, which carry two types of ligands on their shells, can bind to more molecular markers compared to single targeted microbubbles. This increases the signal intensity from derived bound microbubbles, which, for example, may increase the detection of small tumors.
P1: OTA/XYZ P2: ABC c11 JWBS037-Chen February 2, 2011
21:47
Printer Name: Yet to Come
ULTRASOUND CONTRAST AGENTS
255
both early cancer detection and treatment monitoring of cancer using molecular ultrasound imaging [62].
Molecular Ultrasound Imaging of Inflammation Inflammation is a common physiological process occurring in a vast number of diseases such as ischemia (inadequate blood supply to tissue), arthrosclerosis, or inflammatory bowel disease to mention only a few. One crucial component of inflammation is the activation of leukocytes in the blood pool and their transmigration to the extravascular compartment. The recruitment and transmigration of leukocytes are mediated by the interaction between adhesion molecules on the leukocytes and on the endothelial cell surface [20]. Several molecules including E- and P-selectin mediate the initial capture and consecutive rolling of leukocytes on the inner vessel wall. A firm arrest of the rolling leukocytes, the necessary precondition for transmigration, is promoted by a second group of adhesion molecules, namely, the intercellular adhesion molecule-1 (ICAM- 1) and the vascular cell adhesion molecule-1 (VCAM-1). All inflammatory markers are expressed rapidly during the inflammatory process and have been reported to correlate to a certain degree to the stage of inflammation. Both passive and active targeting has been explored for imaging inflammation using molecular ultrasound. Passive targeting does not provide molecular-level information, but rather enables qualitative detection of inflammation as microbubbles attach to (and eventually phagocytosed by) activated leukocytes. This leads to damping, but the microbubbles remain acoustically responsive and still produce an acoustic signal [74]. One strategy for passive targeting is to incorporate negatively charged phosphatidylserine into the shell, which promotes microbubble attachment to activated leukocytes. This technique was used to image tissue necrosis factor-␣-treated inflamed cremaster muscle in mice and to assess the severity and extent of postischemic myocardial inflammation in dogs [74]. For active targeting, monoclonal antibodies that bind inflammatory markers like E- and P-selectin, ICAM-1, and VCAM-1 have been outfitted onto microbubbles for molecular ultrasound imaging of inflammation [8]. P-selectin, which is expressed immediately after an ischemic stimulus, was used to assess postischemic injury in the mouse kidney [63]. P-selectin targeted microbubbles were also used to identify recently ischemic myocardium in a myocardial ischemia reperfusion model in mice [64]. A similar study used sialyl Lewis (a P-selectin ligand) conjugated microbubbles in a rat model for myocardial ischemia reperfusion to visualize “the ischemic memory” noninvasively using molecular ultrasound imaging [65], highlightening the potential of molecular ultrasound imaging as a rapid and straight forward bedside test for screening patients with atypical chest pain for a recent ischemic event. The gut-specific marker “mucosal addressin cellular adhesion molecule-1” (MAdCAM1) was used as a target to image a mouse model for inflammatory bowel disease [66]. Bachmann et al. [66], using transabdominal ultrasound, demonstrated a significant accumulation of MAdCAM-1 targeted microbubbles as compared to nonspecific ones in focal areas of ileal inflammation, thus producing stronger acoustic signals. VCAM-1 antibodyconjugated microbubbles were utilized by Kaufmann et al. [67] to demonstrate a correlation between different stages of arteriosclerosis and the retention of microbubbles. The inflammation disease model system used was ApoE deficient mice, which form inflammatory plaques in the aorta when fed a high-cholesterol diet. Two studies have been reported to investigate transplant rejection in vivo in animal models. The first study employed passively leukocyte-targeted microbubbles with phosphatidylserine to assess acute allograft rejection
P1: OTA/XYZ P2: ABC c11 JWBS037-Chen February 2, 2011
256
21:47
Printer Name: Yet to Come
MICROPARTICLE- AND NANOPARTICLE-BASED CONTRAST-ENHANCED ULTRASOUND IMAGING
in a rat cardiac transplantation model [68]. The degree of rejection in transplanted hearts could be directly revealed by the magnitude of intramyocardial infiltration of macrophages and T lymphocytes due to accumulation of targeted microbubbles at the site of inflammation. The second study used microbubbles conjugated to ICAM-1 antibody for selective imaging of cardiac transplant rejection in rats; approximately a tenfold increase in contrast intensity was observed in rejecting myocardium [33].
Molecular Ultrasound Imaging of Thrombus Formation A thrombus or blood clot is achieved via the aggregation of platelets that form a platelet plug, and the activation of the humoral coagulation system. A thrombus is normal in cases of injury, but pathologic in instances of thrombosis. Research in detection of thrombus formation by ultrasound imaging has been ongoing to develop contrast agents that could detect diseases such as stroke, myocardial infarct, and deep vein thrombosis. Various molecular markers conjugated to microbubbles, PFC nanoparticles, and liposomes were developed for animal studies: for platelet targeting, antibodies against GPIIb/IIIa receptors expressed on activated platelets or RGD peptides recognized by the active binding site of GPIIb/IIIa were developed, and various monoclonal antibodies were generated against markers associated with atheroma development such as fibrin, ICAM-1, VCAM-1, and tissue factor. Microbubble binding studies on arteriolar and venular clots in a mouse cremaster muscle model were conducted by Schumann et al. [76]. They confirmed binding of targeted microbubbles in both venules and arterioles. Hamilton et al. [71] used these targeted liposomes for intravascular ultrasound imaging of injured vessels of miniswine as a model system used to create various stages of atheroma. Different types of targeted liposome were generated with conjugation of anti-ICAM-1, anti-VCAM-1, anti-fibrin, and anti-tissue factor antibodies. These targeted liposomes demonstrated targeted enhancement in the vessel walls 5 min after intravenous administration. Similarly, Demos et al. [53] injected echogenic liposome targeted to atherosclerotic plaque created in the Yucatan miniswine animal model. The liposomes were conjugated with anti-fibrinogen or anti-ICAM-1 antibody. The liposomes attached to thrombi and to atherosclerotic arterial wall. Mean acoustic intensities of blood alone and blood with nontargeted and targeted agents showed that targeted liposomes dramatically increased the echogenicity of blood. Lanza et al. [38] employed targeted PFC nanoparticles for intravascular ultrasound and MRI contrast imaging. The PFC nanoparticles were outfitted with biotin–avidin coupling system along with anti-fibrin monoclonal antibodies. Frequency of thrombus detection increased from 2% to 83% after single administration and increased to 96% after second administration of targeted PFC nanoparticles with increased acoustic reflectivity of carotid thrombi of dogs. In a related study, a stretch induced tissue factor was imaged in balloon stretched pig carotid arteries by administering a tissue factor targeted antibody on biotinylated PFC nanoparticles. After administration of the targeted nanoparticles, the arteries were imaged with a 20-MHz intravascular ultrasound system, and the targeted contrast more than doubled the gray scale intensity of the injured areas [72]. Unger et al. [69] have developed microbubbles labeled with RGD analog that enhanced the echogenicity of induced thrombi in dogs. Molecular Ultrasound Imaging of Lymph Nodes The feasibility of using targeted microbubbles to image peripheral lymph nodes under normal conditions in animal models of mice and dogs was tested by Hauff et al. [70] by the use of stimulated acoustic emission (SAE). SAE involves color or power Doppler imaging with the transmission power set high enough to ensure microbubble disruption on the first pulse. This causes a transient high-
P1: OTA/XYZ P2: ABC c11 JWBS037-Chen February 2, 2011
21:47
Printer Name: Yet to Come
CHALLENGES AND FUTURE DIRECTIONS OF MOLECULAR ULTRASOUND IMAGING
257
amplitude response. Since an ultrasound Doppler system correlates the signal backscattered from a target within a number of successive pulses, the loss of signal correlation caused by the transient bubble collapse is interpreted by the machine as a random Doppler shift, resulting in a mosaic of colors at the location of microbubbles even without flow. SAE works particularly well with the air-filled microbubbles, as used by Hauff and colleagues. It has proved a powerful method of studying passively targeted agents [77] but has not previously been used to image active targeting. Microbubbles were conjugated to L-selectin specific ligand MECA-79. L-selectin is expressed on circulating lymphocytes and is involved in homing of the lymphocytes to lymph nodes. In these experiments, mice were sacrificed after intravenous administration of L-selectin targeted microbubbles. The lymph nodes were removed and examined by using harmonic color Doppler ultrasound in a tank containing degassed water. The lymph nodes of all the mice showed enhanced acoustic signal due to accumulation of L-selectin targeted microbubbles. In another experiment, anesthetized dogs were scanned with ultrasound in vivo after administration of L-selectin targeted microbubbles intravenously. The targeted microbubbles accumulated significantly in healthy lymph nodes. Thus L-selectin ligand-specific US contrast agent could be a candidate for an indirect method of lymphography for the safe and less invasive US identification of lymph nodes—for example, when performing ultrasound-guided biopsy [70].
11.4 CHALLENGES AND FUTURE DIRECTIONS OF MOLECULAR ULTRASOUND IMAGING Ultrasound imaging lacked effective contrast agent to render it a molecular ultrasound imaging tool until recently. With the introduction of microbubble contrast agents, great progress has been made in the field of contrast enhanced molecular ultrasound imaging. Furthermore, the development of targeted contrast agents has made it possible to image the pathophysiology of many diseases at the molecular and cellular levels using molecular ultrasound imaging and has opened up new avenues for using molecular imaging for therapy. A lot of investigative research has been done to detect intravascular events that play a role in cardiovascular and cerebrovascular diseases, including inflammatory responses, angiogenesis, and thrombus formation by targeted contrast-enhanced ultrasound in the last few years. Yet the field currently is restricted to animal models of human disease. Humanizing antibodies used for target detection using targeted microbubbles can be very expensive. Furthermore, the current biotin–streptavidin system used for preclinical trials to attach ligands onto the shell of ultrasound contrast agents is not suitable for clinical use, due to the immunogenic and allergic risk that streptavidin presents to humans, especially in the case of repeated use. Investigation of synthetic novel peptides that recognize a variety of molecular markers are under way and may provide a cost-effective alternative, circumventing the problem of allergic reactions in humans [14, 15]. Recently, it has been shown that a novel, clinically translatable contrast microbubble targeted to human KDR (which corresponds to VEGFR2 in mice) were designed using a small peptide directly integrated into the shell of contrast microbubbles (thereby avoiding binding chemistry that would preclude a clinical translation into patients) [15]. Molecular ultrasound imaging signal using this human KDR-targeted contrast agent showed significantly higher imaging signal in human colon cancer xenografts in mice than control microbubbles and allowed monitoring of antiangiogenic therapy in vivo [15]. This novel contrast agent was the first that was designed specifically for a clinical translation into patients; first in
P1: OTA/XYZ P2: ABC c11 JWBS037-Chen February 2, 2011
258
21:47
Printer Name: Yet to Come
MICROPARTICLE- AND NANOPARTICLE-BASED CONTRAST-ENHANCED ULTRASOUND IMAGING
human clinical trials are expected for the near future assessing the potential of molecular ultrasound imaging in a clinical environment [15]. Several new microbubble constructs such as the described human KDR-targeted microbubbles, comprehensive scanner systems, and imaging protocols are currently under development to expand this technology to the clinical arena. Targeted contrast-enhanced ultrasound imaging could potentially be used in a wide range of applications, including diagnostics, image-guided biopsies, and treatment follow-up. Lastly, microbubbles today are still restricted to intravascular space; research into nanoparticles would likely help to overcome this aspect. Finally, extensive research is being directed toward development of the next generation of microbubbles, which are capable of encapsulating therapeutic agents and releasing them when exposed to high MI ultrasound waves. Microbubbles in combination with a therapeutic agent provide the vehicle for targeting molecular events and thus combining imaging with pathophysiology and ultimately therapy. Therapeutic agents could include genes, thrombolytics, and oncological drugs, and this technique has the clinical potential to increase therapeutic efficacy while decreasing systemic side effects. Clinical use of microbubbles has faced problems with regard to safety issues. In 2007 the United States Food and Drug Administration (FDA) ordered black box warnings for two contrast agents—Definity, formerly sold by Bristol-Myers Squibb, and Optison, sold by GE Healthcare—following reports of 11 deaths allegedly associated with the two agents. Optison consists of perflutren protein-type A microspheres, whereas Definity (available as Luminity in Europe and Australia) is a preparation of liposome-encapsulated microspheres containing perflutren. Both contrast agents were initially approved for clinical use in patients for echocardiography to enhance suboptimal echocardiograms to opacify the left ventricular chamber and to improve the delineation of the left ventricular endocardial border. The related ban nearly entirely halted the use of Definity in the United States. In 2008 the FDA eased off the stringent black box warnings that greatly restricted use of Definity and Optison. The FDA announcement followed publication of a retrospective study in the Journal of the American College of Cardiology (2007; 1704–1706). The review, involving 18,671 consecutive patients at St. Luke’s Mid-America Health Institute in Kansas City, Missouri, found no increased mortality risk for patients who received Definity-enhanced echocardiography compared with patients who were imaged without the agent. Some FDA-mandated labeling restrictions remain: Definity is now contraindicated for patients with known or suspected right-to-left, bidirectional, or transient right-to-left cardiac shunts and hypersensitivity to perflutren, an ingredient of the microspherical agents. Intra-arterial injection is still banned. So far, microbubble ultrasound contrast media are not approved for noncardiac applications in the United States [78]. However, there is an ongoing Phase III clinical trial in thr USA testing the diagnostic accuracy of nontargeted contrast-enhanced ultrasound imaging (using the microbubble SonoVue) for liver lesion characterization compared to CT, MRI or histology as the golden standard. Following successful completion of this clinical trial, US FDA approval of nontargeted contrast-enhanced ultrasound imaging for radiological indications is expected. Contrast-enhanced ultrasound imaging with nontargeted microbubbles is already widely used in Canada, Europe, and Asia. and microbubbles have been shown to be nontoxic with an extremely low adverse event rate as low as 0.13% (29 per 23,988 examinations) [79]. In conclusion, molecular ultrasound is an emerging molecular imaging approach finding its niche among other molecular imaging modalities. High spatial and temporal resolution, real-time imaging, noninvasiveness, relatively low costs, lack of ionizing irradiation and wide availability among the imaging community throughout the world are important advantages that will further define the role of this novel imaging technique both in preclinical
P1: OTA/XYZ P2: ABC c11 JWBS037-Chen February 2, 2011
21:47
Printer Name: Yet to Come
REFERENCES
259
research and clinical applications. In addition, ongoing improvements in ultrasound technology and sophisticated contrast agent design with novel high-affinity targeting ligands using clinically translatable binding chemistry, and further improvements in biodistribution of ultrasound contrast agents beyond the vasculature will further expand the clinical role of molecular ultrasound for imaging diseases at the molecular level in medicine [80, 81, 82].
REFERENCES 1. Hoskins, P. R.; Thrush, A.; Martin, K.; Whittingham, T. A. Diagnostic Ultrasound. Greenwich Medical Media Ltd, London, 2003. 2. Hauff, P.; Foster, S. Ultrasound Basics. Springer-Verlag: Berlin, 2008. 3. Dogra, V.; Rubens, D. J. Ultrasound Secrets. Hanley & Belfus Medical Publishers: Philadelphia, PA, 2003. 4. Aljarboua, Z. Ultrasound effect on biological tissue. Open Educational Resource. 2008. 5. Willmann, J. K.; van Bruggen, N.; et al. Molecular imaging in drug development. Nat. Rev. Drug Discov. 2008, 7(7), 591–607. 6. Rumack, C. M.; Wilson, S. R.; Charboneau, J. W.; Johnson, J. Diagnostic Ultrasound. Mosby: New York, 1998. 7. Leighton, G. The Acoustic Bubble. Academic Press, Inc: London, UK, 1997, 1–117. 8. Lindner, J. R. Molecular imaging with contrast ultrasound and targeted microbubbles. J. Nucl. Cardiol. 2004, 11(2), 215–221. 9. Soye, J. A.; Mullan, C. P.; et al. The use of contrast-enhanced ultrasound in the characterisation of focal liver lesions. Ulster Med. J. 2007, 76(1), 22–25. 10. Stanley, B. Safe use of ultrasound contrast agents. Ultrasound Med. Biol. 2007, 33(2), 171–172. 11. Klibanov, A. L. Ligand-carrying gas-filled microbubbles: ultrasound contrast agents for targeted molecular imaging. Bioconjug. Chem. 2005, 16(1), 9–17. 12. Junru, Wu, W. L. N. Ultrasound, cavitation bubbles and their interaction with cells. Adv. Drug Deliv. Rev. 2008, 60, 1103–1116. 13. Klibanov, A. L. Molecular imaging with targeted ultrasound contrast microbubbles. Ernst Schering Res. Found. Workshop 2005, 49, 171–191. 14. Willmann, J. K.; Kimura, R. H.; Deshpande, N.; Lutz, A. M.; Rosenberg, J.; Cochran, J. R.; Gambhir, S. S. Novel integrin-binding ultrasound contrast agent for molecular ultrasound imaging of tumor angiogenesis. J. Nucl. Med. 2010, 51, 433–440. 15. Pysz, M. A.; Foygel, K.; Rosenberg, J.; Gambhir, S. S.; Schneider, M.; Willmann, J. K. Antiangiogenic cancer therapy: monitoring with molecular US and a clinically translatable contrast agent (BR55). Radiology 2010, 256(2), 519–527. 16. Pochon, S.; Tardy, I.; Bussat, P.; Bettinger, T.; Brochot, J.; von Wronski, M.; Passantino, L.; Schneider, M. BR55: a lipopeptide-based VEGFR2-targeted ultrasound contrast agent for molecular imaging of angiogenesis. Invest Radiol. 2010, 45(2), 89–95. 17. Shah, P. M.; Gramiak, R.; et al. Determinants of atrial (S4) and ventricular (S3) gallop sounds in primary myocardial disease. N. Engl. J. Med. 1968, 278(14), 753–758. 18. Burns, M. Marketing challenges continue for ultrasound contrast media. Biotech Systems Inc. 2004, http://www.biotechsystems.com 19. Lanza, G. M.; Wickline, S. A. Targeted ultrasonic contrast agents for molecular imaging and therapy. Curr. Probl. Cardiol. 2003, 28(12), 625–653. 20. Lindner, J. R.; Song, J.; et al. Noninvasive ultrasound imaging of inflammation using microbubbles targeted to activated leukocytes. Circulation 2000, 102(22), 2745–2750.
P1: OTA/XYZ P2: ABC c11 JWBS037-Chen February 2, 2011
260
21:47
Printer Name: Yet to Come
MICROPARTICLE- AND NANOPARTICLE-BASED CONTRAST-ENHANCED ULTRASOUND IMAGING
21. Fisher Science Education, Hanover Park, IL. www.fishersci.com 22. Oussoren, C.; Zuidema, J.; Crommelin, D. J.; Storm, G. Lymphatic uptake and biodistribution of liposomes after subcutaneous injection. II. Influence of liposomal size, lipid compostion and lipid dose. Biochm. Biophys. Acta. 1997, 1328, 261–272. 23. Klibanov, A. L. Targeted delivery of gas-filled microspheres, contrast agents for ultrasound imaging. Adv. Drug Deliv. Rev. 1999, 37, 139–157. 24. Klibanov, A. L.; Rasche, P. T.; et al. Detection of individual microbubbles of ultrasound contrast agents: imaging of free-floating and targeted bubbles. Invest. Radiol. 2004, 39(3), 187–195. 25. McCulloch, M.; Gresser, C.; et al. Ultrasound contrast physics: a series on contrast echocardiography, article 3. J. Am. Soc. Echocardiogr. 2000, 13(10), 959–967. 26. Leong-Poi, H.; Song, J.; et al. Influence of microbubble shell properties on ultrasound signal: implications for low-power perfusion imaging. J. Am. Soc. Echocardiogr. 2002, 15(10 Pt 2), 1269–1276. 27. Chomas, J. E.; Pollard, R. E.; et al. Contrast-enhanced US of microcirculation of superficially implanted tumors in rats. Radiology 2003, 229(2), 439–446. 28. Pollard, R. E.; Sadlowski, A. R.; et al. Contrast-assisted destruction-replenishment ultrasound for the assessment of tumor microvasculature in a rat model. Technol. Cancer Res. Treat. 2002, 1(6), 459–470. 29. Willmann, J. K.; Paulmurugan, R.; et al. US imaging of tumor angiogenesis with microbubbles targeted to vascular endothelial growth factor receptor type 2 in mice. Radiology 2008, 246(2), 508–518. 30. Walday, P.; Tolleshaug, H.; et al. Biodistributions of air-filled albumin microspheres in rats and pigs. Biochem. J. 1994, 299(Pt 2), 437–443. 31. Perkins, A. C.; Frier, M.; et al. Human biodistribution of an ultrasound contrast agent (Quantison) by radiolabelling and gamma scintigraphy. Br. J. Radiol. 1997, 70(834), 603–611. 32. Willmann, J. K.; Cheng, Z.; et al. Targeted microbubbles for imaging tumor angiogenesis: assessment of whole-body biodistribution with dynamic micro-PET in mice. Radiology 2008, 249(1), 212–219. 33. Weller, G. E.; Lu, E.; et al. Ultrasound imaging of acute cardiac transplant rejection with microbubbles targeted to intercellular adhesion molecule—1. Circulation 2003, 108(2), 218–224. 34. Korpanty, G.; Carbon, J. G.; et al. Monitoring response to anticancer therapy by targeting microbubbles to tumor vasculature. Clin. Cancer Res. 2007, 13(1), 323–330. 35. Palmowski, M.; Huppert, J.; et al. Molecular profiling of angiogenesis with targeted ultrasound imaging: early assessment of antiangiogenic therapy effects. Mol. Cancer Ther. 2008, 7(1), 101–109. 36. Gref, R.; Domb, P.; Quellec, P.; Blunk, T.; Muller, R. H.; Verbavatz, J. M.; Langer, R. The controlled intravenous delivery of drugs using PEG-coated sterically stabilized nanospheres. Adv. Drug Deliv. Rev. 1995, 16, 215–233. 37. Brigger, I.; Dubernet, C.; et al. Nanoparticles in cancer therapy and diagnosis. Adv. Drug. Deliv. Rev. 2002, 54(5), 631–651. 38. Lanza, G. M.; Wallace, K. D.; et al. High-frequency ultrasonic detection of thrombi with a targeted contrast system. Ultrasound Med. Biol. 1997, 23(6), 863–870. 39. Pana, D.; Wickline, S. A.; Caruthers, S. D. Nanomedicine: perspective and promises with liganddirected molecular imaging. Eur. J. Radiol. 2009, 70(2), 274–285. 40. Lanza, G. M.; Trousil, R. M.; Wallace, K. D.; Scott M. J.; Miller J. G.; Eisenberg P. R.; Gaffney P. J.; Wickline S. A. In vitro characterization of a novel, tissue-targeted ultrasonic contrast system with acoustic microscopy. J. Acoust. Soc. Am. 1998, 104, 3665–3672. 41. Kripfgans, O. D.; Fabiilli, M. L.; Cardon, P. L.; et al. On the acoustic vaporization of micrometersized droplets. J. Acoust. Soc. Am. 2004, 116(1), 272–281.
P1: OTA/XYZ P2: ABC c11 JWBS037-Chen February 2, 2011
21:47
Printer Name: Yet to Come
REFERENCES
261
42. Amirriazi, S.; Eghtedari, M.; Jin, S.; et al. Increasing ultrasound sensitivity to perfluorocarbon emulsion droplets. World Molecular Imaging Congress: Montreal, Canada, 2009 (abstract). 43. Joseph, P. M.; Kundel, H. L.; Mukherji, B.; Sloviter, H. A.; Magnetic resonance imaging of fluorine in rats infused with artificial blood. Invest. Radiol. 1985, 20, 504–509. 44. Mukherji, H. A. S. a. B. Prolonged retention in the circulation of emulsified lipid-coated perfluorochemicals. Prog. Clin. Biol. Res. 1983, 122, 181–187. 45. Liu, J.; Levine, A. L.; Li, J.; et al. Biodegradable nanoparticles for targeted ultrasound imaging of breast cancer cells in vitro. Phys. Med. Biol. 2007, 52(16), 4739–4747. 46. Rapoport, N.; Gao, Z.; et al. Multifunctional nanoparticles for combining ultrasonic tumor imaging and targeted chemotherapy. J. Natl. Cancer Inst. 2007, 99(14), 1095–1106. 47. Liu, J.; Mattoon, J. S.; Yamaguchi, M.; Lee, R. J.; Pan, X.; Rosol, T. J. Nanoparticles as image enhancing agents for ultrasonography. Phys. Med. Biol. 2006, 51(9), 2179–2189. 48. Maeda, H.; Wu, J.; et al. Tumor vascular permeability and the EPR effect in macromolecular therapeutics: a review. J. Control. Release 2000, 65(1–2), 271–284. 49. Nolte, I.; Vince, G. H.; Maurer, M.; Herbold, C.; Goldbrunner, R.; Solymosi, L.; Stoll, G.; Bendszus, M. Iron particles enhance visualization of experimental gliomas with high resolution sonography. Am. J. Neuroradiol. 2005, 26, 1469–1474. 50. Linker, R. A.; Kroner, A.; et al. Iron particle-enhanced visualization of inflammatory central nervous system lesions by high resolution: preliminary data in an animal model. Am. J. Neuroradiol. 2006, 27(6), 1225–1229. 51. Torchilin, V. P. Liposomes as Carriers of Contrast Agents for In Vivo Diagnostics. Elsevier Science: New York, 1998. 52. MacDonald, R. Applications of Freezing and Thawing in Liposome Technology. CRC Press: Boca Raton, FL, 1993. 53. Demos, S. M.; Onyuksel, H.; et al. In vitro targeting of antibody-conjugated echogenic liposomes for site-specific ultrasonic image enhancement. J. Pharm. Sci. 1997, 86(2), 167–171. 54. Senior, J. H. Fate and behaviour of liposomes in vivo: a review of controlling factors. Crit. Rev. Ther. Drug Carrier Syst. 1987, 3, 123–193. 55. Oku, N.; Tokudome, Y.; et al. Evaluation of drug targeting strategies and liposomal trafficking. Curr. Pharm. Des. 2000, 6(16), 1669–1691. 56. Marik, J.; Tartis, M. S.; et al. Long-circulating liposomes radiolabeled with [18 F]fluorodipalmitin ([18 F]FDP). Nucl. Med. Biol. 2007, 34(2), 165–171. 57. Ellegala, D. B.; Leong-Poi, H.; et al. Imaging tumor angiogenesis with contrast ultrasound and microbubbles targeted to alpha(v)beta3. Circulation 2003, 108(3), 336–341. 58. Weller, G. E.; Wong, M. K.; et al. Ultrasonic imaging of tumor angiogenesis using contrast microbubbles targeted via the tumor-binding peptide arginine–arginine–leucine. Cancer Res. 2005, 65(2), 533–539. 59. Willmann, J. K.; Lutz, A. M.; et al. Dual-targeted contrast agent for US assessment of tumor angiogenesis in vivo. Radiology 2008, 248(3), 936–944. 60. Rychak, J. J.; Graba, J.; et al. Microultrasound molecular imaging of vascular endothelial growth factor receptor 2 in a mouse model of tumor angiogenesis. Mol. Imaging 2007, 6(5), 289–296. 61. Lyshchik, A.; Fleischer, A. C.; et al. Molecular imaging of vascular endothelial growth factor receptor 2 expression using targeted contrast-enhanced high-frequency ultrasonography. J. Ultrasound Med. 2007, 26(11), 1575–1586. 62. Deshpande, N.; Ren, Y.; Foygel, K.; Rosenberg, J.; Willmann, J. K. Tumor angiogenic marker expression levels during tumor growth: longitudinal assessment with molecular ultrasound imaging. Radiology (in press) 63. Lindner, J. R.; Song, J.; et al. Ultrasound assessment of inflammation and renal tissue injury with microbubbles targeted to P-selectin. Circulation 2001, 104(17), 2107–2112.
P1: OTA/XYZ P2: ABC c11 JWBS037-Chen February 2, 2011
262
21:47
Printer Name: Yet to Come
MICROPARTICLE- AND NANOPARTICLE-BASED CONTRAST-ENHANCED ULTRASOUND IMAGING
64. Kaufmann, B. A.; Lewis, C.; Xie, A.; et al. Detection of recent myocardial ischaemia by molecular imaging of P-selectin with targeted contrast echocardiography. Eur. Heart J. 2007, 28(16), 2011. 65. Villanueva, F. S.; Lu, E.; et al. Myocardial ischemic memory imaging with molecular echocardiography. Circulation 2007, 115(3), 345–352. 66. Bachmann, C.; Klibanov, A. L.; et al. Targeting mucosal addressin cellular adhesion molecule (MAdCAM)-1 to noninvasively image experimental Crohn’s disease. Gastroenterology 2006, 130(1), 8–16. 67. Kaufmann, B. A.; Sanders, J. M.; et al. Molecular imaging of inflammation in atherosclerosis with targeted ultrasound detection of vascular cell adhesion molecule—1. Circulation 2007, 116(3), 276–284. 68. Kondo, I.; Ohmori, K.; et al. Leukocyte-targeted myocardial contrast echocardiography can assess the degree of acute allograft rejection in a rat cardiac transplantation model. Circulation 2004, 109(8), 1056–1061. 69. Unger, E. C.; McCreery, T. P.; et al. In vitro studies of a new thrombus-specific ultrasound contrast agent. Am. J. Cardiol. 1998, 81(12A), 58G–61G. 70. Hauff, P.; Reinhardt, M.; et al. Molecular targeting of lymph nodes with L-selectin ligand-specific US contrast agent: a feasibility study in mice and dogs. Radiology 2004, 231(3), 667–673. 71. Hamilton, A. J.; Huang, S. L.; et al. Intravascular ultrasound molecular imaging of atheroma components in vivo. J. Am. Coll. Cardiol. 2004, 43(3), 453–460. 72. Lanza, G. M.; Abendschein, D. R.; et al. In vivo molecular imaging of stretch-induced tissue factor in carotid arteries with ligand-targeted nanoparticles. J. Am. Soc. Echocardiogr. 2000, 13(6), 608–614. 73. Bloch, S. H.; Dayton, P. A.; et al. Targeted imaging using ultrasound contrast agents. Progess and opportunities for clinical and research applications. IEEE Eng. Med. Biol. Mag. 2004, 23(5), 18–29. 74. Kiessling, F.; Huppert, J.; et al. Functional and molecular ultrasound imaging: concepts and contrast agents. Curr. Med. Chem. 2009, 16(5), 627–642. 75. Lindner, J. R.; Dayton, P. A.; Coggins, M. P.; Ley, K.; Song, J.; Ferrara, K.; Kaul, S. Noninvasive imaging of inflammation by ultrasound detection of phagocytosed microbubbles. Circulation 2000, 102, 531–538. 76. Schumann, P. A.; Christiansen, J. P.; et al. Targeted-microbubble binding selectively to GPIIb IIIa receptors of platelet thrombi. Invest. Radiol. 2002, 37(11), 587–593. 77. Blomley, M. J.; Albrecht, T.; et al. Improved imaging of liver metastases with stimulated acoustic emission in the late phase of enhancement with the US contrast agent SH U 508A: early experience. Radiology 1999, 210(2), 409–416. 78. Brice, J. “FDA eases black box restrictions for ultrasound contrast media.” Diagnostic Imaging, 2008, http://www.diagnosticimaging.com. 79. Kim, T. K.; M.; Jang, H.-J.; Wilson, S. R. Microbubble contrast agents for ultrasound imaging—safety and efficacy in abdominal and vascular imaging. US Radiology 2008, 1(1), 54–57. 80. Deshpande, N.; Needles, A.; Willmann, J. K. Molecular ultrasound imaging: current status and future directions. Clin. Radiol. 2010, 65, 567–581. 81. Pysz, M.; Gambhir, S. S.; Willmann, J. K. Molecular Imaging: current status and emerging strategies. Clin. Radiol. 2010, 65, 500–515. 82. Deshpande, N.; Pysz, M.; Willmann, J. K. Molecular ultrasound assessment of tumor angiogenesis. Angiogenesis 2010 (Epub ahead of print).
P1: OTA/XYZ P2: ABC c12 JWBS037-Chen February 2, 2011
21:49
Printer Name: Yet to Come
CHAPTER 12
Ultrasound-Based Molecular Imaging Using Nanoagents SRIVALLEESHA MALLIDI, MOHAMMAD MEHRMOHAMMADI, KIMBERLY HOMAN, BO WANG, MIN QU, and TIMOTHY LARSON Department of Biomedical Engineering, University of Texas at Austin, Austin, Texas, USA
KONSTANTIN SOKOLOV Department of Biomedical Engineering, University of Texas at Austin, Austin, Texas, and Department of Medical Physics, University of Texas M.D. Anderson Cancer Center, Houston, Texas, USA
STANISLAV EMELIANOV Department of Biomedical Engineering, University of Texas at Austin, Austin, Texas, USA
12.1 INTRODUCTION The need to understand the anatomical and functional aspects of the human body has led to the development of various noninvasive imaging modalities such as X-ray computed tomography (CT), ultrasonography (US), and magnetic resonance imaging (MRI). These imaging modalities are used extensively to diagnose pathologies such as cancer. In particular, ultrasonography, or ultrasound imaging, has gained popularity due to its excellent temporal resolution, reasonable penetration depth, portability, and low cost. Furthermore, ultrasound imaging is a nonionizing technique with no known adverse bioeffects. Currently, ultrasound imaging is being used in various medical fields ranging from obstetric medicine to cardiovascular applications. Obtaining an ultrasound image can be explained as a three-step process. First, the ultrasonic transducer generates pulses of ultrasound waves that are sent through a patient’s body. These waves then interact with organ boundaries and complex tissues, producing echoes due to reflection or scattering. Finally, the backscattered echoes are detected by the same transducer used for transmission of ultrasound waves. Digital postprocessing of these ultrasound signals results in formation of a gray scale image of the body or tissue cross section. Thus contrast in ultrasound images is due to the difference in the acoustic impedance of the tissues being imaged.
Nanoplatform-Based Molecular Imaging Edited by Xiaoyuan Chen C 2011 John Wiley & Sons, Inc. Copyright
263
P1: OTA/XYZ P2: ABC c12 JWBS037-Chen February 2, 2011
264
21:49
Printer Name: Yet to Come
ULTRASOUND-BASED MOLECULAR IMAGING USING NANOAGENTS
Advancements in the field of molecular biology (i.e., discovery of various diseasespecific biomarkers) coupled with development of ultrasound biomicroscopy [1] (imaging performed at the acoustic frequency range of 40–100 MHz) has prompted development of site-specific ultrasound contrast agents that can provide molecular information in the context of a high-resolution anatomical map of the body. Indeed, microbubbles. For example, microbubbles have enhanced ultrasound contrast due to their acoustic impedance difference with tissue. When targeted to various biomarkers in the vascular lumen, these microbubbles can provide some molecular specific ultrasound imaging contrast. However, the gas-filled microbubbles have short lifespans (on the order of minutes) in the body, and due to their large size (∼0.5–500 m diameter) are mostly used as intravascular tracers [2–7]. Other types of ultrasound contrast agents such as liposomes have been developed for molecular imaging [4, 8]. Liposomes have longer circulation time compared to microbubbles and they can also readily be conjugated to various biomarkers. However, passage of liposomes (∼800 nm in diameter) through endothelial gap junctions in the leaky vascular of pathologies such as cancer is size prohibitive. The vasculature of most cancers have endothelial gap junctions from 300 to 800 nm in size, and therefore liposomes would likely not reach the tumor interstitial space [9]. Recently developed ultrasound nanoparticle contrast agents such as perfluorocarbon nanoparticles [4] and silica nanoparticles [10] could extravasate through the leaky vasculature of a tumor into the interstitial space but are less echogenic compared to microbubbles. Metal based nanoparticles (∼5–100 nm in diameter) are being used extensively as molecular specific contrast agents for various imaging modalities such as optical imaging and MRI. Many researchers have demonstrated that metal nanoparticles extravasate and accumulate in tumors due to the enhanced permeability and retention (EPR) effect [11, 12]. This effect is caused by the leaky nature of tumor vessels. Thus by an injection of correctly sized metal nanoparticles, passive accumulation in tumors can be achieved. Furthermore, the metal nanoparticles can be made pathology-specific by bioconjugating them with monoclonal antibodies or antibody fragments. These bioconjugates can be attached either directly to the metal or covalently bound via linker segments. Metal nanoparticles (NPs) such as gold NPs have well-known bioconjugation protocols. Metal NPs are at the same size scale as large protein complexes and so can be targeted to subcellular structures. However, these metal nanoparticles cannot be used as ultrasound contrast agents because their size is much below the resolution of clinically available ultrasound imaging systems. On the other hand, molecular imaging with the metal NPs is possible using ultrasound based imaging modalities, namely, photoacoustic imaging [13, 14] and magneto-motive ultrasound imaging (MMUS) [15, 16]; that is, molecular and functional information could be obtained in the context of the anatomical map of the tissue. For example, gold nanoparticles exhibiting surface plasmon resonance properties and superparamagnetic iron oxide (SPIO) nanoparticles are used as photoacoustic and magnetoacoustic contrast agents, respectively. The synergy among the ultrasound, photoacoustic, and MMUS imaging modalities is schematically represented in Figure 12.1. The ultrasound transducer and the receiver electronics of an ultrasound imaging system are common for both photoacoustic and magnetoacoustic imaging. In photoacoustic imaging, the tissue is illuminated by nanosecond pulsed laser light and the subsequently emitted photoacoustic waves are detected by the ultrasound transducer (Fig. 12.1). The use of photoabsorbers such as gold nanoparticles enhances the contrast in photoacoustic imaging. In MMUS imaging, ultrasound images of the tissue labeled with magnetic nanoparticles are acquired under magnetic field excitation.
P1: OTA/XYZ P2: ABC c12 JWBS037-Chen February 2, 2011
21:49
Printer Name: Yet to Come
PHOTOACOUSTIC IMAGING
Image display
265
Contrast mechanism RF pulse transmit
Amplitude of received RF echoes
Ultrasound imaging
Amplitude of tissue displacement
Magneto-motive ultrasound imaging Amplitude of photoacoustic transients
Tissue Acoustic Impedance Receive echo
RF pulse transmit
Receive echo
Photoacoustic transients
Motion induced in tissue labeled with magnetic NPs
Harmonic or Pulsed magnetic field
Tissue Optical Absorption
Pulsed Laser
Photoacoustic imaging
FIGURE 12.1 Schematic representation of contrast mechanism and image display in ultrasound, magneto-motive ultrasound, and photoacoustic imaging modalities.
The ultrasound images are then digitally postprocessed to observe the magnetically induced tissue motion (Fig. 12.1). In this chapter, we briefly review the fundamentals of photoacoustic and magnetoacoustic imaging modalities and also demonstrate their molecular imaging capabilities using metal nanoparticles. Furthermore, the feasibility of combining the photoacoustic and magnetoacoustic imaging techniques will be discussed. Finally, the advantages, limitations, and future prospects of ultrasound based photoacoustic and magnetoacoustic imaging modalities are illustrated.
12.2 PHOTOACOUSTIC IMAGING Photoacoustic imaging, also known as optoacoustic and thermoacoustic imaging, involves three steps [17–19]: (1) short laser pulses irradiate the tissue, (2) the tissue absorbs the laser energy; undergoes thermoelastic expansion, and subsequently generates a photoacoustic pressure wave; and (3) the generated pressure wave is detected by an ultrasound transducer. Note that the thermoelastic expansion occurs because the laser pulse duration is shorter than the thermal relaxation time of the tissue (thermal confinement condition). The acoustic pressure P(z) generated at a certain depth z using laser illumination of wavelength can be expressed as [17, 18] P(z) =
cs2 CP
a ()F(z, )
(12.1)
where  is the thermal expansion coefficient, cs is the speed of sound, CP is the heat capacity at constant pressure, a is the optical absorption coefficient, and F(z) is the laser fluence
P1: OTA/XYZ P2: ABC c12 JWBS037-Chen February 2, 2011
266
21:49
Printer Name: Yet to Come
ULTRASOUND-BASED MOLECULAR IMAGING USING NANOAGENTS
at depth z. The photoacoustic pressure wave is a broadband radiofrequency (RF) signal and can be detected using an ultrasound transducer that determines the resolution of the photoacoustic imaging. A resolution on the order of tens of micrometers can be achieved using a high-frequency ultrasound transducer. The detected photoacoustic RF signals or transients are then processed and displayed as a photoacoustic image. The expression (cs2 /CP ) in Eq. (12.1) is called the Gruneisen coefficient, a temperaturedependent parameter. For a constant temperature and laser fluence irradiating the tissue, the photoacoustic signal strength is proportional to the optical absorption coefficient of the tissue. Indeed, the distribution of optical absorption properties of the tissues can be obtained by analyzing photoacoustic images captured at multiple wavelengths. For example, functional information such as the oxygenation of blood or the presence of atherosclerotic plaques may be extracted from multiwavelength photoacoustic images [20, 21]. Photoacoustic imaging exploits both the high contrast associated with optical imaging techniques (due to optical absorption properties) and the spatial resolution of ultrasound imaging (due to detection by the ultrasonic transducer). Unlike optical imaging modalities, where the penetration depth is limited by optical backscattering from the tissue, photoacoustic imaging can image deeper since it detects sound versus light. Moreover, greater penetration depth in tissue can be achieved using near-infrared (NIR) wavelengths because endogenous chromophores such as melanin and blood absorb less light in the NIR range. Similar to optical imaging techniques, the use of exogenous contrast agents or photoabsorbers with higher optical absorption in the optical NIR window could facilitate the detection of pathologies in photoacoustic imaging. Metal nanoparticles, which have greater absorbance compared to conventional dyes such as indocyanine green, qualify as contrast agents in photoacoustic imaging. A variety of shapes and sizes of metal nanoparticles including gold or silver nanospheres, rods, and shells can be used as photoabsorbers [14, 22–25]. It’s apparent that by varying the shape and aspect ratio of nanostructures, particles can be manufactured to absorb light at a desired wavelength across a wide spectrum including the near-infrared spectrum, where the absorption of light by tissue is minimal. Moreover, the change in the optical absorption properties due to the plasmon coupling effect of closely spaced nanoparticle assemblies can also be detected using photoacoustic imaging [14]. 12.2.1 Description of Ultrasound Based Photoacoustic Imaging System A block diagram of the combined ultrasound and photoacoustic imaging system is presented in Figure 12.2. The imaging system primarily involves an integrated probe consisting of a ultrasonic and laser light delivery [13]. The light is delivered to the tissue through a fiberoptic bundle as shown in photographs of an integrated probe with a single element high-frequency transducer or an array transducer, respectively. The laser light delivery could also be done through a combination of various optical elements such as an axicon and prisms [18, 26, 27]. The integrated probe could be handheld or controlled using a positioning stage for mechanical scanning. For special applications such as intravascular photoacoustic imaging, an intravascular ultrasound transducer can be integrated with side fire fiber as shown in Figure 12.2c. The combined imaging system can work either in the ultrasound or photoacoustic imaging mode. In the ultrasound mode, the transducer transmits the rf pulse and receives the backscattered echo from the tissue. In photoacoustic mode, the short laser pulse excites the tissue and the generated acoustic transients are received by the ultrasonic transducer. The
P1: OTA/XYZ P2: ABC c12 JWBS037-Chen February 2, 2011
21:49
Printer Name: Yet to Come
PHOTOACOUSTIC IMAGING
267
Trigger Ultrasound Pulser Ultrasound imaging Mode selection
Positioning stage or handheld
Integrated probe Receive Electronics
Transducer
Digital processing of RF data
Image Display
Fiber bundle Photoacoustic imaging (a)
Trigger
Laser
Transducer
(b)
(c)
Optical Fiber
(d)
FIGURE 12.2 (a) Block diagram of combined ultrasound and photoacoustic imaging system. (b–d) Photographs of an integrated probe consisting of a high-frequency single element transducer (b), an intravascular ultrasound transducer (c), and a linear array transducer (d) with optical delivery system.
ultrasound and photoacoustic rf data can be acquired sequentially from the same imaging plane using a single element transducer or an array transducer. In either case, spatially coregistered ultrasound and photoacoustic images can be obtained [13, 28].
12.2.2 Enhancement of Photoacoustic Contrast Using Silver Nanocages In most works to date, gold nanoparticles have been used, almost exclusively, to enhance photoacoustic contrast. Here we present an alternative: silver nanoagents. Being in the noble metal category, both silver and gold exhibit plasmonic resonance in the visible to NIR spectrum of light. Silver, however, has slightly better light absorption and therefore in theory should be a stronger photoacoustic contrast agent [29, 30]. In our attempts to explore this theory, silver nanocages built around a silica core were developed with core sizes ranging from 180 to 520 nm (Fig. 12.3a). These nanocages absorb light broadly across NIR wavelengths. To test their contrast properties in tissue, these nanocages were injected into an ex vivo porcine pancreas and imaged photoacoustically. The result of the photoacoustic and ultrasound imaging performed is shown in Figure 12.3b. Specifically, the combined ultrasound and photoacoustic system with a 7.5-MHz linear array ultrasound transducer and 800-nm laser illumination was employed to image silver nanocages injected into an ex vivo porcine pancreas. The 183-nm silica core, silver outer cage particles (100 L
P1: OTA/XYZ P2: ABC c12 JWBS037-Chen February 2, 2011
268
21:49
Printer Name: Yet to Come
ULTRASOUND-BASED MOLECULAR IMAGING USING NANOAGENTS
(a) Ultrasound
Photoacoustic
Combined
Image
Image
Image
(b)
(c)
(d)
FIGURE 12.3 (a) SEM image of silver nanocages. (b) Ultrasound, (c) photoacoustic, and (d) combined images of ex vivo pancreas tissue injected with silver nanocages. Adapted with permission from Ref. 55.
of 1010 particles/mL) were injected via syringe, approximately 8–10 mm below the pancreas surface. Spatially coregistered ultrasound and photoacoustic rf data was captured using a Cortex ultrasound imaging system (Winprobe Corporation, North Palm Beach, FL, USA). The ultrasound image (Fig. 12.3b) defines the pancreas area, the photoacoustic image (Fig. 12.3c) shows the signal received from the nanocages (white inset), and the combined image (Fig. 12.3d) clearly depicts the location of the nanocages against the background ultrasonic image of the organ. Thus if these nanocages were injected systematically and accumulated in a cancerous area, then combined ultrasound and photoacoustic images could be used to locate them inside the tissue, helping clinicians to better define diseased areas. In this example silver nanoparticles were used instead of the well-characterized gold nanoparticles. Silver nanoparticles have both advantages and disadvantages. One known advantage is silver’s ability to absorb and scatter light better than gold [29]. One disadvantage is silver’s reactivity. Gold is relatively inert, owing to its biocompatibility. Under certain conditions, silver is more reactive and has been shown to be cytotoxic in some in vitro studies [31]. However, silver is well known for its antimicrobial properties and has been used for decades in the treatment of burns [32]. In recent years, silver is being used to line catheters to decrease infection levels [33, 34]. Thus, there is a resurgence in the use of silver in biomedical applications and future studies will be needed to determine its true efficacy and biocompatibility.
12.2.3 Monitoring Accumulation of Gold Nanoparticles in Tumor Using Photoacoustic Microscopy Photoacoustic imaging can also be used to monitor accumulation of nanoparticles in the tumor site over time. A tumor xenograft with A431 cells (human epithelial cancer cell line) was performed in nude mice. After the tumor reached approximately 10 mm in diameter, gold nanoparticles (40 nm in diameter) specifically targeted to the epidermal growth factor receptor [35] (EGFR) are imaged using single element transducer (25 MHz central frequency) and 532-nm wavelength illumination. The combined ultrasound and photoacoustic images of the tumor before and after the injection of gold nanoparticles are shown in Figure 12.4. Clearly, the accumulation of gold NPs over time is evident (Fig. 12.4f–h).
P1: OTA/XYZ P2: ABC c12 JWBS037-Chen February 2, 2011
21:49
Printer Name: Yet to Come
MAGNETO-MOTIVE ULTRASOUND IMAGING
Before injection of Au NPs
269
After injection of Au NPs 0 hrs
2 hrs
4 hrs
(a)
(b)
(c)
(d)
(e)
(f)
(g)
(h)
Ultrasound images
Photoacoustic images
FIGURE 12.4 (a) Monitoring the accumulation of 40-nm gold nanoparticles in tumor using combined ultrasound (a–d) and photoacoustic (e–h) imaging at 532-nm wavelength illumination. The images represent a 13-mm × 11.25-mm field of view.
12.3 MAGNETO-MOTIVE ULTRASOUND IMAGING Magnetoacoustic imaging is a technique that employs high-frequency ultrasound to visualize structural and physiological properties of magnetically labeled tissues. The magneto-acoustic imaging technique capitalizes on significant contrast between magnetic susceptibility of normal tissue constituents and magnetic nanoparticles as well as on deep penetration of magnetic fields into human tissue. Specifically, the magnetoacoustic imaging procedure involves two steps: (1) magnetic nanoparticles such as superparamagnetic iron oxide (SPIO) nanoparticles are used to specifically label pathological tissues, and (2) a pulsed or time-varying magnetic field excites the magnetic nanoparticles to induce motion in the magnetically labeled tissues, along with simultaneous detection of the motion using noninvasive, deep penetrating ultrasound imaging techniques [16, 24, 36–38]. In the last few decades, much research has been devoted to the synthesis of magnetic nanoparticles. Several types of commercial or custom designed nanoparticles such as the superparamagnetic iron oxide (SPIO) nanoparticles were approved by the FDA and has been widely used in clinical applications as MRI contrast agent [36, 39]. Magnetic nanoparticles have been synthesized with a number of different compositions and phases, including iron oxides (Fe3 O4 and ␥ -Fe2 O3 ), pure metals (Fe,Co), spinel-type ferromagnets (MgFe2 O4 , MnFe2 O4 , and CoFe2 O4 ), and alloys (CoPt3 and FePt). A review of the synthesis and functionalization of magnetic nanoparticles has been provided elsewhere [40, 41]. The displacement occurring in the magnetically labeled tissue is dependent on several parameters such as the applied magnetic field and the susceptibility of the magnetic NPs. To understand and estimate the induced motion in the magnetically labeled tissue, the characteristics of the nanoparticles and the magneto-motive force acting on the magnetic nanoparticles need to be analyzed. The magneto-motive force acting on a particle can be expressed as Fm = (m • ∇)B,
(12.2)
P1: OTA/XYZ P2: ABC c12 JWBS037-Chen February 2, 2011
270
21:49
Printer Name: Yet to Come
ULTRASOUND-BASED MOLECULAR IMAGING USING NANOAGENTS
where m is the magnetic moment. Considering a z-directional magnetic flux B = (0,0,Bz ) and the magnetic moment m of paramagnetic particles as (0,0,mz ), the magneto-motive force F m can then be expressed as Fm = (m z • ∇)Bz ,
(12.3)
The magnetic moment mz experienced by a magnetic nanoparticle located in a weakly diamagnetic medium like tissue can be written m z = Vm Mz
(12.4)
where V m is the volume of the magnetic portion of the nanoparticle and can be described as Vm = Vnp × f m , where V np is the total size of the nanoparticle and f m is dimensionless factor called the fraction of magnetite and represents the volumetric ratio of magnetic material in a nanoparticle. The volumetric magnetization M can be written as Mz = ( np − medium )Hz . The magnetic susceptibility of the medium, that is, human tissue, is on order of 10−6 (−11 × 10−6 ≤ Tissue ≥ −7 × 10−6 ) while the susceptibility of magnetite nanoparticles (Fe3 O4 ) is 70 and it is 250 and 600 for other nickel and cobalt magnetic agents [42, 43]. Hence medium is assumed to be negligible. Another assumption is that the magnetic flux density (B) doesn’t change significantly over the nanoparticle due to its small size; the volumetric magnetization M can be expressed as Mz = np
Bz 0
(12.5)
Therefore from Eqs. (12.3) and (12.4), Fm =
Vnp f m np (Bz • ∇)Bz . 0
(12.6)
Since (Bz • ∇)Bz = 12 ∇(Bz • Bz ) = Bz
∂ Bz , ∂z
the magneto-motive force acting on a nanoparticles F m can be expressed as Fm =
Vnp f m np ∂ Bz Bz 0 ∂z
(12.7)
Clearly, from Eq. (12.7) it can be seen that there is a significant contrast in the force experienced by the normal and iron-laden tissue. The magneto-motive force is proportional to size (V np f m ) and the susceptibility ( np ) of nanoparticles. Increasing the size of the nanoparticles has some limitations for specific labeling or cellular uptake as larger nanoparticles are harder to diffuse out of the leaky vessels. Therefore the magnetic susceptibility ( np ) of nanoparticles plays an important role in determining the sensitivity of the magnetoacoustic imaging technique. Iron oxide nanoparticles have high magnetization up to 70 emu/gram; Fe, but other magnetic materials like nickel and cobalt or some alloys can have higher magnetization [44–46]. However, safety and toxicity issues of these materials are still subject to more investigation. Encapsulation of the toxic magnetic alloy core into
P1: OTA/XYZ P2: ABC c12 JWBS037-Chen February 2, 2011
21:49
Printer Name: Yet to Come
MAGNETO-MOTIVE ULTRASOUND IMAGING
271
a carbon or polymer shell could be one of the solutions to reduce the cytotoxicity of the metal nanoparticles [47, 48]. The excitation magnetic field can be either continuous-time (harmonic) or pulsed mode. Pulsed excitation has several advantages over harmonic excitation such as shorter operation time and increased magnetic flux density, thus allowing the imaging of deeper tissue structures [15]. Pulsed excitation also has less severe thermal management constraints. On the other hand, a harmonic system creates motion at a predetermined frequency and possibly can be used to identify and filter out the sources of tissue motion due to cardiac and respiratory systems. 12.3.1 Description of Magneto-motive Ultrasound Imaging System The block diagram of the magnetoacoustic imaging system is shown in Figure 12.5. The system consists of two major parts: (1) the magnetic field generation and (2) the ultrasound imaging and data acquisition. According to Ampere’s law, the magnetic field of a solenoid coil, B, has the same time characteristics as the supplied current. Hence a controllable current amplifier or a flash circuit could be used to generate a harmonic or pulsed magnetic field. A conical iron core can also be incorporated into the solenoid to maximize and localize the magnetic field strength applied to tissue specimens [16, 36]. The displacement or the induced motion in the tissue can be monitored using an ultrasound imaging system equipped with an ultrasonic array transducer or with a single element high-frequency transducer. In magnetoacoustic imaging, it is important to capture ultrasound RF data before, during and after the application before, during and after the application of the magnetic excitation to extract the relative motion with respect to the stationary reference. The induced motion can
Trigger generator
Ultrasound Transducer
Controllable delay Positioning Control
Control Unit
Ultrasound Pulser/Receiver
Sample
Stepper Motor Magnetic coil Iron Core
Current Amplifier
Waveform Generator
Signal Acquisition A/D converter Speckle tracking module US/pMMUS image formation
Processing Unit
Cooling casing
FIGURE 12.5 Block diagram of the combined ultrasound and magneto-motive ultrasound imaging system.
P1: OTA/XYZ P2: ABC c12 JWBS037-Chen February 2, 2011
272
21:49
Printer Name: Yet to Come
ULTRASOUND-BASED MOLECULAR IMAGING USING NANOAGENTS
be detected using various quantitative and qualitative techniques including phase-shift or block-matching speckle tracking, Doppler ultrasound, and color/power Doppler. 12.3.2 Demonstration of Magneto-motive Ultrasound Imaging Using Magnetic Iron Oxide Nanoparticles (MIONs) The ability of magnetoacoustic imaging was demonstrated using tissue phantoms fabricated with the J774A.1 cell line (mouse monocytes–macrophages). Like most cells of macrophage phenotype, the J774A.1 cells rapidly uptake dextran-coated nanoparticles [49–51]. Two tissue phantoms were prepared for the study: (1) a control phantom with macrophage cells only and (2) a labeled phantom prepared with MION loaded macrophage cells. Briefly, the cells were cultured in Dulbecco’s Modified Eagle Media (DMEM), supplemented with 5% fetal bovine serum (FBS) at 37 ◦ C in 5% CO2 . To label cells with MIONs, they were incubated with a suspension of nanoparticles in culturing media. After 24 h of incubation with the nanoparticles, the cells were harvested and suspended in a warm (35 ◦ C) gelatin solution (8% w/v) containing 0.5% silica particles (30 m in diameter). The concentration of the control (unloaded) cells in gelatin was approximately equal to those in the loaded cells. The gelatin suspension with cells (control or labeled) was then pipetted into rubber spacers placed in a petri dish. The gelatin solution was allowed to harden at room temperature for approximately 10 min. In addition, a 1–2-mm thick pure gelatin layer was placed on top of the gelatin layer with cells. Finally, to facilitate imaging, the petri dish was filled with 1× PBS solution (phosphate buffered saline) to maintain the appropriate pH in the medium surrounding the tissue phantoms. The cells from each tissue phantom were imaged optically using a Leica DM 6000 upright microscope in epi-illuminated darkfield mode (Fig. 12.6a,b). Images were collected through a 20×, 0.5 NA darkfield objective and detected using an ultrasensitive 12-bit CCD camera. The unlabeled macrophage cells appear bluish white due to their intrinsic light scattering properties (Fig. 12.6a). The macrophage cells labeled with iron nanoparticles appear as orange regions due to light scattering of iron oxide nanoparticles (Fig. 12.6b). The B-scan and the color Doppler images of the tissue phantoms were obtained using a commercially available ultrasound imaging system (Sonic RP by Ultrasonix, Inc.) equipped with a linear array transducer (38-mm aperture, 5-MHz central frequency). The B-scan images
(a)
(c)
(e)
(d)
(f)
20 m
(b)
FIGURE 12.6 Optical dark-field images of (a) nonlabeled and (b) magnetically labeled macrophages. B-scan ultrasound images (c,d) and color Doppler images (e,f) of nonlabeled and labeled macrophages embedded in 8% gelatin background.
P1: OTA/XYZ P2: ABC c12 JWBS037-Chen February 2, 2011
21:49
Printer Name: Yet to Come
COMBINED PHOTOACOUSTIC AND MAGNETO-MOTIVE ULTRASOUND IMAGING
273
cannot directly differentiate between the samples as the difference in image contrast is only derived from nonspecific variations in the phantom materials. The color Doppler image of the phantom with labeled cells clearly indicates the presence of tissue motion and is different compared to that of the control phantom (no motion). The results confirm that magnetoacoustic imaging can readily detect the presence of MIONs in the cells and tissue [16].
12.4 COMBINED PHOTOACOUSTIC AND MAGNETO-MOTIVE ULTRASOUND IMAGING Currently, in the field of molecular imaging, emphasis is being laid on multimodality nano contrast agents that can provide complementary functional information regarding pathologies such as cancer. For example, core–satellite structured dual functional nanoparticles comprised of a dye-doped silica “core” and multiple “satellites” of magnetic nanoparticles were utilized for optical and MR imaging of neuroblastoma cells [52]. Hybrid gold coated iron oxide nanoparticles were also used for combined optical and MR imaging of cancer cells [53]. A combination of MR and optical imaging techniques provides both anatomical and functional information on pathology. However, obtaining the images using the two techniques at the same spatial cross section could be challenging as the imaging equipments required vary significantly. On the other hand, photoacoustic and magnetomotive ultrasound imaging modalities can utilize these hybrid nanostructures possessing both optical absorption properties and magnetic properties as contrast agents. Moreover, photoacoustic and magneto-motive ultrasound imaging techniques utilize the same receiver electronics as ultrasound imaging (Fig. 12.1), and hence can be transparently integrated to provide complementary functional and morphological information at the same imaging cross section. Tissue mimicking samples made with polyvinyl alcohol (PVA) were used to demonstrate the feasibility of combined ultrasound, photoacoustic, and magneto-motive ultrasound imaging. The procedure to make PVA inclusions has been described elsewhere [28]. Briefly, the first inclusion (Fig. 12.7a) was made with 8% PVA and contained 15 micro-meter silica powder (0.5% w/v) for ultrasound contrast, 40-nm diameter spherical gold nanoparticles for optical contrast, and 20-nm (8-nm magnetic core) diameter iron oxide nanoparticles
FIGURE 12.7 Schematic representation (left) of phantom used for combined ultrasound, photoacoustic, and magneto-motive imaging. The phantom consisted of PVA inclusions with (a) mixture of Au and Fe3 O4 iron nanoparticles and (b) no nanoparticles. (c) Ultrasound, (d) photoacoustic, and (e) magneto-motive images of the PVA inclusions. The images represent a 2.5 mm × 3.5 mm field of view. Adapted with permission from Ref. 56.
P1: OTA/XYZ P2: ABC c12 JWBS037-Chen February 2, 2011
274
21:49
Printer Name: Yet to Come
ULTRASOUND-BASED MOLECULAR IMAGING USING NANOAGENTS
for magnetic contrast. The second inclusion (Fig. 12.7b) was also fabricated using 8% PVA, but it contained only silica powder and no nanoparticles—that is, the inclusion had only ultrasound scatters. The inclusions were placed in a water tank attached to a threedimensional (3D) positioning stage to facilitate mechanical scanning. A schematic representation of the experimental setup used for combined ultrasound photoacoustic and MMUS imaging is shown in Figure 12.7 (left panel). The integrated probe (ultrasound transducer and optical fiber delivery system) and the magnetic solenoid are placed on opposite sides of the sample. A single element focused transducer operating at 48-MHz central frequency was used to detect ultrasound and photoacoustic signals. A Q-switched Nd:YAG laser operating at 532-nm wavelength was used to generate photoacoustic transients. A magnetic pulse excitation of approximately 0.5 tesla was delivered to the sample via the conical iron core incorporated in a solenoid. At a particular spatial location, ultrasonic A-lines (with and without magnetic excitation) followed by photoacoustic A-lines were acquired. The sample was mechanically moved to the next spatial location using the 3D positioning stage to acquire the next set of ultrasonic and photoacoustic A-lines. The step size of the mechanical scan was determined by the lateral resolution of the ultrasound transducer. During offline processing of the acquired RF data, a digital bandpass filter was applied to increase the signal-to-noise ratio. The ultrasound and photoacoustic signals were extracted from the A-line records. The analytic signals were obtained by applying a Hilbert transform on the filtered ultrasound and photoacoustic signals and the images were displayed after spatial interpolation. To obtain a MMUS image, the displacement of the inclusions due to pulsed magnetic excitation was determined by correlating ultrasonic A-line records acquired with and without magnetic excitation. The procedure was repeated for each A-line record acquired at different spatial locations and the displacements obtained were plotted as an image after spatial interpolation. The ultrasound, photoacoustic, and MMUS images of the inclusions are shown in Figure 12.7c–e, respectively. The spatial location and the anatomical shape of the two inclusions are clearly indicated in the ultrasound image. Note how, in the photoacoustic image, only one of the inclusions produced photoacoustic signals. This observation was expected since the first inclusion contained gold nanoparticles while no photoacoustic absorbers were present in the second inclusion. The displacement of the inclusions in the MMUS image is displayed on a color map, where white indicates maximum displacement of 100 m and black indicates zero displacement. Clearly, the inclusion with iron nanoparticles displaced more under magnetic excitation than the second inclusion with no iron nanoparticles. The results in Figure 12.7 clearly indicate the feasibility of combining ultrasonic, photoacoustic, and MMUS imaging. In the current experimental setup, the magnetic coil and the integrated probe are placed on opposite sides of the sample. This configuration might not be feasible for in vivo applications. A transducer with an optical fiber in the center for laser light delivery [54] and magnetic coil surrounding it (Fig. 12.5) could make an effective probe for the combined ultrasound, photoacoustic, and MMUS imaging.
12.5 CONCLUSIONS The photoacoustic and magnetoacoustic imaging modalities can be integrated with the widely available ultrasound imaging system with minimum effort. The excellent features of the combined ultrasound, photoacoustic, and magnetoacoustic imaging
P1: OTA/XYZ P2: ABC c12 JWBS037-Chen February 2, 2011
21:49
Printer Name: Yet to Come
REFERENCES
275
system—noninvasive, nonionizing, descent spatial resolution, and extended penetration depth—can all be obtained at low cost. Moreover, these complementary imaging modalities can simultaneously provide anatomical, optical, and biomechanical properties of the tissue. Given the availability of complex nanostructures with high magnetic susceptibility and optical absorption properties, we anticipate the combination of ultrasound guided photoacoustic and magnetoacoustic imaging techniques could be a valuable tool for early detection of pathologies such as cancer. Further studies are required to evaluate this molecular imaging technique in vivo.
ACKNOWLEDGMENTS Partial support from the National Institutes of Health under grants EB008101 and EB008821 is gratefully acknowledged. The authors would like to thank Dr. Salavat Aglyamov and Dr. Andrei Karpiouk of the University of Texas at Austin for their valuable inputs regarding the development of combined imaging system.
REFERENCES 1. Foster, F. S.; Pavlin, C. J.; Lockwood, G. R.; Ryan, L. K.; Harasiewicz, K. A.; Berube, L.; Rauth, M. A. Principles and applications of ultrasound backscatter microscopy. IEEE Trans. Ultrason Ferr. 1993, 40, 608–617. 2. Dayton, P. A.; Ferrara, K. W. Targeted imaging using ultrasound. J. Magn. Reson. Imaging 2002, 16, 362–377. 3. Kaufmann, B. A.; Lindner, J. R. Molecular imaging with targeted contrast ultrasound. Curr. Opin. Biotechnol. 2007, 18, 11–16. 4. Lanza, G. M.; Wickline, S. A. Targeted ultrasonic contrast agents for molecular imaging and therapy. Curr. Problems Cardiol. 2003, 28, 625–653. 5. Liang, H. D.; Blomley, M. J. K. The role of ultrasound in molecular imaging. Br. J. Radiol. 2003, 76, S140–S150, 6. Klibanov, A. L. Targeted microbubbles: ultrasound contrast agents for molecular imaging. In Nanoparticles in Biomedical Imaging: Emerging Technologies and Applications, J. W. M. Bulte and M. M. J. Modo, Ed., Springer, New York, 2008, pp. 327–341. 7. Tsutsui, J.; Xie, F.; Porter, R. The use of microbubbles to target drug delivery. Cardiovasc. Ultrasound, 2004, 2, 23. 8. Demos, S. M.; Alkan-Onyuksel, H.; Kane, B. J.; Ramani, K.; Nagaraj, A.; Greene, R.; Klegerman, M.; McPherson, D. D. In vivo targeting of acoustically reflective liposomes for intravascular and transvascular ultrasonic enhancement. J. Am. Coll. Cardiol. 1999, 33, 867–875. 9. Brigger, I.; Dubernet, C.; Couvreur, P. Nanoparticles in cancer therapy and diagnosis. Adv. Drug Deliv. Rev. 2002, 54, 631–651. 10. Jun, L.; Andrea, L. L.; John, S. M.; Mamoru, Y.; Robert, J. L.; Xueliang, P.; Thomas, J. R. Nanoparticles as image enhancing agents for ultrasonography. Phys. Med. Biol. 2006, 2179. 11. Hobbs, S. K.; Monsky, W. L.; Yuan, F.; Roberts, W. G.; Griffith, L.; Torchilin, V. P.; Jain, R. K. Regulation of transport pathways in tumor vessels: role of tumor type and microenvironment. Proc. Natl. Acad. Sci. U.S.A. 1998, 95, 4607–4612. 12. Kong, G.; Braun, R. D.; Dewhirst, M. W. Hyperthermia enables tumor-specific nanoparticle delivery: effect of particle size. Cancer Res. 2000, 60, 4440–4445.
P1: OTA/XYZ P2: ABC c12 JWBS037-Chen February 2, 2011
276
21:49
Printer Name: Yet to Come
ULTRASOUND-BASED MOLECULAR IMAGING USING NANOAGENTS
13. Emelianov, S. Y.; Aglyamov, S. R.; Karpiouk, A. B.; Mallidi, S.; Park, S.; Sethuraman, S.; Shah, J.; Smalling, R. W.; Rubin, J. M.; Scott, W. G. Synergy and applications of combined ultrasound, elasticity, and photoacoustic imaging. Presented at IEEE International Ultrasonics Symposium, Vancouver, Canada, 2006. 14. Mallidi, S.; Larson, T.; Aaron, J.; Sokolov, K.; Emelianov, S. Molecular specific optoacoustic imaging with plasmonic nanoparticles. Optics Express. 2007, 15, 6583–6588. 15. Mehrmohammadi, M.; Oh, J.; Ma, L.; Ryoo, S.; Yantsen, E.; Mallidi, S.; Johnston, K. P.; Sokolov, K.; Milner, T. E.; Emelianov, S. Y. Pulsed magneto-motive ultrasound imaging. Presented at HSEMB, Houston, 2008. 16. Mehrmohammadi, M.; Oh, J.; Ma, L.; Yantsen, E.; Larson, T.; Mallidi, S.; Park, S.; Johnston, K. P.; Sokolov, K.; Milner, T.; Emelianov, S. Imaging of iron oxide nanoparticles using magnetomotive ultrasound. Presented at Proceedings of the 2007 IEEE Ultrasonics Symposium, New York, 2007. 17. Oraevsky, A.; Karabutov, A. Optoacoustic tomography. In Biomedical Photonics Handbook; Vo-Dinh, T., Ed.; CRC Press: Boca Raton, FL, 2003; Vol. PM125, Chap. 34, pp. 34/1–34/34. 18. Xu, M.; Wang, L. V. Photoacoustic imaging in biomedicine. Rev. Sci. Instrum. 2006, 77, 041101. 19. Robert, A. K.; Pingyu, L.; Richard, F. Y.; Appledorn, C. R. Photoacoustic ultrasound (PAUS)—Reconstruction tomography. Med. Phys. 1995, 22, 1605–1609. 20. Sethuraman, S.; Amirian, J. H.; Litovsky, S. H.; Smalling, R. W.; Emelianov, S. Y. Spectroscopic intravascular photoacoustic imaging to differentiate atherosclerotic plaques. Optics Express 2008, 16, 3362–3367. 21. Zhang, H. F.; Maslov, K.; Stoica, G.; Wang, L. V. Functional photoacoustic microscopy for high-resolution and noninvasive in vivo imaging. Nat. Biotechnol. 2006, 24, 848–851. 22. Copland, J. A.; Eghtedari, M.; Popov, V. L.; Kotov, N.; Mamedova, N.; Motamedi, M.; Oraevsky, A. A. Bioconjugated gold nanoparticles as a molecular based contrast agent: implications for imaging of deep tumors using optoacoustic tomography. Mol. Imaging Biol. 2004, 6, 341–349. 23. Loo, C.; Lin, A.; Hirsch, L.; Lee, M. H.; Barton, J.; Halas, N.; West, J.; Drezek, R. Nanoshellenabled photonics-based imaging and therapy of cancer. Technol Cancer Res. Treat 2004, 3, 33–40. 24. Oh, J.; Feldman, M. D.; Kim, J.; Kang, H. W.; Sanghi, P.; Milner, T. E. Magneto-motive detection of tissue-based macrophages by differential phase optical coherence tomography. Lasers Surg. Med. 2007, 39, 266–272. 25. Yang, X.; Skrabalak, S. E.; Li, Z. Y.; Xia, Y.; Wang, L. V. Photoacoustic tomography of a rat cerebral cortex in vivo with au nanocages as an optical contrast agent. Nano Lett. 2007, 7, 3798–3802. 26. Mallidi, S.; Larson, T.; Aaron, J.; Sokolov, K.; Emelianov, S. Molecular specific optoacoustic imaging with plasmonic nanoparticles. Optics Express 2007, 15, 6583–6588. 27. Niederhauser, J. J.; Jaeger, M.; Lemor, R.; Weber, P.; Frenz, M. Combined ultrasound and optoacoustic system for real-time high-contrast vascular imaging in vivo. IEEE Trans. Med. Imaging 2005, 24, 436–440. 28. Mallidi, S.; Aglyamov, S. R.; Karpiouk, A. B.; Park, S.; Emelianov, S. Y. Functional and morphological ultrasonic biomicroscopy for tissue engineers. Presented at SPIE Medical Imaging Symposium: Ultrasonic Imaging and Signal Processing, San Diego, California, 2006. 29. Jackson, J. B.; Halas, N. J. Silver nanoshells: variations in morphologies and optical properties. J. Phys. Chem. B. 2001, 105, 2743–2746. 30. Jiang, Z.-J.; Liu, C.-Y. Seed-mediated growth technique for the preparation of a silver nanoshell on a silica sphere. J. Phys. Chem. B. 2003, 107, 12411–12415. 31. Poon, V. K.; Burd, A. In vitro cytotoxity of silver: implication for clinical wound care. Burns 2004, 30, 140–147.
P1: OTA/XYZ P2: ABC c12 JWBS037-Chen February 2, 2011
21:49
Printer Name: Yet to Come
REFERENCES
277
32. Atiyeh, B. S.; Costagliola, M.; Hayek, S. N.; Dibo, S. A. Effect of silver on burn wound infection control and healing: Review of the literature. Burns 2007, 33(2), 139–148. 33. Goldschmidt, H.; Hahn, U.; Salwender, H. J.; Haas, R.; Jansen, B.; Wolbring, P.; Rinck, M.; Hunstein, W. Prevention of catheter-related infections by silver coated central venous catheters in oncological patients. Zentralbl. Bakteriol. 1995, 283, 215–223. 34. Liedberg, H.; Lundeberg, T. Silver coating of urinary catheters prevents adherence and growth of Pseudomonas aeruginosa. Urol. Res. 1989, 17, 357–358. 35. Kumar, S.; Aaron, J.; Sokolov, K. Directional conjugation of antibodies to nanoparticles for synthesis of multiplexed optical contrast agents with both delivery and targeting moieties. Nat. Protoc. 2008, 3, 314–320. 36. Oh, J.; Feldman, M. D.; Kim, J.; Condit, C.; Emelianov, S.; Milner, T. E. Detection of magnetic nanoparticles in tissue using magneto-motive ultrasound. Nanotechnology 2006, 17, 4183–4190. 37. Oh, J.; Sanghi, P.; Kim, J.; Kemp, N.; Emelianov, S.; Feldman, M. D.; Milner, T. E. Magnetomotive detection of superparamagnetic iron oxide nanoparticles using optical coherence tomography and ultrasound. Presented at ASLMS 26th Annual Meeting, Boston, 2006. 38. Oldenburg, A. L.; Gunther, J. R.; Boppart, S. A. Imaging magnetically labeled cells with magnetomotive optical coherence tomography. Optics Lett. 2005, 30, 747–749. 39. Pankhurst, Q. A.; Connolly, J.; Jones, S. K.; Dobson, J. Applications of magnetic nanoparticles in boimedicine. J. Phys. 2003, 36, R167–R181. 40. Lu, A. H.; Salabas, E. L.; Schuth, F. Magnetic nanoparticles: synthesis, protection, functionalization, and application. Angew. Chem. Int. Ed. Engl. 2007, 46, 1222–1244. 41. Laurent, S.; Forge, D.; Port, M.; Roch, A.; Robic, C.; Vander Elst, L.; Muller, R. N. Magnetic iron oxide nanoparticles: synthesis, stabilization, vectorization, physicochemical characterizations, and biological applications. Chem. Rev. 2008, 108, 2064–2110. 42. Schenck, J. F. MR safety at high magnetic fields. Magn. Reson. Imaging Clin. North Am. 1998, 6, 715–730. 43. Schenck, J. F. Safety of strong, static magnetic fields. J. Magn. Reson. Imaging 2000, 12, 2–19. 44. Bai, J.; Wanga, J.-P. High-magnetic-moment core–shell-type FeCo–Au/Ag nanoparticles. Appl. Phys. Lett. 2005, 87, 152502 45. Hsiao, J.-K. et al. Labelling of cultured macrophages with novel magnetic nanoparticles. J. Magn. Magn. Mater. 2006, 304, e4–e6. 46. Schenck, J. F. The role of magnetic susceptibility in magnetic resonance imaging: MRI magnetic compatibility of the first and second kinds. Med. Phys. 1996, 23, 815–850. 47. Sun, X.; Gutierrez, A.; Yacaman, M. J.; Dong, X.; Jin, S. Investigations on magnetic properties and structure for carbon encapsulated nanoparticles of Fe, Co, Ni. Mater. Sci. Eng. 2000, A286, 157–160. 48. Wei, X. W.; Zhu, G. X.; Xia, C. J.; Ye, Y. A solution phase fabrication of magnetic nanoparticles encapsulated in carbon. Nanotechnology 2006, 17, 4307–4311. 49. Trivedi, R. A.; U-King-Im, J.-M.; Graves, M. J.; Cross, J. J.; Horsley, J.; Goddard, M. J.; Skepper, J. N.; Quartey, G.; Warburton, E.; Joubert, I.; Wang, L.; Kirkpatrick, P. J.; Brown, J.; Gillard, J. H. In vivo detection of macrophages in human carotid atheroma: temporal dependence of ultrasmall superparamagnetic particles of iron oxide-enhanced MRI. Stroke 2004, 35, 1631–1635. 50. Beduneau, A.; Ma, Z.; Grotepas, C. B.; Kabanov, A.; Rabinow, B. E.; Gong, N.; Mosley, R. L.; Dou, H.; Boska, M. D.; Gendelman, H. E. Facilitated monocyte–macrophage uptake and tissue distribution of superparmagnetic iron-oxide nanoparticles. PLoS ONE 2009, 4, e4343, 51. Kooi, M. E.; Cappendijk, V. C.; Cleutjens, K. B. J. M.; Kessels, A. G. H.; Kitslaar, P. J. E. H. M.; Borgers, M.; Frederik, P. M.; Daemen, M. J. A. P.; van Engelshoven, J. M. A. Accumulation
P1: OTA/XYZ P2: ABC c12 JWBS037-Chen February 2, 2011
278
52.
53.
54.
55.
56.
21:49
Printer Name: Yet to Come
ULTRASOUND-BASED MOLECULAR IMAGING USING NANOAGENTS
of ultrasmall superparamagnetic particles of iron oxide in human atherosclerotic plaques can be detected by in vivo magnetic resonance imaging. Circulation 2003, 107, 2453–2458. Lee, J.-H., Jun, Y.-W., Yeon, S.-I., Shin, J.-S.; Cheon, J. Dual-mode nanoparticle probes for high-performance magnetic resonance and fluorescence imaging of neuroblastoma13. Angew. Chem. Int. Ed. 2006, 45, 8160–8162. Larson, T. A.; Bankson, J.; Aaron, J.; Sokolov, K. Hybrid plasmonic magnetic nanoparticles as molecular specific agents for MRI/optical imaging and photothermal therapy of cancer cells. Nanotechnology 2007, 325101. Roy, G. M. K.; Erwin, H.; Wiendelt, S.; Ton, G. v. L.; Frits, F. M. d. M. Photoacoustic imaging of blood vessels with a double-ring sensor featuring a narrow angular aperture. J. Biomed. Optics 2004, 9, 1327–1335. Homan, K.; Shah, J.; Gomez, S.; Gensler, H.; Karpiouk, A.; Brannon-Peppas, L.; Emelianov, S. Silver nanosystems for photoacoustic imaging and image-guided therapy. J. Biomed. Optics 2010, 15, 021316. Qu, M.; Mallidi, S.; Mehrmohammadi, M.; Ma, L. L.; Johnston, K.; Sokolov, K.; Emelianov, S. Y. Phantom study with combined photoacoustic and magneto-acoustic imaging technique, Proceedings of the 31st Annual International IEEE EMBS Conference, 2009, 4763–4766.
P1: OTA/XYZ P2: ABC c13 JWBS037-Chen January 25, 2011
12:49
Printer Name: Yet to Come
CHAPTER 13
MRI Contrast Agents Based on Inorganic Nanoparticles HYON BIN NA and TAEGHWAN HYEON National Creative Research Initiative Center for Oxide Nanocrystalline Materials, and School of Chemical and Biological Engineering, Seoul National University, Seoul, South Korea
13.1 INTRODUCTION Magnetic resonance imaging (MRI) is currently one of the most powerful diagnosis tools in medical science [1]. MRI produces images through monitoring the relaxation processes of water protons under a magnetic field. With this technique, it is possible to obtain realtime images of the internal anatomy and physiology of living organisms in a noninvasive manner. Since it can give anatomic images of soft tissue with high resolution, it has been the preferred tool for imaging the brain and the central nervous system, for assessing cardiac function, and for detecting tumors. Although MRI itself gives detailed images, making a diagnosis based purely on the resulting images may not be accurate since normal tissues and lesions often show small differences in relaxation time. There are several strategies to obtain high-resolution MR images such as the use of high magnetic field and the design of coil systems. In economical and practical terms, it is more feasible to develop supplements that can maximize the ability of imaging tools. One of the most effective supplements is a chemical compound known as a contrast agent that is introduced to a living body for the improvement of visibility in the image. In particular, biological and functional information can be obtained in image form as a result of the interrelation of the contrast agent and the biological system. Therefore a MRI contrast agent is an essential research field in biological and medical sciences to supply a vision for the analysis of biological information and the diagnosis of diseases. Most of the presently available MRI contrast agents are paramagnetic complexes, usually gadolinium (Gd3+ ) chelates [2]. Among them, Gd-DTPA has been the most widely used. Its main clinical applications are focused on detecting the breakage of the blood–brain barrier (BBB) and on changes in vascularity, flow dynamics, and perfusion. Twenty years ago, a different class of contrast agent, superparamagnetic iron oxide (SPIO), was developed, and it has received great attention as a liver contrasting agent [3]. It was the first nanoparticulate MRI contrast agent and is still used clinically. Gd-based contrast agents Nanoplatform-Based Molecular Imaging Edited by Xiaoyuan Chen C 2011 John Wiley & Sons, Inc. Copyright
279
P1: OTA/XYZ P2: ABC c13 JWBS037-Chen January 25, 2011
280
12:49
Printer Name: Yet to Come
MRI CONTRAST AGENTS BASED ON INORGANIC NANOPARTICLES
enhance the signal in T1-weighted images [2]. On the other hand, SPIO provides a very strong contrast effect in T2-weighted images due to its different contrasting mechanism [3]. Furthermore, its nanoparticulate properties represented by the nanosized dimension and shape allow different biodistribution and opportunity beyond the conventional imaging of chemical agents. The recent development of molecular and cellular imaging, which enables visualization of the disease-specific biomarkers at the molecular and cellular levels, has led to increased recognition of nanoparticles as MRI contrast agents, where nanoparticulate iron oxide has been the prevailing and the only clinically used nanoparticulate agent. As a result of the tremendous progress in nanotechnology, many researchers have developed new nanoparticulate MRI contrast agents that have further improved contrasting abilities and have extra functions. In the following sections, we review the progress in inorganic nanoparticles as MRI contrast agents [4]. In particular, this chapter is focused on the core nanoparticles related to their contrast mechanisms. First, we discuss the T2 contrast agent based on the superparamagnetic property which is the main part in nanoparticular MRI contrast agents. Newly developed nanoparticulate T1 contrast agents are introduced in the latter part of the chapter. 13.2 BASIC PRINCIPLES AND CLASSES OF MRI CONTRAST AGENTS “Contrast” refers to the signal differences between adjacent regions in images, and when the target of the image is the living body they could be “tissue and tissue,” “tissue and vessel,” and “tissue and bone.” Contrast agents make an enhancement of contrast around those interests. Contrast agents for X-ray and CT show contrasting effects according to the electron density difference, and they make direct contrast effects on their locations. However, the contrast mechanism is more complicated for MRI where the contrast enhancement occurs as a result of the interaction between the contrast agents and neighboring water protons, which can be affected by many intrinsic and extrinsic factors such as proton density and MRI pulse sequences. The basic principle of MRI is based on nuclear magnetic resonance (NMR) together with the relaxation of proton spins in a magnetic field [1]. When the nuclei of protons are exposed to a strong magnetic field, their spins align themselves either parallel or antiparallel to the magnetic field. During their alignment, the spins precess under a specified frequency known as the Larmor frequency (0 ) (see Fig. 13.1a). When the “resonance” frequency in the radiofrequency (rf) range is introduced to the nuclei, the protons absorb energy and are excited to the antiparallel state. After the disappearance of the rf pulse, the excited nuclei relax to their initial, lower-energy state (Fig. 13.1b). There are two different relaxation pathways. The first, called longitudinal relaxation or T1 relaxation, involves the decreased net magnetization (Mz ) recovering to the initial state (Fig. 13.1c). The second, called transverse relaxation or T2 relaxation, involves the induced magnetization on the perpendicular plane (Mxy ) disappearing by the dephasing of the spins (Fig. 13.1d). Based on their relaxation processes, the contrast agents are classified as T1 contrast agents and T2 contrast agents. Commercially available T1 contrast agents are usually paramagnetic complexes while T2 contrast agents are based on iron oxide nanoparticles, which are the most representative nanoparticulate agents. 13.3 T2 NANOPARTICULATE MRI CONTRAST AGENTS Most of reported nanoparticualte MRI contrast agents were T2 contrast agents based on their superparamagnetic properties. Thus there has been much research on their contrast
P1: OTA/XYZ P2: ABC c13 JWBS037-Chen January 25, 2011
12:49
Printer Name: Yet to Come
T2 NANOPARTICULATE MRI CONTRAST AGENTS
(a)
(b) z ω0
p rf
ul
se ω
ω0 = γB0
B0
Mz
ω0: Larmor frequency
γ : gyromagnetic ratio B0 : Magnetic field
281
y
=
ω0
Mz Mxy
x (c)
(d)
Mz
Mxy
M0 M0(1 – e–1)
Mxye–1
T1’ T1
t
T2’ T2
t
FIGURE 13.1 Basic principle of magnetic resonance imaging. (a) Spins align parallel or antiparallel to the magnetic field and precess under resonance frequency (Larmor frequency, 0 ). (b) After induction of a rf pulse, the magnetization of spins changes. Excited spins take the relaxation process of (c) T1 relaxation and (d) T2 relaxation.
effects and clinical trials, and we first introduce T2 contrast agents. The main enhancement of T2 contrast agents is the acceleration on the T2 relaxation process, spin–spin relaxation. When the rf pulse is applied to spins, transverse magnetization on the xy plane (Mxy ) perpendicular to the direction of the static magnetic field is generated (Fig. 13.1b). Net magnetization M, as a vector, has the components Mz and Mxy , which make the interrelated process of spins. The change in Mz is due to energy transfer, whereas that in Mxy is due to the process of spin dephasing, that is, the randomization of the magnetization of excited spins with the same phase coherence immediately after the application of the rf pulse. Their phase coherence in the xy plane disappears due to the difference of magnetic field experienced by the protons. The magnetic field difference is produced by the system performance in shimming and the magnetic properties of imaging objects. Although the inhomogeneity of the static magnetic field by the system imperfection can be reduced by a variety of tools including the shimming coils and shimming algorithms and the usage of the spin echo sequence to reverse this effect, it affects the decay of transverse magnetization. As other source of field inhomogeneity, the magnetic properties of imaging objects can cause phase incoherence. The spin–spin interaction between the hydrogen nuclei or electrons causes a loss of transverse coherence, which makes the true and characteristic T2 relaxation of tissues. For example, the proton interaction of macromolecules in tissue can induce a local magnetic field, as well as a change in the actual magnetic field in their vicinity. Furthermore, the local magnetic field gradient can be induced from the differences in the magnetic susceptibility between the adjacent and different tissues or by contrast agents. Therefore transverse relaxation is affected by inhomogeneous magnetic field produced from tissue-inherent factors or external sources and the total relaxation time, T2∗ , is
P1: OTA/XYZ P2: ABC c13 JWBS037-Chen January 25, 2011
282
12:49
Printer Name: Yet to Come
MRI CONTRAST AGENTS BASED ON INORGANIC NANOPARTICLES
described by 1 1 = + ␥ BS T2∗ T2
(13.1)
where ␥ BS represents the relaxation by the field inhomogeneities and is called susceptibility effects. The magnetization of paramagnetic materials, such as gadolinium complexes, is directly dependent on the number of ions and they have no magnetization in the absence of an external magnetic field. However, ferromagnetic iron oxide has a very large magnetic susceptibility, which can persist even upon removal of the external magnetic field. nanosized iron oxide particles are superparamagnetic, losing their magnetization in the absence of an external magnetic field. However, when an external magnetic field is applied, they exhibit a strong magnetization, which can cause microscopic field inhomogeneity and activate the dephasing of protons. Therefore iron oxide nanoparticles shorten T 2 and T2∗ relaxation times of the neighboring regions and produce a decreased signal intensity in T2- and T2*-weighted MR images. 13.3.1 Dextran-Coated Iron Oxide Nanoparticles Since their first use as MRI contrast agents 20 years ago, iron oxide nanoparticles (usually magnetite (Fe3 O4 ) or maghemite (␥ -Fe2 O3 )) have become extremely popular due to their dramatic ability to shorten T2∗ relaxation times in the liver, spleen, and bone marrow, by selective uptake and accumulation in the cells of the reticuloendothelial system (RES) [3]. With their high magnetization, their selective signal loss allowed for a new class of MRI contrast agents in the world of dominant T1 contrast agents based on ionic complexes. Since the magnetic property of the nanoparticles and their biological distribution are directly dependent on their size, they have been classified by size as follows: (1) micrometer-sized paramagnetic iron oxide (MPIO; several micrometers), (2) superparamagnetic iron oxide (SPIO; hundreds of nanometers); and ultimately, (3) ultrasmall superparamagnetic iron oxide (USPIO: less than 50 nm) [4]. Among them, two classes are widely used in MRI: SPIO and USPIO. There are several approved products in the SPIO family: Feridex® (Berlex) in the United States or Endrem® (Guerbet) in Europe, and Resovist® (Schering) in Europe and Asia. Smaller nanoparticles, USPIO, have similar composition and originate from SPIO. At the beginning they were prepared through size fraction of an SPIO mixture. Nowadays, uniform USPIOs are produced with the improvement of a synthetic technique (Combidex® (Advanced Magnetics) in the United States and Sinerem® (Guerbet) in Europe). Products with USPIO are under consideration for clinical uses by the U.S. Food and Drug Administration (FDA). The most representative and traditional method to prepare SPIOs and USPIOs is the reduction and coprecipitation reaction of a mixture of ferrous and ferric salts by addition of an alkaline solution under vigorous stirring or sonication [5]. This precipitation reaction is performed in the presence of stabilizers such as hydrophilic polymers, dextran derivatives. Figure 13.2 shows the representative synthetic scheme of SPIO and USPIO, and an electron microscopic image of ferumoxides (Feridex® or Endrem® ) and MION. Because there were insufficient controls in the synthetic process, resulting particles were formed with a broad range of sizes. Furthermore, these particles consisted of multiple iron oxide cores within a dextran stabilization shell. The size of particles is the main factor that controls their
P1: OTA/XYZ P2: ABC c13 JWBS037-Chen January 25, 2011
12:49
Printer Name: Yet to Come
T2 NANOPARTICULATE MRI CONTRAST AGENTS
283
FIGURE 13.2 (a) Scheme of the preparation of SPIO and USPIO. (b,c) Transmission electron microscope (TEM) images of (b) ferumoxides [5a] and (c) MION [7] GFC*: gel filtration chromatography. (Reproduced with permission of Elsevier Inc. and the Massachusetts Medical Society.)
biological characteristics such as blood half-life and biodistribution. Because SPIO has relatively large overall size and related opsonization by phagocytic cells located in the RES, it shows fast clearance from the body and short lifetime. Small nanoparticles have a longer plasma circulation time due to their slow excretion by the liver. Therefore iron oxide nanoparticles for molecular imaging are usually in the class of USPIO (<50 nm). Weissleder and co-workers first demonstrated MRI applications of USPIO [6, 7]. As mentioned before, early USPIOs were prepared via size fraction of an SPIO mixture by size-exclusion gel chromatography. The progress of synthetic methods has brought various USPIO products such as Combidex® , MION (monocrystalline iron oxide nanoparticles), and CLIO (crosslinked iron oxide). In particular, the amination of dextran laid down the foundations that allowed USPIO to become a functionalized and smart imaging platform for molecular imaging. As shown in Figure 13.2, USPIO is much smaller than SPIO with single iron oxide core. This decrease in size results in lower magnetism due to a loss of ferromagnetic coupling and an increase of quadrupole-coupled paramagnetic doublets [8]. Therefore the relaxation properties of SPIO and USPIO are very different. The efficiency of the relaxation property, relaxivity, shows the dependency on the size of nanoparticles because the enhancement of relaxation is related to the magnetic property, especially in spin–spin relaxation (detailed explanation is given in the later sections). The relaxivity coefficient r2 of SPIO is around 150 mmol−1 s−1 , whereas that of USPIO is a much smaller value of around 60 mmol−1 s−1 . However, it is hard to say that SPIO is a more efficient contrast agent than USPIO based only on the relaxivity value because their biological behaviors are very different due to their overall sizes. Clinical uses of SPIO and USPIO are often directed by size-dependent distribution in the body rather than their relaxivities. SPIO is selectively taken up by the Kupffer cells in the liver, spleen, and bone marrow, and the usual clinical targets of SPIO have been liver diseases [3]. If the normal liver architecture is destroyed by a hepatic disease, such as a primary liver tumor or liver metastasis, the region will have a lack of Kupffer cells. Due to the negligible uptake by the abnormal part, the SPIO presents a strong contrast between normal and abnormal tissue, thereby enabling clear detection of the abnormal tissue. However, SPIO’s rapid uptake by
P1: OTA/XYZ P2: ABC c13 JWBS037-Chen January 25, 2011
284
12:49
Printer Name: Yet to Come
MRI CONTRAST AGENTS BASED ON INORGANIC NANOPARTICLES
(a)
(b)
(c)
FIGURE 13.3 (a) Mechanism of action of ultrasmall superparamagnetic iron oxide nanoparticles. (b,c) MR images of a metastatic lymph node (b) before and (c) after use of USPIO [7]. (Reproduced with permission of the Massachusetts Medical Society.)
the liver leads to its fast excretion from blood plasma, which induces a limitation in the SPIO’s ability to depict the molecular and biological functions of other organs and legions. A typical clinical application of USPIOs is lymph node imaging [7]. The detection of lymph nodes is critical for accurate tumor staging and subsequent therapeutic planning. Since USPIOs are very small, they can be extravasated from the blood vessels into interstitial spaces and they are transported to lymph nodes through lymphatic vessels (Fig. 13.3a). As nodes that have malignant cells cannot undergo phagocytosis by nodal macrophages, nanoparticles are taken up only by normal nodes. Their accumulation within normal nodes produces significant susceptibility effects, and the specific uptake allows more sensitive detection between normal and malignant cells. Therefore they are expected to improve the diagnosis of metastatic tumors, and they are currently under investigation for human applications. Weissleder and co-workers reported the successful detection of lymph node metastases in patients with prostate cancer using USPIO (Fig. 13.3b) [7]. USPIOs have
P1: OTA/XYZ P2: ABC c13 JWBS037-Chen January 25, 2011
12:49
Printer Name: Yet to Come
T2 NANOPARTICULATE MRI CONTRAST AGENTS
285
also been tested as a blood pool agent because they are readily distributed in intravascular extracellular space. For example, many clinicians have tried to use USPIO for angiography, functional MRI (fMRI), and passive targeted imaging of tumors. However, the aforementioned applications are based mainly on the biological behavior of iron oxide nanoparticles. In the view of molecular imaging, recent interest has focused on active imaging with maximization of nanoparticulate features. As mentioned in the Introduction, nanoparticles have advantages over paramagnetic complex agents. Nanosized particles can easily be taken up by both macrophage cells and nonphagocytic cells. Moreover, targeted imaging is possible because nanoparticles have a large surface area that can be conjugated with biological and targeting probes such as antibodies, oligonucleotides, aptamers, and other imaging probes. Iron oxide nanoparticles have shown outstanding results in the field of target-specific in vivo or in vitro imaging, in particular, monitoring the migration and tracking of cells, and disease targeted imaging. Cell therapy is a personalized treatment that has been used for more than 50 years. In particular, recent research on stem cells gives hope to those suffering from incurable diseases. For successful cell therapy and the monitoring of cell tracking and differentiation in vivo, it is critical to develop methods for the noninvasive assessment of the fate and distribution of cells. Because the magnetic nanoparticles’ large susceptibility provides highly sensitive tracking with very low detection limits, their use enables the labeling of target cells. MRI anatomic imaging using SPIO and USPIO has been employed for phagocytic cells, such as neutrophils, macrophages, and monocytes, and for nonphagocytic cells, such as lymphocytes, glioma cells, and stem cells [4]. Initial demonstrations of cellular imaging were macrophage imaging including hepatic imaging and lymph node imaging through intravenous injection of SPIO or USPIO mentioned earlier. Because monocytes and macrophages can phagocytize nanosized particles, those phagocytic cells and related diseases are targets of in vivo labeling and imaging. For instance, experimental autoimmune encephalomyelitis (EAE) shows the development of macrophages around disrupted blood–brain barrier (BBB), and transplant rejection induces inflammation and macrophages. USPIO has been applied in these in vivo labeling with its long lifetime [4]. Recent research has focused on in vivo MRI imaging of ex vivo labeled cells with SPIO and USPIO for highly efficient and selective labeling on targets. Moore and co-workers prepared labeled pancreatic islet cells through the incubation islets with UPIO stabilized by a modified dextran, which carries fluorochromes for near-infrared optical imaging [9]. They demonstrated in vivo tracking of transplanted pancreatic islet cells using MRI in diabetic mice in real time. In the initial studies, those cells were labeled by simple incubation with a high concentration of native nanoparticles. Because simple incubation has the limitations of long incubation time and low labeling efficiency in the case of the nonphagocytic cells, various chemical, mechanical, and biological approaches were developed to deliver the nanoparticles efficiently into the cells [10–13]. First, the use of transfection agents and electroporation were adapted from the methods used to introduce DNA or plasmids [11]. Bulte and co-workers reported the efficient labeling of stem cells with magnetic iron oxide through electromagnetoporation, which combines the electromechanical permeability changes of the cellular membrane and the accelerated movements of iron oxide by the magnetic field [12]. Another approach is the modification of the nanoparticles by the attachment of biomaterials that facilitate particle binding to the anionic cellular membrane, followed by internalization. Viruses, peptides, proteins, and antibodies have been conjugated or coated with USPIO and SPIO [13]. In particular, dextran-stabilized MIONs were popularly conjugated with biomaterials because MIONs can readily be functionalized
P1: OTA/XYZ P2: ABC c13 JWBS037-Chen January 25, 2011
286
12:49
Printer Name: Yet to Come
MRI CONTRAST AGENTS BASED ON INORGANIC NANOPARTICLES
with amine groups by epichlorohydrin treatment. Josephson et al. [13a] reported that crosslinked iron oxide nanoparticles conjugated with HIV-Tat proteins (CLIO-Tat) showed efficient nonphagocytic cell labeling through activated macropinocytosis. The facile process of functionalizing dextran-stabilized iron oxide nanoparticles with amine groups (CLIO) established a platform for active molecular imaging. Many kinds of biomolecules including antibodies, proteins, peptides, polysaccharides, and aptamers were covalently bound on the surface of the iron oxide nanoparticles for their site-specific accumulation at the targets of interest [14–16]. This functionalization of nanoparticles with bioactive materials moved MRI into the next stage of active targeted imaging. Kang et al. [14] reported antibody-conjugated CLIO and in vitro targeted MR imaging of E-selectin in endothelial cells. Furthermore, amine groups can be conjugated with small molecules, such as optical dyes for multimodal imaging, and synthetic ligands to develop targeted imaging and therapeutic materials with rapid screening and imaging [15, 16]. Weissleder and co-workers developed biocompatible magnetic relaxation switches (MRSs) to detect molecular interactions, such as DNA–DNA, protein–protein, protein–small molecule, and enzyme reactions [16], based on the observation that self-assembled magnetic nanoparticles exhibited enhanced spin–spin relaxation times (T 2 ) compared to the individual magnetic nanoparticle dispersions. 13.3.2 New T2 MRI Contrast Agents For last 20 years, dextran-stabilized iron oxide nanoparticles (SPIO, USPIO) have been applied as MRI contrast agents for conventional clinical imaging and molecular imaging. As mentioned before, they are not uniform particles and even have multiple cores inside a particle due to the limitations of their synthetic processes. For molecular and cellular studies, imaging with high resolution and high sensitivity are required to fulfill the quantization and interpretation issues. The T2 contrasting effect of agents is dependent on their magnetic properties, and those of nanosized materials are strongly dependent on the quality of the nanoparticle. However, a coprecipitation reaction in water at a relatively low temperature produces the formation of polydisperse and poorly crystalline nanoparticles with inferior magnetic properties. With the progress of nanotechnology, various nanoparticles have been prepared with improved uniformity and crystallinity. Among them, magnetic nanoparticles have received lots of attention from many scientists because of their potential for various application fields including ferrofluid magnetic refrigeration systems and magnetic carriers for drugs and catalysis [17]. Solvothermal or thermal decomposition processes have been reported as synthetic strategies for high-quality nanoparticles. However, they need to be modified to fulfill specific requirements for biomedical imaging. Of course, nanoparticles have to possess suitable properties for the contrasting mechanism. Second, nanoparticles should be compatible with biological systems. Finally, nanoparticles should have reactive moieties on the surface for conjugation with biological active materials. In general, the contrasting ability originates from the core materials, whereas the biocompatibility and conjugation capability are related to the surface properties. Recently, numerous studies have been conducted to develop new MRI contrast agents based on magnetic nanoparticles with core materials of improved magnetic properties and suitable surface characteristics.
Synthesis of Uniform Nanoparticles Various magnetic nanoparticles have been researched as signal-enhancing cores for T2 contrast effect, and iron oxide and related ferrite
P1: OTA/XYZ P2: ABC c13 JWBS037-Chen January 25, 2011
12:49
Printer Name: Yet to Come
T2 NANOPARTICULATE MRI CONTRAST AGENTS
287
nanoparticles are the most popular class for T2 agents. Interestingly, they have attracted many scientists not only in the imaging area but in various fields of fundamental and applied sciences because of their unique magnetic properties, which cannot be achieved by their bulk counterparts [17, 18]. Therefore many material scientists have done research on their syntheses and properties. There are several key issues in the synthesis of nanoparticles including magnetic nanoparticles: (1) particle size distribution (uniformity), (2) particle size control, (3) crystallinity and crystal structure, and (4) large-scale production for applications [18a]. That is, the final goal is to produce size-tunable nanoparticles in large quantities. For the last few decades, many scientists prepared nanoparticles through two different approaches: the “top–down” approach, which is molecular-scale lithography via physical methods, and the “bottom–up” approach, which employs solution-phase colloidal chemistry [17, 18]. The advantage of the physical methods is the production of a large quantity of nanoparticles, but it is very difficult to control the size and distribution of nanoparticles. In contrast, colloidal chemical synthetic methods can be used to synthesize uniform nanoparticles with controlled particle size and shape, although generally only subgram quantities are produced. Very recently, colloidal chemical methods were developed to synthesize a large quantity of uniform nanoparticles [19]. In this chapter, we briefly discuss these colloidal chemical methods for nanoparticles. The synthesis of metal and metal oxide nanoparticles, which are major candidates for the core parts of MRI contrast agents, has been well documented in several review articles [18]. The mechanism studies revealed that uniform-sized nanocrystals can be produced when the burst of nucleation, which enables the separation of nucleation and growth processes, is combined with the subsequent diffusion controlled growth process through which the size focusing works [18a,c]. Several chemical synthetic methods have been used to synthesize uniform nanoparticles of metals and metal oxides. The “hot-injection” method, which involves the injection of organometallic precursors in surfactant solutions at high temperature followed by aging at slightly lower temperature, has been used extensively to synthesize uniform nanocrystals of CdSe, Co, and Fe. However, incomplete separation between nucleation and growth processes often resulted in a mixture of nanoparticles with various sizes, and a size-selection procedure is necessary to prepare uniform nanoparticles (Fig. 13.4a). Moreover, for the burst of nucleation process, highly reactive precursors, which are very unstable and even unavailable for many materials, are required. This hot-injection process is performed at high temperature and under extremely vigorous conditions, which can become very dangerous. Consequently, this hot-injection method cannot be used for large-scale synthesis of uniformly sized nanoparticles. The Hyeon group developed a “heat-up” process to synthesize various uniform oxide nanoparticles, in which reaction mixtures composed of metal precursors, surfactants, and solvents are slowly heated from room temperature to high temperature followed by aging [18c]. Kinetics studies showed that the overall process of the size distribution control consists of three steps: the accumulation of monomers from the thermal decomposition of metal precursors, burst nucleation, and diffusion controlled growth for size focusing to produce uniform nanoparticles (Fig. 13.4b). The Hyeon group reported the synthesis of highly uniform iron and iron oxide nanoparticles without a size-selection process via the heat-up method [20a]. In this report, an iron precursor (Fe(CO)5 ) was mixed with a surfactant (oleic acid) in solvent (octylether) and they were slowly heated to reflux. Sun and co-workers synthesized monodisperse magnetite nanoparticles by a similar heatup process using iron acetylacetonate as a precursor and 1,2-hexadecandiol as a mild
P1: OTA/XYZ P2: ABC c13 JWBS037-Chen January 25, 2011
288
12:49
Printer Name: Yet to Come
MRI CONTRAST AGENTS BASED ON INORGANIC NANOPARTICLES
(a)
(b)
FIGURE 13.4 Schematic illustration of (a) hot-injection method [18a] and (b) heat-up method. (Reproduced with permission of the Royal Society of Chemistry.)
reductant in the presence of oleic acid and oleylamine [20b]. They also synthesized metal ferrite nanoparticles such as MnFe2 O4 and CoFe2 O4 [20b]. In 2004, the Hyeon group reported large-scale synthesis of uniform metal oxide nanocrystals via the heat-up method using metal–oleate complexes as the precursors [19a]. Metal–oleate complexes are easily prepared from the reaction between metal chlorides and sodium oleate. This modified heat-up method has been applied to the synthesis of uniform nanocrystals of various materials such as hexagonal-cone-shaped ZnO and pencil-shaped CoO [20c,d]. Especially, iron oxide nanocrystals synthesized via this method show remarkable size uniformity. Because of its simplicity, the heat-up method using a metal–oleate precursor is highly advantageous for scale-up, as demonstrated by the production of iron oxide nanocrystals in tenths of grams from a single batch reaction without the deterioration of the size uniformity.
Size-Dependent Magnetic Properties and T2 Contrast Effects Magnetic nanoparticles have mainly been applied as T2 contrast agents. Acceleration of spin–spin relaxation (T2 shortening) by magnetic nanoparticles results from the dephasing of the magnetic moments due to the magnetic field gradients created by the small magnetized particles. In this process, the major relaxation mechanism is the dipolar outer-sphere interaction between the water proton spins and the magnetic moment of the magnetic nanoparticles. Therefore, as shown in the model suggested by Koenig and Keller, spin–spin relaxation is
P1: OTA/XYZ P2: ABC c13 JWBS037-Chen January 25, 2011
12:49
Printer Name: Yet to Come
T2 NANOPARTICULATE MRI CONTRAST AGENTS
289
dependent on the magnetic moment of the nanoparticles () [21]: R2 =
1 a ␥ 2 2 CNP J (, D ) = T2 dNP D
(13.2)
where a is a constant, dNP the diameter of the nanoparticle, D is the diffusion coefficient, is the magnetic moment of the nanoparticles, ␥ is the gyromagnetic ratio of the water proton, CNP is the concentration of the nanoparticles, and J(, D ) is the spectral density function. In other words, to be efficient T2 contrast agents, nanoparticles should possess a large magnetization. Although magnetism is an intrinsic property of bulk materials, the magnetic properties of nanoparticles are strongly dependent on their size, shape, and surface property [22]. A dramatic increase in surface area induces preponderant surface canting effects and consequently increases the magnetization, as described by the following equation [23]: m S = MS [(r − d)/r ]3
(13.3)
where mS is the saturation magnetization of the nanoparticle, M S is the saturation magnetization of the bulk material, r is the size of the nanoparticle, and d is the thickness of the disordered surface layer. To make efficient T2 contrast agents, we have to control the magnetic properties of the nanoparticles through the designed control of (1) the intrinsic material properties such as material composition and crystal structure, and (2) extrinsic factors such as size and shape [23c]. Most T2 contrast agents are based on iron oxide nanoparticles and are composed of aggregated nanoparticles with multiple iron oxide (4–5 nm) cores and dextran coating. USPIOs with a single core (MION, CLIO) have low relaxivities due to their small core size. As described earlier, uniform and highly crystalline iron oxide nanoparticles have been synthesized by various synthetic methods in the last decade [18–20]. Recently, some of these uniform-sized iron oxide nanoparticles have been used for MRI contrast agents. Cheon and co-workers systematically studied the relationships among size, magnetism, and relaxivity of uniform-sized iron oxide nanoparticles (see Fig. 13.5a) [22b, 23c]. Larger iron oxide nanoparticles (MEIO) possess higher magnetization values, exhibit stronger T2 contrast effects, and possess higher magnetization values. The r2 relaxivity coefficient is 78 mM−1 s−1 for 4-nm MEIO and this gradually increases 106, 130, and to 218 mM−1 s−1 for 6-, 9-, and 12-nm size MEIO nanoparticles, respectively. Large iron oxide nanoparticles have a large magnetization and high r2 relaxivity, which makes it possible to increase the sensitivity in T2/T2*-weighted images and to ease the toxicity concern by decreasing the agent dose. Because magnetism is an intrinsic property of materials, magnetic nanoparticles with high magnetism have recently been synthesized to develop new MRI contrast agents with improved relaxation properties and biocompatibility. Alloy materials can be candidates of more efficient T2 contrast agents. Various bimetallic ferrite nanoparticles including CoFe2 O4 , MnFe2 O4 , and NiFe2 O4 have been tested as T2 contrast agents (Fig. 13.5b and Table 13.1) [23c, 24]. In particular, the MnFe2 O4 nanoparticles (Mn-MEIO) have been found to have a very high magnetization values of 110 emu/g and large r2 relaxivity of 358 mM−1 s−1 [24]. This relaxivity is about six times higher than that of CLIO nanoparticles.
290 Dextran Carboxydextran Dextran Dextran DMSA[b] DMSA[b]
DMSA[b] DMSA[b] PEG PEG
Fe3 O4 , ␥ -Fe2 O3 Fe3 O4 Fe3 O4 Fe3 O4 Fe3 O4 MnFe2 O4
CoFe2 O4 NiFe2 O4 Fe3 O4 FePt (fcc)
Ferumoxides (Feridex) Ferucarbotran (Resovist) Ferumoxtran (Combidex) CLIO-Tat
WSIO (MEIO)
MnMEIO
CoMEIO NiMEIO Au-Fe3 O4 Au-FePt
4.96 4 5.85 5 4 6 9 12 6 9 12 12 12 20 6
Diameter of Core [nm] 160 60 35 30
Hydrodynamic Diameter [nm]
61 60 25 43 80 101 68 98 110 99 85
45
Magnetization [emu/g][a]
120 186 65 62 78 106 130 218 208 265 358 172 152 114 59
r2 [mM-1 s-1 ]
1.5 1.5 3.0 3.0
1.5
1.5
1.5 1.5 1.5 1.5
B0 [T]
5a 5b 5b 13a 22b, 23b 22b, 23b 22b, 23b 22b, 23b 23b, 24 23b, 24 23b, 24 23b, 24 23b, 24 41a 41b
Reference
12:49
[a] Magnetic properties were measured at 1.5 T external field [b] 2,3-dimercaptosuccinic acid.
Surface
Core Material
Name
TABLE 13.1 Properties of T2 Contrast Agents
P1: OTA/XYZ P2: ABC c13 JWBS037-Chen January 25, 2011 Printer Name: Yet to Come
P1: OTA/XYZ P2: ABC c13 JWBS037-Chen January 25, 2011
12:49
Printer Name: Yet to Come
T2 NANOPARTICULATE MRI CONTRAST AGENTS
(a)
291
(b) 6 nm
9 nm
12 nm mass magnetization (emu/g)
4 nm
25 nm
110 105 100 95 90 85
250 MEIO
Co-MEIO
Ni-MEIO
200 150 100 50 0
150 5
size
10
(nm)
100
50 15
0
ms
)
Fe)
/g( emu
Relaxivity coefficient (mM–1s–1)
r 2(mM–1s–1)
Mn-MEIO
400 300 200 100 0 Mn-MEIO
MEIO
Co-MEIO
Ni-MEIO
(
FIGURE 13.5 (a) The size-dependent magnetic property and r2 relaxivity of monodisperse iron oxide nanoparticles. (b) The magnetic property and r2 relaxivity of several bimetallic ferrite nanoparticles [23b]. (Reproduced with permission of the American Chemical Society.)
Biocompatible Packing and Modification of T2 Nanoparticulate Contrast Agents For many biomedical applications, in particular, for in vivo imaging, nanoparticles should possess a good colloidal stability and low toxicity in a biological environment. Furthermore, surface modification is important for the nanoparticles to impart additional functions because bioactive materials are conjugated through the reactive groups on the surface. Dextran-stabilized iron oxide nanoparticles such as CLIO and MION were synthesized by coprecipitation methods, which prepare iron oxide nanoparticles with biocompatible dextran directly, and they did not need any additional step to stabilize in aqueous media. On the other hand, because conventional dextran is very stable and inert, it is very important to have reactive moieties for further functionalizations of nanoparticles. For example, amination has been investigated from the very early days of USPIO. A representative method is the crosslinking with epichlorohydrin and the resulting product is CLIO [6]. Although high-quality magnetic nanoparticles with uniform particle size and high crystallinity were synthesized from thermal decomposition of metal precursors in organic media, they are hydrophobic and consequently are not sufficiently stable in aqueous media for biomedical applications. Therefore modifying these nanoparticles is essential to endow them with hydrophilic properties so that they can be used extensively for various biomedical applications [25]. Polymer and silica derivatives are typical materials for the surface modification of nanoparticles for biomedical applications, to make inorganic nanoparticles compatible and stable, and what’s more, they are functionalized easily. Recently, various surface modification methods were developed and two representative strategies are ligand exchange with water dispersible ligands and encapsulation with biocompatible shells (Fig. 13.6). The ligand exchange method [26] has several advantages including simple procedure, thin passivating layer, and small overall size. Nanoparticles prepared in organic solvents
P1: OTA/XYZ P2: ABC c13 JWBS037-Chen January 25, 2011
292
12:49
Printer Name: Yet to Come
MRI CONTRAST AGENTS BASED ON INORGANIC NANOPARTICLES
Ligand exchange with hydrophilic ligands
Hydrophobic shell
Encapsulation with hydrophilic surface
FIGURE 13.6 Two strategies for biocompatible nanoparticles: (a) ligand exchange with hydrophilic ligands via substitution reaction of stabilizing ligands on nanoparticles reaction and (b) Encapsulation by hydrophilic matrix such as polymers and silica derivatives.
are stabilized by surfactants that have their hydrophilic head groups bound onto the surface of the nanoparticles and the hydrophobic hydrocarbon tails facing the solvent. Since the atoms on the surface of the nanoparticles have an affinity for the functional groups, surface functionalization with proper hydrophilic ligands allows the phase transfer of nanoparticles from organic media to aqueous media. Various kinds of nanoparticles can undergo substitution reactions with excess bifunctional ligands that consist of a strong binding moiety to the surface of the nanoparticles and relatively low binding hydrophilic group. Because metal oxide nanoparticles have a relatively unreactive surface compared to other kinds of nanoparticles such as chalcogenide-based quantum dots, there are very limited kinds of functional groups available. For example, metal oxide nanoparticles do not bind strongly to alkanethiols that serve as excellent ligands for nanoparticles of chalcogenides and metals. Consequently, there has been intensive recent research on the development of new ligand systems that can bind strongly to metal oxide nanoparticles, in particular, iron oxide nanoparticles. Dopamine was used as a ligand to immobilize functional molecules on iron oxide nanoparticles [26a,b], whereas Peng and co-workers took advantage of the interaction between hydroxamic acid and iron oxide [26c]. Cheon and co-workers were able to form waterdispersible iron oxide nanoparticles using a small organic molecule, 2,3-dimercaptosuccinic acid (DMSA), which has a bidentate carboxyl group [22b]. The resulting iron oxide nanoparticles were dispersible in water and they were successfully used for in vivo MRI. To improved the biocompatibility of the nanoparticles, the backbones of the ligands were designed to contain biocompatible polymers such as polyethylene glycol (PEG). Sun and co-workers demonstrated that a dopamine derivatized PEG (DPA-PEG) was an efficient ligand to stabilize 9-nm Fe3 O4 nanoparticles. The resulting nanoparticles were stable in biological media and showed very low uptake by macrophage cells [26b]. PEG-derivatized phosphine oxide ligands (PO-PEGs), consisting of a biocompatible PEG tail group and surface coordinating phosphine oxide head group, could stabilize several kinds of oxide nanoparticles in aqueous media [26d]. They were prepared by simple reaction between
P1: OTA/XYZ P2: ABC c13 JWBS037-Chen January 25, 2011
12:49
Printer Name: Yet to Come
T2 NANOPARTICULATE MRI CONTRAST AGENTS
293
POCl3 and PEGs. Polymeric or dendrimeric ligands have also been used for the stabilization of iron oxide nanoparticles in aqueous media derived from their multiple functional groups and bulky structure [26e]. These ligands interact through a number of bonds on the nanoparticles by ligand exchange. The second strategy for surface modification is the formation of biocompatible shells or matrices around the nanoparticles. The advantages of this strategy include easy modification and expansion in those shell structures. Although overall sizes of products are relatively larger than those from the ligand exchange method, they can contain other kinds of material together and form multifunctional structures. There are a variety of biocompatible shell formation methods; they can broadly be classified according to the shell materials and the encapsulation processes. Typical shell materials are silica and polymers. Crosslinked amorphous silica shell can stabilize oxide nanoparticles. Although the toxicity of nanosized silica is still in debate, there have been promising in vitro and in vivo reports on silica-encapsulated nanoparticles [27–29]. Ying and co-workers reported silica-coated nanoparticles produced by base-catalyzed silica formation in a reverse microemulsion [28a,b]. Although the silica layer is relatively thick (10–30 nm), the encapsulation process is simple and various kinds of nanoparticles can be encapsulated. It also allows the encapsulation of various kinds of nanoparticles simultaneously within one shell. Fluorescent dye molecules can be incorporated in silica shells during the condensation reaction of silane reagents. Lee and co-workers reported magnetic fluorescent core–shell nanoparticles (Fig. 13.7a) [28c]. Cobalt ferrite (CoFe2 O4 ) nanoparticles were synthesized by coprecipitation in aqueous media and a dye-doped silica shell was formed around a nanoparticle via the modified St¨ober method. The surface of the silica shells was further functionalized with PEG and conjugated with an antibody to target cancer cells. Resulting core–shell nanoparticles exhibited T2 contrast enhancing properties in the phantom test and the confocal images of the cancer cells treated with the nanoparticles showed strong fluorescence. Huang and co-workers also synthesized core–shell nanoparticles composed of iron oxide nanoparticles and FITC-doped silica shells [28d]. They synthesized the iron oxide nanoparticles by the thermal decomposition method and stabilized with surfactants such as oleic acid and oleylamine. Because resulting iron oxide nanoparticles were hydrophobic, they were coated with dye-doped silica shells via a reverse microemulsion system. The authors demonstrated human stem cell tracking using these multi-imaging nanoparticles. The probes were easily internalized into human mesenchymal stem cells (hMSCs), and they were injected subcutaneously into the dorsal flank of a mouse and could be visualized in a clinical 1.5-T MRI imager. The Hyeon group reported the fabrication of magnetic fluorescent mesoporous silica nanoparticles composed of mesoporous silica matrix embedded with iron oxide nanoparticles and quantum dots [29a]. Hydrophobic magnetic and fluorescent nanoparticles were transferred into the aqueous phase using cetyl trimethylammonium bromide (CTAB) as a secondary surfactant. Because CTAB also acted as a structure directing agent of the mesostructures, the subsequent silica sol-gel reaction resulted in the formation of mesoporous silica spheres embedded with magnetic and fluorescent nanoparticles. Their high surface area and large pore volumes allowed for the loading of chemical drugs and controlled release. There are several other reports on composite materials of magnetic nanoparticles and mesoporous silica materials [29b–f]. However, they could not be used for in vivo MR imaging because of their large overall size of >100 nm. Very recently, the Hyeon group also synthesized discrete and monodisperse core–shell mesoporous silica NPs smaller than 100 nm by using single Fe3 O4 nanoparticles as cores [29g]. The uniform 3-nm mesoporous
P1: OTA/XYZ P2: ABC c13 JWBS037-Chen January 25, 2011
294
12:49
Printer Name: Yet to Come
MRI CONTRAST AGENTS BASED ON INORGANIC NANOPARTICLES
(a)
(b)
FIGURE 13.7 Silica-encapsulated iron oxide nanoparticles. (a) TEM images and schematic illustration of CoFe2 O4 –silica (core–shell) functionalized with organic dyes and antibodies [28c]. (b) The synthesis of magnetite nanoparticle/mesoporous silica core–shell and their in vivo multimodal imaging (MRI and optical imaging) [29g]. (Reproduced with permission of Wiley-VCH Verlag GmbH & Co. KGaA.)
shell can contain fluorescent dye and anticancer drug and the resulting composite magnetic silica nanoparticles were applied to simultaneous in vivo magnetic resonance (MR) and fluorescence imaging, and drug delivery vehicle (Fig. 13.7b). Various natural and synthetic polymers have been used extensively in the preparation of biocompatible nanoparticles. The immobilization of polymers can reduce the safety and toxicology concerns of nanoparticles for the clinical applications. Because the silica stabilized nanoparticles often experience precipitation and gel formation, additional biocompatible polymers (e.g., PEG) have been immobilized on the surface of the silica shell to improve the colloidal stability [28, 29]. Furthermore these polymer–nanoparticle hybrids can be utilized for multifunctional biomedical applications with simultaneous drug delivery and imaging capability [29]. Polyesters, such as poly(d,l-lactide-co-glycolide) (PLGA), poly(d,l-lactide) (PLA), and poly(glycolide) (PGA), have been most popularly used for these applications [30]. Gao and co-workers employed amphiphilic block copolymers of methoxy- and maleimideterminated poly(ethylene glycol)-block-poly(d,l-lactide) (PEG-PLA) to fabricate polymer
P1: OTA/XYZ P2: ABC c13 JWBS037-Chen January 25, 2011
12:49
Printer Name: Yet to Come
T2 NANOPARTICULATE MRI CONTRAST AGENTS
295
micelles [30b]. Hydrophobic iron oxide nanoparticles and hydrophobic drug (doxorubicin, DOX) were spontaneously incorporated into the hydrophobic PLA core part and hydrophilic PEG was exposed to the aqueous environment. The resulting particles were relatively large (>50 nm) because several iron oxide nanoparticles were clustered in the hydrophobic core part. They were functionalized with RGD peptide, and they targeted cancer cells in in vitro T2 MRI and showed therapeutic effects via the release of loaded doxorubicin. Kim et al. [30c] fabricated multifunctional PLGA nanoparticles with particle size of 100–200 nm by simultaneously immobilizing magnetite nanoparticles, quantum dots, and anticancer drug (doxorubicin) in PLGA matrix via a conventional oil-in-water emulsion–evaporation process [30c]. Using the multifunctional polymer nanoparticles, they demonstrated simultaneous cancer-targeted MR imaging and optical imaging, as well as drug delivery. In addition, the loaded magnetite nanoparticles facilitated the magnetic guiding of the polymer particles, thereby increasing the synergetic targeting efficiency. Jiang and co-workers fabricated hollow Fe3 O4 –polymer hybrid nanospheres by the addition of Fe3 O4 nanoparticles to an aqueous solution of polymer–monomer pairs composed of the cationic chitosan polymer and the anionic acrylic acid monomer, followed by polymerization of acrylic acid and selective crosslinking of chitosan at the end of polymerization [30d,e]. The phantom test of magnetic resonance imaging showed that the synthesized hybrid hollow nanospheres had a significant magnetic resonance signal enhancement in T2-weighted image [30e]. PEG, a representative biocompatible polymer, has received great attention due to its nonfouling property, which supports a resistance to protein adsorption and an ability to bypass the RES and natural barriers such as the nasal mucosa [31]. PEGs have been used extensively as stabilizing materials for many nanoparticles in biomedical applications, in particular, in long circulating in vivo imaging systems. Because PEG itself is very inert, surface-anchorable materials such as copolymers, phospholipids, and silica are combined with PEGs to encapsulate the nanoparticles. Dubertret et al. [32a] reported that PEGphospholipid block copolymers could form a stable micelle structure on quantum dots via the hydrophobic interaction between hydrophobic tail groups of the surfactants and phospholipid parts. The outer surface of the nanoparticles is comprised of a dense PEG layer, which is stable in biological media. This process is highly reliable and can generally be applicable to many other kinds of nanoparticles. Using a very similar strategy, various water-dispersible metal oxide nanoparticles, including iron oxide nanoparticles, were generated [32b]. Because these PEG-phospholipid block copolymers are very expensive, they cannot be applied for the large-scale preparation. Amphiphilic di- and triblock copolymers have also been used as stabilizing shell materials for water-dispersible nanoparticles. Their hydrophobic blocks can strongly interact with the hydrophobic surface of the nanoparticles, whereas the outer hydrophilic blocks can make the nanoparticles dispersible in water [33]. These block copolymers are relatively inexpensive and can be derivatized with other functional groups for additional functionalization. One disadvantage of using these block copolymers is that their large shell thickness is derived from the high molecular weight, which limits their many potential applications. Oligomeric and dendritic molecules were also used as the shell materials for water-dispersible nanoparticles because they form thinner shells while preserving their stability in aqueous media. For example, Yang and co-workers demonstrated that the cyclic oligosaccharides that have hydrophobic cavities and hydrophilic rims can transfer the nanoparticles from an organic to an aqueous phase [34]. Weller and co-workers showed that the relaxivities of magnetic nanoparticles were dependent not only on the size of the core nanoparticles but also on the types of shells [35].
P1: OTA/XYZ P2: ABC c13 JWBS037-Chen January 25, 2011
296
12:49
Printer Name: Yet to Come
MRI CONTRAST AGENTS BASED ON INORGANIC NANOPARTICLES
(b)
(a)
900 40 nm
+ encapsulation
15 nm
relaxivity [mM-1s-1]
ge
an
ch
ex
r2 exchange r2* micelles r2 micelles r2* exchange r2* encapsulated r2 encapsulated r2* Resovist r2 Resovist
800 700 600 500 400 300 200 100
O mi + ce lle s
0 2 250 nm
4
6
8
10
12
14
16
18
20
d [nm]
FIGURE 13.8 (a) Three kinds of methods to prepare water-dispersed manganese ferrite nanoparticles. (b) The r2 relaxities of water-dispersed manganese ferrite nanoparticles with three kinds of shell systems [35]. (Reproduced with permission of the American Chemical Society.)
The MnFe2 O4 nanoparticles synthesized in high boiling ether solvent were transferred to water using three different approaches, including ligand exchange to form a watersoluble polymer shell, embedding into an amphiphilic polymer shell, and encapsulating in large micelles formed by lipids. In the first two polymer-based systems, nanoparticles are individually dispersed to form homogeneous dispersion with a hydrodynamic radius of 30–40 nm, whereas aggregated nanoparticles were randomly distributed inside the micelles with a hydrodynamic radius of 250 nm. Interestingly, the relaxivity, r2*, is much higher for the micellar system than for the polymer-stabilized particles using the same-sized nanoparticles (Fig. 13.8). Nanosized and biocompatible iron oxide nanoparticles have been applied extensively to the diagnosis of cancers. They can be accumulated spontaneously in tumor sites via the enhanced permeability and retention (EPR) effect, which is the enhanced accumulation of macromolecular species including nanoparticles in tumor tissues that have abnormal blood vessels [36]. Consequently, iron oxide nanoparticles were successfully used to image tumors without any targeting probes, which is called passive targeting [37]. For more efficient targeted imaging, the surface of the iron oxide nanoparticles needs to be conjugated with active targeting probes such as antibodies and proteins. Magnetic nanoparticles have been conjugated with various bioactive materials such as antibodies, oligonucleotides, peptides, and proteins. Cheon and co-workers prepared Herceptin-conjugated iron oxide nanoparticles, and they were delivered selectively and imaged tumor cells by interactions with the human epidermal growth factor receptor (Her2/neu), which is usually overexpressed in breast cancers [22b]. The Cheon group also demonstrated that Herceptin-conjugated manganese ferrite (MnMEIO) nanoparticles with high magnetization showed more sensitive in vivo cancer targeted imaging with large r2 relaxivity (Figure 13.9) [24]. This result demonstrated that advanced MRI contrast agents consisting of high magnetic moment nanoparticles and appropriate targeting agents could enable the ultrasensitive detection of various types of cancer in T2/T2*-weighted MRI. As another demonstration of targeted MRI, Gao and co-workers conjugated a cancer-targeting antibody, anticarcinoembryonic antigen (CEA) monoclonal antibody rch 24 onto uniform PEG-coated
P1: OTA/XYZ P2: ABC c13 JWBS037-Chen January 25, 2011
12:49
Printer Name: Yet to Come
T2 NANOPARTICULATE MRI CONTRAST AGENTS
(a)
(d)
(b)
(e)
(c)
(f)
297
FIGURE 13.9 In vivo MR detection of cancer after administration of magnetic nanoparticles– Herceptin conjugates. Manganese ferrite nanoparticles (MnMEIO) (a–c) show higher signal enhancement than crosslinked iron oxide (CLIO) (d–f) [24]. (Reproduced with permission of Nature Publishing Group.)
iron oxide nanoparticles and successfully performed targeted MR imaging for human colon carcinoma tumors [38]. Zhang and co-workers demonstrated that PEG-coated iron oxide nanoparticles conjugated with targeting peptide (chlorotoxin) were preferentially accumulated within gliomas and exhibited highly contrast-enhanced MR imaging [39].
T2 Nanoparticulate Contrast Agents of Unique Structures Recently, new multifunctional nanomedical platforms have been fabricated by combining various nanostructured materials with different functions, making it possible to accomplish multimodal imaging and simultaneous diagnosis and therapy [40]. Heterodimers of magnetic nanoparticles and other nanoparticles can serve as multimodal imaging agents such as MRI contrast agents and optical probes [41]. Dumbbell-like nanoparticles composed of magnetic nanoparticles and gold nanoparticles have been used for dual MR and optical imaging. The Sun group reported the dual modal imaging properties of Fe3 O4 –Au dumbbell nanoparticles, which were prepared by the growth of Fe3 O4 on as-prepared Au nanoparticles in the presence of oleic acid and oleylamine (Fig. 13.10) [41a]. Fe3 O4 and Au components had different surface properties, and they were modified by dopamine and thiol groups, respectively. The epidermal growth factor receptor antibody (EGFRA) was conjugated on the surface of Fe3 O4 , and the resulting heterostructured nanoparticles were successfully applied to the cancer-targeted MR imaging. Furthermore, the unique surface plasmon property of the Au component enabled reflection imaging. The Cheon group demonstrated dual modal imaging using heterodimeric FePt–Au nanoparticles, which were synthesized by the catalytic growth of Au on the surface of the FePt nanoparticles [41b]. Antibody-conjugated FePt–Au nanoparticles were used as both T2 MRI contrast agent and biosensor, and neutravidin-conjugated nanoparticles acted as detecting probes on the biotin patterned
P1: OTA/XYZ P2: ABC c13 JWBS037-Chen January 25, 2011
298
12:49
Printer Name: Yet to Come
MRI CONTRAST AGENTS BASED ON INORGANIC NANOPARTICLES
(a)
(c)
(b)
FIGURE 13.10 (a) TEM image of Au–Fe3 O4 nanoparticles. (b) T2-weighted MR image of A431 cells labeled with Au–Fe3 O4 nanoparticles. (c) Reflection image of A431 cells labeled with Au–Fe3 O4 nanoparticles [41a]. (Reproduced with permission of Wiley-VCH Verlag GmbH & Co. KGaA.)
biochip. Ying and co-workers reported Fe3 O4 –CdSe heterodimer nanoparticles by growing CdSe on the surface of the as-synthesized Fe3 O4 nanoparticles [41c]. Although there are no MRI demonstrations in their paper, it might be a candidate as an MRI contrast agent. The Hyeon group also reported that heterostructured nanoparticles composed of various combinations of a metal (Au, Ag, Pt, or Ni) and an oxide (Fe3 O4 or MnO) were readily synthesized from thermal decomposition of mixtures of metal–oleate complexes and metal–oleylamine complexes [41d]. Nanoparticles of Au–MnO and Au–Fe3 O4 have the potential to serve in many multifunctional biomedical applications, such as multimodal imaging or detection probes. The Hyeon group fabricated biocompatible hollow iron oxide nanocapsules and demonstrated their in vitro T2 MRI and drug delivery capabilities to cancer cells. They were fabricated from akagenite (-FeOOH) nanorods via the wrap–bake–peel process, which involves silica coating, heat treatment, and finally the removal of the silica layer. During heat treatment at 500 ◦ C, first in air and then in 10% hydrogen atmosphere the ␥ -FeOOH nanorods were transformed into magnetite (Fe3 O4 ) capsules. Removing the silica shells resulted in the formation of water-dispersible hollow Fe3 O4 nanocapsules. Large hollow pores could be loaded with chemical drugs that were released in cancer cells, and the magnetite shells were used as the T2 MRI contrast agent (Fig. 13.11) [42].
13.3.3 T1 Nanoparticulate MRI Contrast Agents Over the last 20 years, most nanoparticulate contrast agents have been T2 contrast agents using iron oxide nanoparticles. However, these magnetic nanoparticle-based T2 contrast agents have several disadvantages that limit their extensive clinical applications. First, they are negative contrast agents, which give a signal decreasing effect. The resulting dark signal could be confused with other pathogenic conditions and makes images of lower contrast than T1 contrasted images. Moreover, the high susceptibility of the T2 contrast agents induces distortion of the magnetic field on neighboring normal tissues. This distortion of the background is called the susceptibility artifact or “blooming effect,” which generates obscure images and demolishes the background around the lesions [4]. Because of the
P1: OTA/XYZ P2: ABC c13 JWBS037-Chen January 25, 2011
12:49
Printer Name: Yet to Come
T2 NANOPARTICULATE MRI CONTRAST AGENTS
299
(a) β-FeOOH
β-FeOOH/sillica
Haematite/sillica Heat treatment
Sillica coating
Haematite
Removal of sillica using NaOH
100 nm
Reduction
Removal of sillica using NaOH
Magnetite
Magnetite/sillica
(b)
(c)
120 PEG-MNC-D0X Fe-DOX
0
80
160
320
640
T 2-weighted image T 2 (ms)
viability (x)
100
[Fe] (µM)
100 nm
80 60 40 20
328
122
62
27
11
0
0.001
0.010
0.100
1.000
10.00
DOX (µM)
FIGURE 13.11 (a) Schematic illustration of the procedure for the synthesis of uniform and waterdispersible iron oxide nanocapsules and their TEM images. (b) T2-weighted MR images of the magnetite nanocapsules. (c) In vitro cytotoxicity of free DOX and DOX loading nanocapsules against SKBR-3 cells [42]. (Reproduced with permission of the Nature Publishing Group.)
limitations of these T2 agents, most extensively and clinically used MRI contrast agents are based on gadolinium complex-based T1 agents. T1 relaxation is the process of equilibration of the net magnetization (Mz ) after the introduction of an rf pulse. This change of Mz is a consequence of energy transfer between the proton spin system and the nearby matrix of molecules. All biological systems are composed of various molecules and organisms, and they have different relaxation behaviors and different T 1 relaxation times. The presence of paramagnetic ions near the tissue enhances its relaxation and shortens the T1 relaxation time. In particular, transition and lanthanide metal ions with a large number of unpaired electrons, such as Gd3+ , Mn2+ , and Fe3+ , show very effective relaxation [2]. T1 contrast agents enhance T1 relaxation, which makes a signal enhancement on images. Compared to T2 contrast agents, the major advantage of T1 contrast agents is positive contrast imaging by signal enhancement, which can maximize the forte of MRI (i.e., anatomic imaging with high spatial resolution). Furthermore, their bright signal can be distinguished clearly from other pathogenic or biological conditions. As T1 contrasting agents are basically paramagnetic, they do not disrupt the magnetic homogeneity over the large dimension, which can disturb other anatomic backgrounds. Since Gd3+ has seven unpaired electrons with a large magnetic moment, most T1 contrast agents are Gd3+ -based agents. However, due to the toxicity of heavy metal ions, the conventional contrast agents are in the form of ionic complexes with chelating ligands,
P1: OTA/XYZ P2: ABC c13 JWBS037-Chen January 25, 2011
300
12:49
Printer Name: Yet to Come
MRI CONTRAST AGENTS BASED ON INORGANIC NANOPARTICLES
which are thermodynamically and kinetically stable and less toxic. There is, however, no biochemistry based on gadolinium(III) ion in natural human system. In spite of their fewer unpaired electrons and lower magnetic moments, manganese(II), iron(III), and copper(II) ions could be alternative candidates. Manganese(II) ion, in particular, plays various important roles in many biological processes such as cofactors of enzymes and a release controller of neurotransmitters. Although there are some manganese(II) complex contrast agents, Mn2+ itself, in the form of MnCl2 solution, has been most frequently used. It has shown very prominent contrasting effects that can reveal detailed physiological and biological information, and it constitutes a new imaging category known as manganese-enhanced MRI (MEMRI). In particular, MEMRI can visualize the anatomic structure of the brain and its neuronal activity [43], which cannot be obtained with any of the gadolinium(II)-based contrast agents. Unfortunately, however, MEMRI can only be applied in animal studies because Mn2+ ions cause hepatic failure and have cardiac toxicity. As shown above, the present T1 contrast agents are based on paramagnetic ions and are used in the form of ion complexes. They have short life spans in the body and work in a nonspecific manner. Most T1 contrast agents reside within the extracellular space and usually interact with the blood so that they have some limitations as molecular probes for longer time tracking. As shown in recent studies of iron oxide nanoparticles, nanoparticulate agents are very promising for molecular and cellular imaging, which aims to visualize the disease-specific biomarkers at the molecular and cellular levels, respectively. However, the negative contrasting effect and magnetic susceptibility artifacts of iron oxide nanoparticles can be significant drawbacks when using iron oxide nanoparticles. This is because the resulting dark signal can mislead the clinical diagnosis in T2-weighted MRI as the signal is often confused with the signals from bleeding, calcification, or metal deposits, and the susceptibility artifacts distort the background image [4]. Recently, intensive research has been devoted to developing new T1 contrast agents that overcome the above-mentioned drawbacks of Gd3+ ion- and Mn2+ ion-based T1 contrast agents and SPIO-based T2 contrast agents. Briefly, these new classes of contrast agents should satisfy the following characteristics: (1) positive (T1) contrast ability, (2) easy intracellular uptake and accumulation for imaging cellular distribution and functions, (3) a nanoparticulate form for easy surface modification and efficient labeling with biological and bioactive materials, and (4) favorable pharmacokinetics and dynamics for easy delivery and efficient distribution to the biomarkers with minimal side effects. The first class of particulate T1 contrast agents is based on nanostructured frames that have many anchoring sites for paramagnetic ions [44]. Those particles can carry a large number of paramagnetic payloads and produce strong T1 contrast. Various platforms, such as silicas, dendrimers, perfluorocarbons, emulsions, and nanotubes, have been used. Lanza and co-workers used perfluorocarbon nanoparticle incorporated paramagnetic Gd–DTPA complexes for the sensitive detection of fibrin and for the molecular detection of angiogenesis [44a,b]. Lin and co-workers developed silica nanoparticles [44c], metal–organic framework [44d], and mesoporous silica (Fig. 13.12) [44e] to contain a large number of Gd3+ ions and to develop multifunctional properties. Carbon nanotubes could act as the framework that holds Gd3+ ions, either on the surface [44f] or in the structural defect sites [44g]. In these materials, many metal ions are concentrated in a defined volume and their biological behavior and relaxivities are different from those of the complex agents. Basically, this type of contrast agent is an extension of the paramagnetic complex agents so that the maximum number of ions is limited by the density of anchoring groups on the surface. Furthermore, the synthetic procedures are generally very complicated and expensive. The most significant limitation is their large overall size of >100 nm, which is much larger than
P1: OTA/XYZ P2: ABC c13 JWBS037-Chen January 25, 2011
12:49
Printer Name: Yet to Come
T2 NANOPARTICULATE MRI CONTRAST AGENTS
(a)
301
(b)
FIGURE 13.12 (a) Schematic illustration of Gd–hybrid mesoporous silica nanospheres. (b) Precontrast (left) and postcontrast (right) T1-weighted mouse MR images [44e]. (Reproduced with permission of the American Chemical Society.)
the inorganic nanoparticles, such as USPIO. To prevent easy excretion by the RES and to compete with the small-sized T2 contrast agents, their overall size should be <100 nm, which is satisfied by inorganic nanoparticles. Very recently, paramagnetic inorganic nanoparticles were developed as new T1 contrast agents. To be effective T1 contrast agents, the ratio between the transverse and longitudinal relaxivities (expressed as r2/r1), which is a defining parameter indicating whether the contrast agent can be employed as a positive or negative agent, has to be low [21]. That is, nanoparticles have a large paramagnetic property (large r1) with negligible magnetic anisotropy (small r2). Nanoparticles of the many compounds of transition metals and lanthanide metals could be good candidates for T1 MRI contrast agents because the surface of the nanoparticles contains a large amount of metal ions with high magnetic moments. The most obvious nanoparticles for T1 agents are gadolinium-based nanoparticles because many gadolinium complexes have been used extensively as clinical MRI contrast agents. Nanoparticles of gadolinium oxide (Gd2 O3 ) [45], gadolinium fluoride (GdF3 ) [46], and gadolinium phosphate (GdPO4 ) [47] have been investigated as MRI contrast agents. They exhibited signal-enhancing contrast in T1-weighted images with low r2/r1 values. The Gd2 O3 nanoparticle-based agents are generally composed of small core nanoparticles of <5 nm and stabilizing shells of dextran [45a], PEG [45b], and silica [45c]. Bridot et al. [45c] reported the use of biocompatible gadolinium oxide nanoparticles with polysiloxane shells containing a fluorescent dye (GadoSiPEG2C) for in vivo dual imaging of magnetic resonance and fluorescence imaging. Water-dispersible GdF3 (or GdF3 /LaF3 ) nanoparticles were prepared with either positively charged surface by conjugation with 2-aminoethyl phosphate groups (GdF3 /LaF3 :AEP) or negatively charged surface by coating with citrate groups (GdF3 :cit) [46]. Dextran-coated GdPO4 nanoparticles (PGP/dextranK01) were synthesized by a hydrothermal process in the presence of dextran [47]. These Gd-based nanoparticles had significantly low r2/r1 values and they showed strong positive contrast effects. However, uniformly sized nanoparticles of gadolinium or related lanthanide compounds have not yet been demonstrated. Very recently, MnO nanoparticles were developed as a new T1 MRI contrast agent for body organs such as the brain, liver, and kidney [48]. In particular, clear T1-weighted MR images of the brain structures, depicting fine anatomic structures, were obtained (Fig. 13.13). This clear anatomic imaging of the various brain structures can be utilized
P1: OTA/XYZ P2: ABC c13 JWBS037-Chen January 25, 2011
302
12:49
Printer Name: Yet to Come
MRI CONTRAST AGENTS BASED ON INORGANIC NANOPARTICLES
(a)
(b)
(c)
FIGURE 13.13 (a,b) Typical coronal (top), axial (middle), and sagittal (bottom) views of a T1weighted MONEMRI before (a) and after (b) the administration of the MnO nanoparticles. (c) Breast cancer cells were selectively enhanced in T1-weighted MRI by the Herceptin-functionalized MnO nanoparticles [48]. (Reproduced with permission of Wiley-VCH Verlag GmbH & Co. KGaA.)
for many applications, such as basic neuroscience research and the diagnosis of clinical neurological diseases (e.g., neurodegenerative diseases). Furthermore, functionalized MnO nanoparticles, by conjugation with a tumor-specific antibody, were used for the selective imaging of the breast cancer cells in a metastatic brain tumor (Fig. 13.13c). Cancer cells were selectively enhanced in T1-weighted MRI because the functionalized MnO nanoparticles with tumor specific-antibody were delivered to and accumulated in the cancer cells with a clear marginal detectability without destroying the anatomic background. This clear marginal detectability while preserving the anatomic background images is a marked superiority of using a T1 contrast agent over a T2 contrast agent.
P1: OTA/XYZ P2: ABC c13 JWBS037-Chen January 25, 2011
12:49
Printer Name: Yet to Come
T2 NANOPARTICULATE MRI CONTRAST AGENTS
303
Gilad et al. [49] reported the labeling and tracking of rat glioma cells with either MnO nanoparticles or SPIO (Figure 13.14). Cells labeled with MnO nanoparticles and SPIO were each transplanted in contralateral brain hemispheres of a rat and tracked by in vivo MRI. SPIO-labeled cells showed a strong negative contrast in T2/T2*-weighted MRI. However, it is difficult to distinguish them from blood/hemosiderin-associated hypointense regions. MnO-labeled transplanted cells had a higher R1 than the brain tissue surrounding the transplanted cells. Therefore MnO-labeled cells could be successfully detected and distinguished with positive contrast for in vivo T1-weighted MRI. The authors showed a MR “double labeling” technique using opposite contrasts simultaneously on two sites, where one is labeled with MnO and the other with SPIO nanoparticles. This labeling method can be applied to the tracking of cell populations that are injected in different locations or at different time points by complementing and amplifying two kinds of contrasting effect. Very recently, hollow structured manganese oxide nanoparticles were reported. The Hyeon group synthesized various hollow oxide nanoparticles via nanoscale acid etching using MnO as the starting material and alkylphosphonic acid impurity as the etchant [50a]. These hollow nanoparticles were paramagnetic at room temperature, and when dispersed in water they showed spin relaxation enhancement effect. Hollow manganese oxide nanoparticles were synthesized by selective removal of the core of nanoparticles in an
FIGURE 13.14 In vivo MR images of labeled rat glioma cells 24 h after transplantation in the striata of rat brain. (a–c): Spin echo images. (d–f): R1 maps. (g–i): R2 maps. (j–l): R1/R2 merged maps. MnO- and SPIO-labeled cells (a,b), and one with SPIO-labeled cells and unlabeled cells (c; control) [49]. (Reproduced with permission of Wiley-Liss, Inc.)
P1: OTA/XYZ P2: ABC c13 JWBS037-Chen January 25, 2011
304
12:49
Printer Name: Yet to Come
MRI CONTRAST AGENTS BASED ON INORGANIC NANOPARTICLES
acidic buffer and these hollow manganese oxide nanoparticles were applied simultaneously as MRI contrast agent and drug delivery vehicle [50a].
13.4 CONCLUSION In this chapter, we report the progress for MRI contrast agents based on inorganic nanoparticles. Over the last 20 years, various inorganic nanoparticles have been investigated as MRI contrast agents through the collaborative efforts of scientists working in material science, biology, and medicine. In particular, magnetic iron oxide nanoparticles have been used extensively as T2 MRI contrast agents due to their strong ability to shorten T2* relaxation times. Dextran-stabilized SPIOs have been used clinically for the diagnosis of liver diseases because SPIOs are selectively taken up by the Kupffer cells. Smaller iron oxide nanoparticles, USPIOs, have been applied for lymph node imaging and are under consideration for clinical uses. More recently, uniform nanoparticles with high crystallinity have been synthesized from the thermal decomposition of metal precursors at high temperature in the presence of surfactants in organic media. After proper surface modification to impart biocompatibility, these uniform ferrite nanoparticles have been employed successfully as new T2 MRI contrast agents with improved relaxation effects and multifunctional properties. Furthermore, the functionalization with bioactive materials enabled these nanoparticlebased contrast agents to be applicable to targeted imaging via the site-specific accumulation of nanoparticles at the targets of interest. Recently, extensive research has been conducted to develop nanoparticle-based T1 contrast agents to overcome the drawbacks of iron oxide nanoparticle-based negative T2 contrast agents. Gadolinium complexes immobilized in various nanostructured materials, including nanoporous silicas, dendrimers, perfluorocarbon nanoparticles, and nanotubes, have been investigated as new T1 MRI contrast agents due to their ability to carry a large number of paramagnetic payloads and produce strong T1 contrast. More recently, inorganic nanoparticles such as Gd2 O3 , GdF3 , and GdPO4 have been investigated as signal-generating cores of T1 relaxation. Uniformly sized MnO nanoparticles were developed as a new T1 MRI contrast agent for body organs such as the brain, liver, and kidney. In particular, clear T1-weighted MR images of the brain structures, depicting fine anatomic structures, were obtained. The development of multifunctional platforms for either multimodal imaging or simultaneous imaging and therapy has been intensively pursued through the combinations of various nanomaterials. Most of these newly developed nanoparticle-based MRI contrast agents are still in the stage of in vitro testing or preliminary animal studies. There are many issues to be clearly addressed prior to their extensive use in clinical diagnosis, including toxicological effects, long-term stability, and pharmacokinetics. Intensive interdisciplinary collaborative research is essential to achieve the ultimate goal of using nanoparticle-based MRI contrast agents for molecular and active imaging that can be applied to personalized diagnosis.
REFERENCES 1. Brown, M. A.; Semelka, R. C. MRI: Basic Principles and Applications. Wiley-Liss: Hoboken, NJ, 2003. 2. Caravan, P.; Ellison, J. J.; McMurry, T. J.; Lauffer, R. B. Chem. Rev. 1999, 99, 2293.
P1: OTA/XYZ P2: ABC c13 JWBS037-Chen January 25, 2011
12:49
Printer Name: Yet to Come
REFERENCES
305
3. (a) Mendonca Dias, M. H.; Lauterbur, P. C. Magn. Reson. Med. 1986, 3, 328. (b) Semelka, R. C.; Helmberger, T. K. Radiology 2001, 218, 27. 4. (a) Bulte, J. W. M.; Kraitchman, D. L. NMR Biomed. 2004, 17, 484. (b) Na, H. B.; Song, I. C.; Hyeon, T. Adv. Mater. 2009, 21, 2133. 5. (a) Jung, C. W.; Jacobs, P. Magn. Reson. Imaging 1995, 13, 661. (b) Wang, Y. X.; Hussain, S, M.; Krestin, G. P. Eur. Radiol. 2001, 11, 2319. (c) Wunderbaldinger, P.; Josephson, L.; Weissleder, R. Acad. Radiol. 2002, 9, S304. 6. (a) Weissleder, R.; Elizondo, G.; Wittenberg, J.; Rabito, C. A.; Bengele, H. H. Radiology 1990, 175, 489. (b) Palmacci, S.; Josephson, L. Synthesis of polysaccharide covered superparamagnetic oxide colloids, U.S. Patent 5,262,176, 1993. 7. Harisinghani, M. G.; Barentsz, J.; Hahn, P. F.; Deserno, W. M.; Tabatabaei, S.; van de Kaa, C. H.; de la Rosette, J.; Weissleder, R. N. Engl. J. Med. 2003, 348, 2491. 8. Lawaczeck, R.; Menzel, M.; Pietsch, H. Appl. Organometal. Chem. 2004, 18, 506. 9. Evgenov, N. V.; Medarova, Z.; Dai, G.; Bonner-Weir, S.; Moore, A. Moore. Nat. Med. 2006, 12, 144. 10. Bulte, J. W. M.; Douglas, T.; Witwer, B.; Zhang, S.-C.; Strable, E.; Lewis, B. K.; Zywicke, H.; Miller, B.; van Gelderen, P.; Moskowitz, B. M.; Duncan, I. D.; Frank, J. A. Nat. Biotechnol. 2001, 19, 1141. 11. Arbab, A. S.; Yocum, G. T.; Kalish, H.; Jordan, E. K.; Anderson, S. A.; Khakoo, A. Y.; Read, E. J.; Frank, J. A. Blood 2004, 104, 1217. 12. Walczak, P.; Kedziorek, D. A.; Gilad, A. A.; Lin, S.; Bulte, J. W. M. Magn. Reson. Med. 2005, 54, 769. 13. (a) Josephson, L.; Tung, C.-H.; Moore, A.; Weissleder, R. Bioconjug. Chem. 1999, 10, 186. (b) Hawrylak, N.; Ghosh., P.; Broadus, J.; Schlueter, C.; Greenough, W. T.; Lauterbur, P. C. Exp. Neurol. 1993, 121, 181. (c) Lewin, M.; Carlesso, N.; Tung, C.-H.; Tang, X.-W.; Cory, D.; Scadden, D. T.; Weissleder, R. Nat. Biotechnol. 2000, 18, 410. (d) Ahrens, E. T.; Feili-Hariri, M.; Xu, H.; Genove, G.; Morel, P. A. Magn. Reson. Med. 2003, 49, 1006. (e) de Vries, I. J.; Lesterhuis, W. J.; Barentsz, J. O.; Verdijk, P.; van Krieken, J. H.; Boerman, O. C.; Oyen, W. J.; Bonenkamp, J. J.; Boezeman, J. B.; Adema, G. J.; Bulte, J. W.; Scheenen, T. W.; Punt, C. J.; Heerschap, A.; Figdor, C. G. Nat. Biotechnol. 2005, 23, 1407. 14. Kang, H. W.; Josephson, L.; Petrovsky, A.; Weissleder, R.; Bogdanov, A. Jr. Bioconjug. Chem. 2002, 13, 122. 15. Josephson, L.; Kircher, M. F.; Mahmood, U.; Tang, Y.; Weissleder, R. Bioconjug. Chem. 2002, 13, 554–560. 16. (a) Perez, J. M.; Josephson, L.; O’Loughlin, T.; H¨ogemann, D.; Weissleder, R. Nat. Biotechnol. 2002, 20, 816. (b) Weissleder, R.; Kelly, K.; Sun, E. Y.; Shtatland, T.; Josephson, L. Nat. Biotechnol. 2005, 23, 1418. 17. (a) Fertman, V. E. Magnetic Fluids Guide Book: Properties and Applications. Hemisphere Publishing: New York, 1990. (b) Schmid, G. Nanoparticles: From Theory to Application. WileyVCH: Weinheim, Germany, 2004. (c) O’Handley, R. C. Modern Magnetic Materials: Principles and Applications. Wiley-Interscience: Hoboken, NJ, 1999. (d) Fendler, J. H. Nanoparticles and Nanostructured Films. Wiley-VCH: Weinheim, Germany, 1998. 18. (a) Hyeon, T. Chem. Commun. 2003, 927. (b) Jun, Y.-W.; Choi, J.-S.; Cheon, J. Angew. Chem. Int. Ed. 2006, 45, 3414. (c) Park, J.; Joo, J.; Gu K won, S.; Jang, Y.; Hyeon, T. Angew. Chem. Int. Ed. 2007, 46, 4630. (d) Kwon, S. G.; Hyeon, T. Acc. Chem. Res. 2008, 41, 1696. 19. (a) Park, J.; An, K.; Hwang, Y.; Park, J. G.; Noh, H. J.; Kim, J. Y.; Park, J. H.; Hwang, N. M.; Hyeon, T. Nat. Mater. 2004, 3, 891. (b) Joo, J.; Yu, T.; Kim, Y. W.; Park, H. M.; Wu, F.; Zhang, J. Z.; Hyeon, T. J. Am. Chem. Soc. 2003, 125, 6553. (c) Yu, T.; Joo, J.; Hyeon, T. Angew. Chem. Int. Ed. 2005, 44, 7411. (d) Kim, J. I.; Lee, J.-K.; Adv. Funct. Mater. 2006, 16, 207.
P1: OTA/XYZ P2: ABC c13 JWBS037-Chen January 25, 2011
306
12:49
Printer Name: Yet to Come
MRI CONTRAST AGENTS BASED ON INORGANIC NANOPARTICLES
20. (a) Hyeon, T.; Lee, S. S.; Park, J.; Chung, Y.; Na, H. B. J. Am. Chem. Soc. 2001, 123, 12798. (b) Sun, S.; Zeng, H.; Robinson, D. B.; Raoux, S.; Rice, P. M.; Wang, S. X.; Li, G. J. Am. Chem. Soc. 2004, 126, 273. (c) Choi, S. H.; Kim, E. G.; Park, J.; An, K.; Lee, N.; Kim, S. C.; Hyeon, T. J. Phys. Chem. B 2005, 109, 14792. (d) An, K.; Lee, N.; Park, J.; Kim, S. C.; Hwang, Y.; Park, J. G.; Kim, J. Y.; Park, J. H.; Han, M, J.; Yu, J.; Hyeon, T. J. Am. Chem. Soc. 2006, 128, 9753. 21. (a) Koenig, S. H.; Keller, K. E. Magn. Reson. Med. 1995, 34, 227. (b) Kellar, K. E.; Fujii, D. K.; Gunther, W. H. H.; Briley-Saebo, K.; Spiller, M.; Koenig, S. H. Magn. Reson. Mater. Phys. 1999, 8, 207. 22. (a) Park, J.; Lee, E.; Hwang, N.-M.; Kang, M.; Kim, S. C.; Hwang, Y.; Park, J.-G.; Noh, H.-J.; Kim, J.-Y.; Park, J.-H.; Hyeon, T. Angew. Chem. Int. Ed. 2005, 44, 2872. (b) Jun, Y.-W.; Huh, Y.-M.; Choi, J.-S.; Lee, J.-H.; Song, H.-T.; Kim, S.; Yoon, S.; Kim, K.-S.; Shin, J.-S.; Suh, J.-S.; Cheon, J. J. Am. Chem. Soc. 2005, 127, 5732. 23. (a) Morales, M. P.; Veintemillas-Verdaguer, S.; Montero, M. I.; Serna, C. J. Chem. Mater. 1999, 11, 3058. (b) Jun, Y.-W.; Seo, J.-W.; Cheon, J. Acc. Chem. Res. 2008, 41, 179. (c) Jun, Y.-W.; Lee, J.-H.; Cheon, J. Angew. Chem. Int. Ed. 2008, 47, 5122. 24. Lee, J.-H.; Huh, Y.-M.; Jun, Y.-W.; Seo, J.-W.; Jang, J.-T.; Song, H.-T.; Kim, S.; Cho, E.-J.; Yoon, H.-G.; Suh, J.-S.; Cheon, J. Nat. Med. 2007, 13, 95. 25. (a) Kumar, S. S. R.; Hormes, J.; Leuschner, C., Eds. Nanofabrication Towards Biomedical Applications: Techniques, Tools, Applications, and Impact. Wiley-VCH: Weinheim, Germany, 2005. (b) Kumar, C., Ed. Biofunctionalzation of Nanomaterials. Wiley-VCH: Weinheim, Germany, 2005. 26. (a) Xu, C.; Xu, K.; Gu, H.; Zheng, R.; Liu, H.; Zhang, X.; Guo, Z.; Xu, B. J. Am. Chem. Soc. 2004, 126, 9938. (b) Xie, J.; Xu, C.; Kohler, N.; Hou, Y.; Sun, S. Adv. Mater. 2007, 19, 3163. (c) Kim, M.; Chen, Y.; Liu, Y.; Peng, X. Adv. Mater. 2005, 17, 1429. (d) Na, H. B.; Lee, I. S.; Seo, H.; Park, Y. I.; Lee, J. H.; Kim, S.-W.; Hyeon, T. Chem. Commun. 2007, 5167. (e) Kim, S.-W.; Kim, S.; Tracy, J. B.; Jasanoff, A.; Bawendi, M. G. J. Am. Chem. Soc. 2005, 127, 4556. 27. Piao, Y.; Burns, A.; Kim, J.; Wiesner, U.; Hyeon, T. Adv. Funct. Mater. 2008, 18, 3745. 28. (a) Yi, D. K.; Selvan, S. T.; Lee, S. S.; Papaefthymiou, G. C.; Kundaliya, D.; Ying, J. Y. J. Am. Chem. Soc. 2005, 127, 4990. (b) Selvan, S. T.; Patra, P. K.; Ang, C. Y.; Ying, J. Y. Angew. Chem. Int. Ed. 2007, 46, 2448. (c) Yoon, T.-J.; Yu, K. N.; Kim, E.; Kim, J. S.; Kim, B. G.; Yun, S.-H.; Sohn, B.-H.; Cho, M.-H.; Lee, J.-K.; Park, S. B. Small 2006, 2, 209. (d) Lu, C.-W.; Hung, Y.; Hsiao, J.-K.; Yao, M.; Chung, T.-H.; Lin, Y.-S.; Wu, S.-H.; Hsu, S.-C.; Liu, H.-M.; Mou, C.-Y.; Yang, C.-S.; Huang, D.-M.; Chen, Y.-C. Nano Lett. 2007, 7, 149. 29. (a) Kim, J.; Lee, J. E.; Lee, J.; Yu, J. H.; Kim, B. C.; An, K.; Hwang, Y.; Shin, C.-H. O.; Park, J.-G.; Kim, J.; Hyeon, T. J. Am. Chem. Soc., 2006, 128, 688. (b) Giri, S.; Trewyn, B. G.; Stellmaker, M. P.; Lin, V. S.-Y. Angew. Chem. Int. Ed. 2005, 44, 5038. (c) Zhao, W.; Gu, J.; Zhang, L.; Chen, H.; Shi, J. J. Am. Chem. Soc. 2005, 127, 8916. (d) Lin, Y.-S.; Wu, S.-H.; Hung, Y.; Chou, Y.-H.; Chang, C.; Lin, M.-L.; Tsai, C.-P.; Mou, C.-Y. Chem. Mater. 2006, 18, 5170. (e) Deng, Y.; Qi, D.; Deng, C.; Zhang, X.; Zhao, D. J. Am. Chem. Soc. 2008, 130, 28. (f) Gorelikov, I.; Matsuura, N. Nano Lett. 2008, 8, 369. (g) Kim, J.; Kim, H. S.; Lee, N.; Kim, T.; Kim, H.; Yu, T.; Song, I. C.; Moon, W. K.; Hyeon, T. Angew. Chem. Int. Ed. 2008, 47, 8438. 30. (a) Panyam, J.; Labhasetwar, V. Adv. Drug Delivery Rev. 2003, 55, 329. (b) Nasongkla, N.; Bey, E.; Ren, J.; Ai, H.; Khemtong, C.; Guthi, J. S.; Chin, S.-F.; Sherry, A. D.; Boothman, D. A.; Gao, J. Nano Lett. 2006, 6, 2427. (c) Kim, J.; Lee, J. E.; Lee, S. H.; Yu, J. H.; Lee, J. H.; Park, T. G.; Hyeon, T. Adv. Mater. 2008, 20, 478. (d) Ding, Y.; Hu, Y.; Jiang, X.; Zhang, L.; Yang, C. Angew. Chem. Int. Ed. 2004, 43, 6393. (e) Ding, Y.; Hu, Y.; Zhang, L.; Chen, Y.; Jiang, X. Biomacromolecules 2006, 7, 1766. 31. (a) Golander, C. G.; et al., In Poly(ethylene glycol) Chemistry; Harris, J. M.; Ed., Plenum: New York, 1992, p. 221. (b) Gref, R.; Minamitake, Y.; Peracchia, M. T.; Trubetskoy, V.; Torchilin, V.; Langer, R. Science 1994, 263, 1600. (c) Bazile, D.; Prud’homme, C.; Bassoullet, M.; Marlard, M.;
P1: OTA/XYZ P2: ABC c13 JWBS037-Chen January 25, 2011
12:49
Printer Name: Yet to Come
REFERENCES
32.
33.
34. 35.
36. 37. 38. 39. 40. 41.
42. 43. 44.
45.
46. 47.
307
Spenlehauer, G.; Veillard, M. J. Pharm. Sci. 1995, 84, 493. (d) Tobio, M.; Gref, R.; Sanchez, A.; Langer, R.; Alonso, M. J. Pharm. Res. 1998, 15, 270. (a) Dubertret, B.; Skourides, P.; Norris, D. J.; Noireaux, V.; Brivanlou, A. H.; Libchaber, A. Science 2002, 298, 1759. (b) LaConte, L. E.W.; Nitin, N.; Zurkiya, O.; Caruntu, D.; O’Connor, C. J.; Hu, X.; Bao, G. J. Magn. Reson. Imaging 2007, 26, 1634. (a) Qin, J.; Laurent, S.; Jo, Y. S.; Roch, A.; Mikhaylova, M.; Bhujwalla, Z. M.; Muller, R. N.; Muhammed, M. Adv. Mater. 2007, 19, 1874. (b) Gonzalesa, M.; Krishnan, K. M. J. Magn. Magn. Mater. 2007, 311, 59. Wang, Y.; Wong, J. F.; Teng, X.; Lin, X. Z.; Yang, H. Nano Lett. 2003, 3, 1555. Tromsdorf, U. I.; Bigall, N. C.; Kaul, M. G.; Bruns, O. T.; Nikolic, M. S.; Mollwitz, B.; Sperling, R. A.; Reimer, R.; Hohenberg, H.; Parak, W. J.; Fo1rster, S.; Beisiegel, U.; Adam, G.; Weller, H. Nano Lett. 2007, 7, 2422. Maeda, H.; Fang, J.; Inutsuka, T.; Kitamoto, Y. Int. Immunopharmacol. 2003, 3, 319. Lee, H.; Lee, E.; Kim, D. K.; Jang, N. K.; Jeong, Y. Y.; Jon, S. J. Am. Chem. Soc. 2006, 128, 7383. Hu, F.; Wei, L.; Zhou, Z.; Ran, Y.; Li, Z.; Gao, M. Adv. Mater. 2006, 18, 2553. Sun, C.; Veiseh, O.; Gunn, J.; Fang, C.; Hansen, S.; Lee, D.; Sze, R.; Ellenbogen, R. G.; Olson, J.; Zhang, M. Small 2008, 4, 372. Kim, J.; Piao, Y.; Hyeon, T. Chem. Soc. Rev. 2009, 38, 372. (a) Xu, C.; Xie, J.; Ho, D.; Wang, C.; Kohler, N.; Walsh, E. G.; Morgan, J. R.; Chin, Y. E.; Sun, S. Angew. Chem. Int. Ed. 2008, 47, 173. (b) Choi, J.-S.; Jun, Y.-W.; Yeon, S.-I.; Kim, H. C.; Shin, J.-S.; Cheon, J. J. Am. Chem. Soc. 2006, 128, 15982. (c) Selvan, S. T.; Patra, P. K.; Ang, C. Y.; Ying, J. Y. Angew. Chem. Int. Ed. 2007, 46, 2448. (d) Choi, S.-H.; Na, H. B.; Park, Y. I.; An, K.; Kwon, S. G.; Jang, Y.; Park, M.; Moon, J.; Son, J. S.; Song, I. C.; Moon, W. K.; Hyeon, T. J. Am. Chem. Soc. 2008, 130, 15573. Piao, Y.; Kim, J.; Na, H. B.; Kim, D.; Baek, J. S.; Ko, M. K.; Lee, J. H.; Shokouhimehr, M.; Hyeon, T. Nat. Mater. 2008, 7, 242. Silva, C.; Lee, J. H.; Aoki, I.; Koretsky, A. P. NMR Biomed. 2004, 17, 532. (a) Flacke, S.; Fischer, S.; Scott, M. J.; Fuhrhop, R. J; Allen, J. S.; McLean, M.; Winter, P.; Sicard, G. A.; Gaffney, P. J.; Wickline, S. A.; Lanza, G. M. Circulation 2001, 104, 1280. (b) Winter, P. M.; Morawski, A. M.; Caruthers, S. D.; Fuhrhop, R. W.; Zhang, H.; Williams, T. A.; Allen, J. S.; Lacy, E. K.; Robertson, J. D.; Lanza, G. M.; Wickline, S. A. Circulation, 2003, 108, 2270. (c) Rieter, W. J.; Kim, J. S.; Taylor, K. M. L.; An, H.; Lin, W.; Tarrant, T.; Lin, W. Angew. Chem. Int. Ed. 2007, 46, 3680. (d) Rieter, W. J.; Taylor, K. M. L.; An, H.; Lin, W.; Lin, W. J. Am. Chem. Soc. 2006, 128, 9024. (e) Taylor, K. M. L.; Kim, J. S.; Rieter, W. J.; An, H.; Lin, W.; Lin, W. J. Am. Chem. Soc. 2008, 130, 2154. (f) Richard, C.; Doan, B.-T.; Beloeil, ´ Scherman, D. Nano Lett. 2008, 8, 232. (g) Sitharaman, B.; Kissell, J.-C.; Bessodes, M.; T´oth, E.; K. R.; Hartman, K. B.; Tran, L. A.; Baikalov, A.; Rusakova, I.; Sun, Y.; Khant, H. A.; Ludtke, S. ´ Helm, L.; Merbachd, A. E.; Wilson, L. J. Chem. Commun. 2005, J.; Chiu, W.; Laus, S.; T´oth, E.; 3915. (a) McDonald, M. A.; Watkin, K. L. Acad. Radiol. 2006, 13, 421. (b) Fortin, M.-A.; Petoral., R. M., Jr.; S¨oderlind, F.; Klasson, A.; Engstr¨om, M.; Veres, T.; K¨all, P.-O.; Uvdal, K. Nanotechnology 2007, 18, 395501. (c) Bridot, J.-L.; Faure, A.-C.; Laurent, S.; Rivi`ere, C.; Billotey, C; Hiba, B.; Janier, M.; Josserand, V.; Coll, J.-L.; Elst, L. V.; Muller, R.; Roux, S.; Perriat, P.; Tillement, O. J. Am. Chem. Soc. 2007, 129, 5076. Evanics, F.; Diamente, P. R.; van Veggel, F. C. J. M.; Stanisz, G. J.; Prosser, R. S. Chem. Mater. 2006, 18, 2499. Hifumi, H.; Yamaoka, S.; Tanimoto, A.; Citterio, D.; Suzuki, K. J. Am. Chem. Soc. 2006, 128, 15090.
P1: OTA/XYZ P2: ABC c13 JWBS037-Chen January 25, 2011
308
12:49
Printer Name: Yet to Come
MRI CONTRAST AGENTS BASED ON INORGANIC NANOPARTICLES
48. Na, H. B.; Lee, J. H.; An, K.; Park, Y. I.; Park, M.; Lee, I. S.; Nam, D.-H.; Kim, S. T.; Kim, S.-H.; Kim, S.-W.; Lim, K.-H.; Kim, K.-Soo; Kim, S.-O.; Hyeon, T. Angew. Chem. Int. Ed. 2007, 46, 5397. 49. Gilad, A. A.; Walczak, P.; McMahon, M. T.; Na, H. B.; Lee, J. H.; An, K.; Hyeon, T.; van Zijl, P. C. M.; Bulte, J. W. M. Magn. Reson. Med. 2008, 60, 1. 50. (a) An, K.; Kwon, S. G.; Park, M.; Na, H. B.; Baik, S.-I.; Yu, J. H.; Kim, D.; Son, J. S.; Kim, Y. W.; Song, I. C.; Moon, W. K.; Park, H. M.; Hyeon, T. Nano Lett. 2008, 8, 4252. (b) Shin, J.; Anisur, R. M.; Ko, M. K.; Im, G. H.; Lee, J. H.; Lee, I. S. Angew. Chem. Int. Ed. 2009, 48, 321.
P1: OTA/XYZ P2: ABC c14 JWBS037-Chen February 1, 2011
0:6
Printer Name: Yet to Come
CHAPTER 14
Cellular Magnetic Labeling with Iron Oxide Nanoparticles ´ SEBASTIEN BOUTRY, SOPHIE LAURENT, LUCE VANDER ELST, and ROBERT N. MULLER Department of General, Organic and Biomedical Chemistry, NMR and Molecular Imaging Laboratory, University of Mons, Mons, Belgium
14.1 INTRODUCTION For molecular magnetic resonance imaging (MRI), an amplification of the molecular labeling is required. This can be achieved, for example, by avidin–biotin systems in the extracellular compartment [1]. Cellular internalization of the contrast agent is a way to increase the labeling of targeted cells. The challenge for cellular MRI is obtaining a sufficient uptake of MRI contrast agents in cells, especially nonphagocytic cells, in order to render them distinct from the surrounding tissue. Although this approach gives less specific results than molecular MRI, it offers suitable systems for the noninvasive and repeated in vivo monitoring (identification and tracking) of cells by MRI. Magnetic labeling can be realized in vitro or in vivo through specific (involving a cell surface receptor) or nonspecific cellular internalization pathways with contrast agents considered as positive (complexes of paramagnetic ions such as gadolinium(III) or manganese(II); shortening the T1 of their adjacent hydrogen nuclei as a predominant effect and brightening the regions taking them up on a T1-weighted MR image), and often with contrast agents considered as negative (superparamagnetic particles based on a monocrystalline or polycrystalline iron oxide core, shortening the T2 and T2* much more than the T1 of hydrogen nuclei in their neighborhood and generating a signal darkening on T2 (T2* )-weighted MR images). Many studies have indeed been performed in vitro and in vivo with iron oxide nanoparticles [2, 3]. Another mandatory point about cellular MRI is that the loading of cells with contrast agents does not have to affect normal cell function [3]. This chapter mainly focuses on the use of native (unmodified) or unspecifically cell-targeted iron oxides as cell markers. Some examples of cellular labeling performed with other MRI contrast agents, or obtained via specific internalization pathways, are also given.
Nanoplatform-Based Molecular Imaging Edited by Xiaoyuan Chen C 2011 John Wiley & Sons, Inc. Copyright
309
P1: OTA/XYZ P2: ABC c14 JWBS037-Chen February 1, 2011
310
0:6
Printer Name: Yet to Come
CELLULAR MAGNETIC LABELING WITH IRON OXIDE NANOPARTICLES
14.2 SPECIFIC MAGNETIC LABELING OF CELLS This type of magnetic labeling occurs via receptor-mediated endocytosis (Fig. 14.1), which is a highly specific and efficient mechanism of internalization. The transferrin receptor (TfR) pathway has been used to label cells in several studies. Transferrin (Tf), an iron chelating protein, interacts with TfR and enters cells by receptor-mediated endocytosis, supplying them with iron from stores in the liver. As iron is an essential element involved in cell proliferation, a high number of TfRs is expressed in proliferating cells [4]. Thus, this mechanism of iron uptake appears as an interesting marker for tumor cells, or an interesting way to label progenitor cells. Furthermore, the high turnover rate of the TfR avoids the phenomenon of receptor saturation [5]. Oligodendrocyte progenitors (CG-4) have been magnetically labeled in culture with monocrystalline iron oxide nanocompound (MION) covalently linked to an antibody (OX-26) specific for TfR. The binding of OX-26 to TfR induced the internalization of the nanoparticles. Labeled cells were tracked by MRI after transplantation into the spinal cord of myelin-deficient rats and were found to synthesize myelin [6]. In another study, the in vitro receptor-mediated endocytosis of Tf covalently linked to ultrasmall superparamagnetic iron oxide (USPIO) (Tf-USPIO) by human epidermoid carcinoma A431 was observed. Tf-USPIO was also injected intravenously (200 mol Fe/kg) into rats bearing implanted mammary carcinoma. Despite the fact that most of the injected dose was captured by the liver and the spleen, MR imaging results showed a signal reduction in the tumor only with Tf-USPIO and not with the parent compound or a nonspecific albumin-bound USPIO, suggesting a transferrin-mediated endocytosis of USPIO by the tumor [5]. In another approach, rat 9L gliosarcoma cells were transfected with a modified gene coding for the human TfR, to constitutively express high levels of the receptor protein. The engineered cells were implanted in mice flank and a strong negative contrast was detected in the resulting tumors after intravenous injection of MION conjugated to human holo-transferrin (Tf-MION) (3 mg Fe/animal). This result suggested an efficient uptake of Tf-MION by tumor cells and provided the possibility to image transgene expression [7]. The approach, consisting of the manipulation of iron metabolism genes to provide cellular iron accumulation and subsequent MRI contrast, was further developed in a study where coexpression of transgenic human TfR and human ferritin H-subunit was induced in a stably transfected neural stem cell line (C17) [8]. Ferritin (FT) is composed of H- and L-subunits. Twenty-four subunits form the apoferritin, which is able to incorporate up to 4500 iron atoms. The H-subunit has a ferroxidase activity that promotes iron oxidation and incorporation [9, 10]. Thus FT-bound iron has a semicrystalline structure and acts as an endogenous cellular negative contrast agent. These studies demonstrated that transfected cells grown in iron-rich medium incorporated significantly more iron than control cells, inducing a contrast effect in MRI that was preserved after transplantation
Extracellular compartment Cell membrane
Intracellular compartment CA Endocytic vesicle formation
CA Specific ligand of the cell surface receptor Cell surface receptor
FIGURE 14.1 Schematic representation of the receptor-mediated endocytosis of a contrast agent (CA).
P1: OTA/XYZ P2: ABC c14 JWBS037-Chen February 1, 2011
0:6
Printer Name: Yet to Come
NONSPECIFIC MAGNETIC LABELING OF CELLS
311
of cells into the mouse brain. After uptake through TfR, iron was stored by FT without affecting cell viability [8]. The usefulness of FT as an endogenous contrast agent has been reported also after inducing the overexpression of this protein by C6 rat glioma cells. Indeed, intracellular iron is redistributed in excess FT, which generates negative contrast. As iron chelation induces a transient decrease of intracellular iron concentration, TfR is expressed in compensation to increase the iron uptake and restore iron homeostasis. This newly endocytosed iron will also be sequestered by overexpressed FT, leading to an enhancement of the contrast effect [11]. The internalization pathway of a vitamin has also been used often to shuttle iron oxides in cells. The folate (or vitamin B9 ) receptor has been targeted with folic acid-linked MRI contrast agents in order to induce an uptake of the cell tag by cancer cells overexpressing the receptor [12]. The internalization of superparamagnetic nanoparticles bearing chlorotoxin in the cytoplasm of 9L glioma cells has been reported. This 36 amino acid peptide indeed renders the nanoprobe capable of targeting tumor cells expressing membrane-bound matrix metalloproteinase-2 [13]. Prostate cancer cells could be labeled via receptor-mediated endocytosis with iron oxides conjugated to an antibody specific for the prostate-specific membrane antigen (PSMA) [14]. Intracellular accumulation of iron oxide nanoparticles in cells has been achieved also by targeting hormone receptors. The single chain anti-epidermal growth factor receptor (EGFR) antibody has been bound to iron oxide nanoparticles to serve as an intracellular marker of EGFR-expressing cancer cells [15]. Luteinizing hormone releasing hormone (LHRH) and luteinizing hormone/chorionic gonadotropin (LH/CG)conjugated iron oxide nanoprobes have been used to label breast cancer cells [16].
14.3 NONSPECIFIC MAGNETIC LABELING OF CELLS This type of magnetic labeling can occur via natural mechanisms of endocytosis, which are not receptor dependent. In this context of nonspecific magnetic labeling of cells, most of the studies concern the in vitro labeling of nonphagocytic cells potentially useful for cell therapy. Among these cells of interest we find mesenchymal cells, embryonic stem cells, neural stem cells, oligodendrocyte progenitor cells, endothelial progenitor cells, and muscle progenitor cells. In other words, undifferentiated or less differentiated cells that, after their reintroduction in a living organism, could hopefully reach the damaged area (ischemic lesion) of an organ and differentiate to replace dead cells or, more specifically for endothelial precursors, reach a tumor were angiogenesis is occurring. Many in vitro cellular magnetic labeling studies are performed also on CD34+ hematopoietic stem cells or immune cells (lymphocytes) because the study of their in vivo trafficking could help us to better understand the immune response and, more particularly, to improve cell-based cancer therapies [2, 3, 17]. Thus visualization of the in vivo behavior of these cells is crucial and depends on their prelabeling in culture. Furthermore, the study of contrast agent accumulation in phagocytes and tumor cells can become a tool for the biological characterization of pathologies [18, 19]. 14.3.1 Fluid-Phase Endocytosis of Contrast Agents With or Without the Help of a Transfection Agent: Influence of the Contrast Agent Surface Charge Gadolinium-based contrast agents have been developed for the in vitro labeling of cells. Gd chelates coupled to fluorescent rhodamine dextran (GRID), or the combined use of Gd
P1: OTA/XYZ P2: ABC c14 JWBS037-Chen February 1, 2011
312
0:6
Printer Name: Yet to Come
CELLULAR MAGNETIC LABELING WITH IRON OXIDE NANOPARTICLES -
COO-
OOC N
N Eu3+
N
N
CH3
-
OOC OH
FIGURE 14.2 Structure of Eu-HP-DO3A. R and europium (Eu, fluorescent) complexed by the chelate HP-DO3A (ProHance , Bracco Diagnostics, Milan, Italy) (Fig. 14.2), constitute bimodal agents. Indeed, the detection of reimplanted labeled cells can be achieved by in vivo MRI and corroborated by fluorescent histology [3, 20, 21]. Lanthanide ions such as Eu(III) or terbium(III) (Tb(III)) have been used to build complexes called paramagnetic chemical exchange saturation transfer (PARACEST) agents. When the mobile (exchangeable) protons of these complexes are irradiated at their absorption (resonance) frequency, they transfer saturated magnetization to the surrounding bulk water, acting so as negative contrast agents. Thus contrast alteration on MR images is based on a phenomenon called chemical exchange dependent saturation transfer (CEST). As the resonance frequency of the exchangeable protons of the complex is influenced differently according to the type of paramagnetic ion used, PARACEST agents can generate their own specific contrast when irradiated at the specific absorption frequency of their mobile protons. In other words, cells labeled in vitro with different PARACEST agents, by internalization of the complex through pinocytosis, could be distinguished and separately tracked in vivo (Fig. 14.3a) [22]. Several attempts have been made to improve the in vitro cellular uptake of Gd chelates. Gd-DTPA has been combined with transfection agents such as calcium phosphate or the lipofection agent LipofectamineTM , usually facilitating the delivery of functional DNA into the cell [23]. Unmodified dextran-coated iron oxide nanoparticles can be used to label cells in vitro. Lymphocytes or T cells in culture have been loaded with this type of contrast agent through fluid-phase endocytosis [24, 25]. Rat bone marrow mesenchymal cells and mouse
Cell penetrating peptide Cell membrane
Extracellular compartment d
a
CA
Intracellular compartment
CA c Endocytic vesicle formation
CA b
CA
Cell transfection reagent
FIGURE 14.3 Schematic representation of the uptake of contrast agents (CAs) by cells through nonspecific (receptor-independent) endocytic processes: (a) fluid-phase endocytosis (pinocytosis) of the native CA, (b) endocytosis after adsorption of the CA mixed with a cell transfection reagent to the cell surface, (c) endocytosis after adsorption of the native CA to the cell surface, and (d) endocytosis after adsorption of a “cell penetrating peptide”-conjugated CA conjugated to the cell surface.
P1: OTA/XYZ P2: ABC c14 JWBS037-Chen February 1, 2011
0:6
Printer Name: Yet to Come
NONSPECIFIC MAGNETIC LABELING OF CELLS
313
embryonic stem cells were simply incubated with superparamagnetic iron oxide (SPIO) R ) at a relatively high concentration (around 100 g Fe/mL in the cell culture (Endorem medium for 48–72 h of incubation) [26–29]. In both cases, the in vivo fate of labeled cells was successfully MR imaged in vivo. However, in many studies, the cellular uptake R ) is optimized by mixing the superparamagnetic contrast agent with a of SPIO (Feridex cationic transfection agent such as poly-l-lysine. It is hypothesized that these dextran-coated SPIO nanoparticles can be coated with the positively charged transfection agent through R R (or Endorem ) electrostatic interactions. The citrate-containing formulation of Feridex may allow for the presence of a few carboxyl groups at the particle surface, which was indeed R ) [30, 31]. Thanks to found to be negatively charged (zeta potential of -41 mV for Feridex the positive charges of poly-l-lysine, the complex rapidly binds to the negatively charged cell membrane and is shuttled into cells through endosomes, producing a cellular magnetic labeling equivalent to the loading obtained by simply adding SPIO to the culture medium of cells (10–20 pg Fe/cell). However, the iron incubation concentration is lower (25 g Fe/mL (vs. ∼100 g Fe/mL)) and the magnetic labeling duration can potentially be shorter (2–48 h, depending on the cell type (vs. 48–72 h)). In the case of human mesenchymal stem cells, an iron content of 16 pg/cell has been found after a 2-h incubation [3, 32, 33]. The same approach has been developed with other commercial transfection agents, which have been combined with iron oxide nanoparticles in order to form complexes possessing a net positive charge that can associate with the negatively charged cell surface [34]. The resulting increase of contrast agent uptake has allowed for an improvement of the magnetic labeling of cultured cells. Another polycationic amine, protamine sulfate (Food and Drug Administration apR ) proved), has been used instead of poly-l-lysine to form complexes with SPIO (Feridex + TM and label human mesenchymal stem cells or CD34 cells [35]. Superfect consists of activated dendrimers with a spherical architecture and branches radiating from a central core and terminated by positively charged amino groups [36]. This commercially available transfection reagent has also been suitable for the magnetic labeling of mammalian R ). An iron content of 30 pg Fe/cell has been obtained for hucells with SPIO (Feridex R complex man mesenchymal stem cells incubated for 2 h with the SuperfectTM –Feridex R (25 g Fe/mL), a value twice higher than with poly-l-lysine–Feridex complex under the same conditions [33, 35]. The mechanisms initiating the endocytosis of dendrimers and polyamine-complexed iron oxide nanoparticles are thought to be of an electrostatic origin. Indeed, the adsorption of the positively charged complex on the negatively charged cell surface is supposed to induce membrane bending and membrane disruption. These phenomena induce invaginations of the cell membrane, leading to the encapsulation of the complexes in endosomes (Fig. 14.3b) [3, 32]. A carboxylated dendrimer has also been used as a coating for iron oxide nanoparticles. The carboxylate groups gave an anionic surface to the dendrimers. The resulting magnetodendrimer, called MD-100 (Fig. 14.4), has successfully served as a magnetic marker for different cultured cell types, originating from mouse, rat, and human, including stem cells. The uptake mechanism of MD-100 is also believed to result from electrostatic interactions between the dendrimer and the cell surface. The highly charged polymers bind on multiple sites on the cell membrane, inducing bending and disruption of the plasma membrane. An amount of 9 pg Fe/cell has been reached for CG-4 rat oligodendrocyte progenitors after 48 hours of incubation with MD-100 (25 g Fe/mL) [3, 37, 38]. A binding of nanoparticles bearing negative charges to the cell surface, which is also mainly negatively charged, appears as contradictory, and the reason for such an event is unknown. Nevertheless, a high level of
P1: OTA/XYZ P2: ABC c14 JWBS037-Chen February 1, 2011
314
0:6
Printer Name: Yet to Come
CELLULAR MAGNETIC LABELING WITH IRON OXIDE NANOPARTICLES
γ-Fe2O3 magnetodendrimers
FIGURE 14.4 Structure of magnetodendrimers (adapted from [32]).
cell internalization, comparable to the one obtained with the magnetodendrimer MD-100, has been obtained also with anionic iron oxide nanoparticles. The negative surface charge of these nanoparticles, due to the unbound carboxylate groups of their dimercaptosuccinic acid coating, has been hypothesized to allow them to interact strongly and nonspecifically with cationic sites of the plasma membrane, which are much more scarcely distributed than anionic areas. This mechanism involving an adsorption of the nanoparticles on the cell surface through electrostatic interactions, preceding their internalization, is called adsorptive endocytosis (Fig. 14.3c) [39]. Commercially available anionic nanoparticles (carboxydextran-coated SPIO R R )) have been found to be more efficient than dextran-coated SPIO (Feridex ) to (Resovist label cultured cells. Furthermore, the effectiveness of the intracellular uptake of nanoparticles has been found to be dependent on the density of carboxyl groups at the nanoparticle surface (with there being an optimum density), as well as on the cell type. The uptake of anionic nanoparticles has indeed been observed to be more pronounced in mesenchymal stem cells than in human cervical carcinoma (HeLa) cells, and a lower density of carboxyl groups was necessary to obtain an effective uptake of nanoparticles by mesenchymal stem R was captured in a higher amount cells compared to the latter cancer cells. Indeed, Feridex by mesenchymal stem cells than by HeLa cells. It was hypothesized that the citrate used R might allow for the presence of a few carboxyl groups at in the preparation of Feridex R the surface of nanoparticles, which would be sufficient to trigger an uptake of Feridex by mesenchymal stem cells, but not by HeLa cells [30]. The magnetic labeling of cultured nonphagocytic adherent cells with iron oxide nanoparticles has been found to be dependent on the relative concentration of the magnetic tag and of the cells in culture, on the nanoparticle hydrodynamic diameter, and on the coating charge (Fig. 14.5) [40]. 14.3.2 Cellular Delivery of Contrast Agents by Cationic Liposomes The same approach as for dendrimers and polyamines has been developed with liposomal transfection reagents [34]. Several types of superparamagnetic iron oxide nanoparticles, R R R , Endorem , Sinerem (dextran-coated), or P7228 (covered by an including Feridex anionic dextran derivative), have been incorporated into cationic transfection liposomes (such as LipofectamineTM or FuGENETM ). The lipid complex, carrying a net positive charge, binds to the negatively charged cell surface (Fig. 14.3b). The delivery of the contrast agent to the cytoplasm then depends on the liposome–cell membrane fusion. It has been demonstrated that this fusion only occurred after uptake of the liposome into the endocytic
P1: OTA/XYZ P2: ABC c14 JWBS037-Chen February 1, 2011
0:6
Printer Name: Yet to Come
NONSPECIFIC MAGNETIC LABELING OF CELLS
315
FIGURE 14.5 T2-weighted MR images (TR/TE: 3000/15 ms, 16th echo (240 ms), 5 × 105 cells/mL in 2% gelatin) of 3T6 fibroblasts, after 48 h of SPIO labeling with 50 g Fe/mL of R R or Resovist (upper row samples). Samples imaged on the bottom row (from left to Endorem R R at 0.01 mM Fe, and 2% right: Resovist nanoparticles at 0.01 mM Fe, unlabeled cells, Endorem gelatin) are for comparison (from [40]).
pathway [41]. These cationic liposome–iron oxide nanoparticle complexes have been used to magnetically label several types of mammalian cells including mouse embryonic stem cells, human mesenchymal stem cells, rat oligodendrocyte progenitor CG-4 cells, rabbit skeletal myoblasts, human hematopoietic progenitor cells, mouse lymphocytes, or HeLa cells [29, 33, 42, 43]. The iron content of human mesenchymal stem cells was 7.6 pg/cell R complex. In other after a 2-h incubation with 25 g Fe/mL of LipofectamineTM –Feridex R complex used terms, this loading was twice lower than with the poly-l-lysine-Feridex under the same conditions [33]. 14.3.3 Cellular Delivery of Contrast Agents Through a Cell Penetrating Peptide Another nonspecific way of cellular internalization exists, since some natural proteins, such as the human immunodeficiency virus (HIV) Tat (transactivator of transcription) protein, have the ability to penetrate cell membranes directly. The basic domain of the HIV Tat protein, formed of an arginine-rich sequence, has been identified as the domain responsible for the translocation of this protein. The highly cationic peptide derived from this basic domain is called a cell-penetrating peptide because of its ability to cross the plasma membrane and consequently to drag the rest of the protein with it [44, 45]. The mechanism by which this task is performed is based on an interaction of the peptide with glycosaminoglycans attached to cell surface heparan sulfate proteoglycans, initiating an adsorptive endocytic process (Fig. 14.3d) [46, 47]. Gd chelates and iron oxide nanoparticles have been conjugated to the Tat basic domain peptide (or more simply, Tat peptide) in order to efficiently transport the contrast agents into cells. Mammalian cells have been labeled successfully with Tat peptide-derivatized Gd-DOTA,[48, 49] or superparamagnetic contrast agents linked to the Tat peptide [50–52]. Human CD34+ hematopoietic stem cells labeled for 1 h with the Tat peptide-conjugated iron oxide nanoparticles (100 g Fe/mL) contained from 10 up to 30 pg Fe/cell [50].
P1: OTA/XYZ P2: ABC c14 JWBS037-Chen February 1, 2011
316
0:6
Printer Name: Yet to Come
CELLULAR MAGNETIC LABELING WITH IRON OXIDE NANOPARTICLES
14.3.4 Instant Cellular Delivery of Contrast Agents by Electroporation This method, called magnetoelectroporation (MEP), circumvents a prolonged incubation of cells and the use of a transfection agent, and allows achieving an instant magnetic labeling of cells. Mesenchymal stem cells and neural stem cells have been labeled with R SPIO (Feridex ) by the MEP method. The nanoparticles were taken up in endosomes and the iron content of cells (picogram range/cell) was comparable to values obtained with transfection agents. Furthermore, when used under properly calibrated conditions (a pulse of 130 V for 17 ms), MEP is claimed to have no effect on cell viability, proliferation, or differentiation [53]. 14.3.5 Fluid-Phase Endocytosis of Contrast Agents with Micron-Sized Particles Some advantages have been found in the use of micrometer-sized particles of iron oxide (MPIO) to label cells, as compared to the dextran-coated iron oxide nanoparticles SPIO or USPIO. For the latter, the endocytosis of millions of nanometer-sized particles is required to achieve sufficient contrast, and cell division can dilute the label beyond the detection threshold. MPIO (Bangs Laboratories, Fishers, Indiana, USA) are magnetite cores encapsulated with styrene/divinyl benzene. Various sizes exist (from 0.76 to 5.80 m in diameter). The smallest ones (0.76–1.63 nm) also contain a fluorescein-5-isothiocyanate analog (Dragon Green). Human CD34+ hematopoietic progenitor cells and porcine mesenchymal stem cells have been reported to efficiently internalize 0.9-m MPIO through endocytosis under incubation conditions (concentration and duration) comparable to those used with dextrancoated nanoparticles (e.g., 18 h with 45.6 g Fe/mL), and to lead to an iron load of hundreds of picograms per cell [54]. The styrene/divinyl benzene coating is thought to be inert inside cells and cannot be degraded within the cell, which is not the case with the dextran coating [55]. As the T2* effect of an iron oxide particle increases with its size, these MPIO have a larger T2* effect than SPIO in MRI, making possible the detection of single particles by in vitro MRI with a 50-m resolution in agarose phantoms or in cultured cells [56]. The possibility to MR image labeled cells by detecting single endocytosed particles overcomes the problems related to the dilution of the magnetic label induced by cell divisions [57]. Furthermore, it has been shown that mouse hepatocytes, prelabeled with 1.63-m MPIO, were detectable by MRI as single cells in the liver of recipient mice 1 month after their transplantation into the spleen [58].
14.4 MAGNETIC LABELING OF PHAGOCYTES R ) undergo a higher It has been demonstrated that dextran-coated SPIO (Endorem R ) in vitro, macrophage uptake than the smaller-sized dextran-coated USPIO (Sinerem confirming their respective in vivo biodistribution; SPIO are rapidly cleared from the circulation as a result of their capture mainly by liver Kuppfer cells, and USPIO have a longer blood half-life, allowing them to reach macrophages situated in less accessible territories, such as lymph nodes [59]. Scavenger receptors (SRs) are relatively nonspecific receptors mainly expressed by macrophages. These receptors bind to a broad range of negatively charged macromolecules (class A) or modified low-density lipoprotein (LDL) (class A and class B). The class A SR
P1: OTA/XYZ P2: ABC c14 JWBS037-Chen February 1, 2011
0:6
Printer Name: Yet to Come
MAGNETIC LABELING OF PHAGOCYTES
317
Extracellular compartment Opsonin-independent receptor
Phagocyte membrane extension
CA a
Cell membrane
CA Intracellular compartment
b
Opsonin-dependent receptor Opsonin
FIGURE 14.6 Schematic representation of the uptake of MRI contrast agents (CAs) by phagocytic cells: (a) endocytosis initiated by a receptor mediating an opsonin-independent recognition of the CA as a phagocytic target, and (b) endocytosis initiated by a receptor mediating a recognition of the CA as a phagocytic target through opsonins adsorbed to its surface.
(SR-A) types I and II bind anionic proteins, lipids, polynucleotides, and polysaccharides including dextran sulfate or fucose sulfate (fucoidan) and bacterial lipopolysaccharides [59]. The SR-A types mediate an opsonin-independent recognition of microorganisms or apoptotic cells before their phagocytosis (Fig. 14.6a) [60]. Competition experiments have suggested that the endocytosis of SPIO by cultured macrophages was mediated by the SR-A types. Indeed, ligands of the SR-A (fucose sulfate and polyinosinic acid) were effective inhibitors of the SPIO capture by cultured mouse peritoneal macrophages [59]. Phagocytosis concerns the uptake by macrophages of large material, such as bacteria (∼1 m) or senescent eukaryotic cells (∼10 m), which are much larger than the hydrodynamic diameter of iron oxide nanoparticles. The larger-sized dextran-coated SPIO (80–150 nm) can nevertheless initiate their macrophage endocytosis by interacting with the SR-A [59]. The macrophage uptake of dextran-coated USPIO (20–40 nm) occurs primarily through a pinocytosis mechanism and leads to a weaker internalization of nanoparticles than with dextran-coated SPIO [59, 61]. However, it has been demonstrated that citrate-coated very small superparamagnetic iron oxide nanoparticles with a hydrodynamic diameter of 8 nm (VSOP-C125) were efficiently taken up by macrophages through a mechanism involving actin, which is the cytoskeleton component determinant for phagocytosis. This result suggests that the pronounced negative surface charge of the nanoparticles could help their macrophage endocytosis by allowing an interaction with a receptor initiating phagocytosis. Another important point in this context is that the clustering of the nanoparticles, leading to the formation of larger-sized particulate aggregates, can also result in different uptake stimuli for macrophages (i.e., phagocytosis rather than pinocytosis) [61]. Particles can also bind to nonspecific scavenger receptors after adsorption to their surface of proteins from the heat-inactivated fetal bovine serum supplement added in culture medium [62]. In fresh plasma, MION have been covered by vitronectin, by fibronectin, and by fragments of the complement fraction C3 (C3b and its fragment C3dg) bound to IgG heavy or light chain [18]. This opsonization induced the recognition of MION by specific receptors on the macrophages (Fig. 14.6b) [63]. Indeed, IgG heavy chain-complexed dimeric forms of C3b and C3dg are ligands for the complement receptors CR1 and CR2, respectively [64]. The vitronectin receptor and the 2 integrins, such as the complement receptors CR3 and CR4, usually bind the opsonin iC3b (fragment of C3b). They can also mediate the nonopsonic recognition of phagocytic targets [60, 63]. Iron oxide nanoparticles have been targeted to a 2 integrin for the in vitro magnetic labeling of dendritic cells before their reimplantation and in vivo MRI monitoring. By conjugation to an antibody, iron oxide nanoparticles
P1: OTA/XYZ P2: ABC c14 JWBS037-Chen February 1, 2011
318
0:6
Printer Name: Yet to Come
CELLULAR MAGNETIC LABELING WITH IRON OXIDE NANOPARTICLES
have been targeted to CD11c, the alpha subunit of the 2 integrin CR4 (CD11c/CD18), and allowed to obtain an efficient intracellular labeling in vitro through receptor-mediated endocytosis [64]. Gd-based contrast agents (Gd-DTPA) have been linked to long fatty acid chains in order to form 0.5–1-m insoluble particles. These foreign particles were captured by dendritic cells in vitro through phagocytosis and intracellular lipases cleaved the long aliphatic side chains, restoring the solubility of Gd-DTPA and consequently its contrast agent activity inside the cells [65].
14.5 TOXICITY From the toxicity point of view, iron oxides are generally well tolerated by cells. No significant short-term toxicity of iron oxide nanoparticles has been reported [3, 26, 33, 37, 42]. Several studies have shown that the endosomal incorporation of iron oxide nanoparticles R complexed did not affect cell viability or proliferation, but it was found that Feridex to poly-l-lysine had an effect on the human mesenchymal stem cells’ differentiation into chondrocytes [66]. Furthermore, the same complex significantly affected the viability of marrow stromal cells as compared to unlabeled control cells when incubations were done for long periods (24 or 72 h) with relatively high concentrations (100 and 250 g Fe/mL). R –poly-l-lysine complexes at 25 g Fe/mL did On the contrary, a labeling with Feridex not induce any significant mortality of mesenchymal stem cells, even after a 72-h incubation period, and long-term toxicity studies performed after a 16–18-h incubation period of mesenchymal stem cells with the same amount of complex did not show any significant cell death as compared to unlabeled control cells until 15 days postlabeling [67]. The metabolic fate of internalized iron has not been well determined in current studies. It is assumed that iron atoms are integrated in the normal iron turnover inside the cell [3]. Nonetheless, incubation with iron oxide at high concentrations (1 mgFe/mL for 24 h) has been found to cause the generation of free radicals, a decrease in proliferation, and cell death [43]. VSOP were massively taken up by macrophages and were reported to induce a transient oxidative stress R in these cells [68]. Human neural precursor cells that were labeled for 72 h with Sinerem (800 gFe/mL) did not show any loss of short-term viability. However, a decrease in their subsequent survival and a reduction in their neurosphere-forming capacity were observed when cells were dissociated and plated in culture under standard growth conditions after incubation with the USPIO (more than 80% of cell death at the first subculturing passage postlabeling). This phenomenon was avoided with lower incubation concentrations (e.g., 400 gFe/mL) and shorter incubation times (e.g., 48 h). Poly-l-lysine has been found to R R and Endorem without impairing cell improve the labeling efficiency of both Sinerem survival, even surprisingly increasing the short-term survival of the human neural precurR , and 200 and 400 sor cells at high iron incubation doses (100 gFe/mL for Endorem R R gFe/mL for Sinerem ) [69]. Higher Sinerem incubation concentrations (up to 22 mg Fe/mL) and a shorter incubation time (12 h) have been used to perform a cationic liposome R )-helped labeling of D3 embryonic stem cells [70]. Toxicity was mainly in(Metafectene R /mL). duced by the transfection agent (50% of immediate cell death at 25 L Metafectene R A small influence of Sinerem alone on cell proliferation was only observed at the highest iron incubation concentration. In the same study, comparisons between different cell lines (D3 embryonic stem cells, C17.2 neuronal progenitor cells, and dendritic cells) were also performed: just as the uptake of a given superparamagnetic cell tag is cell line specific, the tolerance to the transfection agent was also observed to be dependent on the cell type [70].
P1: OTA/XYZ P2: ABC c14 JWBS037-Chen February 1, 2011
0:6
Printer Name: Yet to Come
APPLICATIONS OF CELLULAR MAGNETIC LABELING WITH UNTARGETED NANOPARTICLES
319
14.6 APPLICATIONS OF CELLULAR MAGNETIC LABELING WITH UNMODIFIED OR UNSPECIFICALLY CELL-TARGETED MRI CONTRAST AGENTS The main application of cellular magnetic labeling is the in vivo tracking of cells by the noninvasive MRI technique, principally in the context of cell therapies. This method provides the possibility to visualize the migration of cells such as lymphocytes during the immune response. It also allows for the monitoring of the in vivo fate and behavior of cells (e.g., stem or progenitor cells) transfused in a diseased living organism, or directly transplanted in a damaged organ. Cellular imaging of phagocytes infiltrated in inflammatory areas has also been reported, as well as the magnetic labeling of tumors. In many studies, iron oxide nanoparticles have been chosen for cell labeling in vitro or in vivo and have been effective as magnetic tags to follow the cells in vivo by MRI [3]. 14.6.1 MRI Monitoring of Cells Transplanted or Transfused In Vivo After In Vitro Magnetic Labeling The in vivo distribution of magnetically labeled cells from the hematopoietic lineage has been studied by MRI. Hybridomas (a fusion between B lymphocytes and myeloma cells) internalized anionic nanoparticles in vitro and were intraperitoneally injected in the mouse. MRI suggested a homing of the labeled cells in the spleen 24 h postinjection. These cells were also found to proliferate in the spleen [71]. Magnetically labeled human hematopoietic progenitor cells were intravenously injected in athymic mice. MRI monitoring showed signal attenuations in spleen, liver, and bone marrow. Histology confirmed the homing of the labeled cells in these recipient organs [72]. Tat peptide-derivatized iron oxide nanoparticles have also been applied to image cell trafficking in vivo after intravenous injection of CD34+ cells into immunodeficient mice. These cells, prelabeled with superparamagnetic–Tat peptide conjugates, were detected in the bone marrow by MRI [50]. The MRI tracking of prelabeled lymphocytes (T cells) has also been performed in the context of the mouse experimental autoimmune encephalomyelitis (EAE) model, allowing researchers to collect information about the pathogenesis of this dysmyelinating disease [73]. Because of its high relevance for immune cell-based anticancer therapies, the in vivo trafficking of tumor-targeted lymphocytes has been studied in several models. Iron oxide nanoparticles conjugated to a large number of Tat peptides were used to label CD8+ cytotoxic T lymphocytes stimulated with the ovalbumin-derived immunogenic peptide. Ovalbumin-transfected B16 melanoma cells were subcutaneously implanted in mice and intraperitoneally injected with the labeled cells. The tumoral recruitment of these cells was visualized by MRI and confirmed by histological analysis [74]. In a similar type of animal model, ovalbumin-specific splenocytes, labeled with iron oxide nanoparticles, were injected into mice with ovalbumin-expressing tumors. MRI results suggested that cells went first to the spleen (homing, 24 h postinjection) before being recruited by the tumor (48–72 h postinjection) [75]. The in vivo MRI monitoring of the accumulation of immune cells in tumors was also reported in a study performed with human natural killer cells directed against HER2/neu receptors, expressed by mammary tumors. Cells were magnetically labeled with iron oxide nanoparticles and intravenously injected into mice bearing HER2/neu-positive mammary tumors. The tumoral signal decay in MRI, confirmed by histology results, demonstrated the presence of the labeled natural killer cells in the tumors [76]. These different reports have shown the usefulness of MRI as
P1: OTA/XYZ P2: ABC c14 JWBS037-Chen February 1, 2011
320
0:6
Printer Name: Yet to Come
CELLULAR MAGNETIC LABELING WITH IRON OXIDE NANOPARTICLES
a noninvasive method for the monitoring of cell-based immunotherapies. Furthermore, a human approach to therapeutic cell tracking appeared as clinically safe. Dendritic cells, which can play the role of an antitumor cellular vaccine since they are antigen-presenting cells, were SPIO-labeled and detected by MRI after injection in the lymph nodes of patients suffering from melanoma. MRI showed how accurate the nodal delivery was of the iron oxide-loaded autologous dendritic cells [77]. Many attempts of stem cell transplantation have been performed with the aim of repairing central nervous system (brain or spinal cord) disorders, and promoting anatomical and functional recovery. The magnetic prelabeling of these cells allows visualization of their presence and their migration by MRI. In order to provide a strategy of cell therapy to reverse dysmyelinating diseases, stem cell-derived oligodendrocyte progenitors (CG-4) were magnetically labeled and were observed to migrate after transplantation in the spinal cord of 7-day-old myelin deficient rats. Furthermore, histological analysis identified iron-loaded cells as oligodendrocytes, and detected newly formed myelin in areas corresponding to the localization of labeled cells on MR images [6]. Oligodendroglial progenitors derived from neural stem cells were also magnetically labeled in vitro and transplanted intraventricularly in the brain of neonatal rats for a successful in vivo MRI tracking [37]. The possibility to repair spinal cord injuries has been investigated also by studying the migration and behavior of transplanted labeled stem cells. Marrow stromal cells (multipotent progenitor R ) and intravenously cells extracted from bone marrow) were labeled with SPIO (Endorem injected into rats 1 week after induction of a spinal cord injury. It was observed that the paraplegic rats showed good scores of recovery from the locomotor and the hind limb sensitivity points of view during the 5 weeks following marrow stromal cell injection. Hypointense areas observed in ex vivo MRI suggested a migration of the SPIO-labeled cells to the lesion site, and histology confirmed the presence of iron-loaded cells in the lesion. Furthermore, a reduction of the lesion size was noticed in marrow stromal cell-injected animals, suggesting tissue regeneration due to these cells. As the regeneration of axons can be affected by the formation of scar tissue in the injured region, biocompatible macroporous hydrogels (pore size: 10–50 m) have been used to anatomically replace a removed part of the spinal cord and prevent scarring. A block of hydrogel seeded with SPIO-labeled marrow stromal cells was inserted in the place of the right half of a spinal cord segment and was visible by MRI. Hydrogels create an environment where intrinsic growth factors can diffuse and promote marrow stromal cell migration. Indeed, 6 weeks after implantation, the hydrogel was adherent to the spinal segments and neurofilaments were stained inside by immunohistochemistry, demonstrating the growth of axons into the hydrogel [28]. Another central nervous system injury requiring a replacement of lost cells is cerebral stroke. In the context of brain lesions, the fate (migration and differentiation) of in vivo implanted stem cells, after magnetic labeling in culture, has been monitored by the noninvasive MRI technique. The behavior of marrow stromal cells, embryonic stem cells (ESCs), and CD34+ hematopoietic progenitor cells in the injured brain has been studied using a model of cortical photochemical lesion. On T2-weighted MR images, SPIO-labeled marrow stromal cells were visible as a hypointense signal at the level of their implantation site, in the hemisphere contralateral to the lesion. A signal darkening was noticed within 1 and 3 weeks in the damaged cortical tissue, around the necrotic part of the lesion, suggesting a migration of labeled cells. Immunohistochemistry revealed that a small proportion (less than 3%) of the large population of iron-loaded cells present in the lesion expressed the neuronal marker NeuN. Intravenously injected SPIO-labeled marrow stromal cells infiltrated the damaged hemisphere at the border of the lesion. The presence of labeled cells was visible by MRI
P1: OTA/XYZ P2: ABC c14 JWBS037-Chen February 1, 2011
0:6
Printer Name: Yet to Come
APPLICATIONS OF CELLULAR MAGNETIC LABELING WITH UNTARGETED NANOPARTICLES
321
and confirmed by histology [26, 28]. The same type of experiment was performed with ESCs. However, to avoid the risk of forming embryonic tumors once grafted in vivo, ESCs were differentiated into neural stem cells before implantation. These ESCs were SPIOlabeled and intravenously or intracerebrally (contralateral to the lesion) injected into rats 1 week after the induction of a photochemical cortical lesion. In both cases, MRI revealed a hypointense signal in the lesion 1–2 weeks after injection or implantation. As ESCs were modified to express a fluorescent marker (green fluorescent protein, GFP), fluorescence microscopy performed on histological slices confirmed the presence of labeled cells in this injured site. MRI and histology results also suggested that intracerebrally implanted cells were migrating across the corpus callosum to reach the lesion. Iron-loaded ESCs found in the injured region after intravenous injection were observed to achieve a final differentiation and were identified as mostly astrocytes (70%), but also oligodendrocytes (1%) and neurons (5%) [27, 28]. Under the same experimental conditions, magnetically labeled human CD34+ cells were observed to reach the lesioned target 24 h after contralateral R ) implantation [28]. In other studies, ESCs magnetically labeled with USPIO (Sinerem combined to lipofection agent have been implanted in the brain of rats suffering from a focal cerebral ischemia. Two weeks after this induced stroke, labeled cells were implanted in the contralateral hemisphere and observed by MRI. A line of hypointense signal was visualized along the corpus callosum 6 days after grafting of the cells, suggesting a migration of the cells toward the lesion. Indeed, 2 days later, the ischemic hemisphere became darker, suggesting that labeled cells had moved from the injection site and accumulated in the damaged area. Once again, other techniques, such as immunocytochemistry, confirmed the presence of implanted cells in the areas of MR signal darkening. Furthermore, changes in the labeled cells’ shape suggested that their differentiation occurred during their migration. Neuron-like cells were observed in the ischemic area [42]. This phenomenon of transplanted stem cells attracted by an ischemic site in the brain has also been observed with neural stem cells labeled using a bimodal contrast agent, the GRID, which allowed for the detection of cells by MRI and fluorescent microscopy [78]. The migration of iron oxide-labeled marrow stromal cells and neural progenitor cells toward another type of brain injury, a gliosarcoma, also has been observed by MRI after the injection of cells in a tail vein or in a subarachnoid space: the cisterna magna [79]. Stem cell transplantation can also be important as a therapy for neurodegenerative diseases. GRID-labeled neural stem cells were intracerebrally transplanted in a rat model of Huntington disease and observed to infiltrate the damaged hemisphere, migrating from their contralateral transplantation site [3]. Another possibility to track magnetically labeled stem cells in the context of ischemic injury has been tested in a model of swine myocardial infarction. Mesenchymal stem cells labeled with fluorescent micrometer-sized particles of iron oxide (MPIO) were percutaneously injected into the intact or infarcted myocardium. Signal darkening was detected in the beating normal or infarcted heart using cardiac MRI (fast gradient echo (FGE) or steady-state free precession (SSFP)) until 3 weeks after the cells’ implantation (confirmed ex vivo by MRI and histology). Fluorescence microscopy performed on histological slices showed that implanted mesenchymal stem cells were elongated and aligned with the host myocardium, suggesting their in vivo differentiation [80]. Using a similar model, SPIO-labeled mesenchymal stem cells were injected in the infarct and could be visualized by in vivo MRI until 1 week postimplantation. However, MRI also suggested that the size of the lesion increased and that the negative contrast decreased. Possible explanations for the loss of contrast were division of mesenchymal stem cells inducing a SPIO dilution, a migration of labeled cells, or eventually both [81]. In the same myocardial infarction
P1: OTA/XYZ P2: ABC c14 JWBS037-Chen February 1, 2011
322
0:6
Printer Name: Yet to Come
CELLULAR MAGNETIC LABELING WITH IRON OXIDE NANOPARTICLES
model, SPIO-labeled myogenic precursor cells were injected in the infarcted area of pig’s heart. The necrotic region was identified in the left ventricle through endoventricular electromechanical mapping, as the part of the myocardium with the most reduced electrical and mechanical activity. Implantation sites of labeled myogenic precursor cells were detected in vivo by cardiac MRI and ex vivo by iron staining on histological slices [82]. Other experiments have been performed in rats or mice. Rat smooth muscle cells labeled with dimercaptosuccinic acid-coated particles (anionic particles) were implanted in rat hearts with ischemic injury. MR imaging of implantation sites was performed ex vivo on the heart excised from 2 to 48 h after grafting of cells. An anatomical correlation between the hypointense signal in myocardium (in T2- or T2* -weighted MRI) and the labeled cells detected in histology or immunohistochemistry was observed [83]. SPIO-labeled ESCs have also been visualized in vivo by T2-weighted MRI of the left ventricle of the intact rat heart [84]. Magnetically labeled cardiac progenitor cells have been transplanted into the healthy or injured myocardium of mice. High-resolution MRI allowed the precise determination of the location of labeled cells in the mouse’s heart. It was observed that a cardiac ischemic lesion appeared as a hyposignal area, similarly sized on both T2* -weighted and proton density-weighted MR images. However, after transplantation of iron oxide-loaded cardiac progenitor cells to the infarcted mouse myocardium, the signal attenuation area noticed on proton density MR images reflected only the lesion size, while it was larger on T2* -weighted MR images because of the higher detection sensitivity of magnetically labeled cells allowed by the latter parameter weighting [85]. In order to target other potentially damaged organs with stem cells, the fate of SPIOlabeled mesenchymal stem cells injected into the portal veins of rats treated with carbon tetrachloride to induce centrolobular liver necrosis, as well as into the renal arteries of healthy rats, has been studied by MRI and correlated to the histological observation of the iron-loaded cells’ distribution in the liver or kidney [86]. However, further experiments have shown that SPIO-labeled mesenchymal stem cells systemically injected (i.e., intravenously and not through the feeding artery) in rats with kidney mesangiolysis homed to the injured kidney but were below the in vivo MRI detection threshold. Only ex vivo images of the kidneys showed areas of lower intensity that were correlated to iron staining by histology [87]. Type 1 diabetes mellitus has been treated experimentally with another kind of cell therapy. Instead of stem cells, pancreatic islets containing -cells (producing insulin) were transplanted into the liver of diabetic rats through injection in the portal vein. The pancreatic R ) and their hepatic location was islets were magnetically labeled with SPIO (Resovist visualized by MRI. The labeled pancreatic islets, situated in the liver sinusoids, were observed as hypointense areas homogeneously distributed in the liver. Furthermore, a normal glycemia was restored in previously diabetic rats 1 week after transplantation of the islets [88]. 14.6.2 MRI Monitoring of Cells After Magnetic Labeling In Vivo Visualizing cells in vivo by MRI without any in vitro magnetic prelabeling requires a sufficient cellular uptake of the MRI contrast agent provided by the blood circulation after intravenous administration.
Inflammation Hepatic MRI using iron oxides is based on the capture of nanoparticles by the resident macrophages of the liver: the Kupffer cells. As inflammatory diseases are associated with monocyte/macrophage activity, this macrophage targeting approach has been
P1: OTA/XYZ P2: ABC c14 JWBS037-Chen February 1, 2011
0:6
Printer Name: Yet to Come
APPLICATIONS OF CELLULAR MAGNETIC LABELING WITH UNTARGETED NANOPARTICLES
323
further developed to visualize inflammatory lesions in the body. Thus following intravenous injection of iron oxide nanoparticles, inflammatory foci can be visualized by MRI. This has been achieved in experimental models of brain inflammatory diseases. After adminisR ) at a concentration of 300 mol Fe/kg, macrophage activity tration of USPIO (Sinerem was imaged in rats with experimental autoimmune encephalomyelitis (EAE), thanks to the capture of the nanoparticles by peripheral macrophages (blood monocytes or lymph node macrophages) subsequently invading the inflamed brain. However, the injected iron dose was about seven times higher than the appropriate dose for human use (45 mol Fe/kg (2.6 mg/kg)). This was explained by the fact that a high iron dose is necessary to saturate the uptake capability of liver, spleen, or bone marrow macrophages. Furthermore, R in rats is only 2 h, while it is 20–24 h at the normal dose, the blood half-life of Sinerem in humans [89]. In rat models of cerebral stroke, infiltration of macrophages in ischemic lesions has also been visualized by MRI following USPIO uptake. The localization of these hypointense areas on T2-weighted MR images was correlated to the histological detection of iron and to the immunohistological labeling of macrophages postmortem [90, 91]. In this context of brain damages associated with inflammatory events, macrophage activity has been studied by MRI after USPIO infusion also in human patients suffering from stroke [92]. In vivo cellular MR imaging of USPIO-labeled macrophages has found successful applications in other animal models of disease, such as in kidney injuries (nephropathies) induced by a nephrotoxic drug (puromycin aminonucleoside) or in organ graft rejection (e.g., after renal allograft transplantation without any immunosuppressive drug) [3, 93, 94]. The accumulation of long-circulating USPIO in monocytes and macrophages recruited to atherosclerotic plaques has been demonstrated also in hyperlipidemic rabbits [95, 96]. Furthermore, USPIO-enhanced MRI has also been used for in vivo detection of macrophages in human plaques. The strongest signal decreases, due to iron-loaded macrophage accumulation, were observed in ruptured and rupture-prone human atherosclerotic lesions rather than in stable lesions, allowing for a differentiation between low-risk and high-risk plaques on MR images [3, 97]. MPIO also have been intravenously injected into rats with the aim of labeling macrophage infiltrates in the rejection site of cardiac graft, allowing for the noninvasive monitoring of organ transplant rejection by MRI [98].
Tumors The in vivo tumoral accumulation of iron oxide nanoparticles has been investigated given the great interest of detecting tumors by MRI. In a rat model of intracerebrally implanted glioma, MION were intravenously injected (10 mg Fe/kg) and induced a signal increase of the tumor on T1-weighted MR images and a signal decrease on T2-weighted MR images, particularly at the tumoral periphery, 24 h after injection. Microscopic and histological examinations showed that the nanoparticles were intracellularly and extracellularly distributed in the tumor, with a higher level of accumulation in the periphery. Indeed, the highest microvascular density of a tumor is found in this area, which is consequently a region of preferential extravasation. Tumor-associated macrophages and vascular endothelial cells were observed to take up the nanoparticles, but the majority of nanoparticle-containing cells were tumor cells. Furthermore, the cellular uptake of the nanoparticles seems to be directly related to the tumor growth—in other words, directly related to the cellular cycle [19]. The same type of MION distribution in the tumor cells adjacent to the hyperpermeable tumor–brain interface has been reported in another study [99]. USPIO were injected intravenously into rats bearing intracerebral tumors (LX-1 small-cell lung carcinoma) and did not show the same distribution pattern as MION. Within 24 h postinjection, T1-weighted MR images showed an enhancement of the signal at the periphery of the tumor, surrounding a
P1: OTA/XYZ P2: ABC c14 JWBS037-Chen February 1, 2011
324
0:6
Printer Name: Yet to Come
CELLULAR MAGNETIC LABELING WITH IRON OXIDE NANOPARTICLES
brighter region corresponding to a necrotic core. Histological analysis allowed for the detection of iron in and around necrotic areas, and at the tumor margin, but only in macrophages, reactive astrocytes, or gliosis (scar made of a dense fibrous network of neuroglia). Regarding LX-1 small-cell lung carcinoma, other tumor types exhibited a weaker T1-enhancement pattern with USPIO and consequently showed a sparser iron staining in histology, but confirmed the distribution of the nanoparticles in cells infiltrated at the brain–tumor interface R ), rapidly and in areas of necrosis. It was noticed that bigger nanoparticles (SPIO, Feridex captured by spleen or liver macrophages, were ineffective for imaging brain tumors [100]. MION and USPIO are long-circulating iron oxide nanoparticles, able to passively target tumors by extravasation through the leaky tumoral capillaries. This interstitial accumulation of USPIO has also been suggested by a tumoral signal enhancement in T1-weighted MRI, 12 h after intravenous injection of the nanoparticles in mice bearing nonnecrotic subcutaneously implanted tumors (human prostatic adenocarcinoma). However, a signal decrease of the tumor was generated in T2-weighted MRI by dimercaptosuccinic acidcoated anionic iron oxide nanoparticles, 24 h postinjection. It suggested an intracellular tumoral distribution of these nanoparticles [101]. Magnetic labeling of tumors has also been achieved by magnetically targeting liposomes loaded with iron oxide nanoparticles to human prostatic adenocarcinoma implanted in mice. A preferential accumulation of these polyethylene glycol-coated long-circulating magnetoliposomes (200 nm in diameter) was induced by placing a magnet on the skin above the tumor during the circulation of the intravenously injected magnetoliposomes. MRI showed a strong signal darkening of the tumor exposed for 24 h to the magnet, compared to nonexposed tumors. It was observed that, due to a higher global magnetic moment, the magnetoliposomes could be magnetically driven to a tumor more efficiently than nonencapsulated USPIO. The mechanism of tumoral accumulation of magnetoliposomes is diffusion to the interstitium, through the leaky vasculature. Histological analysis revealed the presence of magnetoliposomes mainly in the most highly vascularized zones: in capillaries, interstitium, and cells (macrophages, fibroblasts, endothelial cells), but not in tumor cells. Iron was also found in liver and spleen [102]. Paramagnetic contrast agents have also been used as tumor-specific contrast agents in vivo. Gd-based metalloporphyrins, such as Gadophrin-2 (Bayer Schering Pharma AG), can selectively accumulate in tumors because of their high affinity to necrotic tissues. It has been suggested that their ability to bind plasma albumin is responsible for their slow extravasation and subsequent accumulation into the necrotic tumor interstitium [103, 104].
14.7 CONCLUSION AND PERSPECTIVES Iron oxide particles, made of polymer- or monomer-coated iron oxide crystals, are well known in clinical applications as intravenously injected negative MRI contrast agents. In vivo, they are sooner (for larger ones) or later (for smaller ones) taken up by cells of the mononuclear phagocyte system. Iron oxide particles have been presented as an interesting tool for cellular magnetic labeling because of their greater effect on the MR signal as compared to paramagnetic ions, allowing their MRI detection at relatively low concentrations (nanomolar) in tissues and thus opening the field of cellular MRI research. The ability of iron oxide particles to be taken up by phagocytic cells after intravenous injection has been further exploited to image macrophage-invaded areas in the context of inflammatory pathologies or tumors. MRI monitoring of nonphagocytic cells is often possible by a magnetic labeling of these cells performed in vitro, prior to their in vivo implantation. In
P1: OTA/XYZ P2: ABC c14 JWBS037-Chen February 1, 2011
0:6
Printer Name: Yet to Come
REFERENCES
325
other terms, iron oxide particles must be internalized by cultured cells in an optimal way. In many cases, particles are used unmodified (polymer- or monomer-coated iron oxide crystals) or they can be unspecifically cell-targeted (mixed with DNA transfection reagents, conjugated to a cell-penetrating peptide), triggering their endocytosis via nonspecific pathways. Iron oxide nanoparticles can also be linked to a ligand specific for a certain receptor at the cell surface, allowing their receptor-mediated uptake. In brief, the magnetic labeling of cells with MRI contrast agents, in particular, with iron oxide nanoparticles, is receiving increasing attention since in vivo applications exist in different pathological contexts. Since R R R (Endorem ), Resovist ) were they are taken up by phagocytic cells, SPIO (Feridex originally developed and approved for clinical MRI to increase the lesion-to-liver contrast in malignancies where Kupffer cells are absent. It is worth mentioning that these SPIO were recently taken off the market due to lack of sales in that context. Nevertheless, they were used for cellular magnetic labeling studies in patients, suggesting that a great deal of future research and development will be driven in the field of cellular magnetic labeling and cellular MRI [105]. REFERENCES 1. Artemov, D.; Mori, N.; Okollie, B.; Bhujwalla, Z. M. MR molecular imaging of the Her-2/neu receptor in breast cancer cells using targeted iron oxide nanoparticles, Magn. Reson. Med. 2003, 49, 403–408. 2. Bulte, J. W. M.; Modo, M. M. J. Nanoparticles in Biomedical Imaging, Emerging Technologies and Applications. Springer: New York, 2007. 3. Modo, M.; Hoehn, M.; Bulte, J. W. Cellular MR imaging, Mol. Imaging 2005, 4, 143–164. 4. Hopkins, C. R.; Trowbridge, I. S. Internalization and processing of transferrin and the transferrin receptor in human carcinoma A431 cells. J. Cell Biol. 1983, 97, 508–521. 5. Kresse, M.; Wagner, S.; Pfefferer, D.; Lawaczeck, R.; Elste, V.; Semmler, W. Targeting of ultrasmall superparamagnetic iron oxide (USPIO) particles to tumor cells in vivo by using transferrin receptor pathways. Magn. Reson. Med. 1998, 40, 236–242. 6. Bulte, J. W.; Zhang, S.; van Gelderen, P.; Herynek, V.; Jordan, E. K.; Duncan, I. D.; Frank, J. A. Neurotransplantation of magnetically labeled oligodendrocyte progenitors: magnetic resonance tracking of cell migration and myelination. Proc. Natl. Acad. Sci. U.S.A. 1999, 96, 15256–15261. 7. Weissleder, R.; Moore, A.; Mahmood, U.; Bhorade, R.; Benveniste, H.; Chiocca, E. A.; Basilion, J. P. In vivo magnetic resonance imaging of transgene expression. Nat. Med. 2000, 6, 351–355. 8. Deans, A. E.; Wadghiri, Y. Z.; Bernas, L. M.; Yu, X.; Rutt, B. K.; Turnbull, D. H. Cellular MRI contrast via coexpression of transferrin receptor and ferritin. Magn. Reson. Med. 2006, 56, 51–59. 9. Harrison, P. M.; Arosio, P. The ferritins: molecular properties, iron storage function and cellular regulation. Biochim. Biophys. Acta 1996, 1275, 161–203. 10. Treffry, A.; Zhao, Z.; Quail, M. A.; Guest, J. R.; Harrison, P. M. Dinuclear center of ferritin: studies of iron binding and oxidation show differences in the two iron sites. Biochemistry. 1997, 36, 432–441. 11. Cohen, B.; Dafni, H.; Meir, G.; Harmelin, A.; Neeman, M. Ferritin as an endogenous MRI reporter for noninvasive imaging of gene expression in C6 glioma tumors. Neoplasia. 2005, 7, 109–117. 12. Wang, Z. J.; Boddington, S.; Wendland, M.; Meier, R.; Corot, C.; Daldrup-Link, H. MR imaging of ovarian tumors using folate-receptor-targeted contrast agents. Pediatr. Radiol. 2008, 38, 529–537.
P1: OTA/XYZ P2: ABC c14 JWBS037-Chen February 1, 2011
326
0:6
Printer Name: Yet to Come
CELLULAR MAGNETIC LABELING WITH IRON OXIDE NANOPARTICLES
13. Sun, C.; Veiseh, O.; Gunn, J.; Fang, C.; Hansen, S.; Lee, D.; Sze, R.; Ellenbogen, R. G.; Olson, J.; Zhang, M. In vivo MRI detection of gliomas by chlorotoxin-conjugated superparamagnetic nanoprobes. Small 2008, 4, 372–379. 14. Serda, R. E.; Adolphi, N. L.; Bisoffi, M.; Sillerud, L. O. Targeting and cellular trafficking of magnetic nanoparticles for prostate cancer imaging. Mol. Imaging 2007, 6, 277–288. 15. Yang, L.; Mao, H.; Wang, Y. A.; Cao, Z.; Peng, X.; Wang, X.; Duan, H.; Ni, C.; Yuan, Q.; Adams, G.; Smith, M. Q.; Wood, W. C.; Gao, X.; Nie, S. Single chain epidermal growth factor receptor antibody conjugated nanoparticles for in vivo tumor targeting and imaging. Small 2009, 5, 235–243. 16. Leuschner, C.; Kumar, C. S.; Hansel, W.; Soboyejo, W.; Zhou, J.; Hormes, J. LHRH-conjugated magnetic iron oxide nanoparticles for detection of breast cancer metastases. Breast Cancer Res. Treat. 2006, 99, 163–176. 17. Anderson, S. A.; Glod, J.; Arbab, A. S.; Noel, M.; Ashari, P.; Fine, H. A.; Frank, J. A. Noninvasive MR imaging of magnetically labeled stem cells to directly identify neovasculature in a glioma model. Blood 2005, 105, 420–425. 18. Moore, A.; Weissleder, R.; Bogdanov, A. Uptake of dextran-coated monocrystalline iron oxides in tumor cells and macrophages. J. Magn. Reson. Imaging 1997, 7, 1140–1145. 19. Moore, A.; Marecos, E.; Bogdanov, A.; Weissleder, R. Tumoral distribution of long-circulating dextran-coated iron oxide nanoparticles in a rodent model. Radiology 2000, 214, 568–574. 20. Modo, M.; Cash, D.; Mellodew, K.; Williams, S. C.; Fraser, S. E.; Meade, T. J.; Price, J.; Hodges, H. Tracking transplanted stem cell migration using bifunctional, contrast agent-enhanced, magnetic resonance imaging. Neuroimage 2002, 17, 803–811. 21. Crich, S. G.; Biancone, L.; Cantaluppi, V.; Duo, D.; Esposito, G.; Russo, S.; Camussi, G.; Aim´e, S. Improved route for the visualization of stem cells labeled with a Gd-/Eu-chelate as dual (MRI and fluorescence) agent. Magn. Reson. Med. Engl. 2004, 51, 938–944. 22. Aim´e, S.; Carrera, C.; Delli Castelli, D.; Crich, S. G.; Terreno, E. Tunable imaging of cells labeled with MRI-PARACEST agents. Angew. Chem. Int. Ed. 2005, 44, 1813–1815. 23. Rudelius, M.; Daldrup-Link, H. E.; Heinzmann, U.; Piontek, G.; Settles, M.; Link, T. M.; Schlegel, J. Highly efficient paramagnetic labelling of embryonic and neuronal stem cells. Eur. J. Nucl. Med. Mol. Imaging 2003, 30, 1038–1044. 24. Schoepf, U.; Mercos, E. M.; Melder, R. J.; Jain, R. K.; Weissleder, R. Intracellular magnetic labeling of lymphocytes for in vivo trafficking studies. Biotechniques 1998, 24, 642–646, 648–651. 25. Yeh, T. C.; Zhang, W.; Ildstad, S. T.; Ho, C. Intracellular labeling of T-cells with superparamagnetic contrast agents. Magn. Reson. Med. 1993, 30, 617–625. 26. Jendelova, P.; Herynek, V.; DeCroos, J.; Glogarova, K.; Andersson, B.; Hajek, M.; Sykova, E. Imaging the fate of implanted bone marrow stromal cells labeled with superparamagnetic nanoparticles. Magn. Reson. Med. 2003, 50, 767–776. 27. Jendelova, P.; Herynek, V.; Urdzikova, L.; Glogarova, K.; Kroupova, J.; Andersson, B.; Bryja, V.; Burian, M.; Hajek, M.; Sykova, E. Magnetic resonance tracking of transplanted bone marrow and embryonic stem cells labeled by iron oxide nanoparticles in rat brain and spinal cord. J. Neurosci. Res. 2004, 76, 232–243. 28. Sykova, E.; Jendelova, P. Magnetic resonance tracking of implanted adult and embryonic stem cells in injured brain and spinal cord. Ann. N.Y. Acad. Sci.. 2005, 1049, 146–160. 29. Daldrup-Link, H. E.; Rudelius, M.; Oostendorp, R. A.; Settles, M.; Piontek, G.; Metz, S.; Rosenbrock, H.; Keller, U.; Heinzmann, U.; Rummeny, E. J.; Schlegel, J.; Link, T. M. Targeting of hematopoietic progenitor cells with MR contrast agents. Radiology 2003, 228, 760–767. 30. Mail¨ander, V.; Lorenz, M. R.; Holzapfel, V.; Musyanovych, A.; Fuchs, K.; Wiesneth, M.; Walther, P.; Landfester, K.; Schrezenmeier, H. Carboxylated superparamagnetic iron oxide
P1: OTA/XYZ P2: ABC c14 JWBS037-Chen February 1, 2011
0:6
Printer Name: Yet to Come
REFERENCES
31.
32. 33.
34.
35.
36. 37.
38. 39.
40.
41. 42.
43.
44.
45.
46.
327
particles label cells intracellularly without transfection agents. Mol. Imaging Biol. 2008, 10, 138–146. Kalish, H.; Arbab, A. S.; Miller, B. R.; Lewis, B. K.; Zywicke, H. A.; Bulte, J. W.; Bryant, L. H., Jr; Frank, J. A. Combination of transfection agents and magnetic resonance contrast agents for cellular imaging: relationship between relaxivities, electrostatic forces, and chemical composition. Magn. Reson. Med. 2003, 50, 275–282. Bulte, J. W. Magnetic nanoparticles as markers for cellular MR imaging. J. Magn. Magn. Mater. 2005, 289, 423–427. Frank, J. A.; Miller, B. R.; Arbab, A. S.; Zywicke, H. A.; Jordan, E. K.; Lewis, B. K.; Bryant, L. H.; Bulte, J. W. Clinically applicable labeling of mammalian and stem cells by combining superparamagnetic iron oxides and transfection agents. Radiology 2003, 228, 480–487. Arbab, A. S.; Yocum, G. T.; Wilson, L. B.; Parwana, A.; Jordan, E. K.; Kalish, H.; Frank, J. A. Comparison of transfection agents in forming complexes with ferumoxides, cell labeling efficiency, and cellular viability. Mol. Imaging 2004, 3, 24–32. Arbab, A. S.; Yocum, G. T.; Kalish, H.; Jordan, E. K.; Anderson, S. A.; Khakoo, A. Y.; Read, E. J.; Frank, J. A. Efficient magnetic cell labeling with protamine sulfate complexed to ferumoxides for cellular MRI. Blood 2004, 104, 1217–1223. R Transfection Reagent Handbook, For High Transfection Efficiencies in a Broad Superfect Range of Cell Lines. Qiagen Valencia, CA, 2000. Bulte, J. W.; Douglas, T.; Witwer, B.; Zhang, S. C.; Strable, E.; Lewis, B. K.; Zywicke, H.; Miller, B.; van Gelderen, P.; Moskowitz, B. M.; Duncan, I. D.; Frank, J. A. Magnetodendrimers allow endosomal magnetic labeling and in vivo tracking of stem cells. Nat. Biotechnol. 2001, 19, 1141–1147. Klajnert, B.; Bryszewska, M. Dendrimers: properties and applications. Acta. Biochim. Pol. 2001, 48, 199–208. Wilhelm, C.; Billotey, C.; Roger, J.; Pons, J. N.; Bacri, J. C.; Gazeau, F. Intracellular uptake of anionic superparamagnetic nanoparticles as a function of their surface coating. Biomaterials 2003, 24, 1001–1011. Boutry, S.; Brunin, S.; Mahieu, I.; Laurent, S.; Vander Elst, L.; Muller, R. N. Magnetic labeling of non-phagocytic adherent cells with iron oxide nanoparticles: a comprehensive study. Contrast Media Mol. Imaging 2008, 3, 223–232. Wrobel, I.; Collins, D. Fusion of cationic liposomes with mammalian cells occurs after endocytosis. Biochim. Biophys. Acta 1995, 1235, 296–304. Hoehn, M.; Kustermann, E.; Blunk, J.; Wiedermann, D.; Trapp, T.; Wecker, S.; Focking, M.; Arnold, H.; Hescheler, J.; Fleischmann, B. K.; Schwindt, W.; Buhrle, C. Monitoring of implanted stem cell migration in vivo: a highly resolved in vivo magnetic resonance imaging investigation of experimental stroke in rat. Proc. Natl. Acad. Sci. U.S.A. 2002, 99, 16267–16272. van den Bos, E. J.; Wagner, A.; Mahrholdt, H.; Thompson, R. B.; Morimoto, Y.; Sutton, B. S.; Judd, R. M.; Taylor, D. A. Improved efficacy of stem cell labeling for magnetic resonance imaging studies by the use of cationic liposomes. Cell Transplant. 2003, 12, 743–756. Vives, E.; Brodin, P.; Lebleu, B. A truncated HIV-1 Tat protein basic domain rapidly translocates through the plasma membrane and accumulates in the cell nucleus. J. Biol. Chem. 1997, 272, 16010–16017. Potocky, T. B.; Menon, A. K.; Gellman, S. H. Cytoplasmic and nuclear delivery of a TAT-derived peptide and a beta-peptide after endocytic uptake into HeLa cells. J. Biol. Chem. 2003, 278, 50188–50194. Vives, E.; Richard, J. P.; Rispal, G.; Lebleu, B. TAT peptide internalization: seeking the mechanism of entry. Curr. Protein. Pept. Sci. 2003, 4, 125–132.
P1: OTA/XYZ P2: ABC c14 JWBS037-Chen February 1, 2011
328
0:6
Printer Name: Yet to Come
CELLULAR MAGNETIC LABELING WITH IRON OXIDE NANOPARTICLES
47. Iozzo, R. V. Heparan sulfate proteoglycans: intricate molecules with intriguing functions. J. Clin. Invest. 2001,108, 165–167. 48. Bhorade, R.; Weissleder, R.; Nakakoshi, T.; Moore, A.; Tung, C. H. Macrocyclic chelators with paramagnetic cations are internalized into mammalian cells via a HIV-tat derived membrane translocation peptide. Bioconjug. Chem. 2000, 11, 301–305. 49. Prantner, A. M.; Sharma, V.; Garbow, J. R.; Piwnica-Worms, D. Synthesis and characterization of a Gd-DOTA-D-permeation peptide for magnetic resonance relaxation enhancement of intracellular targets. Mol. Imaging 2003, 2, 333–341. 50. Lewin, M.; Carlesso, N.; Tung, C. H.; Tang, X. W.; Cory, D.; Scadden, D. T.; Weissleder, R. Tat peptide-derivatized magnetic nanoparticles allow in vivo tracking and recovery of progenitor cells. Nat. Biotechnol. 2000, 18, 410–414. 51. Josephson, L.; Tung, C. H.; Moore, A.; Weissleder, R. High-efficiency intracellular magnetic labeling with novel superparamagnetic-Tat peptide conjugates. Bioconjug. Chem. 1999, 10, 186–191. 52. Tung, C. H.; Weissleder, R. Arginine containing peptides as delivery vectors. Adv. Drug. Deliv. Rev. 2003, 55, 281–294. 53. Walczak, P.; Kedziorek, D. A.; Gilad, A. A.; Lin, S.; Bulte, J. W. Instant MR labeling of stem cells using magnetoelectroporation. Magn. Reson. Med. 2005, 54, 769–774. 54. Hinds, K. A.; Hill, J. M.; Shapiro, E. M.; Laukkanen, M. O.; Silva, A. C.; Combs, C. A.; Varnet, T. R.; Balaban, R. S.; Koretsky, A. P.; Dunbar, C. E. Highly efficient endosomal labeling of progenitor and stem cells with large magnetic particles allows magnetic resonance imaging of single cells. Blood 2003, 102, 867–872. 55. Shapiro, E. M.; Skrtic, S.; Koretsky, A. P. Sizing it up: cellular MRI using micron-sized iron oxide particles. Magn. Reson. Med. 2005, 53, 329–338. 56. Weisskoff, R. M.; Zuo, C. S.; Boxerman, J. L.; Rosen, B. R. Microscopic susceptibility variation and transverse relaxation: theory and experiment. Magn. Reson. Med. 1994, 31, 601– 610. 57. Shapiro, E. M.; Skrtic, S.; Sharer, K.; Hill, J. M.; Dunbar, C. E.; Koretsky, A. P. MRI detection of single particles for cellular imaging. Proc. Natl. Acad. Sci. U.S.A. 2004, 101, 10901–10906. 58. Shapiro, E. M.; Sharer, K.; Skrtic, S.; Koretsky, A. P. In vivo detection of single cells by MRI. Magn. Reson. Med. 2006, 55, 242–249. 59. Raynal, I.; Prigent, P.; Peyramaure, S.; Najid, A.; Rebuzzi, C.; Corot, C. Macrophage endocytosis of superparamagnetic iron oxide nanoparticles: mechanisms and comparison of ferumoxides and ferumoxtran-10. Invest. Radiol. 2004, 39, 56–63. 60. Palecanda, A.; Paulauskis, J.; Al-Mutairi, E.; Imrich, A.; Qin, G.; Suzuki, H.; Kodama, T.; Tryggvason, K.; Koziel, H.; Kobzik, L. Role of the scavenger receptor MARCO in alveolar macrophage binding of unopsonized environmental particles. J. Exp. Med. 1999, 189, 1497–1506. 61. Fleige, G.; Seeberger, F.; Laux, D.; Kresse, M.; Taupitz, M.; Pilgrimm, H.; Zimmer, C. In vitro characterization of two different ultrasmall iron oxide particles for magnetic resonance cell tracking. Invest. Radiol. 2002, 37, 482–488. 62. Champion, J. A.; Mitragotri, S. Role of target geometry in phagocytosis. Proc. Natl. Acad. Sci. U.S.A. 2006, 103, 4930–4934. 63. Jelezarova, E.; Luginbuehl, A.; Lutz, H. U. C3b2-IgG complexes retain dimeric C3 fragments at all levels of inactivation. J. Biol. Chem. 2003, 278, 51806–51812. 64. Ahrens, E. T.; Feili-Hariri, M.; Xu, H.; Genove, G.; Morel, P. A. Receptor-mediated endocytosis of iron-oxide particles provides efficient labeling of dendritic cells for in vivo MR imaging. Magn. Reson. Med. 2003, 49, 1006–1013.
P1: OTA/XYZ P2: ABC c14 JWBS037-Chen February 1, 2011
0:6
Printer Name: Yet to Come
REFERENCES
329
65. Himmelreich, U.; Aim´e, S.; Hyeronymus, T.; Justicia, C.; Uggeri, F.; Zenke, M.; Hoehn, M. A responsive MRI contrast agent to monitor functional cell status. Neuroimage 2006, 32, 1142–1149. 66. Kostura, L.; Kraitchman, D. L.; Mackay, A. M.; Pittenger, M. F.; Bulte, J. W. Feridex labeling of mesenchymal stem cells inhibits chondrogenesis but not adipogenesis or osteogenesis. NMR Biomed. 2004, 17, 513–517. 67. Arbab, A. S.; Bashaw, L. A.; Miller, B. R.; Jordan, E. K.; Bulte, J. W.; Frank, J. A. Intracytoplasmic tagging of cells with ferumoxides and transfection agent for cellular magnetic resonance imaging after cell transplantation: methods and techniques. Transplantation 2003, 76, 1123–1130. 68. Stroh, A.; Zimmer, C.; Gutzeit, C.; Jackstadt, M.; Marschinke, F.; Jung, T.; Pilgrimm, H.; Grune, T. Iron oxide particles for molecular magnetic resonance imaging cause transient oxidative stress in rat macrophages. Free Radic. Biol. Med. 2004, 36, 976–984. 69. Neri, M.; Maderna, C.; Cavazzin, C.; Deidda-Vigoriti, V.; Politi, L. S.; Scotti, G.; Marzola, P.; Sbarbati, A.; Vescovi, A.L.; Gritti, A. Efficient in vitro labeling of human neural precursor cells with superparamagnetic iron oxide particles: relevance for in vivo cell tracking. Stem. Cells 2008, 26, 505–516. 70. K¨ustermann, E.; Himmelreich, U.; Kandal, K.; Geelen, T.; Ketkar, A.; Wiedermann, D.; Strecker, C.; Esser, J.; Arnhold, S.; Hoehn, M. Efficient stem cell labeling for MRI studies. Contrast Media Mol. Imaging 2008, 3, 27–37. 71. Smirnov, P.; Gazeau, F.; Lewin, M.; Bacri, J. C.; Siauve, N.; Vayssettes, C.; Cuenod, C. A.; Clement, O. In vivo cellular imaging of magnetically labeled hybridomas in the spleen with a 1.5-T clinical MRI system. Magn. Reson. Med. 2004, 52, 73–79. 72. Daldrup-Link, H. E.; Rudelius, M.; Piontek, G.; Metz, S.; Brauer, R.; Debus, G.; Corot, C.; Schlegel, J.; Link, T. M.; Peschel, C.; Rummeny, E. J.; Oostendorp, R. A. Migration of iron oxide-labeled human hematopoietic progenitor cells in a mouse model: in vivo monitoring with 1.5-T MR imaging equipment. Radiology 2005, 234, 197–205. 73. Anderson, S. A.; Shukaliak-Quandt, J.; Jordan, E. K.; Arbab, A. S.; Martin, R.; McFarland, H.; Frank, J. A. Magnetic resonance imaging of labeled T-cells in a mouse model of multiple sclerosis. Ann. Neurol. 2004, 55, 654–659. 74. Kircher, M. F.; Allport, J. R.; Graves, E. E.; Love, V.; Josephson, L.; Lichtman, A. H.; Weissleder, R. In vivo high resolution three-dimensional imaging of antigen-specific cytotoxic T-lymphocyte trafficking to tumors. Cancer Res. 2003, 63, 6838–6846. 75. Smirnov, P.; Lavergne, E.; Gazeau, F.; Lewin, M.; Boissonnas, A.; Doan, B. T.; Gillet, B.; Combardi`ere, C.; Cl´ement, O. In vivo cellular imaging of lymphocyte trafficking by MRI: a tumor model approach to cell-based anticancer therapy. Magn. Reson. Med. 2006, 56, 498–508. 76. Daldrup-Link, H. E.; Meier, R.; Rudelius, M.; Piontek, G.; Piert, M.; Metz, S.; Settles, M.; Uherek, C.; Wels, W.; Schlegel, J.; Rummeny, E. J. In vivo tracking of genetically engineered, anti-HER2/neu directed natural killer cells to HER2/neu positive mammary tumors with magnetic resonance imaging. Eur. Radiol. 2005, 15, 4–13. 77. de Vries, I. J.; Lesterhuis, W. J.; Barentsz, J. O.; Verdijk, P.; van Krieken, J. H.; Boerman, O. C.; Oyen, W. J.; Bonenkamp, J. J.; Boezeman, J. B.; Adema, G. J.; Bulte, J.W.; Scheenen, T. W.; Punt, C. J.; Heerschap, A.; Figdor, C. G. Magnetic resonance tracking of dendritic cells in melanoma patients for monitoring of cellular therapy. Nat. Biotechnol. 2005, 23, 1407–1413. 78. Modo, M.; Mellowdew, K.; Cash, D.; Fraser, S. E.; Meade, T. J.; Price, J.; Williams, S. C. Mapping transplanted stem cell migration after a stroke: a serial, in vivo magnetic resonance imaging study. Neuroimage 2004, 21, 311–317.
P1: OTA/XYZ P2: ABC c14 JWBS037-Chen February 1, 2011
330
0:6
Printer Name: Yet to Come
CELLULAR MAGNETIC LABELING WITH IRON OXIDE NANOPARTICLES
79. Zhang, Z.; Jiang, Q.; Jiang, F.; Ding, G.; Zhang, R.; Wang, L.; Zhang, L.; Robin, A. M.; Katakowski, M.; Chopp, M. In vivo magnetic resonance imaging tracks adult neural progenitor cell targeting of brain tumor. Neuroimage 2004, 23, 281–287. 80. Hill, J. M.; Dick, A. J.; Raman, V. K.; Thompson, R. B.; Yu, Z. X.; Hinds, K. A.; Pessanha, B. S.; Guttman, M. A.; Varney, T. R.; Martin, B. J.; Dunbar, C. E.; McVeigh, E. R.; Lederman, R. J. Serial cardiac magnetic resonance imaging of injected mesenchymal stem cells. Circulation 2003, 108, 1009–1014. 81. Kraitchman, D. L.; Heldman, A. W.; Atalar, E.; Amado, L. C.; Martin, B. J.; Pittenger, M. F.; Hare, J. M.; Bulte, J. W. In vivo magnetic resonance imaging of mesenchymal stem cells in myocardial infarction. Circulation 2003, 107, 2290–2293. 82. Garot, J.; Unterseeh, T.; Teiger, E.; Champagne, S.; Chazaud, B.; Gherardi, R.; Hittinger, L.; Gueret, P.; Rahmouni, A. Magnetic resonance imaging of targeted catheter-based implantation of myogenic precursor cells into infarcted left ventricular myocardium. J. Am. Coll. Cardiol. 2003, 41, 1841–1846. 83. Rivi`ere, C.; Boudghene, F. P.; Gazeau, F.; Roger, J.; Pons, J. N.; Laissy, J. P.; Allaire, E.; Michel, J. B.; Letourneur, D.; Deux, J. F. Iron oxide nanoparticle-labeled rat smooth muscle cells: cardiac MR imaging for cell graft monitoring and quantitation. Radiology 2005, 235, 959–967. 84. Tallheden, T.; Nanmark, U.; Lorentzon, M.; Rakotonirainy, O.; Soussi, B.; Waagstein, F.; Jeppsson, A.; Sjogren-Jansson, E.; Lindhal, A.; Omerovic, E. In vivo MR imaging of magnetically labeled human embryonic stem cells. Life Sci. 2006, 79, 999–1006. 85. Kustermann, E.; Roell, W.; Breitbach, M.; Wecker, S.; Wiedermann, D.; Buehrlr, C.; Welz, A.; Hescheler, J.; Fleischmann, B. K.; Hoehn, M. Stem cell implantation in ischemic mouse heart: a high-resolution magnetic resonance imaging investigation. NMR Biomed. 2005, 18, 362–370. 86. Bos, C.; Delmas, Y.; Desmouliere, A.; Solanilla, A.; Hauger, O.; Grosset, C.; Dubus, I.; Ivanovic, Z.; Rosenbaum, J.; Charbord, P.; Combe, C.; Bulte, J. W.; Moonen, C. T.; Ripoche, J.; Grenier, N. In vivo MR imaging of intravascularly injected magnetically labeled mesenchymal stem cells in rat kidney and liver. Radiology 2004, 233, 781–789. 87. Hauger, O.; Frost, E. E.; van Heeswijk, R.; Deminiere, C.; Xue, R.; Delmas, Y.; Combe, C.; Moonen, C. T.; Grenier, N.; Bulte, J. W. MR evaluation of the glomerular homing of magnetically labeled mesenchymal stem cells in a rat model of nephropathy. Radiology 2006, 238, 200–210. 88. Jirak, D.; Kriz, J.; Herynek, V.; Andersson, B.; Girman, P.; Burian, M.; Saudek, F.; Hajek, M. MRI of transplanted pancreatic islets. Magn. Reson. Med. 2004, 52, 1228–1233. 89. Dousset, V.; Doche, B.; Petry, K. G.; Brochet, B.; Delalande, C.; Caille, J. M. Correlation between clinical status and macrophage activity imaging in the central nervous system of rats. Acad. Radiol. 2002, 9 (Suppl 1), S156–S159. 90. Rausch, M.; Sauter, A.; Frohlich, J.; Neubacher, U.; Radu, E. W.; Rudin, M. Dynamic patterns of USPIO enhancement can be observed in macrophages after ischemic brain damage. Magn. Reson. Med. 2001, 46, 1018–1022. 91. Schroeter, M.; Saleh, A.; Wiedermann, D.; Hoehn, M.; Jander, S. Histochemical detection of ultrasmall superparamagnetic iron oxide (USPIO) contrast medium uptake in experimental brain ischemia. Magn. Reson. Med. 2004, 52, 403–406. 92. Saleh, A.; Schroeter, M.; Jonkmanns, C.; Hartung, H. P.; Modder, U.; Jander, S. In vivo MRI of brain inflammation in human ischaemic stroke. Brain 2004, 127, 1670–1677. 93. Hauger, O.; Delalande, C.; Trillaud, H.; Deminiere, C.; Quesson, B.; Kahn, H.; Cambar, J.; Combe, C.; Grenier, N. MR imaging of intrarenal macrophage infiltration in an experimental model of nephrotic syndrome. Magn. Reson. Med. 1999, 41, 156–162. 94. Beckmann, N.; Cannet, C.; Fringeli-Tanner, M.; Baumann, D.; Pally, C.; Bruns, C.; Zerwes, H. G.; Andriambeloson, E.; Bigaud, M. Macrophage labeling by SPIO as an early marker of
P1: OTA/XYZ P2: ABC c14 JWBS037-Chen February 1, 2011
0:6
Printer Name: Yet to Come
REFERENCES
95. 96.
97.
98.
99.
100.
101.
102.
103. 104. 105.
331
allograft chronic rejection in a rat model of kidney transplantation. Magn. Reson. Med. 2003, 49, 459–467. Weinmann, H. J.; Ebert, W.; Misselwitz, B.; Schmitt-Willich, H. Tissue-specific MR contrast agents. Eur. J. Radiol. 2003, 46, 33–44. Ruehm, S. G.; Corot, C.; Vogt, P.; Kolb, S.; Debatin, J. F. Magnetic resonance imaging of atherosclerotic plaque with ultrasmall superparamagnetic particles of iron oxide in hyperlipidemic rabbits. Circulation 2001, 103, 415–422. Kooi, M. E.; Cappendijk, V. C.; Cleutjens, K. B.; Kessels, A. G.; Kitslaar, P. J.; Borgers, M.; Frederik, P. M.; Daemen, M. J.; van Engelshoven, J. M. Accumulation of ultrasmall superparamagnetic particles of iron oxide in human atherosclerotic plaques can be detected by in vivo magnetic resonance imaging. Circulation 2003, 107, 2453–2458. Wu, Y. L.; Ye, Q.; Foley, L. M.; Hitchens, T. K.; Sato, K.; Williams, J. B.; Ho, C. In situ labeling of immune cells with iron oxide particles: an approach to detect organ rejection by cellular MRI. Proc. Natl. Acad. Sci. U.S.A. 2006, 103, 1852–1857. Zimmer, C.; Wright, S. C., Jr; Engelhardt, R. T.; Johnson, G. A.; Kramm, C.; Breakefield, X. O.; Weissleder, R. Tumor cell endocytosis imaging facilitates delineation of the glioma–brain interface. Exp. Neurol. 1997, 143, 61–69. Muldoon, L. L.; Sandor, M.; Pinkston, K. E.; Neuwelt, E. A. Imaging, distribution, and toxicity of superparamagnetic iron oxide magnetic resonance nanoparticles in the rat brain and intracerebral tumor. Neurosurgery 2005, 57, 785–796. Brillet, P. Y.; Gazeau, F.; Luciani, A.; Bessoud, B.; Cuenod, C. A.; Siauve, N.; Pons, J. N.; Poupon, J.; Cl´ement, O. Evaluation of tumoral enhancement by superparamagnetic iron oxide particles: comparative studies with ferumoxtran and anionic iron oxide nanoparticles. Eur. Radiol. 2005, 15, 1369–1377. Fortin-Ripoche, J. P.; Martina, M. S.; Gazeau, F.; Menager, C.; Wilhelm, C.; Bacri, J. C.; Lesieur, S.; Cl´ement, O. Magnetic targeting of magnetoliposomes to solid tumors with MR imaging monitoring in mice: feasibility. Radiology 2006, 239, 415–424. Artemov, D.; Bhujwalla, Z. M.; Bulte, J. W. Magnetic resonance imaging of cell surface receptors using targeted contrast agents. Curr. Pharm. Biotechnol. 2004, 5, 485–494. Hofmann, B.; Bogdanov, A.; Marecos, E.; Ebert, W.; Semmler, W.; Weissleder, R. Mechanism of gadophrin-2 accumulation in tumor necrosis. J. Magn. Reson. Imaging 1999, 9, 336–341. Bulte, J. W. M.; In vivo MRI cell tracking: clinical studies. AJR. Am. J. Roentgenol. 2009, 193, 3314–3325.
P1: OTA/XYZ P2: ABC c15 JWBS037-Chen January 25, 2011
23:25
Printer Name: Yet to Come
CHAPTER 15
Nanoparticles Containing Rare Earth Ions: A Tunable Tool for MRI ` C. RIVIERE Laboratoire de Physique de la Matiere ` Condensee ´ et Nanostructures, Universite´ de Lyon, Lyon, France
S. ROUX Laboratoire de Physico-Chimie des Materiaux Luminescents, Universite´ de Lyon, Lyon, France ´
R. BAZZI Laboratoire Physico-Chimie des Electrolytes, Collo¨ıdes et Sciences Analytiques, Universite´ Pierre et Marie Curie, Paris, France
J.-L. BRIDOT Service de Chimie Gen Organique et Biomedicale, Laboratoire de RMN et d’Imagerie ´ erale, ´ ´ Moleculaire, Universite´ de Mons-Hainaut, Mons, Belgium ´
C. BILLOTEY Laboratoire CREATIS–Animage, Universite´ Claude Bernard, Lyon, France
P. PERRIAT Groupe d’Etudes de Metallurgie Physique et de Physique des Materiaux, Universite´ Claude ´ ´ Bernard, Lyon, France
O. TILLEMENT Laboratoire de Physico-Chimie des Materiaux Luminescents, Universite´ de Lyon, Lyon, France ´
15.1 INTRODUCTION Thanks to its highly resolved three-dimensional (3D) noninvasive imaging capabilities, magnetic resonance imaging (MRI) appears as one of the most powerful clinical diagnostic techniques. MRI can not only give precise anatomical information, but it can also provide functional measurements. MRI takes advantage of the magnetic resonance of water molecules that compose our bodies. The contrast obtained in MR images is due to the differences in the density and relaxation time of water protons within the defined
Nanoplatform-Based Molecular Imaging Edited by Xiaoyuan Chen C 2011 John Wiley & Sons, Inc. Copyright
333
P1: OTA/XYZ P2: ABC c15 JWBS037-Chen January 25, 2011
334
23:25
Printer Name: Yet to Come
NANOPARTICLES CONTAINING RARE EARTH IONS: A TUNABLE TOOL FOR MRI
volume of investigation. While the density of water protons will depend on the level of hydration of the different structures of the body, the relaxation time will depend on the environment of water molecules and the degree of mobility of the water molecules within this environment. MRI contrast is thus dependent on chemical, physical, and biological properties of the investigated tissue. Any changes in the tissue (protein density, edema, tumor, etc.) will affect the magnetic interactions between water protons, leading to changes in MRI contrast properties of that tissue. The relaxation of water protons is also dependent on the strength of the static magnetic field and on the pulse sequence of radiofrequency (rf) waves. Thanks to the development of various types of sequences, it is possible to get images mainly reflecting water proton density, transverse relaxation time (T2 imaging), or longitudinal relaxation time (T1 imaging). The signal tends to increase with decreasing relaxation time T1 and decrease with decreasing relaxation time T2. The interpretation of the resulting images therefore leads to the delineation and identification of most tissues. To increase MRI sensitivity, contrast agents have been developed that modify the relaxation time of water protons in their vicinity, hence modifying the resulting signal intensity detected in this region. Contrast agents are divided into two groups, the negative and the positive ones. Negative contrast agents are mainly superparamagnetic iron oxide nanoparticles, which induce a large shortening of the transverse relaxation time T2 of surrounding water protons, resulting in a decrease of tissue signal intensity, and hence darkening of the MR images. Positive contrast agents are mainly paramagnetic chelates, which induce a large shortening of the longitudinal relaxation time T1 of surrounding water protons, resulting in an increase of tissue signal intensity, and hence brightening of the MR images. Even if quantitative results can be obtained in theory for both negative and positive contrast agents, it requires development of important data treatment and physical interpretation for negative contrast agents, whereas it is much more straightforward for positive contrast agents. Among positive contrast agents, gadolinium(III), a rare earth compound, is the most widely used paramagnetic element, due to its seven unpaired electrons and relatively long electronic relaxation. Contrast properties are dependent on two key features: the water-exchange rate between bulk water and water bound to the Gd ions and the rotational correlation time of the Gd containing entities. As free gadolinium is extremely toxic, gadolinium ions are caged within chelates ® structures, such as DTPA (diethylenetriaminepentaacetic acid, Magnevist ), and DOTA ® (1,4,7,10-tetraazacyclododecane-1,4,7,10-tetraacetic acid, Dotarem ) [1]. As many nanostructures presented in this chapter are based on the use of such complexes, the structure of gadolinium chelates is described in detail in Section 15.2. This caging limits the number of coordinated water molecules associated with each Gd ion, and hence limits the modification of signal intensity that could be induced with such a paramagnetic element. Another limitation is due to the small size (<1 nm) of clinically used Gd complexes, which leads to short rotational correlation time, hence limiting the modification of proton relaxivity. Their small size also reduces the blood half-life of Gd complexes, and blood-pool imaging is only accessible during the first few minutes after intravenous injection of the contrast agent. Moreover, as it is quickly eliminated from the body, both passive (by taking advantage of the enhanced permeability and retention (EPR) effect in tumoral tissues) and active targeting (by grafting relevant target-specific ligand on the surface of the contrast agents) are limited. In that context, many nanostructures have been developed in the past few decades (1) to increase MRI sensitivity while keeping a positive contrast enhancement
P1: OTA/XYZ P2: ABC c15 JWBS037-Chen January 25, 2011
23:25
Printer Name: Yet to Come
STRUCTURE OF GADOLINIUM CHELATES AND CHEMICAL MODIFICATION
335
to get quantitative results more easily; and (2) to tune the biodistribution of the contrast agent for tissue-targeting purpose. Moreover, these nanostructures can often support additional attractive features to combine imaging and therapeutic functionalities and/or to induce a contrast in different complementary imaging techniques (optic, MRI, PET, ultrasound, etc.). This chapter specifically focuses on the different nanostructure-based strategies employed to increase the sensitivity and specificity of the positive contrast agents. To increase the sensitivity of rare earth based contrast agents, three parameters are usually considered: (1) increasing the number of Gd per contrast agent, (2) increasing the rotational correlation time by increasing molecular weight and size, and (3) increasing the number of coordinated water molecules. The sensitivity of a given contrast agent is classically quantified by measurements of the longitudinal and transverse relaxivity values, r1 and r2 , which refer to the amount of increase in 1/T1 and 1/T2, respectively, per millimolar of agent (often given as per mM of Gd). The ratio r2 /r1 is used to classify the contrast agents as positive T1 agents, which usually have r2 /r1 ratios of 1–2, or as negative T2 agents, which have r2 /r1 ratios as high as 10 or more. For nanostructure elements, it is not only the relaxivity per Gd that defines the efficiency of the contrast agent but also the number of chelated Gd per nano-object. It is then also interesting to compute the relaxivity values per mM of nano-objects, instead of per mM of Gd. Section 15.3 is dedicated to a review of different strategies to entrap the rare earth chelates in 3D nanostructures, while Section 15.4 is devoted to the use of nanostructures with rare earth surface functionalization, including the development of crystalline nanoparticles based on inorganic or organometallic compounds of rare earth elements. In Section 15.5, we introduce the use of such MR contrast agents for both imaging and therapy.
15.2 STRUCTURE OF GADOLINIUM CHELATES AND CHEMICAL MODIFICATION The structure of gadolinium chelates was extensively described in several reviews, which are taken as the authority in the MRI field [1, 2]. Generally, gadolinium chelates are composed of a gadolinium(III) ion buried in a ligand, which is designed for occupying at most eight binding sites at the metal center. The remaining coordination sites are occupied by solvent water molecules. The presence of the water molecules in the coordination sphere of Gd(III) ion and its exchange with the water molecules of the surrounding solution are essential for affecting the relaxation time of solvent protons. The hydration of the gadolinium(III) complex is an important parameter since the higher the number of water molecules in the coordination sphere, the more pronounced are the effects on the relaxation and therefore on the contrast of the MR images. However, the gain in relaxivity is often depreciated by the loss of stability of the complex, which leads to the undesirable liberation of toxic Gd3+ in the body. The majority of clinically approved contrast agents that are commonly used for medical imaging are gadolinium(III) chelates whose ligand is based on a linear or cyclic polyaminocarboxylate motif [1, 2]. Among them, linear DTPA and cyclic DOTA are well suited for the immobilization of gadolinium(III) onto the nanoparticles (Fig. 15.1). First, Gd-DTPA and Gd-DOTA (actually Gd(DTPA)2− and Gd(DOTA)− ) form
P1: OTA/XYZ P2: ABC c15 JWBS037-Chen January 25, 2011
336
23:25
Printer Name: Yet to Come
NANOPARTICLES CONTAINING RARE EARTH IONS: A TUNABLE TOOL FOR MRI
O
O
OH O
HO HO
N
N O
O OH
OH
N O
O OH
HO
N
N
N
N
O HO
DTPA
OH
DOTA
O
FIGURE 15.1 The structures of commercially available ligands DTPA (left) and DOTA (right).
stable octacoordinated chelates. The presence of several carboxylic acid moieties in these ligands can be exploited for the covalent immobilization of gadolinium(III) chelates. Their grafting can be performed directly onto the surface or indirectly by using a crosslinker or by modification of the ligand. It will depend on the chemical composition of the nanoparticle surface. Since an amide group is, regarding hydrolysis, more resistant than an ester function, the immobilization of gadolinium(III) chelates, whatever the grafting mode, requires a condensation reaction between COOH groups of the ligand and an amine function (Fig. 15.2). This reaction is generally promoted by a water-soluble carbodiimide (EDC) and N-hydroxysuccinimide (NHS), which are added to the carboxylic acid containing molecules before the reaction with amine [3]. The conversion of one or two acid groups into amide is accompanied by a lowering of the stability constant of the resulting gadolinium chelate because the coordination strength of amide is lower than the one of carboxylate. Despite this, their use for imaging application is possible because, in most cases, the decrease of the stability constant is of the same magnitude as the one observed for the clinically approved Gd-DTPA-bis(amide) (DTPA-BMA:Gd, K = 1016.85 ). In order to circumvent the decrease of stability constant, the carboxylic acid moieties must therefore be preserved. This implies that the carboxylic groups of the ligands cannot be used for the covalent grafting of the gadolinium complexes. In this case, the immobilization of gadolinium chelates requires the presence of a reactive group on the diethylenetriamine backbone besides the carboxylic acid groups. Brechbiel and co-workers developed the synthesis of DTPA and DOTA derivatives comprised of a grafting group (amine, isothiocyanate) carried by the ethylene chain (Fig. 15.3) [4, 5]. Despite the advantages that these functionalized ligands can afford for the covalent grafting of gadolinium(III) chelates, they are rarely applied, in contrast to DTPA and DOTA, because of their time-consuming multistep synthesis.
15.3 NANOSTRUCTURES AS 3D SCAFFOLD FOR RARE EARTH CONTRAST AGENT DELIVERY Various strategies have been developed to concentrate paramagnetic contrast agent within a 3D scaffold and thereby to increase MRI sensitivity per contrast agent. One approach is to encapsulate contrast agents within lipid-based nanostructures such as liposomes, or within porous mesoporous nanostructures. More recently, carbon nanotubes and metallofullerenes have also been studied as potentially attractive scaffolds for rare earth contrast agents. Those three approaches will be developed in the following sections.
P1: OTA/XYZ P2: ABC c15 JWBS037-Chen January 25, 2011
23:25
Printer Name: Yet to Come
337
NANOSTRUCTURES AS 3D SCAFFOLD FOR RARE EARTH CONTRAST AGENT DELIVERY
O
O
O
O
2RNH2 +
OH
N
O
N
N
O O
OH
O N
RHN
O N
O
O O
O
HO RNH2
HO N
N
+ O
O
O
OH
N O
OH
EDC +NHS
HO
N
N O
O
O OH
OH
O
OH N
N
O OH
OH
HO
O EDC +NHS
OH
+
HO
N
N
NHR
N O
O RNH2
NHR
OH
OH
HO
O N
O
N
N
N
N
NHR
O HO
HO
O
SCN
O
H N
RHN C COOH
RNH2
HOOC
N
N
N
HOOC
RNH2
N
N
N
N
RHN
C
H N
N
HOOC
COOH COOH
COOH
HOOC
S COOH
N
HOOC
COOH N
+ HOOC
HOOC
COOH COOH
HOOC SCN
COOH
S
+
N
N
N
N COOH
FIGURE 15.2 Reaction of DTPA, DOTA, and DTPA and DOTA-analog with amine groups.
15.3.1 Lipid-Based Structures In “lipid-based structures,” nano-objects possessing an outer membrane and/or an inner cavity will be described. This refers mainly to liposome-based and lipoprotein-based contrast agents. Liposomes are particularly attractive, being phospholipid-based biocompatible structures of adjustable size distribution. They can coencapsulate hydrophobic (within the phospholipid bilayer), hydrophilic (inside the vesicle), and amphiphilic molecules, such as therapeutic chemicals, genes, and contrast agents [6–13]. In order to avoid opsonization and recognition by the reticulo endothelium system (RES), liposomes are usually coated with polyethylene glycol (PEG). The furtiveness
P1: OTA/XYZ P2: ABC c15 JWBS037-Chen January 25, 2011
338
23:25
Printer Name: Yet to Come
NANOPARTICLES CONTAINING RARE EARTH IONS: A TUNABLE TOOL FOR MRI
X
O HO HO
O
O N
N O
O
O
O OH
OH
N O
N
O
N
N
O
O
OH
OH
O
X= NCS, NH2
NCS or NH2-Bn-DTPA
DTPA-BA
COOH
HOOC X
N
N
N
N COOH
HOOC X= NCS, NH2
NCS or NH2-Bn-DOTA
FIGURE 15.3 The structure of the DTPA and DOTA analog bearing a highly reactive isothiocyanate (NCS), amino (NH2 ), or anhydride group.
of such long-circulating liposomes makes them promising passive drug delivery devices ® [13–16]. Stealth liposomes are used commercially to deliver chemotherapeutic drugs such as doxorubicin and cisplatin (Doxil). The use of liposomes as paramagnetic ion carriers has been investigated since the early 1990s. Two issues are to be addressed in the design and synthesis of these biomimeticbased structures: the rare earth payload and the signal intensity enhancement induced by the liposome-associated paramagnetic ions. Even if the relaxivities are improved with such structures, contrast enhancement is usually not as high as the one predicted by the number of Gd complexes that can be trapped in or within these biomimetic nano-objects (Table 15.1). In all these studies, the moderate increase in contrast properties is attributed to the moderate degree of freedom of the rare earth ions, which cannot fully interact with the water protons in the environment. Several attempts to address these two specific issues (payload and signal enhancement) have been carried out over the past two decades. As far as relativities are concerned, liposomes bearing rare earth ions must be divided into two classes, depending on the location of the paramagnetic entities: (1) encapsulated in the liposome cavity and (2) in the membrane.
Paramagnetic Complexes Encapsulated in the Liposome Cavity The use of liposomes as T1 contrast agent was first investigated by encapsulating Gd-DTPA complexes within the liposomes cavity [17]. Even if the relaxivity per contrast agent object was increased with this strategy, the liposome encapsulation significantly lowered the T1 relaxivity (in mM−1 · s−1 of Gd) compared to free Gd complexes. The relaxivity obtained for liposomes varied linearly with the surface-to-volume ratio: the smaller the liposomes, the greater the relaxivity. Even the smallest liposomes (70 nm) have a relaxivity nearly twofold lower than the one obtained with the free complex. This has been attributed to the limited access of Gd atoms to the bulk water molecules because of the liposomal bilayer,
®
10.3
3.4 3.4 9.2 3.9 6.8
3.4 3.4 3.9 6.8
4.3 7
9.1
4.8 3.8
Clinical Gd Complexes
9.1
45,000 43,000 1800 510 15–20 27,000 ?
125 117 130 40 9 100 20–80 nm/ 3–10 nm
22 170
39 10.4
7.2 1 17
41 ?
49,7 ?
? ? 26
1.3 ? 1.9 ?
2.5 × 104 ? 1.1 × 106 ?
2 × 104 150–200 5.9 × 105 ?
? ? 1.5
1.1 1
1.4 1.1
1.1
? ? 4.7 × 104
4.3 7
4.8 3.8
10.3
3.2 × 105 4,3 × 104 3.1 × 104
Nanostructures as 3D Scaffolds for Rare Earth Contrast Agent Delivery
1 1 1 1 1
1
r1 r2 r1 r2 Nb (Gd) (mM−1 · s−1 ) (mM−1 · s−1 ) (mM−1 · s−1 ) (mM−1 · s−1 ) r1 /r2
Relaxivities per Nanostructure
0.47 T 1.5 T
0.47 T 1.5 T
1.4 T 0.47 T 1.5 T
1T 1T 3T 1T 3T
1T
Field (T)
(Continued)
48 39
117 28
19 18 23
157 157 144 157 51
156
Reference
23:25
Porous polymerosomes Gd-encapsulated Stealth liposomes DPPC and Gd-DTPA-bisamide derivative Shell-crosslinked nanoparticles Lipoprotein micelles Gd in the lipid membrane Nanorods Gadonanotubes
Gd-DOTA (Dotarem ) ® Gd-DTPA (Magnevist ) Gd-DTPA ® Gadodiamide (Omniscan ) Gd-Si-DTTA complex
Gd3+ free
Size (nm)
Relaxivities per Gd
TABLE 15.1 Comparative Relaxivity Values of the Different Rare Earth Containing Nanostructures
P1: OTA/XYZ P2: ABC c15 JWBS037-Chen January 25, 2011 Printer Name: Yet to Come
339
340
Gd-DOTA-QD Gd-DOTA-QD Au@DTDTPA-Gd150 Au@DTDTPA-Gd50 Au-Gd Gd2 O3 @PEG
Silica nanoparticles Silica nanoparticles Gd-Si-DTTA Silica nanoparticles Gd-Si-DTPA Silica nanoparticles Gd-Si-DTTA LbL technique
Gadofullerenes Gd@C82
67 31
1 1
?
28.8 10.2 23
24,300
79 131
65.5 110.8 34 67 31
7 × 10 2.48 × 105 ? 5
79 131
1.6 × 10 2.7 × 106 ?
16,000 10,200 63,200 10,200 28,000 ? 45 150 50 ?
100 37 40 37 44 10 8 2.4 2.4 75 4.3
19 ? 23 3.9 3.9 23.7 12
9 19.7 7.8 19 55 ? 54 ? ? 89.5
116 60 12.3 55
1.9 × 106 6.1 × 105 7.8 × 105 5.6 × 105 1.6 × 106 6779 2438 ? ? ?
1.4 × 105 2 × 105 4.9 × 105 1.9 × 105 5.3 × 105 365 1019 585 195 ?
6
2.9 19 2.3 ? ? 3.8
13 3 1.6 2.9
1.2 4.2
2.3 11 1.5
r1 r2 r1 r2 Nb (Gd) (mM−1 · s−1 ) (mM−1 · s−1 ) (mM−1 · s−1 ) (mM−1 · s−1 ) r1 /r2
Relaxivities per Nanostructure
Nanostructures with Rare Earth Surface Functionalization for Multimodal Imaging
27 nm Pore size: 2,4 nm 120 nm Pore size: 2 nm 1 nm
Size (nm)
Relaxivities per Gd
3T ? 1.4 T 7T 7T 1.5 T 1.5 T
4.7 T 3T 3T 3T
0.47 T 4.7 T
3T 9.4 T 9.4 T
Field (T)
52 87 88 101,102 101,102 105 140
158 51 51 52
34 34
49
47
Reference
23:25
Mesoporous silica nanoparticles
Mesoporous silica nanospheres
TABLE 15.1 (Continued)
P1: OTA/XYZ P2: ABC c15 JWBS037-Chen January 25, 2011 Printer Name: Yet to Come
Hollow sphere gadolinium oxide nanoparticles: shell thickness 21 nm, interior cavity 91 nm Porous sphere gadolinium oxide nanoparticles: pores 3–12 nm GdF3 nanoparticles@cit GdF3 nanoparticles@AEP Gadolinium carbonate (Gd2 O(CO3 )2 · H2 O) particles—spheres
Gd2 O3 @DMSA-PEG Gd2 O3 @CA-PEG Gd2 O3 core/silica shell nanoparticles: shell thickness 2 nm Gd2 O3 core/silica shell nanoparticles: shell thickness 2 nm Gd2 O3 core/silica shell nanoparticles: shell thickness 2 nm Dysprosium oxide nanoparticles core/silica shell
16.8 3.17 2.71 16.5
2110 8900 ? ? ?
? 6.3 × 106 3.25 × 105 ?
3.8 4.6 1 nm core/70 total 1 nm core/60 total 130
187 130 51 500
17.7
?
?
4.4
8.8
420
14.2 7.8 8.8
4.3 4.3 2.2
? ? 210.7
2.6 × 105
3.9 × 104
2.107 8.8 × 105 ?
?
?
?
? ? ?
?
?
675
190
6.1 × 104
1.9 × 104
?
4.8 × 103
3.7 × 103
? ? 13
1.9
1.5
?
?
6.6
3.3
1.3
14 T 14 T 3T
3T
3T
17.6 T
7T
7T
7T
124 124 126
144
144
142
142
131
131
140 140 131
23:25
31.4
26.6
?
?
28.9
28.8
11.4
1.5 T 1.5 T
P1: OTA/XYZ P2: ABC c15 JWBS037-Chen January 25, 2011 Printer Name: Yet to Come
341
P1: OTA/XYZ P2: ABC c15 JWBS037-Chen January 25, 2011
342
23:25
Printer Name: Yet to Come
NANOPARTICLES CONTAINING RARE EARTH IONS: A TUNABLE TOOL FOR MRI
which lowers the rate of water exchange between the bulk and the interior of the liposome. The presence of cholesterol in the liposomal bilayer was found to decrease even further the relaxivity by reducing the permeability coefficient of water molecules. Another disadvantage of such an approach was the possible leakage of low molecular weight compounds from the vesicles in vivo. More recently, PEGylated Gd-encapsulated liposomes (117 nm) presenting longcirculating properties and high Gd content were investigated [18]. The decrease in relaxivity compared to the free Gd complexes was overcome by an efficient payload (43,000 Gd per liposomes on average), leading to efficient and long-lasting contrast properties, which are extremely attractive properties for blood-pool imaging and MR angiography. Other interesting alternatives are also under investigation, such as dendrimer-attached Gd complexes encapsulated in so-called porous polymerosomes [19]. Porous polymerosomes are 125-nm structures composed of an outer porous membrane that enables better water exchange across the bilayer, stabilized by a polymer shell (an amphiphilic diblock copolymer). High T1 relaxivity obtained with such a nanostructure (r1 (60 MHz, 1.41 T, 40 ◦ C) = 7.2 mM−1 · s−1 versus 3.1 mM−1 · s−1 for free dendrimer-attached Gd complexes and 3.9 mM−1 · s−1 for free Gd-DTPA in the same condition) is assumed to be due to high Gd payload (approximately 45,000 Gd/porous polymerosome) and faster flux of water across the bilayer together with slower rotational correlation lifetime of the encapsulated Gd dendrimers.
Paramagnetic Complexes in the Membrane To overcome the decrease in water accessibility of encapsulated Gd complexes, a second strategy consists of entrapping paramagnetic complexes within the membrane. This strategy has been used successfully with chelating moieties modified with different hydrophobic groups such as alkyl chains of varying lengths attached via amide linkages, Gd-DTPA-stearylamine (Gd-DTPA-SA), GdDTPA-phosphatidyl ethanolamine (DTPA-PE) [20], or more recently Gd-DTPA-polylysine linked to a polychelating amphiphilic polymer (Gd-DTPA-PLL- PAP) [21, 22]. Laurent and co-workers [23] have studied in detail the effects of chain type and length of amphiphilic complexes bearing gadolinium complexes within the phospholipid membrane. The relaxivity of the studied paramagnetic unilamellar DPPC (1,2-sn-glycero-3phosphatidylcholine) liposomes was found to be mainly related to the matching of the carbon chain length and of the chemical structure of the phospholipid and the Gd complex. The best efficiency as an MRI constrast agent was obtained when the membrane is less structured, that is, when the immobilization of the paramagnetic complex within the membrane was lowered, enabling efficient water exchange rate. Fortunately, the incorporation of PEG moieties on the outer surface of the membrane, classically used to increase the contrast agent’s stability and blood half-life, also contributes to an increase in contrast properties, as water molecules are tightly associated with hydrophilic PEG molecule [24]. Sterically stabilized ∼100-nm liposomes with Gd complexes coupled to a fatty acid chain (dimyristoyl-sn-glycero-3-phosphoethanolamineN-diethylene-triaminepentaacetate—DMPE-DTPA) exhibit threefold higher relaxivity at 1.5 T compared to conventional paramagnetic complexes and a prolonged blood halflife (120 min), leading to a progressive signal enhancement in the tumor region (with a maximum obtained 20 h after liposome injection) [25] . Limitations of this approach include the low loading capacity and the high quantity of the lipid in the carrier. One strategy to increase the number of complexes per liposome is the use of amphiphilic polychelating polymers. The advantage of this approach is that one
P1: OTA/XYZ P2: ABC c15 JWBS037-Chen January 25, 2011
23:25
Printer Name: Yet to Come
NANOSTRUCTURES AS 3D SCAFFOLD FOR RARE EARTH CONTRAST AGENT DELIVERY
343
phospholipid can carry a polymer molecule with multiple paramagnetic complexes on the liposome surface. It was found that liposome modification with Gd-DTPA-polylysine can increase the rare earth load per vesicle 5- to 15-fold [21], while decreasing the lipid-to-Gd ratio. Even if this approach does not increase the relaxivity per mole of gadolinium, the relaxivity per nano-objects is increased (Table 15.1). Most recent studies have focused on the use of PEG-liposomal membrane-incorporating polychelating amphiphilic polymer heavily loaded with gadolinium and functionalized with ligand specifically targeting cancer cells for sensitive and specific tumor detection [22].
Stable Lipoprotein Mimetic Micelles A promising alternative approach is the use of lipoprotein-based structures. These structures can be considered as attractive biomimetic nanostructures to target vulnerable plaques (in atherosclerosis), tumor, or other inflammatory related disease. To our best knowledge, the design of such objects is among the most advanced techniques concerning in vivo targeting and imaging of the early stages of atherosclerosis. High-density lipoprotein (HDL) is a native structure that naturally interacts with atherosclerotic plaques [26]. In vivo, HDL helps remove excess cholesterol from macrophages in atherosclerotic plaques. It is composed of a hydrophobic core of triglycerides and cholesterol esters covered in a monolayer of phospholipids into which apolipoprotein A-I (apoA-I) is embedded [27]. After extraction, the apoliproteins are reconstituted with phospholipids, including a phospholipid-based contrast agent such as GdDTPA-DMPE (1,2-dimyristoyl-sn-glycero-3-phosphoethanolamine diethylenetriamine pentaacetic acid) [28] or Gd-DTPA-BSA (bovine serum albumin [29]). Depending on the dialysis method, both spherical [28] and disks-like [29] nano-objects can be obtained, bearing about 20 molecules of Gd. As the protein is endogeneous, it is not recognized by the reticuloendothelial system, resulting in a low level of immunogenic host reaction. In addition, the small size of the final nano-object obtained (7–12 nm in diameter) is much smaller than the one obtained with liposomes and is thus more adapted to penetrate the plaques. As far as contrast properties are concerned, due to the reduced rotation of the Gd complexes, compared with free Gd-DTPA, the obtained relaxivity values are relatively important (two- to threefold increase compared to Gd-DTPA, with r1 = 10.4 · mM−1 · s−1 at 65 MHz [28]). This type of lipoprotein-based contrast agent is readily internalized by macrophage cell lines in vitro [27], resulting in high contrast enhancement of labeled cell pellets. Using such a contrast agent in a mouse model of atherosclerosis (hyperlipidemic mice, Apo-E knockout), in vivo signal enhancement in atherosclerotic plaques rich in macrophages has been reported 24 h post intravenous injection, confirming the targeting efficiency of this contrast agent [28, 29]. Alternative solutions are currently under investigation to optimize this type of structure: (1) the use of synthetic peptides instead of HDL [26] to avoid the lengthy separation procedure required to isolate HDL and also to reduce immunogenicity; and (2) inclusion of an organic core (gold, iron oxide, or quantum dots) for multimodal imaging [27]. In addition, by postfunctionalization with a molecule known to be expressed in large amount in a given state of the development of atherosclerosis or tumor progression (such as VCAM-1, E-selectin, or ␣v 3 integrin), even more specific targeting could be achieved with such a contrast agent [27].
P1: OTA/XYZ P2: ABC c15 JWBS037-Chen January 25, 2011
344
23:25
Printer Name: Yet to Come
NANOPARTICLES CONTAINING RARE EARTH IONS: A TUNABLE TOOL FOR MRI
pH-Sensitive Liposomes In order to detect pathologies associated with pH modifications, pH-sensitive liposomes have been developed recently [30–32]. Basically, the idea is to modify the liposome structure as a function of pH values. This modification will then induce a selective modification in the contrast enhancement, providing a direct visualization of in vivo pH variations. Classically, the system consists of gadolinium diethylenetriamine pentaacetic acid bismethylamide (gadodiamide, GdDTPA-BMA) encapsulated within liposomes. The modification of contrast enhancement can be induced by a destabilization of the liposome’s membrane, hence liberating entrapped Gd complexes. As stated above, the relaxation properties of free Gd complexes are usually higher than the one of Gd complexes entrapped within the liposome cavity. The liberation of Gd complexes due to pH variation will hence increase the contrast enhancement in that particular area. The difficult part resides in the control of the destabilization process, as the liposomes should remain stable within the bloodstream, before reaching the pathological area. Many formulations are currently under investigation. One of the most promising is the dipalmitoylphosphatidylethanolamine/ dipalmitoylglycerosuccinate system (DPPE/DPSG) [31], which remains stable in blood while displaying large pH response. 15.3.2 Carbon Nanotubes and Metallofullerenes Different carbon structures have shown recently their potential as MRI T1 contrast agents. Two types of structures have been shown to have extremely high relaxivity values at clinical magnetic field strength and are currently under investigation: (1) metallofulleres and (2) carbon nanotubes.
Metallofullerenes Metallofullerenes consist of endohedral cages of about 1 nm that encapsulate atoms, clusters, or small molecules. Gadolinium containing metallofullerenes (Gd@C60 or Gd@C82 mainly), so-called gadofullerenes, are emerging as novel caging structures concurrently replacing clinically used gadolinium chelates. They exhibit high relaxivity properties while preventing Gd3+ dissociation in vivo thanks to the confinement of the paramagnetic ion within the fullerene cage. The metallofullerenes are typically obtained thanks to an arc discharge method of metal/graphite composite rods. The resulting soot must then be extracted with carbon disulfide. The soluble fullerene M@C82 is then purified and isolated by two-stage high performance liquid chromatography (HPLC). The obtained toluene solution of M@C82 is then mixed with NaOH solution containing a few drops of tetrabutylammonium hydroxide (TBAH) as catalyst. This step produces a reaction mixture that can easily be isolated by decantation. Aqueous solution of M@C82 (OH)n (gadofullerenols in the case M = Gd) is then precipitated with methanol and the precipitate is further rinsed with methanol to ensure the removal of the catalyst and NaOH. The rinsed precipitate is finally dissolved in water and further purified by dialysis against water. Because this synthesis is labor intensive, has low yield (∼1% wt), and only 4% of sublimable Gd endohedrals are of the soluble Gd@C82 variety, other routes of synthesis have recently been investigated to produce soluble M@C60 , which represents the major portion of all metallofullerenes produced in the traditional arc synthesis method. Indeed, the first generation of M@C60 (M being any metal ion) materials were insoluble in water and air sensitive, which prevents their use in biomedical applications. New techniques of derivatization, solubilization, and stabilization of these species have overcome these limitations [33], by forming soluble carboxylate gadofullerenes, Gd@C60[C(COOH)].
P1: OTA/XYZ P2: ABC c15 JWBS037-Chen January 25, 2011
23:25
Printer Name: Yet to Come
NANOSTRUCTURES AS 3D SCAFFOLD FOR RARE EARTH CONTRAST AGENT DELIVERY
345
The obtained Gd@C82 and Gd@C60 were investigated as potential MRI contrast agents. They both show high water proton relaxivity with r1 , respectively, 20 and 10 times higher than clinically used Gd-DTPA both in vitro (r1 (Gd@C82 ) = 81 mM−1 · s−1 at 1 T, 25 ◦ C [34] and r1 (Gd@C60 ) = 40 mM−1 · s−1 [35], against 4 mM−1 . s−1 for Gd-DTPA in the same condition) and in vivo (enhancement detected after intravenous (IV) injection of a dose 20 times lower than the one currently used for Gd-DTPA [34]). It is difficult to understand the mechanisms underlying this contrast enhancement, as the encapsulated Gd ion is unable to interact directly with water protons. Even if the precise electronic and magnetic properties of gadofullerenols are not known at present, the relaxivity enhancement is attributed to electronic interaction of cages of Gd ions with bulk water protons through the fullerene cage walls and their functionalized groups [35]. Gd@C82 has indeed paramagnetic properties due to the three electrons transferred from Gd to the C82 fullerene cage, leading to seven unpaired electrons in the Gd orbitals inside the C82 cage and one unpaired electron on the cage [36]. The enhanced relaxation of the surrounding water protons is thus likely due to dipolar interactions between unpaired electrons and the large number of water molecules on the surface of the fullerenes [1], together with a decrease in the overall molecular rotational motion. As opposed to Gd chelates, whose relaxation mechanism is mainly based on an “innersphere” mechanism (direct interaction of water protons with Gd ions, that rapidly exchange with bulk water molecules—fast rotational correlation motion), the mechanism for gado–fullerene is entirely “outer-sphere” (related to diffusion of water molecules in the vicinity of the contrast agents) [35]. Although there is no experimental evidence on the dynamic structure of the surrounding water molecules, the presence of hydrogen bonds at the surface of the fullerene, together with their negative surface charge, may lower the rotation of the water molecules in direct contact with the surface of the fullerene and increase the effective hydrodynamic diameter, inducing substantial decrease in the overall molecular rotational motion [37]. Moreover, relaxivity measurements as a function of field strength (NMRD profile) show high-field peaks, typical of slowly rotating objects [35]. NMRD measurements were best fitted with a rotational correlation time two to three orders of magnitude higher than the one obtained with Gd chelates (on the order of nanoseconds for gadofullerenols versus picoseconds for Gd chelates), and a rate of water exchange two to three times greater than the rate of water exchange from the inner sphere water to bulk water for Gd3+ . The relaxivity enhancement has also been related to an increase in hydrodynamic diameters of gadofullerenes (hence a decrease in its tumbling rate) by aggregation processes as a function of pH [38]. Taking advantage of the controlled aggregation process, gadofullerenes appear as interesting pH-sensitive agents. The strong pH dependency of the aggregation (increasing hydrodynamic diameter with decreasing pH), and hence the strong pH dependency of the relaxivities (increasing contrast enhancement with decreasing pH) can be used to induce contrast enhancement in regions with low pH values, such as in tumors [35, 38]. A biodistribution study in a mouse model indicated that, in vivo, the Gd@C82 (OH) aggregates are recognized by the reticuloendothelial system (RES), with strong signal enhancement and Gd concentration in lung, liver, spleen, and kidney after IV administration [34]. In contrast, biodistribution of the Gd@C60 [C(COOH)2 ]10 in a rat model shows a low level of recognition by the RES for that type of gadofullerenols, which is rapidly appearing in the kidneys, with only minimal uptake into the liver and excretion via the bladder within 1 h of injection [33].
P1: OTA/XYZ P2: ABC c15 JWBS037-Chen January 25, 2011
346
23:25
Printer Name: Yet to Come
NANOPARTICLES CONTAINING RARE EARTH IONS: A TUNABLE TOOL FOR MRI
Carbon Nanotubes Since other chapters of this book are dedicated to carbon nanotubes for optical imaging and multimodal imaging, this section is restricted to a short description of interesting results concerning the use of carbon nanotubes as a versatile structure to get a high load of Gd complexes. The so-called gadonanotubes developed by Sitharaman and co-workers [39] is certainly one of the most advanced attempts to use carbon nanotubes as T1 MRI contrast agents. In their study, they used short nanotubes (20–100 nm). Such single-walled carbon nanotubes (SWNTs) have recently been shown to shuttle various cargoes across cellular membrane without cytotoxicity [40–42]. Such gadonanotubes exhibit relaxivity values at clinical field strength 40-fold higher than the current free Gd complexes with r1 ∼ 170 mM−1 · s−1 (at 1.5 T and 40 ◦ C) [39]. Compared to gadofullerenes, the interesting point is the good transport properties of water and protons [38, 43], enabling access of water molecules to Gd complexes within the carbon nanotubes. Proton relaxivity measurements as a function of the magnetic field strength (NMRD profile) of such systems show a profile remarkably different from other paramagnetic agents, notably with constant high relaxivity values at high field. As MRI resolution increases with the magnetic field strength, the use of such a system could circumvent the actual limitation of paramagnetic complexes, which usually show a decrease in relaxivity above 1.5 T. More recently, pH-sensitive gadonanotubes have been developed, which exhibit a twofold change in relaxivity between pH 7.4 and 6 [44]. Such a system is particularly attractive to achieve smart MRI detection. Cancerous and normal tissues do not have the same pH, with lower pH usually associated with cancerous areas. Detecting such pH variation is thus an interesting approach for the detection of cancer. 15.3.3 Mesoporous Nanoparticles Mesoporous nanoparticles are mainly based on silica nanoparticles. Such rigid structure has a large pore volume, with tunable pore size and surface functionalization. The biocompatibility of such a system is now established [45, 46] and mesoporous silica nanoparticles are emerging as ideals agents for biomedical application. Superparamagnetic nanoparticles and paramagnetic complexes were successfully incorporated in these structures for their use as MRI contrast agents. As far as paramagnetic molecules are concerned, different Gd complexes have been used: Gd-DTTA [47] (diethylenetriaminetetraacetate), Gd-DTPA [48, 49], or Gd-DOTA [50]. The synthesis is based on a sol-gel procedure using surfactant such as CTAB (cetyltrimethylammonium bromide) as template. A first route consists of cocondensation under basic conditions of tetraethyl orthosilicate (TEOS) with chelating agent (such as DTPA) conjugated with silane. The obtained solution of mesoporous nanostructure is then mixed with GdCl3 , which enables caging of Gd ions within chelates [48, 49]. A second route consists of condensation of tetraethyl orthosilicate (TEOS) under basic conditions. The obtained mesoporous structures are then coated with Gd chelates conjugated with silane complex via a siloxane linkage [47]. Mesoporous nano-objects of tunable form (nanospheres [47, 49, 50], nanorods of various aspect ratio [48]), size (from 60 to 1100 nm), and pore size (from 2 to 10 nm) can be obtained [47]. As far as MRI contrast properties are concerned, as the accessibility of paramagnetic entities to water molecules is one of the most important features in obtaining high contrast
P1: OTA/XYZ P2: ABC c15 JWBS037-Chen January 25, 2011
23:25
Printer Name: Yet to Come
NANOSTRUCTURES AS 3D SCAFFOLD FOR RARE EARTH CONTRAST AGENT DELIVERY
347
agent relaxivity properties [51, 52], the large surface-to-volume ratio of mesoporous nanoobjects appears as an interesting property for the design of sensitive contrast agents. In fact, Gd containing mesoporous nano-objects present a remarkable relaxivity increase, compared to free Gd complexes or Gd entrapped in plain silica nanoparticles. This enhancement in contrast properties is due to a concomitant effect of the porosity of the structure, which enables free movement of water molecules in and out of the structure, and of the interaction of Gd complexes with the silica structure, which impedes the rotational movement of the lanthanide complexes. This decrease in tumbling rate of the paramagnetic unit consequently increases the relaxivities of water molecules in the vicinity of the Gd complexes. Five- to tenfold increase in longitudinal and transverse relaxivities were reported for mesoporous nanorods [48] of 100 nm with a 3/5 aspect ratio and a pore size of 2–3 nm, as well as on spherical mesoporous nanoparticles of 20–50 nm with pore size of 3 · 5 nm [50]. Lower enhancement on a per-Gd basis was reported for larger nanospheres (600–700 nm with pore size of 8 nm) [50]. The difference in relaxivities between various mesoporous nanostructures is attributed to the different accessibility of the water molecules to the Gd complexes, localized inside or outside the silica pores [50]. It is hypothesized that only Gd complexes grafted outside the nanopores contribute to the observed relaxivity, whereas the complexes grafted inside the nanopores do not modify the relaxation of water molecules. This may be due to a hindered or limited (slow) access of the water molecules to the Gd complexes inside the pores. As the localization of the Gd complex on the inner or outer surfaces of the mesoporous silica particles strongly depends on the pore size, the corresponding relaxivity is similarly affected by the pore size. Gd load per nano-object is also a key issue to increase the relaxivity on a per contrastobject basis. However, the relation between Gd load and relaxivity enhancement is not linear. Lin and co-workers [53] reported higher relaxivities for Gd load ranging from 1.6% to 3.1% wt. The r1 value reaches its maximum for 1.6% Gd load (23.6 mM−1 · s−1 at 9.4 T and 23 ◦ C) and decreases abruptly as the Gd loading increases from 2.3% to 3.1%, falling to relaxivity values close to free Gd-DTPA at 6.8% (4.4 versus 3.9 mM−1 · s−1 for GdDTPA at 9.4 T and 23 ◦ C). At such high Gd loading, the dipole–dipole interaction among the paramagnetic ions is presumed to be more significant, hence shortening the electronic relaxation time. Gd-containing mesoporous nanoparticles have been found to have potential application as blood-pool contrast agents [47], but also as tissue-targeting contrast agents and cell labeling agents. As the use of mesoporous structures as drug carriers has been demonstrated [54, 55], potential application combining drug and sensitive contrast agent targeting are envisaged. Postgrafting of the ligand of interest is possible thanks to a large number of amine groups on the surface of the mesoporous nanoparticles at the end of the synthesis. Tissue targeting is thus possible in theory with such structures; however, no reports of such application have been made to our knowledge. The presence of amine groups has mainly been used for the design of multifunctional mesoporous nanoparticles combining optical and magnetic properties. The fluorescent dye can be incorporated in the mesoporous nanostructures during synthesis, by adding a small amount of organic dyes conjugated to aminopropyltriethoxysilane [47]. Detailed description of this approach is given in Section 15.4.2. Another interesting approach that is developed (Section 15.4.2) is to combine the incorporation of Gd with other lanthanides such as Eu and Tb [53], which possess excellent
P1: OTA/XYZ P2: ABC c15 JWBS037-Chen January 25, 2011
348
23:25
Printer Name: Yet to Come
NANOPARTICLES CONTAINING RARE EARTH IONS: A TUNABLE TOOL FOR MRI
luminescent properties in the visible range. This approach enables the design of multimodal contrast agents that combine high resolution and sensitivity of optical setup, with 3D localization provided by MRI. The use of mesoporous nanoparticles as cell markers has also been studied: these structures are readily internalized by many types of cells [46–48, 56], without the need of any transfection agent. Hsiao and colleagues have recently reported the use of hybrid mesoporous nanoparticles with Gd for stem cell tracking and were able to track as few as 4000 stem cells for 14 days using 1.5-T clinical MRI [49]. However, precise study on the effect of internalization on contrast properties has to be performed for long-term cell trafficking follow-up [57, 58]. 15.4 MULTIFUNCTIONAL NANOSTRUCTURES CONTAINING RARE EARTH ELEMENTS FOR MULTIMODAL IMAGING 15.4.1 Introduction The derivatization of the nanoparticle surface by gadolinium chelates constitutes an attractive route for the elaboration of highly sensitive MRI contrast agents. As already stated in the previous section, this strategy is expected to improve the enhancement of the positive contrast of MR images in comparison to isolated gadolinium chelate because of the large amount of paramagnetic centers per probe and of the reduction of the rotational motion. Since the great attention that nanoparticles received rests on their possibility to gather a large range of properties in a reduced volume, the immobilization of gadolinium chelates on nanoparticles increases the multifunctional character of the nanoparticles. Many works focused on the combination of MRI modality and fluorescent nanoparticles. The combination of fluorescence and magnetic resonance imaging in the same nano-object provides probes that are characterized by a high sensitivity (due to the fluorescence properties) and high resolution (owing to the specifications of MRI). Fluorescence of these paramagnetic nanoparticles can arise from the encapsulation of organic dyes (extrinsic fluorescence) or from intrinsic properties of the nanoparticles. Among the latter, quantum dots are probably the most widely known. Other combinations of imaging techniques can also be envisaged by functionalizing gold nanoparticles with gadolinium chelates since gold nanoparticles are able to induce a contrast enhancement in CT imaging. The association of several imaging modalities is very important because it ensures better reliability of observations and also allows monitoring of the contrast agents at various scales (from the subcellular scale to the whole organism) by different techniques [59–62]. Moreover, an increasing number of works have recently emphasized the therapeutic activity of the nanoparticles, which can be remotely controlled by an external physical stimulus ( e.g., magnetic field, visible and NIR light, X-ray photons, and neutron beam). In other words, harmless nanoparticles can be rendered cytotoxic after energy absorption [63, 64]. For optimal effect, follow-up of these therapeutic agents is required from administration to elimination from the body. For instance, the activation of these nanoparticles must be performed only when they are accumulated preferentially in the pathological zone. The possibility to image the fate of the nanoparticles is consequently a crucial issue. Since most of these nanoparticles are not detected by MRI, the functionalization of their surface by gadolinium chelates is highly expected. However, it depends on the chemical properties of the nanoparticle’s surface. For this reason, the composition of the ligand and/or the surface must be tuned for ensuring an efficient anchoring of the gadolinium chelates.
P1: OTA/XYZ P2: ABC c15 JWBS037-Chen January 25, 2011
23:25
Printer Name: Yet to Come
MULTIFUNCTIONAL NANOSTRUCTURES CONTAINING RARE EARTH ELEMENTS
349
15.4.2 Functionalization of Nanoparticles by Gadolinium(III) Chelates The functionalization of the nanoparticles aims at the synthesis of highly sensitive MRI contrast agents and/or the combination of several imaging techniques and/or the follow-up of therapeutic agents. Most of the particles developed for eventual biomedical applications are essentially composed of silica (actually polysiloxane), quantum dots, and gold nanoparticles. The functionalization of nanoparticles by gadolinium chelate depends on the chemical composition of the surface, which is in most cases imposed by the nature of the nanoparticles.
Fluorescent Silica Nanoparticles and Other Oxide Nanoparticles Silica nanoparticles can easily be produced by applying either the protocol developed by St¨ober or microemulsion techniques. St¨ober’s protocol consists of the alkaline hydrolysis– condensation of tetraethyl orthosilicate (TEOS) poured in a large volume of ethanol [65]. The high dilution and the aqueous NH4 OH to TEOS ratio are key parameters for yielding monodisperse silica spheres in a large range of sizes (from 20 to 800 nm). Control of the size, when microemulsion techniques are applied, is ensured by the volume of micelles, which behave as a nanoreactor [66]. Moreover, these synthesis protocols can be exploited for the encapsulation of metal oxide nanoparticles because of the great affinity between the surface of these particles and polysiloxane shell precursors. Besides their simplicity, both methods are very attractive because multifunctional nanoparticles can be obtained by minor variations of the composition of the polysiloxane precursors and/or by postfunctionalization. Indeed, a large variety of polysiloxane precursors are available for envisaging accurate control of the composition of the silica nanoparticles and for anticipating a functionalization before and/or after the hydrolysis–condensation. The production of multifunctional silica nanoparticles can be performed in several steps for the elaboration of multilayered nanostructures [67]. Each layer and the surface after derivatization contribute to the multifunctional character of the nanoparticles since they can each be designed for expressing a peculiar feature. For instance, aminopropyltriethoxysilane (APTES) was widely used in mixture with TEOS for introducing in silica nanoparticles, in each layer or on the surface, amino groups, which can act as a grafting site for the immobilization of organic dyes, hydrophilic molecules for improving the colloidal stability in aqueous biological media, and biotargeting groups for increasing the accumulation of the nanoparticles in the zone of interest within the organism. Van Blaaderen and co-workers demonstrated that the conjugation of fluorescein isothiocyanate to APTES before the hydrolysis–condensation allows the formation of fluorescent silica nanoparticles in which the organic dyes are distributed in the whole nanoparticles or in the whole thickness of the polysiloxane shell, whereas the postfunctionalization (i.e., after the hydrolysis–condensation) of the amino groups led to the tethering of the organic dyes onto the nanoparticles [67]. Wiesner and co-workers developed the synthesis of highly fluorescent nanoparticles, which are composed of an organic dye (rhodamine B isothiocyanate (RBITC), cyanine 5 NHS ester (Cy5-NHS)) loaded core and a protective shell [68]. The fluorescent cores were obtained by adapting St¨ober’s protocol: the hydrolysis–condensation was performed on a mixture of TEOS and organic dyes conjugated to APTES. In a second step, the protective shell was obtained by hydrolysis–condensation of a mixture of TEOS and APTES whose amino group makes possible the covalent immobilization of neutral polyethylene glycol chains. Such a postfunctionalization favors renal clearance of the nanoparticles and significantly reduces the liver uptake after in vivo injection of these
P1: OTA/XYZ P2: ABC c15 JWBS037-Chen January 25, 2011
350
23:25
Printer Name: Yet to Come
NANOPARTICLES CONTAINING RARE EARTH IONS: A TUNABLE TOOL FOR MRI
highly fluorescent nanoprobes [69]. The microemulsion technique can also be exploited for the synthesis of fluorescent silica nanoprobes. The research group of W. Tan determined the experimental conditions for the production of fluorescent nanoparticles in water in oil (W/O) microemulsion (reverse microemulsion). Water droplets (micelles) are dispersed in the oil (cyclohexane) phase and stabilized by a surfactant layer (Triton X-100) for preventing the coalescence of droplets. The alkaline hydrolysis–condensation of TEOS and organic dye conjugated APTES takes place inside the micelles. This group demonstrated the versatility of this strategy since various organic and inorganic dyes can be covalently or not encapsulated in the amorphous silica matrix [66, 70]. They succeeded in tuning the color of the resulting nanoprobes and therefore in covering the visible spectral domain by embedding in the same nanoparticles several distinct organic dyes but with various ratios [71]. Moreover, these nanoparticles can be postfunctionalized. In other words, the surface of these fluorescent nanoparticles can be derivatized by the covalent immobilization of oligonucleotide for lowering the detection threshold of hybrization detection [70] or by the condensation of negatively charged polysiloxane precursors for improving colloidal stability [66]. This brief overview, which is naturally not exhaustive, aims at drawing attention to the possible ways of providing multifunctional silica nanoparticles. They were exploited for the synthesis of fluorescent and paramagnetic silica nanoparticles. Fluorescent, radio-opaque, and paramagnetic silica nanoparticles were synthesized by applying reverse microemulsion (water-in-oil system) techniques. Soluble ruthenium(II) based fluorophores are encapsulated in a silica core, which is further embedded in paramagnetic and aminated shells inside water droplets of the reverse microemulsion. The paramagnetic character of these nanoparticles was conferred by the capture of gadolinium(III) ions since the fluorescent core is covalently functionalized by N-(trimethoxysilylpropyl) ethyldiamine (TSPETE) [72]. The water solubility of these particles is ensured by the formation of an outer shell obtained by polymerization of TEOS, APTES, and 3-(trihydroxysilyl)propyl methylphosphonate (THPMP). The presence of negatively charged THPMP compensates the positive charge of APTES and generates the colloidal stability in water because the fucntionalization by THPMP provides a high negative charge to the surface of the particles. Moreover, the presence of amine functions, resulting from APTES, can be used for bioconjugation of the particles. Due to their composition, these particles behave as in vitro contrast agents for fluorescence imaging (Ru(bpy) dyes), MRI (TSPETE-Gd), and computed tomography imaging (Ru and Gd elements). However, these particles showed a lower radio-opacity than Omnipaque (iodine-based organic compound), a widely used contrast agent for clinical CT because Ru and Gd exhibit atomic numbers that are smaller than the atomic number of iodine. On the other hand, these particles exert a higher contrast enhancement for MRI than in the case of the commercially available MRI contrast agent Gadoteridol, (Gd-HP-DO3A. Unfortunately, no data about the stability of the complex TSPETE-Gd3+ was available although it is suspected to be relatively low because TSPETE molecules have only five reactive coordination sites. Despite these imperfections, the development of these particles is important because it opens the door to multimodal imaging applications based on gadolinium(III) chelate coated nanoparticles. Lin and colleagues performed the postfunctionalization of fluorescent silica nanoparticles by two different gadolinium chelates, which carry one (Gd-Si-DTTA) or two (Gd-Si-DTPA) propyltrimethoxysilyl moieties [51]. After encapsulation of a luminescent [Ru(bpy)3 ]Cl2 core (bpy = 2,2 -bypyridine) in an amorphous silica matrix, the surface of the resulting nanoparticles was derivatized by condensation of silanol (Si OH) groups present on the nanoparticles and the methoxysilyl groups of the gadolinium chelates. The authors
P1: OTA/XYZ P2: ABC c15 JWBS037-Chen January 25, 2011
23:25
Printer Name: Yet to Come
MULTIFUNCTIONAL NANOSTRUCTURES CONTAINING RARE EARTH ELEMENTS
351
noticed that the functionalization with Gd-Si-DTTA led to the formation of a monolayer of gadolinium chelates, whereas a thick paramagnetic multilayered shell was provided when Gd-Si-DTPA is used for the postfunctionalization of the fluorescent nanoparticles. The final size of the nanoparticles is around 37 and 40 nm, respectively. The covalent immobilization of these gadolinium chelates confers to the nanoparticles the ability to enhance the positive contrast of the MR images since they exhibit higher longitudinal relaxivity values per mM Gd3+ ions than the one for unbound gadolinium complexes. Moreover, each Gd-Si-DTTA chelate bound to the fluorescent silica nanoparticles exhibits higher relaxivities r1 and r2 than that of Gd-Si-DTPA grafted to similar nanoparticles (r1 = 19.7 s−1 · mM−1 Gd3+ and r2 = 60 s−1 · mM−1 Gd3+ vs. r1 = 7.8 s−1 · mM−1 Gd3+ and r2 = 12.3 mM−1 · s−1 Gd3+ ). This can be reasonably explained by a higher water exchange rate since gadolinium chelates Gd-Si-DTTA coordinate two water molecules whereas Gd-Si-DTPA chelates coordinate only one water molecule. In addition, another feature of nanoparticles coated by Gd-Si-DTTA plays in favor of a higher longitudinal relaxivity. Since the surface functionalization of the fluorescent silica nanoparticles by Gd-Si-DTTA yields a monolayered shell in contrast to the derivatization of the nanoparticles by Gd-Si-DTPA, gadolinium(III) ions are therefore more accessible. But the silica nanoparticles coated by Gd-Si-DTPA exhibit a higher longitudinal relaxivity r1 when the comparison is based on the amount of particulate contrast agents since each particle coated by Gd-Si-DTPA contains more gadolinium(III) ions. It must be pointed out that these nanoparticles have an impressive longitudinal relaxivity whatever the coating (r1 = 2.0 ×105 s−1 · mM−1 and r2 = 6.1 × 105 s−1 · mM−1 vs. r1 = 4.9 × 105 s−1 · mM−1 and r2 = 7.8 × 105 s−1 · mM−1 ). The great potential of these luminescent and paramagnetic nanoparticles was confirmed by in vitro experiments, carried out on monocyte cells with Gd-Si-DTTA coated nanoparticles. Their uptake by these cells was successfully and efficiently monitored by fluorescence imaging, flow cytometric experiments and MRI (T1 and T2-weighted images). This study reveals that the internalization of the nanoparticles in the monocyte cells does not alter the viability and that the fluorescent and paramagnetic nanoparticles allow multimodal in vitro imaging. These authors explored another strategy, for developing particulate contrast agents for imaging cells by fluorescence and MRI. From fluorescent nanoparticles coated by Gd-SiDTTA (see above), they applied a layer-by-layer (LbL) techniques for immobilizing a higher amount of gadolinium chelates [52]. Since Gd-Si-DTTA coated nanoparticles are highly anionic, a layer of cationic Gd-DOTA oligomers was deposited onto the negatively charged surface via electrostatic interactions. As a result, the particles exhibit a positive charge that allows the further deposition of anionic polystyrenesulfonate (PSS). The LbL assembly strategy, which consists in successive depositions of cationic Gd-DOTA oligomers and anionic PSS, yields luminescent and paramagnetic nanoparticles with alternate multilayer coatings of positively charged Gd-DOTA oligomers and negatively charged PSS. It was demonstrated that the size of the multilayered nanoparticles linearly increases with the number of depositions (from 37 nm (for a single layer) to 43 nm (for seven layers)). A similar evolution is observed for the relaxivities r1 and r2 of the particulate contrast agents because the alternate deposition of an additional cationic layer led to an increase in the amount of gadolinium(III) chelates for each nanoparticle. The per particle r1 values increase from 1.94×105 mM−1 · s−1 for nanoparticles coated by a single Gd-Si-DTTA layer to 5.34 × 105 mM−1 · s−1 for Gd-Si-DTTA coated nanoparticles after the deposition of seven layers (four of cationic Gd-DOTA oligomers and three of anionic PSS), whereas the per particle r2 values increase from 5.61×105 mM−1 · s−1 to 1.55×106 mM−1 · s−1 . The LbL self-assembly thus offers an interesting strategy for increasing nanoparticle MR relaxivities. It also affords the
P1: OTA/XYZ P2: ABC c15 JWBS037-Chen January 25, 2011
352
23:25
Printer Name: Yet to Come
NANOPARTICLES CONTAINING RARE EARTH IONS: A TUNABLE TOOL FOR MRI
opportunity to design these particulate contrast agents for the targeting of cancerous cells, which is highly recommended for the early detection of tumors. Since LbL self-assembled nanoparticles are terminated with anionic PSS polymers, a biotargeting peptide sequence containing arginine–glycine–aspartate (RGD) and seven consecutive lysine (K) residues (K7RGD) can be immobilized onto the surface through electrostatic interaction between positively charged lysine residues of the K7RGD sequence and negatively charged PSS layer. The RGD sequence is well known for its strong interaction with ␣v 3 integrins that are upregulated in growing endothelial and angiogenic cancer cells. Such a functionalization led to the efficient targeting of human colon cancer (HT-29) cells and calf pulmonary artery endothelial (CPAE) cells as revealed by in vitro experiments monitored by fluorescence imaging and MRI. Although these results are very promising, the in vivo application of these contrast agents is not yet validated. Thanks to the trimethoxysilyl group, which allows its grafting onto the oxide surface, Si-Gd-DTTA can be used successfully for rendering mesoporous silica nanoparticles paramagnetic [47]. As described Section 15.3.2, such a postfunctionalization confers to the mesoporous nanoparticles very high longitudinal and transversal relaxivities r1 and r2 , which amounts to 7.0 × 105 mM−1 · s−1 (3 T)/2.48 × 105 mM−1 · s−1 (9.4 T) and 1.6 × 106 mM−1 · s−1 (3 T)/2.7 × 106 mM−1 · s−1 (9.4 T) on a per millimolar particle basis, respectively. These relaxivity values are much larger than the solid silica nanoparticles that are coated with multilayers of the Gd-DTPA derivative (see above [51]). The enhanced MR relaxivity was attributed to the ready access of water molecules through the nanochannels of the mesoporous particles derivatized by gadolinium chelates Si-Gd-DTTA. When the functionalization is performed by a mixture of Si-Gd-DTTA and of rhodamine B conjugated to aminopropyltriethoxysilane (4 mol %, relative to Gd-Si-DTTA), the resulting mesoporous nanoparticles can behave as contrast agents for fluorescence and MR imaging. The uptake of these fluorescent and paramagnetic mesoporous nanoparticles by monocyte cells was revealed both by the fluorescence and the contrast enhancement MR images. Moreover, they appeared well suited for in vivo MRI after intravenous injection in the tail vein since aorta and liver are clearly visible on T1- and T2-weighted images, respectively. These nanoparticles are indeed able to induce a positive and a negative contrast. Even if hepatic uptake renders liver imaging possible thanks to the phagocytosis of the nanoparticles by the resident liver macrophage cells (Kupfer cells), it however leads to a quick clearance from the blood and limits the use of these nanoparticles as in vivo contrast agents. In addition to gadolinium, luminescent rare earth ions (Tb3+ , Eu3+ , etc.) are coordinated by the same ligands and therefore can be immobilized onto the nanoparticles. The postfunctionalization of a fluorescent silica core that contains Ru(bpy) as the fluorophore with silylated polyaminocarboxylate ligands allows the immobilization of both Tb3+ and Gd3+ ions. The presence of Tb3+ confers to the particles emission properties that are complementary to those of Ru2+ entrapped inside the core, whereas gadolinium ions are expected to render the nanoparticles paramagnetic [73]. The silylated gadolinium chelates can also be successfully applied for the functionalization of metal oxide nanoparticles since they are covered by OH groups, which act as anchoring sites. Frias and co-workers reported on the modification of boehmite (AlO(OH)) nanoparticles with silylated gadolinium(III) chelates (derivatives of DO3A) [74]. These particles are therefore able to enhance the positive contrast of MR images. As expected, these ligands, which are covalently bound to the nanoparticle surface through the condensation between methoxy groups of silylated ligands and OH groups of the surface, are able to capture luminescent lanthanide ions (Eu3+ and Tb3+ ).
P1: OTA/XYZ P2: ABC c15 JWBS037-Chen January 25, 2011
23:25
Printer Name: Yet to Come
353
MULTIFUNCTIONAL NANOSTRUCTURES CONTAINING RARE EARTH ELEMENTS
HOOC
COOH N
N
N
N
HOOC NH
O NH
HOOC
HOOC
HOOC
N
N
N
N
COOH
C+
NH
HOOC
N
N
N
HS
N
O
N
N
N
N
COOH
O HOOC
N
N
N
N COOH
COOH
HOOC
COOH
COOH
HN NH
N
O NH
M O
H N
COOH
ED
H N
NH2 NH2
H2 N MxOy
APTES + TEOS
O
H2 N
MxOy
NH2
HO
O
NH2
H2N
N
N OH
O
NH2
OH ED
C+
NH
HO
O
N O
HOOC HOOC
OH
O
N
N
HOOC
N
S
COOH
O
NH
N
HN HN
COOH N COOH
N
HOOC
COOH NH
M O
H N HOOC HOOC
N
COOH
N
HOOC HOOC
NH
NH
N O
NH
O
N N
COOH N COOH COOH
FIGURE 15.4 Schematic illustration of the procedure for the synthesis of the silica encapsulated oxide nanoparticles. The DTPA or DOTA was attached to the amine groups of the surface through an amide reaction.
The postfunctionalization of fluorescent silica nanoparticles by silylated gadolinium(III) chelates is very efficient, as demonstrated by the work of Lin’s research group. Since most of the clinically approved gadolinium(III) chelates are based on polyaminocarboxylate motif, they can easily be immobilized on aminated nanoparticles through the formation of amide linkage (Fig. 15.4). Amine functions in nanoparticles are generally introduced by the inorganic copolymerization of TEOS and aminated polysiloxane precursors (APP like APTES or APTMS) or by the derivatization of oxide nanoparticles by APP. Delville and co-workers succeeded in converting alumina and silica nanoparticles into fluorescent and paramagnetic contrast agents [75]. Such a tuning was performed after the coating of the bare nanoparticles with a polysiloxane shell yielded by the hydrolysis–condensation of APTMS. Such an encapsulation allows tethering organic dyes (rhodamine isothiocyanate) and polyaminocarboxylate ligand (DTPA derivative) for the immobilization of gadolinium(III) ions onto the surface of the nanoparticles. DTPA ligand was grafted to the surface either by the reaction between aminated surface and bis-anhydride derivative of DTPA or by the peptide reaction promoted by the carbodiimide EDC in the presence of NHS between aminated surface and DTPA. Despite a low amount of gadolinium chelates immobilized on the nanoparticles (4 gadolinium ions per nanoparticle), a strong positive and negative contrast enhancement is observed. This increase in relaxivity is consistent with the increase in rotational correlation time, which results from both the increase in molecular weight and the maximum compactness of the SiO2 structure as compared to organic molecules. Indeed, the rigid nature of the inorganic nanoparticle tends to avoid the Gd3+ chelate from freely rotating or even folding back. Since the propensity of nanoparticles to be internalized in cells is higher than in the case of molecules, the immobilization of the gadolinium chelates onto the nanoparticles is attractive because an efficient labeling of microglial cells by paramagnetic agents is expected. The uptake of these particles by the cells was confirmed by the strong contrast enhancement and fluorescence. Microglial cells are well known to be
P1: OTA/XYZ P2: ABC c15 JWBS037-Chen January 25, 2011
354
23:25
Printer Name: Yet to Come
NANOPARTICLES CONTAINING RARE EARTH IONS: A TUNABLE TOOL FOR MRI
chemoattracted and strongly accumulate particularly in brain pathological situations, including tumors. These cells can therefore behave as cargo for freighting a large amount of fluorescent and paramagnetic nanoparticles in the zone of interest. As a result, labeled cells can be followed up from the injection site to the target, which will be clearly delineated thanks to the presence of a large amount of contrast agents. This type of detection will open the route for more efficient contrast agents for MR cell tracking. The aforementioned examples obviously show that the encapsulation of nanoparticles in a polysiloxane shell is an efficient route for rendering the nanoparticles paramagnetic through the immobilization of gadolinium ions either by the use of silylated polyaminocarboxylate ligands or by the conjugation of polyaminocarboxylate ligands with amine functions that are present on the surface of the polysiloxane shell if aminated precursors are used. However, this method requires strict control of the experimental conditions to avoid bulk polymerization of the polysiloxane precursors. The surface modification of transition metal oxide nanoparticles can also be performed by an alternative method that relies on the grafting of gadolinium(III) chelates onto naked nanoparticles. These ligands are terminated by an anchoring function, which ensures their immobilization onto the nanoparticles since they are designed for establishing a strong interaction with the surface. The selective reactivity of the surface of TiO2 nanoparticles to ortho-substituted enediols (such as dopamine) was exploited for grafting gadolinium(III) chelates [76]. Prior to the derivatization of the TiO2 nanoparticles, dopamine modified MR contrast agents (DOPA-DO3A) were synthesized by coupling commercially available succinimidyl ester activated 1,4,7,10-tetraazacyclododecane-1,4,7-tetraacetic acid to dopamine hydrochloride. The gadolinium(III) ions were captured by the resulting ligands. As expected, these particles are able to exert a contrast enhancement of MR images since they exhibit a longitudinal relaxivity of 3.5 mM−1 · s−1 per gadolinium ion, which is similar to the value of clinically used small molecule contrast agents. Since this method of functionalization leads to the immobilization of a single monolayer of gadolinium(III) chelates, the longitudinal relaxivity per particles remains moderate (61 mM−1 · s−1 ) in comparison to the shell obtained from silylated ligands (up to 7.0×105 mM−1 · s−1 , see above [47, 51]). In addition to dopamine modified contrast agents, these nanoparticles can also be functionalized by DNA oligonucleotides (DNA-DOPA-DO3A-Gd) for targeted imaging and therapy. DNADOPA-DO3A-Gd nanoparticles were successfully applied for the labelling PC3M cells, which were revealed by two-dimensional X-ray fluorescence maps and by T1-weighted MR images. Since phosphonic acid groups have been found to be very effective in anchoring organic molecules to various metal oxide surfaces, gadolinium(III) complexes of polyaminocarboxylate ligands containing a phosphonate group in a pendant arm were used for the surface modification of TiO2 nanoparticles [77].
Paramagnetic Quantum Dots (QDs) The functionalization of QDs by gadolinium chelates represents another route for the elaboration of fluorescent positive contrast agents combining fluorescence imaging and MRI. QDs refer to a crystallite cluster of II–VI semiconductors (CdS, CdSe, ZnS, CdSe@ZnS, etc.) exhibiting unique optical properties that arise from quantum confinement effects characteristic of these small dimensions (1–8 nm) [78–80]. Their attractive features (high photostability, high quantum yield, controllable and narrow emission band, broad excitation band) should contribute to a breakthrough in fluorescence imaging even if their routine application is still controversial (owing to the toxicity of Cd2+ ions) [81, 82]. Their relatively narrow emission band whose position
P1: OTA/XYZ P2: ABC c15 JWBS037-Chen January 25, 2011
23:25
Printer Name: Yet to Come
MULTIFUNCTIONAL NANOSTRUCTURES CONTAINING RARE EARTH ELEMENTS
355
is controlled either by the particle size or, for a given size, by tuning the composition of the alloyed semiconducting QD (CdSeTe), together with their broad absorption spectra offer a large palette of distinct color enabling multilabeling [83]. Since several synthetic routes have been developed for the preparation of water-soluble QDs with narrow size distribution and high quantum yields, the literature is overflowing with examples of biological materials labeled by QDs [84, 85]. Indeed, the most common synthetic routes of QDs lead to hydrophobic fluorescent nanoparticles since they are coated by trioctylphosphine/ trioctylphosphine oxide (TOP/TOPO). The biological application of QDs requires their functionalization by capping hydrophilic molecules. Besides the water solubility, the functionalization plays two supplementary important roles: capping ligands (1) can serve as anchoring sites for grafting molecules and (2) can ensure, by passivating the QD surface, their protection and consequently the preservation of optical properties. A representative list of caps and the QD dispersal strategies they use is provided in a review written by the research group of Mattoussi [86]. These strategies can be grouped into three major routes (Fig. 15.5). The first one consists in the exchange of a the hydrophobic monolayer of TOP/TOPO by hydrophilic thiol or phosphine terminated ligands (mercaptocarbonic acids, alkylthiol terminated DNA, thioalkylated oligo-ethyleneglycols). This functionalization is driven by mass action and is possible because sulfur or phosphorus atoms can establish strong dative bonds with the chalcogenide species present at the surface of QDs. However, it must be pointed out that the colloidal and chemical stability is considerably improved when hydrophilic ligands used for replacing TOP/TOPO carry at least two thiol or phosphine functions (dihydrolipoic acid derivatives, oligomeric phosphines). The second strategy described in the literature is based on the encapsulation of QDs in a silica shell. As mentioned earlier, silica shells afford multiple ways for further surface modification. The first step of the formation of the silica shell lies in the replacement of TOP/TOPO by mercaptopropyltrimethoxysilane (MPTMS) whose thiol end allows the immobilization onto the QD. The growth of the silica shell is therefore generated by hydrolysis–condensation of a polysiloxane precursor mixture (TEOS, APTES, MPTMS, etc.) from the alkoxysilyl groups of mercaptopropyltrimethoxysilane bound to the QD surface. The third method preserves the native TOP/TOPO on the QDs and uses variants of amphiphilic “diblock” and “triblock” copolymers and phospholipids to tightly interleave/interdigitate the alkylphosphine ligands through hydrophobic attraction, whereas the hydrophilic outer block permits aqueous dispersion and further derivitization. These three strategies (Fig. 15.5) were applied for derivatizing fluorescent QDs with gadolinium chelates. To add the MRI contrast ability, QDs can be functionalized with Gd3+ DOTA (1,4,7,10-tetraazacyclododecane-1,4,7,10-tetraacetic acid) complexes as shown by the study of Jin et al. [87]. For achieving this, the TOP/TOPO layer was first replaced by glutathione, which is composed of a thiol function for the immobilization of the ligand onto the NIR emitting QD, an amine group for further functionalization, and two carboxylic moieties for water compatibility and also for further functionalization. The addition of DOTA NHS ester to glutathione coated QDs led to the covalent grafting of DOTA through the formation of amide bond. In a last functionalization step, QDs are made paramagnetic by the capture of gadolinium(III) ions from DOTA (Gd3+ -DOTA-QD). The functionalization induces an increase of the hydrodynamic diameter from 7 to 10 nm. This strategy seems to be efficient since Gd3+ -DOTA-QD exhibits a relatively high longitudinal relaxivity r1 (365 mM−1 · s−1 ). Moreover, these paramagnetic QDs have a great potential for in vivo imaging since the implantation of a tube filled with Gd3+ -DOTA-QD into the abdomen of a mouse can be detected by NIR-fluorescence imaging and by MRI. Unfortunately, no data
P1: OTA/XYZ P2: ABC c15 JWBS037-Chen January 25, 2011
356
23:25
Printer Name: Yet to Come
NANOPARTICLES CONTAINING RARE EARTH IONS: A TUNABLE TOOL FOR MRI
R
R
R R
R
R R S
S
HS
S ZnS S S S S S CdSe Core S S S S S S S S
R
ZnS
R
HS
HS
R
R
S S ZnS
R
HS S
CdSe Core
S
S
R
R
(RO) Si S
S
R
S
Si(OR)
S (RO) Si
S
S
S S ZnS
S
S
CdSe Core
S
R
S
S
Si(OR)
S
Si(OR)
S S Si(OR)
(RO) Si
S (RO) Si
R R
(RO) Si
(RO) Si
S
Si(OR) Si(OR) Si(OR) Si(OR)
R R
R
R
Si(OR)
(RO) Si S
S
S
R R
(RO) Si (RO) Si (RO) Si
Si(OR)
S S
R
R=CH , C H ...
R=CH , C H
R
S
R S
R
R
S
S
CdSe Core
Si(OR)
R
S
R
R=NH , COOH...
R=NH , COOH... R
S
R
ZnS
R
R
R
R
R
CdSe Core
R=NH , COOH...
R
R
Si(OR)
(RO) Si Si(OR)
(RO) Si (RO) Si
Si(OR) Si(OR) Si(OR)
(RO) Si
Si(OR)
Si(OR)
APTES + TEOS + ...
NH2 H2N
NH2 CdSe Core
H2N
NH2
FIGURE 15.5 Schematic of generic QD functionalization.
are given about the intravenous injection of these nanoparticles and their biodistribution, which is a key parameter for the biomedical application of these QDs. Gerion et al. [88] demonstrated that the functionalization of QDs by gadolinium(III) chelates can be performed, as expected, after their encapsulation in a silica shell. The shell was obtained in two steps. After a priming step replacing trioctylphosphine oxide (TOPO) surfactants on the QD surface with mercaptopropyltrimethoxysilane (MPTMS), polymerization of siloxane was performed in methanol under slightly basic conditions. In a second step, addition of fresh MPTMS and polyethylene glycol(PEG)–propyltrimethoxysilane permitted the introduction of functional (SH) and stabilizing PEG groups on the surface of the SiO2 shell. As reported in the section devoted to fluorescent silica nanoparticles, ligands used for the immobilization of paramagnetic ions can easily be tethered to the polysiloxane shell provided that functional groups are present. In order to preserve the eight coordination sites of DOTA, an aromatic amine was grafted to one of the ethylene bridges of DOTA (NH2 -Bn-DOTA, Fig. 15.3). DOTA derivative with paramagnetic Gd3+ chelates are linked to the thiols of the silica shell embedding the QDs by using the sulfo-SMCC crosslinker. The maleimide group of sulfo-SMCC allows the grafting to the thiol function due to the formation of thioether linkage while the NHS ester group at the opposite end of the crosslinker favors the formation of amide linkage with the aromatic amine of the DOTA derivative. As compared to unbound Gd-DOTA, the Gd3+ ion relaxivity r1 for the gadolinium chelates attached to the silica-coated nanoparticles is increased approximately sixfold (23 mM−1 · s−1 vs. 3.7 mM−1 · s−1 ). Chemical analyses revealed that about 50 Gd-DOTA are grafted onto the nanoparticles. As a result, the paramagnetic QDs (8 nm) exhibit a relatively high longitudinal relaxivity (1019 mM−1 · s−1 , 1.4 T). Preliminary in vivo tests indicate that paramagnetic silica-coated
P1: OTA/XYZ P2: ABC c15 JWBS037-Chen January 25, 2011
23:25
Printer Name: Yet to Come
MULTIFUNCTIONAL NANOSTRUCTURES CONTAINING RARE EARTH ELEMENTS
357
nanoparticles after intravenous injection provide a contrast enhancement in MRI. The accumulation of paramagnetic nanoparticles in the bladder was evidenced by the appearance of brighter zones in the bladder 5 min after the injection. However, these nanoparticles do not accumulate in the other organs since no contrast enhancement was observed elsewhere. This indicates that the nanoparticles freely circulate without undesirable accumulation and most of them are quickly removed by renal excretion. Both previous works led to the development of paramagnetic QDs suitable for MRI applications. Their functionalization by gadolinium(III) chelates rests on replacement of the hydrophobic TOP/TOPO layer by a hydrophilic one. However, the TOP/TOPO layer can be exploited for a further functionalization of QDs as demonstrated by the studies performed by Bakalova and co-workers and by Mulder and co-workers [89–92]. The preservation of the TOP/TOPO layer limits the deleterious effect of polar species on the luminescence since the hydrophobic layer constitutes an efficient barrier. In contrast to the procedure of Gerion et al. [88], which requires the replacement of TOP/TOPO layer by MPTMS, Bakalova et al. [89] demonstrated that QDs can be encapsulated by a silica shell from the hydrophobic layer. Prior to the inorganic polymerization of the polysiloxane network, hydrophobic QDs were covered by vinyltrimethoxysilane. The hydrophobic tail of this precursor (i.e., the vinyl moiety) was entrapped in the hydrophobic TOP/TOPO layer, whereas the methoxysilyl end group pointed toward the outside. These methoxysilyl groups serve as anchoring sites for the polysiloxane shell yielded by hydrolysis–condensation of a mixture of TEOS and APTES. Owing to the amine function, which is a characteristic of the latter, the surface is ready for surface modification. The MRI contrast agent ability was conferred to the QD either by the entrapment of amphilic gadolinium(III) chelates (common name: Resolve Al–Gd) before the encapsulation in the silica shell or by the surface modification of the silica-coated QDs by hydrophilic gadolinium(III) chelates, which are DOTA derivatives (Gd-DOTA-Bn-NCS) (Fig. 15.3). The isothiocyanatobenzyl group ( CH2 N C S) of Gd-DOTA-Bn-NCS carried by one of the ethylene bridges of the macrocycle ensures the covalent grafting of the gadolinium chelates by condensation with the surface amine function leading to the formation of a thio-urea linkage (Fig. 15.2). Whatever method is used to render them paramagnetic, silica-coated QDs are able to induce the shortening of the longitudinal relaxation time T1. This was reflected by the enhancement of the contrast of T1-weighted images. In vitro and in vivo studies showed that they can be applied as probes for multimodal imaging. Indeed, the internalization of these particles can be monitored by both fluorescence imaging and MRI. It was demonstrated that silica-coated QDs do not alter the viability of labeled cells. Furthermore, labeled cells under continuous UV or laser irradiation remain viable, whereas it generates a cytotoxic effect when cells are loaded with uncoated QDs. The silica shell seems to suppress the photosensitization effect, which was previously reported for uncoated QDs. The low toxicity of these contrast agents is encouraging for in vivo applications. However, accumulation in the spleen must be solved since this reflects the uptake by macrophages, which is commonly observed when positively charged particles are injected intravenously. This capture should be overcome by PEGylation. The encapsulation of TOP/TOPO-coated QDs by the silica shell was also performed via a reverse microemulsion technique [90]. But in this case, the silica shell was formed between the QDs and the TOP/TOPO layer. The postfunctionalization of the resulting particles for rendering them paramagnetic and hydrophilic was realized according to the protocol developed by Mulder’s research group for TOP/TOPO-coated QDs [93].
P1: OTA/XYZ P2: ABC c15 JWBS037-Chen January 25, 2011
358
23:25
Printer Name: Yet to Come
NANOPARTICLES CONTAINING RARE EARTH IONS: A TUNABLE TOOL FOR MRI
Hydrophobic TOP/TOPO-coated QDs can be rendered hydrophilic by coating them with amphiphilic PEGylated phospholipids (PEG lipids: PEG-DSPE (1,2-distearoyl-sn-glycero3-phosphoethanolamine-N-[methoxy-(polyethylene glycol)-2000]). These molecules are widely used to stabilize liposomes for pharmaceutical applications because PEG chains form a hydrophilic layer on the liposomal surface. The coating of hydrophobic QDs (or QDs encapsulated in a silica shell) with a mixture of PEG lipids and of paramagnetic lipids (i.e., gadolinium(III) chelate conjugated to hydrophobic stearylamide) yields watersoluble and fluorescent positive contrast agents for fluorescence imaging and MRI. The postfunctionalization of hydrophobic silica-coated QDs with PEG lipids markedly improves the biodistribution and the pharmacokinetics of these contrast agents [94]. Moreover, some PEG lipids are terminated by a maleimide moiety, which allows bioconjugation. The research group of Mulder succeeded in functionalizing these nanoprobes by RGD, and annexin A5 proteins [91,92]. The grafting of RGD, which is well known for targeting ␣v 5 and ␣v 3 integrins, favors the uptake of these paramagnetic QDs by HUVEC due to the overexpression of ␣v 3 integrins. The ability of these contrast agents to target ␣v 5 and ␣v 3 integrins opens the door to the detection of neoangiogenosis, which is associated with the development of tumors, whereas the conjugation of Annexin A5 to these nanoprobes makes them well suited for the detection of apoptosis by fluorescence imaging and MRI. The strategy of the Mulder research group, which is based on the synthesis of lipid-based nanoparticles, is very attractive because it leads to a large variety of contrast agents for multimodality imaging [93]. By a convenient choice of the building blocks (the lipids, the nanoparticles, the biotargeting groups, etc.) the behavior of these platforms can be accurately tuned for a definite application. Recently, they demonstrated that endogenous nanoparticles (high-density lipoproteins, HDLs) can also be modified for multimodality imaging by incorporating paramagnetic and/or fluorescent lipids but also by including within each HDL nanoparticle a hydrophobic inorganic core such as iron oxide, QDs, or gold [27]. Since it is known to migrate to atherosclerotic plaques and remove cholesterol from macrophages, HDL was exploited by Frias and co-workers to develop a probe for molecular imaging of macrophages [28, 29]. The inclusion of an inorganic core affords additional features for imaging purposes. For instance, the presence of a hydrophobic gold core in a paramagnetic and fluorescent HDL allows monitoring gold–HDL by MRI, fluorescence imaging, and X-ray imaging since the high atomic number of the gold element confers to the gold nanoparticles a great ability to absorb X-ray photons.
Paramagnetic Gold Nanoparticles One of the most promising applications of nanosized gold objects lies in their ability to combine imaging and therapy [95]. Two main strategies were developed for designing gold particles that can be used for imaging and destroying tumors. The first one results from the properties of the gold element, whereas the second one rests on the tunable optical properties of gold particles. Gold particles are indeed characterized by strong light diffusion and absorption cross sections, which are sensitive to the size, shape, and dielectric environment [96]. By tuning these parameters, gold-based nano-objects (gold-coated silica particles and gold nanorods) can be synthesized and exploited for NIR imaging and therapy as demonstrated by Halas and co-workers and El-Sayed and co-workers [97, 98]. Whatever the size, shape, and environment of the gold nano-objects, they strongly absorb X-ray photons. Hainfeld et al. [99, 100] demonstrated that ultrasmall hydrophilic gold nanoparticles (1.9 nm) can be monitored by X-ray imaging after intravenous injection to mice and can be applied for radiotherapy. However, X-ray imaging is not completely harmless for envisaging a long-term biodistribution
P1: OTA/XYZ P2: ABC c15 JWBS037-Chen January 25, 2011
23:25
Printer Name: Yet to Come
MULTIFUNCTIONAL NANOSTRUCTURES CONTAINING RARE EARTH ELEMENTS
359
study without interference with radiotherapy. Moreover, this technique suffers from a low sensitivity. In order to circumvent this problem, the use of gold nanoparticles designed for magnetic resonance imaging was proposed [101, 102]. Roux and co-workers succeeded in designing gold nanoparticles for MRI by reducing gold salt in the presence of a dithiolated ligand DTDTPA (Au@DTDTPA). As a result, 2.4-nm sized gold cores embedded in a DTDTPA shell were obtained. DTDTPA, is a dithiolated derivative of DTPA, which is commonly used as a gadolinium chelate (DTPA-Gd) for clinical MRI. The transformation of DTPA into DTDTPA was performed by the condensation of a DTPA bis-anhydride molecule with two aminoethanethiol molecules. It was demonstrated that each gold core is encapsulated in a multilayered shell, which is composed of about 150 DTDTPA molecules. The tethering and formation of this multilayered shell are actually ensured by the presence of two thiol functions per DTDTPA. Since the thiol function exhibits a great affinity for gold, some DTDTPA molecules present in the shell are immobilized onto gold nanoparticles through the strong interaction between gold and sulfur atoms of one thiol function. Thiol functions that were not involved in the grafting of the shell onto the nanoparticles were implied in the formation of disulfide bonds. These bonds make possible the formation of a multilayered shell around each gold core. Due to the presence of the DTDTPA shell, 150 Gd3+ ions can be immobilized on a gold nanoparticle (Au@DTDTPA-Gd150 ). However, a good compromise between colloidal stability and the amount of active ions for MRI was found when the number of gadolinium(III) ions per particle is fixed at 50 (Au@DTDTPAGd50 ). The behavior of Au@DTDTPA-Gd50 in the bloodstream after intravenous injection can easily be monitored by both X-ray imaging (in transmission and computed tomography modes) and MRI (r1 = 3.9 mM−1 · s−1 based on gadolinium concentration and r1 = 195 mM−1 · s−1 based on nanoparticle concentration). Biodistribution studies performed with both medical imaging techniques and with postmortem elemental analysis of the organs revealed that these paramagnetic gold nanoparticles freely circulate in the bloodstream and do not accumulate in liver, spleen, and lungs. In the case of tumor bearing rats (or mice), it must be pointed out, however, that accumulation is observed in the tumor due to the EPR effect. This phenomenon was exploited for radiotherapy [103]. It was demonstrated that the survival of diseased rats was lengthened only when they were treated by X-ray beam after intravenous injection of Au@DTDTPA-Gd50 . Owing to its imaging and therapeutic potentialities, Au@DTDTPA-Gd50 seems to be suited for in vivo medical applications. Gold nanoparticles coated by rare earth chelates were also produced for development of a fluorescent sensor since some rare earth ions are characterized by a very photostable luminescence, which is distinguished from organic dyes and QDs by narrower emission peaks and longer lifetime of the excited state. Gunlaugsson and co-workers demonstrated that the luminescence of europium buried in a heptadendate ligand (a tetraazamacrocycle) can be modulated by exploiting the deleterious quenching effect of water on rare earth luminescence [104]. The luminescence is switched off when two water molecules are present in the coordination sphere, whereas the displacement of water by a diketone derivative bearing an aromatic ring reestablishes the light, which is even enhanced due to the effect of the antenna from the aromatic ring. This principle developed for modulating fluorescence can be applied for “smart” MRI. When water is present in the coordination sphere, these gold particles can exert an enhancement of the positive contrast of MR images, whereas no enhancement is visible when water was displaced by analytes. MRI contrast agents that are sensitive to the presence of definite biomolecules can therefore be imagined.
P1: OTA/XYZ P2: ABC c15 JWBS037-Chen January 25, 2011
360
23:25
Printer Name: Yet to Come
NANOPARTICLES CONTAINING RARE EARTH IONS: A TUNABLE TOOL FOR MRI
Gold nanoparticles can also be rendered paramagnetic by the covalent grafting of gadolinium(III) chelates to biomolecules, which are then immobilized on the particles for targeting them to the zone of interest. For instance, Chung and co-workers applied this strategy for preparing paramagnetic gold nanostructures, which are able to combine dual modality imaging (MRI and optical imaging) and targeted therapy [105]. First, hollow gold nanoparticles coated by thiol ended PEG and cystein-protein-G were synthesized. Each class of molecules grafted to the hollow nanoparticles ensures a specific role. PEG was used for improving the colloidal stability and therefore the particle blood circulation time and the efficiency of particle internalization, while protein G molecules were introduced to optimize the orientation of antibodies (anti-HER2) on the surface of the hollow gold nanoparticles. Meanwhile anti-HER2 antibodies were derivatized by gadolinium chelate thanks to the covalent coupling between the amino group of the antibody and the isothiocyanate group of p-SCN-Bn-DTPA (Figs. 15.2 and 15.3). After chelation of gadolinium(III) ions by DTPA conjugated to anti-HER2, the paramagnetic antibody (anti-HER2-Gd-DTPA) was immobilized onto hollow gold nanoparticles. The resulting particles were successfully applied as both MRI (r1 = 23.7 mM−1 · s−1 ) and optical imaging contrast enhancers of cancer cells since they can be observed by optical microscopy (due to the optical properties of hollow gold nanoparticles) and by MRI. The presence of anti-HER2 on the nanoparticles favors the internalization in SKBR3 cells (breast cancer cells that overexpress epidermal growth factor receptors targeted by anti-HER2). Labeled cells that can therefore be visualized by dark-field-based scattering images and MRI were destroyed after NIR irradiation due to the photothermal effect, which occurs when NIR light is absorbed by the gold nanostructures. The multifunctional hollow gold nanoparticles are also expected to be used for imaging specific cancer sites using MRI, and then for destroying cancer cells selectively via illumination with NIR light. However, their large size should impede in vivo application of these paramagnetic therapeutic agents. For this reason, data regarding their biodistribution and their pharmacokinetic parameters, which are not available, are required. 15.4.3 Gadolinium(III) Containing Crystalline Nanoparticles
Introduction The development of gadolinium(III)-based nanoparticles, which is discussed in this chapter, aims at affording positive contrast agents for MRI with a stronger ability of contrast enhancement (due to both the increase of gadolinium(III) ion number and the decrease of rotational motion), better pharmacokinetic parameters, a better control of biodistribution, and a larger palette of properties than molecular gadolinium complexes provide. To achieve this goal, Gd-DTPA or Gd-DOTA functionalized polymer [106, 107], selfassembled peptide amphiphiles [108,109] or viral capsid [110], Gd-DTPA terminated dendrimer [111, 112], gadolinium complex loaded liposomes [113], high-density lipoprotein nanoparticles [28, 29], micelles [114, 115] or polymeric nanoparticles [116, 117], gadolinium ions entrapped in zeolites [118], fullerenes [35], carbon nanotubes [39], clays [119] or mesoporous silica nanoparticles [53], and gadolinium chelates immobilized on quantum dots [91, 92], on lipid particles [120], and on gold nanoparticles [101, 102] were synthesized and studied. They are all characterized by an increase of the molecular weight and of the amount of Gd(III) ions per contrast agent. As a consequence of their structure, some of them were easily functionalized by biotargeting groups and/or fluorescent molecules, conferring on them additional attractive features [28, 29, 91, 92, 106, 110–113, 120]. Although the number of gadolinium atoms is very high even for the smaller size, the potential of crystalline nanoparticles based on inorganic gadolinium such as gadolinium oxide [121–123],
P1: OTA/XYZ P2: ABC c15 JWBS037-Chen January 25, 2011
23:25
Printer Name: Yet to Come
MULTIFUNCTIONAL NANOSTRUCTURES CONTAINING RARE EARTH ELEMENTS
361
gadolinium fluoride nanoparticles [124], and gadolinium carbonate [125, 126] has only recently been evaluated. Pioneering studies devoted to gadolinium oxide nanoparticles were performed by Roberts and Watkins. They revealed the potential of gadolinium oxide nanoparticles to be applied as contrast agents for MRI. However, these preliminary works suffered from the availability of well-defined hydrophilic gadolinium oxide nanoparticles. The exploitation of the attractive characteristics of gadolinium oxide nanoparticles for MRI became conceivable only when a reproducible and efficient synthesis protocol of naked gadolinium oxide nanoparticles and their functionalization were developed.
Synthesis and Functionalization of Gadolinium Oxide Nanoparticles The research activity of Tillement’s group was decisive in this field. First, a reproducible and reliable synthesis of gadolinium oxide nanoparticles was reported by this group [127, 128]. These particles were synthesized by applying, with some modifications, the polyol routes, which opened the door to the synthesis of a large range of inorganic crystalline nanoparticles. The synthesis of gadolinium oxide nanoparticles is based on the alkaline hydrolysis of gadolinium chloride in diethylene glycol (DEG), a high boiling point diol. Diethylene glycol plays an important role. It allows carrying out the synthesis at 180 ◦ C and to accurately control the size of gadolinium oxide nanoparticles (between 1 and 5 nm according to the experimental conditions) because the high viscosity of DEG and its adsorption on growing particles limit their growth and avoid agglomeration. The alkaline hydrolysis of a mixture of gadolinium chloride and of chloride salt of luminescent rare earth ions (Tb3+ , Eu3+ , Nd3+ , etc.) in boiling DEG provides rare earth (RE) doped gadolinium oxide nanoparticles whose Gd to RE molar ratio is very similar to the initial ratio [127, 128]. Since gadolinium oxide is transparent in the visible range of light, it was demonstrated that these nanoparticles are fluorescent when rare earth ions are present in the crystalline nanoparticles. Despite the remarkable optical properties of luminescent rare earth ions (high photostability, narrow emission peaks, long lifetime of excited state), the in vivo application of luminescent doped gadolinium oxide nanoparticles is impeded by the low quantum yield and by the energy of the excitation wavelength (UV range). However, the in vivo application of gadolinium oxide nanoparticles remains very promising for MRI provided that the surface of the gadolinium oxide core is adapted to a biological environment. The second important contribution of Tillement’s research group lies in the functionalization of these cores. The strategy developed by this group rests on the encapsulation of the gadolinium oxide cores in a polysiloxane shell [129]. This functionalization mode provides several important advantages. Besides a protective effect of the core, which was validated by the preservation of the luminescence in acidic media when the cores are doped by Tb3+ ions, the polysiloxane shell exerts a positive influence on the luminescence of rare earth ions present in the cores since its presence induces the replacement of RE OH bonds, which behaves as quenchers for the luminescence by the RE O Si bonds and an energy transfer from the polysiloxane shell to the luminescent centers [130]. The encapsulation of luminescent rare earth doped gadolinium oxide cores therefore yielded nano-objects that are more fluorescent than the uncoated gadolinium oxide cores are. But the most striking feature of the polysiloxane shell results from its chemical nature, which can provide new properties to the particles. The polysiloxane shell was indeed obtained by hydrolysis–condensation of tetraethyl orthosilicate (TEOS) and by aminopropyltriethoxysilane (APTES). The use of APTES allows the sequential functionalization of the polysiloxane shell: before and after the hydrolysis–condensation step, that is, before and after the formation of the polysiloxane shell because each APTES
P1: OTA/XYZ P2: ABC c15 JWBS037-Chen January 25, 2011
362
23:25
Printer Name: Yet to Come
NANOPARTICLES CONTAINING RARE EARTH IONS: A TUNABLE TOOL FOR MRI
carries an amino (NH2 ) function. Before the hydrolysis–condensation, APTES can be conjugated to organic dyes thanks to the coupling between the amino group of APTES and the isothiocyanate or NHS ester group of the organic fluorophores, while the polysiloxane shell can be postfunctionalized by the covalent grafting of hydrophilic PEG chains [131, 132] and/or of biotargeting groups [129, 133]. The encapsulation of gadolinium oxide cores enlarges the range of properties of these particles because the inner part of the polysiloxane shell is functionalized by organic dyes and the outer part is derivatized by hydrophilic molecules. As expected, these gadolinium oxide nanoparticles encapsulated in a fluorescent and PEGylated polysiloxane shell (GadoSiPEG) can be followed up after intravenous injection to mice and rats by fluorescence imaging (due to cyanine 5 covalently bound to the polysiloxane network) and by MRI (thanks to the presence of gadolinium(III) in the core of each particle). The in vivo imaging experiments revealed that the particles exhibit attractive biodistribution and pharmacokinetics parameters that are accurately controlled by the length of the PEG chains and the nature of the end group [131, 132]. They freely circulate in the bloodstream without undesirable nonspecific accumulation in liver, spleen, and lungs. Moreover, these nanoparticles are rather quickly removed from the body essentially by renal excretion. All these observations denote a safe behavior of GadoSiPEG nanoparticles when they are intravenously injected into small animals. However, the accumulation of these nanoparticles was observed after intravenous injection into gliosarcoma (9L) bearing rats. This was revealed by the obvious delineation of tumor due to the enhancement of the positive contrast induced by the presence of the particles in the tumor region. This accumulation, which can be explained by the EPR effect, was exploited for radiotherapy since the atomic number of the gadolinium element is sufficiently high for generating a dose enhancement of the X-ray beam. The treatment of brain tumor bearing rats by radiotherapy led to a longer survival only when hybrid gadolinium oxide nanoparticles were administered by intravenous injection before exposure to a therapeutic X-ray beam [134, 135]. The therapeutic activity of gadolinium oxide nanoparticles is not limited to radiotherapy. As widely mentioned in the literature, objects containing gadolinium could possibly replace the boron based compounds for neutron capture therapy. The internalization of GadoSiPEG nanoparticles in bioluminescent murine lymphoma cells (EL4-luc), which was monitored by fluorescence (rhodamine B isothiocyanate was covalently bound to the polysiloxane shell), MRI, and elemental analysis (ICP), generates no cytotoxicity and no alteration of the cell proliferation until [Gd]incubation = 0.3 mM. Unloaded cells are also not affected by exposure to a thermal neutron beam if the delivered dose is equal to 3 Gy (or less). But the irradiation of EL4-luc cells after incubation in the presence of GadoSiPEG ([Gd]incubation = 0.05 mM) by harmless thermal neutron beam (3 Gy) generates a great killing effect since all cells are destroyed [136]. Gadolinium oxide cores embedded in a fluorescent and PEGylated polysiloxane shell are very attractive multifunctional nanostructures for in vivo application and, in particular, for imaging guided therapy since they combine fluorescence imaging (organic dyes conjugated to the inner part of the polysiloxane shell), MRI (gadolinium(III) ions in the crystalline core), and therapeutic activity, which exploits some properties of the gadolinium element (radiotherapy and neutron therapy). The potential of gadolinium oxide cores for MRI was confirmed thanks to studies performed by a research group from Link¨oping University (Sweden) [121]. Their works is very interesting and complementary to the work done by Tillement’s group. Indeed, they propose an alternative way for derivatizing the gadolinium oxide cores and therefore for exploiting their paramagnetic character in MRI experiments. Their strategy consists in the capping of
P1: OTA/XYZ P2: ABC c15 JWBS037-Chen January 25, 2011
23:25
Printer Name: Yet to Come
MULTIFUNCTIONAL NANOSTRUCTURES CONTAINING RARE EARTH ELEMENTS
363
gadolinium oxide nanoparticles, which were synthesized according to the polyol route [127, 128], by organic acids or by the grafting of PEG-silane molecules. After their synthesis in DEG, the resulting gadolinium oxide cores are coated by DEG molecules, which can be replaced in hot DEG by citric acid or dimercaptosuccinic acid [137, 138]. This layer of organic acid molecules acts as a primer for further functionalization and, in particular, for the PEGylation, which is required for biological applications. The grafting of nonimmunogenic and nonantigenic PEG is expected to render the gadolinium oxide cores more robust, to improve the colloidal stability in biological media, to favor the uptake by cells, and to enhance blood retention. Citric-acid-coated gadolinium oxide cores were derivatized by PEG chains ended by a thiol (SH) group and an amino (NH2 ) group (Gd2 O3 @CA-PEG). The amino group of PEG permits condensation with carboxylic acid groups of citric acid, which were activated by EDC/NHS mixture prior to the coupling reaction. The thiol group at the opposite extremity can be used for the grafting of molecules bearing a maleimide group [3]. Since each dimercaptosuccinimide (DMSA) molecule grafted to gadolinium oxide cores carries two thiol functions, Mal-PEG-NHS (Mal and NHS for maleimide and N-hydroxysuccinimide groups, respectively) was used for the PEGylation of DMSA capped gadolinium oxide cores. The covalent coupling of PEG is ensured by the formation of a thioether linkage resulting from the condensation of the thiol group of DMSA with the maleimide moiety of Mal-PEG-NHS (Gd2 O3 @DMSA-PEG). The presence of the NHS group affords the possibility to graft aminated molecules (typically biomolecules that are rich in amine functions) onto the PEGylated nanoparticles. The PEGylation can also be performed using PEG-silane [139] since the alkoxysilyl end group of PEG-silane favors its grafting on oxide nanoparticles through condensation between the alkoxysilyl groups and hydroxyl (OH) groups present on the surface of the oxide materials (Gd2 O3 @PEG). These PEGylated oxide nanoparticles, which have been prepared by the research group from Link¨oping University, exhibit, as expected, relatively high longitudinal relaxivity r1 . The relaxivity r1 values of Gd2 O3 @PEG, Gd2 O3 @DMSA-PEG, and Gd2 O3 @CAPEG (based on the amount of gadolinium(III) ions in each sample) were 12.0, 14.2, and 7.8 mM−1 · s−1 , respectively [140]. Except for Gd2 O3 @CA-PEG, these PEGylated gadolinium oxide nanoparticles exhibit a higher relaxivity than that of GadoSiPEG developed by Tillement’s research group (see above). The longitudinal relaxivities of GadoSiPEG nanoparticles containing the smallest gadolinium oxide cores (2.2 nm and 3.8 nm) were 8.8 mM−1 · s−1 whatever the core size whereas r1 dropped to 4.4 mM−1 · s−1 when the size of the core amounted to 4.6 nm [131]. This difference can be explained by a better accessibility of water molecules to the surface of the oxide cores in the case of Gd2 O3 @PEG and Gd2 O3 @DMSA-PEG because the cores are encapsulated in a monolayer, whereas the gadolinium oxide cores of GadoSiPEG are embedded in a relatively thick polysiloxane layer. However, the cause of this difference is probably not so obvious since Gd2 O3 @CAPEG particles exhibit a lower relaxivity than GadoSiPEG particles do despite the thin PEG shell of Gd2 O3 @CA-PEG. It must be pointed out, however, that two different strategies for the functionalization of gadolinium oxide cores are available. They provide similar performance in terms of positive contrast enhancement for MRI. Unfortunately, their behavior after intravenous injection in small animals (rats, mice) cannot be compared because no in vivo imaging experiment was reported with the PEGylated organic acid capped gadolinium oxide cores in contrast to GadoSiPEG. Although the encapsulation of the gadolinium oxide cores by polysiloxane shell seems more difficult to control, it nevertheless allows enlarging more easily the palette of properties of the contrast agents since postfunctionalization of gadolinium oxide cores can be realized sequentially, in contrast to the encapsulation by
P1: OTA/XYZ P2: ABC c15 JWBS037-Chen January 25, 2011
364
23:25
Printer Name: Yet to Come
NANOPARTICLES CONTAINING RARE EARTH IONS: A TUNABLE TOOL FOR MRI
organic acids. Besides MR imaging, which is efficient whatever the functionalization mode, GadoSiPEG nanoparticles can indeed be easily followed up by fluorescence imaging after their uptake by cells or after their intravenous injection into small animals [131, 132, 136, 141], whereas the detection of PEGylated organic acid capped gadolinium oxide nanoparticles by fluorescence imaging rests only on the doping of the core by rare earth ions [140]. Despite its interesting properties, the fluorescence imaging based on the luminescence of rare earth ions (Tb3+ ) is not very well suited for biological applications because the light emission of Tb3+ in gadolinium oxide is induced by the absorption of UV photons, which are harmful for living matter. Moreover, the UV excitation is not suited for fluorescence imaging of small animals since the UV irradiation is almost totally absorbed by the skin. Because of the great chemical similarities of the rare earth elements, the synthesis of gadolinium oxide nanoparticles by applying the polyol route was easily adapted to the synthesis of dysprosium oxide nanoparticles (diameter: 1 nm). It was demonstrated that these particles exhibit a high transverse relaxivity r2 that is dependent on the size [142]. The optimal size for the highest r2 value is between 60 and 70 nm: r2 varies between 190 mM−1 · s−1 (70 nm at 7 T) and 675 mM−1 · s−1 (60 nm at 17.6 T). Other routes for the preparation of gadolinium oxide based contrast agents have recently been developed. Gadolinium oxide nanoparticles in carbon nanotubes were synthesized for MRI application [143]. They were obtained after thermal treatment of gadolinium acetate entrapped in single-wall carbon nanohorns (SWNHs) with holes (NHox). Gd2 O3 -NHox dispersed in agarose gel induces a significant positive contrast enhancement as revealed by T1-weighted images of phantoms. However, no r1 value was given, probably because the variation of 1/T1 as a function of gadolinium concentration is not linear. Although the enhancement induced by Gd2 O3 -NHox appears sufficient for MRI, their application as an in vivo contrast agent seems to be limited due to the lack of colloidal stability in biological fluid and the toxicity of carbon nanotubes. Yeh and co-workers succeeded in synthesizing porous and hollow gadolinium oxide particles [144]. They demonstrated that hollow spheres produced better positive contrast enhancement (r1 = 17.7 mM−1 · s−1 and r2 /r1 = 1.50) than porous particles did (r1 = 16.8 mM−1 · s−1 and r2 /r1 = 1.87). From a chemical point of view the elaboration of such materials is very interesting; however, their application as contrast agents for MRI can be impeded by the too large size of these particles. The sizes of hollow Gd2 O3 spheres (∼130 nm) and of porous Gd2 O3 particles (∼187 nm) are obviously larger than the hydrodynamic diameter recommended for the renal excretion of the nanoparticles in the bloodstream (Dh < 5 nm) according to the studies by Frangioni and co-workers [84].
Synthesis of Gadolinium Containing Crystalline Nanoparticles At the same time, several routes were explored for obtaining alternative crystalline nanomaterials rich in gadolinium(III) by wet chemistry techniques. The potential of these gadolinium containing crystalline nanoparticles for MRI was naturally evaluated since there is a real need for efficient multifunctional contrast agents. Among the variety of materials, fluoride and carbonate particles catch the attention. Fluoride Nanoparticles Containing Gadolinium(III) Van Veggel and colleagues acquired great experience in the synthesis of lanthanide fluoride nanoparticles for luminescence application. They paid particular attention to the synthesis and characterization of gadolinium fluoride. Synthesis of the nanoparticles was carried out for 3–4 hours in water at moderate temperature (75 ◦ C) and rests on the reaction of ammonium hydroxide
P1: OTA/XYZ P2: ABC c15 JWBS037-Chen January 25, 2011
23:25
Printer Name: Yet to Come
MULTIFUNCTIONAL NANOSTRUCTURES CONTAINING RARE EARTH ELEMENTS
365
(NH4 OH) with sodium fluoride and gadolinium(III) nitrate (Gd(NO3 )3 · 6H2 O) in the presence of a hydrophilic stabilizing agent (citric acid or 2-aminoethylphosphate (AEP)). As a result, relatively large gadolinium fluoride nanoparticles capped by citric acid or AEP (GdF3 @cit (mean diameter 129.3 nm) and GdF3 @AEP (mean diameter 51.5 nm), respectively) were collected [124]. The presence of hydrophilic agents on the surface of the particles facilitates their dispersion in aqueous solution and prevents agglomeration. Moreover, citric acid and especially AEP allow the postfunctionalization of the nanoparticles by a biotargeting group (biotin) and by PEG chains [145]. Whatever the capping ligands, GdF3 nanoparticles exhibit moderate longitudinal relaxivities (3.17 and 2.71 mM−1 · s−1 per Gd3+ for GdF3 @cit and GdF3 @AEP, respectively) when it was expressed in terms of the number of gadolinium ions in the nanoparticles. But the relaxivities per mM nanoparticles reach high values that amount to 2.0 × 107 and 8.8 × 105 s−1 · mM−1 for the GdF3 @cit and GdF3 @AEP nanoparticles, respectively. This study suggests that Gd3+ ions located beneath the outer shell of Gd3+ are also contribute to the relaxation. This was confirmed by work focused on GadoSiPEG, which showed that the effect on r1 not only results from the Gd3+ ions present at the surface of the crystalline core but also partially from those inside the core. Indeed, the relaxivities r1 are identical for 2.2- and 3.8-nm Gd2 O3 cores, whereas the smallest particles possess the highest amount of surface Gd3+ ions for the same gadolinium concentration. It means that in both samples all Gd3+ ions have the same influence whatever their location in the crystalline structure and that the range of the influence of Gd3+ ions is at least 1.9 nm. NaYF4 nanocrystals constitute another interesting class of fluoride materials since they can easily be doped by various rare earth ions. These fluoride nanocrystals received much attention for photoluminescence applications because they are considered the most efficient host materials for upconversion phosphors. They exhibit visible emission upon IR excitation when doped with Yb3+ and Er3+ /Tm3+ ions. By an appropriate choice of rare earth ions incorporated in the fluoride matrix, the emission wavelength can be red-shifted relative to the excitation wavelength. Downconversion is therefore achieved and can be exploited for biological imaging. For instance, Chatterjee et al. [146] have recently demonstrated the use of upconversion nanophosphors (NaYF4 : Er3+ , Yb3+ ) for in vitro imaging of cancer cells and in vivo imaging in tissues. It is claimed that the advantages of these nanoprobes include absence of photodamage to living organisms, very low autofluorescence, high detection sensitivity, and high light penetration depth in biological tissues. If many studies devoted to the photoluminescence of rare earth doped NaYF4 nanocrystals were performed, data about the doping of these crystals by gadolinium remains rare despite the great potential of contrast agents with multimodality imaging. Prasad and co-workers contributed to the development of paramagnetic fluoride nanophosphors of size <50 nm. They reported the synthesis of functionalized fluoride nanocrystals, which form a stable aqueous dispersion that can be used for both optical imaging and MRI. Down- and upconverted luminescence from the rare earth ions (Eu3+ , Er3+ , and Yb3+ ) doped into nanometer-sized fluoride matrices allows optical imaging modality, and co-doping with Gd3+ confers to the resulting nanoparticles the ability to enhance the positive contrast of MR images. By a convenient surface derivatization (immobilization of antibodies (anti-claudin and anti-mesothelin) via the covalent coupling with mercaptopropionic acid anchored onto the nanoparticles), these nanoparticles can be selectively taken up by pancreatic cancer cells. The selective internalization in cells was demonstrated only by the light emitted by the nanoparticles (using a confocal microscope). Since the longitudinal relaxivity r1 value of these Gd3+ co-doped fluoride nanocrystals is very low (r1 = 0.14 mM−1 · s−1 and r2 /r1 = 62 1), in comparison to gadolinium oxide
P1: OTA/XYZ P2: ABC c15 JWBS037-Chen January 25, 2011
366
23:25
Printer Name: Yet to Come
NANOPARTICLES CONTAINING RARE EARTH IONS: A TUNABLE TOOL FOR MRI
or gadolinium fluoride (GdF3 ) nanoparticles, the uptake of these nanoprobes cannot be monitored by MRI. The low value of r1 observed in the case of Gd3+ co-doped NaYF4 can be explained by the amount of gadolinium(III) ions in the crystalline fluoride matrix, which is largely lower than in the case of gadolinium oxide or gadolinium fluoride (GdF3 ) nanoparticles. If a small amount of luminescent rare earth ions is sufficient for detection by optical imaging techniques, this study showed clearly that a greater amount of paramagnetic ions is required for inducing positive contrast enhancement in MRI because this medical imaging is less sensitive. Nevertheless, MRI remains a more attractive in vivo imaging technique than fluorescence imaging techniques because the resolution of MRI is better. Moreover, a three-dimensional cartography of living bodies can be achieved by MRI, whereas it is impossible with fluorescence imaging techniques owing to the high absorption of biological tissues.
Carbonate Particles Containing Gadolinium(III) By refluxing aqueous solution containing urea and gadolinium chloride, amorphous gadolinium carbonate (Gd2 O(CO3 )2 · H2 O) particles can be obtained [126]. The shape (spherical, rhombus-, or rice-shaped nanoparticles) and the size can be tuned by varying the urea to gadolinium chloride molar ratio. Relaxation rate measurements revealed that the relaxivities depend on the size and on the shape of the particles. In the case of 500-nm sized spherical Gd2 O(CO3 )2 · H2 O particles, r1 is equal to the 16.5 mM−1 · s−1 (r2 /r1 = 12,7, 3 T). After injection, the behavior of these large particles can be monitored by MRI. As expected for large particles, accumulation in the liver was observed. These particles can be functionalized by an aminated polysiloxane shell thanks to the hydrolysis–condensation of the aforementioned mixture of TEOS and APTES. This polysiloxane shell permits a further functionalization. Due to the presence of the amino groups on the nanoparticles, gold nanoparticles can be immobilized since nitrogen atoms have a great affinity for gold atoms. These gold nanoparticles act as seeds for the growth of a gold layer. Gd2 O(CO3 )2 · H2 O particles can also be encapsulated by a gold shell whose thickness can be accurately controlled [125]. The encapsulation of these gadolinium carbonate particles by a gold shell induces a decrease of the longitudinal relaxivity with the increase of the thickness of the gold shell. However, the gold shell renders the resulting nanoparticles suitable for photothermal therapy [97]. GdO(CO3 )2 · H2 O particles encapsulated in a gold shell are able to combine MRI and therapy as revealed by in vitro experiments. However, the exploitation of this attractive feature requires the reduction of the size of this class of materials.
15.5 CONCLUSION AND OUTCOME Although many outstanding results have been achieved recently, the use of nanoparticles containing rare earth ion as MRI contrast agents is still in its infancy. Many of the studies are currently dealing with in vitro analysis or in vivo “proof-of-principles.” Deeper in vivo studies on relevant animal models and clinical trials will certainly give rise to many optimizations in the design, stability, and contrast enhancement of this new class of contrast agents. In addition to active and passive targeting for molecular imaging, many rare earth ion containing nanoparticles combining both imaging and therapeutic capacities are under
P1: OTA/XYZ P2: ABC c15 JWBS037-Chen January 25, 2011
23:25
Printer Name: Yet to Come
CONCLUSION AND OUTCOME
367
development. These multifunctional nanoparticles open new possibilities for “personalized nanomedicine,” where both diagnosis and treatment can be best fitted to every single patient. As the size range of currently designed nano-objects is in the window of the vascular system (>20 nm), they are mainly targeting vascular related modifications such as angiogenesis associated with tumor progression or atherosclerosis. With the development of smaller nanoparticles containing rare earth ions (<20 nm) that can extravasate from the endothelium to the interstitial space, tumor cell targeting can also be envisioned. Different strategies are currently under investigation to combine active targeting with therapy: (1) entrapping pharmaceutical compounds within the nano-objects to enable selective drug delivery, (2) taking advantage of unique physical properties of these nano-objects to perform selective therapy, or (3) combining both. As the use of lanthanide-based nanoparticles is relatively recent, their use in multifunctional nanoplatforms has undergone little investigation so far. Nevertheless, recently reported work is highly promising. For example, successful attempts to administer drugs locally have been reported for atherosclerosis treatment in a rabbit model using perfluorocarbon nanoparticles (175–220 nm) loaded with Gd ions as the contrast agent and fumagillin as the drug [148]. Other highly promising strategies with minimum side effects rely on the exploitation of the unique physical properties of these nanosized carriers for efficient and selective radiotherapy, phototherapy, or neutron therapy. Zielhuis and colleagues [149] have successfully designed liposomes (130 nm) combining paramagnetic gadolinium ions and radionuclide rare earth ions holmium-166 (beta- and gamma-emitter with highly paramagnetic properties) for both imaging and radiotherapy. Several recent works have also demonstrated the possibility to combine imaging and phototherapy, using rare earth containing nanoparticles with absorbing nanomaterials such as gold [105, 125] or to a lesser extent carbon [150]. The principle relies in the conversion of absorbed light into local heat strong enough to damage nanoparticle containing cells. This is possible thanks to the strong light-absorption properties of nanomaterials (in particular, gold nanomaterials), known as surface plasmon resonance (SPR). The nano-objects are classically designed to tune the SPR peak in the near-infrared wavelength region, where the depth penetration into tissues is higher. The advantages of such a technique are minimum invasivity and an increase in selectivity, as damages are only done in the vicinity of the nano-objects, thus reducing severe injury of healthy tissue. So far, reported studies have used relatively large nanoparticles (50–450 nm [105, 125, 150]), limiting their use for local injection. There is no doubt, however, that these pioneering works will lead to many innovative nano-objects more suitable for cell targeting after intravenous injection in clinical applications. Another attractive feature of gold lies in its high atomic number. As a result, gold absorbs very strongly the X-ray photons. Gold nanoparticles therefore appear promising for use in radiotherapy since their presence in tumor permits the absorption of almost all X-ray photons. Consequently, healthy tissue is more preserved while the tumoricide effect the of X-ray beam is enhanced [99, 100, 103]. Last but not least, neutron therapy appears as an interesting alternative technique that has hardly been exploited. The principle relies on the irradiation of compounds with a high neutron capture cross section (neutrophages) by a thermal neutron beam. The irradiation by a thermal neutron beam or the presence of neutrophage compounds is harmless by itself. Selective destruction of surrounding cells can be achieved by combining nanoparticle targeting with thermal neutron beam irradiation. Most therapeutic protocols are now based on
P1: OTA/XYZ P2: ABC c15 JWBS037-Chen January 25, 2011
368
23:25
Printer Name: Yet to Come
NANOPARTICLES CONTAINING RARE EARTH IONS: A TUNABLE TOOL FOR MRI
molecular boron compounds [151]. However, different groups have reported the possibility to take advantage of ␥ -ray emission and Auger electrons of gadolinium isotopes (155 Gd and 157 Gd) upon interaction with a thermal neutron [136, 152–155]. In particular, the in vitro study of Bridot and colleagues showed the therapeutic potential of hybrid nanoparticles combining high contrast properties and long blood half-life [131] and a size (7 nm in diameter) compatible with partial extravasation from the vasculature and cell targeting. Multifunctional rare earth based nanoparticles are far from clinical use, but different routes to achieve that goal are under investigation by many multidisciplinary joint research programs, starting from chemical design and synthesis to in vitro and in vivo toxicity and therapeutic efficacy analysis. The aim is to be able to get a diagnosis as early as possible and to treat the pathology with high efficacy and limited side effects. Rare earth based multifunctional nanoplatforms are one possible route to achieve high contrast properties (and thus lower the detection sensitivity threshold) as well as high specificity toward the targeted pathological sites. The near future will certainly see the design of many innovative rare earth based multifunctional nanoplatforms for early diagnosis and treatment, most presumably in oncology and atherosclerosis diseases.
REFERENCES 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. 14. 15. 16.
17. 18.
Caravan, P.; Ellison, J. J.; McMurry, T. J.; Lauffer, R. B. Chem. Rev. 1999, 99, 2293–2352. Bottrill, M.; Kwok, L.; Long, N. J. Chem. Soc. Rev. 2006, 35, 557–571. Hermanson, G. T. Bioconjugate Techniques, 2nd ed.; Academic Press: Amsterdam, 2008. Nwe, K.; Xu, H.; Regino, C. A.; Bernardo, M.; Ileva, L.; Riffle, L.; Wong, K. J.; Brechbiel, M. W. Bioconjug. Chem. 2009, 20, 1412–1418. Izard, M. E.; Boniface, G. R.; Hardiman, K. L.; Brechbiel, M. W.; Gansow, O. A.; Walkers, K. Z. Bioconjug. Chem. 1992, 3, 346–350. Luciani, A.; Olivier, J. C.; Clement, O.; Siauve, N.; Brillet, P. Y.; Bessoud, B.; Gazeau, F.; Uchegbu, I. F.; Kahn, E.; Frija, G.; Cuenod, C. A. Radiology 2004, 231, 135–142. Torchilin, V. P. Curr. Pharm. Biotechnol. 2000, 1, 183–215. Tilcock, C. Adv. Drug. Deliv. Rev. 1999, 37, 33–51. Weinmann, H. J.; Ebert, W.; Misselwitz, B.; Schmitt-Willich, H. Eur. J. Radiol. 2003, 46, 33–44. Martina, M.-S.; Fortin, J.-P.; M´enager, C.; Clement, O.; Barratt, G.; Grabielle-Madelmont, C.; Gazeau, F.; Cabuil, V.; Lesieur, S. J. Am. Chem. Soc. 2005, 127, 10676–10685. Lasic, D. D.; Templeton, N. S. Adv. Drug. Deliv. Rev. 1996, 20, 221–266. Lasic, D. D. Trends Biotechnol. 1998, 16, 307–321. Marjan, J. M.; Allen, T. M. Biotechnol. Adv. 1996, 14, 151–175. Allen, T. M.; Hansen, C.; Martin, F.; Redemann, C.; Yau-Young, A. Biochim. Biophys. Acta 1991, 1066, 29–36. Needham, D.; Hristovan, K.; McIntosh, T. J.; Dewhirst, M. W.; Wu, N.; Lasic, D. D. J. Liposome Res. 1992, 2, 411–430. Papahadjopoulos, D.; Allen, T.; Gabizon, A.; Mayhew, E.; Matthay, K.; Huang, S.; Lee, K.; Woodle, M.; Lasic, D.; Redemann, C.; Martin, F. Proc. Natl. Acad. Sci. U.S.A. 1991, 88, 11460–11464. Tilcock, C.; Unger, E.; Cullis, P.; MacDougall, P. Radiology 1989, 171, 77–80. Ayyagari, A. L.; Zhang, X.; Ghaghada, K. B.; Annapragada, A.; Hu, X.; Bellamkonda, R. V. Magn. Reson. Med. 2006, 55, 1023–1029.
P1: OTA/XYZ P2: ABC c15 JWBS037-Chen January 25, 2011
23:25
Printer Name: Yet to Come
REFERENCES
369
19. Cheng, Z.; Tsourkas, A. Langmuir 2008, 24, 8169–8173. 20. Kabalka, G. W.; Davis, M. A.; Holmberg, E.; Maruyama, K.; Huang, L. Magn. Reson. Imaging 1991, 9, 373–377. 21. Weissig, V. V.; Babich, J.; Torchilin, V. P. Colloids Surf. B Biointerfaces 2000, 18, 293–299. 22. Erdogan, S.; Medarova, Z. O.; Roby, A.; Moore, A.; Torchilin, V. P. J. Magn. Reson. Imaging 2008, 27, 574–580. 23. Laurent, S.; Vander Elst, L.; Thirifays, C.; Muller, R. Eur. Biophys. J. 2008, 37, 1007–1014. 24. Trubetskoy, V. S.; Cannillo, J. A.; Milshtein, A.; Wolf, G. L.; Torchilin, V. P. Magn. Reson. Imaging 1995, 13, 31–37. 25. Bertini, I.; Bianchini, F.; Calorini, L.; Colagrande, S.; Fragai, M.; Franchi, A.; Gallo, O.; Gavazzi, C.; Luchinat, C. Magn. Reson. Med. 2004, 52, 669–672. 26. Cormode, D. P.; Briley-Saebo, K. C.; Mulder, W. J.; Aguinaldo, J. G.; Barazza, A.; Ma, Y.; Fisher, E. A.; Fayad, Z. A. Small 2008, 4, 1437–1444. 27. Cormode, D. P.; Skajaa, T.; van Schooneveld, M. M.; Koole, R.; Jarzyna, P.; Lobatto, M. E.; Calcagno, C.; Barazza, A.; Gordon, R. E.; Zanzonico, P.; Fisher, E. A.; Fayad, Z. A.; Mulder, W. J. Nano Lett. 2008, 8, 3715–3723. 28. Frias, J. C.; Williams, K. J.; Fisher, E. A.; Fayad, Z. A. J. Am. Chem. Soc. 2004, 126, 16316–16317. 29. Frias, J. C.; Ma, Y.; Williams, K. J.; Fayad, Z. A.; Fisher, E. A. Nano Lett. 2006, 6, 2220–2224. 30. Lokling, K. E.; Fossheim, S. L.; Klaveness, J.; Skurtveit, R. J. Control. Release 2004, 98, 87–95. 31. Lokling, K. E.; Skurtveit, R.; Dyrstad, K.; Klaveness, J.; Fossheim, S. L. Int. J. Pharm. 2004, 274, 75–83. 32. Lokling, K. E.; Skurtveit, R.; Bjornerud, A.; Fossheim, S. L. Magn. Reson. Med. 2004, 51, 688–696. 33. Bolskar, R. D.; Benedetto, A. F.; Husebo, L. O.; Price, R. E.; Jackson, E. F.; Wallace, S.; Wilson, L. J.; Alford, J. M. J. Am. Chem. Soc. 2003, 125, 5471–5478. 34. Mikawa, M.; Kato, H.; Okumura, M.; Narazaki, M.; Kanazawa, Y.; Miwa, N.; Shinohara, H. Bioconjug. Chem. 2001, 12, 510–514. 35. Toth, E.; Bolskar, R. D.; Borel, A.; Gonzalez, G.; Helm, L.; Merbach, A. E.; Sitharaman, B.; Wilson, L. J. J. Am. Chem. Soc. 2005, 127, 799–805. 36. Kato, H.; Suenaga, K.; Mikawa, W.; Okumura, M.; Miwa, N.; Yashiro, A.; Fujimura, H.; Mizuno, A.; Nishida, Y.; Kobayashi, K.; Shinohara, H. Chem. Phys. Lett. 2000, 324, 255–259. 37. Kato, H.; Kanazawa, Y.; Okumura, M.; Taninaka, A.; Yokawa, T.; Shinohara, H. J. Am. Chem. Soc. 2003, 125, 4391–4397. 38. Sitharaman, B.; Bolskar, R. D.; Rusakova, I.; Wilson, L. J. Nano Lett. 2004, 4, 2373–2378. 39. Sitharaman, B.; Kissell, K. R.; Hartman, K. B.; Tran, L. A.; Baikalov, A.; Rusakova, I.; Sun, Y.; Khant, H. A.; Ludtke, S. J.; Chiu, W.; Laus, S.; Toth, E.; Helm, L.; Merbach, A. E.; Wilson, L. J. Chem. Commun. (Camb.) 2005, 3915–3917. 40. Kam, N. W.; O’Connell, M.; Wisdom, J. A.; Dai, H. Proc. Natl. Acad. Sci. U.S.A. 2005, 102, 11600–11605. 41. Kam, N. W.; Liu, Z.; Dai, H. J. Am. Chem. Soc. 2005, 127, 12492–12493. 42. Kam, N. W.; Dai, H. J. Am. Chem. Soc. 2005, 127, 6021–6026. 43. Hummer, G.; Rasaiah, J. C.; Noworyta, J. P. Nature 2001, 414, 188–190. 44. Hartman, K. B.; Laus, S.; Bolskar, R. D.; Muthupillai, R.; Helm, L.; Toth, E.; Merbach, A. E.; Wilson, L. J. Nano Lett. 2008, 8, 415–419. 45. Huang, D. M.; Hung, Y.; Ko, B. S.; Hsu, S. C.; Chen, W. H.; Chien, C. L.; Tsai, C. P.; Kuo, C. T.; Kang, J. C.; Yang, C. S.; Mou, C. Y.; Chen, Y. C. FASEB J. 2005, 19, 2014–2016.
P1: OTA/XYZ P2: ABC c15 JWBS037-Chen January 25, 2011
370
23:25
Printer Name: Yet to Come
NANOPARTICLES CONTAINING RARE EARTH IONS: A TUNABLE TOOL FOR MRI
46. Chung, T. H.; Wu, S. H.; Yao, M.; Lu, C. W.; Lin, Y. S.; Hung, Y.; Mou, C. Y.; Chen, Y. C.; Huang, D. M. Biomaterials 2007, 28, 2959–2966. 47. Taylor, K. M.; Kim, J. S.; Rieter, W. J.; An, H.; Lin, W. J. Am. Chem. Soc. 2008, 130, 2154–2155. 48. Tsai, C. P.; Hung, Y.; Chou, Y. H.; Huang, D. M.; Hsiao, J. K.; Chang, C.; Chen, Y. C.; Mou, C. Y. Small 2008, 4, 186–191. 49. Hsiao, J. K.; Tsai, C. P.; Chung, T. H.; Hung, Y.; Yao, M.; Liu, H. M.; Mou, C. Y.; Yang, C. S.; Chen, Y. C.; Huang, D. M. Small 2008, 4, 1445–1452. 50. Carniato, F.; Tei, L.; Dastru, W.; Marchese, L.; Botta, M. Chem. Commun. 2009, 1246–1248. 51. Rieter, W. J.; Kim, J. S.; Taylor, K. M.; An, H.; Lin, W.; Tarrant, T. Angew. Chem. Int. Ed. Engl. 2007, 46, 3680–3682. 52. Kim, J. S.; Rieter, W. J.; Taylor, K. M.; An, H.; Lin, W. J. Am. Chem. Soc. 2007, 129, 8962– 8963. 53. Lin, Y. S.; Hung, Y.; Su, J. K.; Lee, R.; Chang, C.; Lin, M. L.; Mou, C. Y. J. Phys. Chem. B 2004, 108, 15608–15611. 54. Lai, C. Y.; Trewyn, B. G.; Jeftinija, D. M.; Jeftinija, K.; Xu, S.; Jeftinija, S.; Lin, V. S. J. Am. Chem. Soc. 2003, 125, 4451–4459. 55. Lu, J.; Liong, M.; Zink, J. I.; Tamanoi, F. Small 2007, 3, 1341–1346. 56. Lu, F.; Wu, S. H.; Hung, Y.; Mou, C. Y. Small 2009, 5, 1408–1413. 57. Daldrup-Link, H. E.; Rudelius, M.; Oostendorp, R. A.; Settles, M.; Piontek, G.; Metz, S.; Rosenbrock, H.; Keller, U.; Heinzmann, U.; Rummeny, E. J.; Schlegel, J.; Link, T. M. Radiology 2003, 228, 760–767. 58. Crich, S. G.; Biancone, L.; Cantaluppi, V.; Duo, D.; Esposito, G.; Russo, S.; Camussi, G.; Aime, S. Magn. Reson. Med. 2004, 51, 938–944. 59. Lewin, M.; Carlesso, N.; Tung, C. H.; Tang, X. W.; Cory, D.; Scadden, D. T.; Weissleder, R. Nat. Biotechnol. 2000, 18, 410–414. 60. Nahrendorf, M.; Zhang, H.; Hembrador, S.; Panizzi, P.; Sosnovik, D. E.; Aikawa, E.; Libby, P.; Swirski, F. K.; Weissleder, R. Circulation 2008, 117, 379–387. 61. Weissleder, R. Science 2006, 312, 1168–1171. 62. Cai, W.; Chen, X. Small 2007, 3, 1840–1854. 63. Ferrari, M. Nat. Rev. Cancer 2005, 5, 161–171. 64. Riehemann, K.; Schneider, S. W.; Luger, T. A.; Godin, B.; Ferrari, M.; Fuchs, H. Angew. Chem. Int. Ed. Engl. 2009, 48, 872–897. 65. St¨ober, W.; Fink, A.; Bohn, E. J. Colloid Interface Sci. 1968, 26, 62–69. 66. Bagwe, R. P.; Hilliard, L. R.; Tan, W. Langmuir 2006, 22, 4357–4362. 67. Van Blaaderen, A.; Vrij, A. Langmuir 1992, 8, 2921–2931. 68. Ow, H.; Larson, D. R.; Srivastava, M.; Baird, B. A.; Webb, W. W.; Wiesner, U. Nano Lett. 2005, 5, 113–117. 69. Burns, A. A.; Vider, J.; Ow, H.; Herz, E.; Penate-Medina, O.; Baumgart, M.; Larson, S. M.; Wiesner, U.; Bradbury, M. Nano Lett. 2009, 9, 442–448. 70. Zhao, X.; Tapec-Dytioco, R.; Tan, W. J. Am. Chem. Soc. 2003, 125, 11474–11475. 71. Wang, L.; Tan, W. Nano Lett. 2006, 6, 84–88. 72. Santra, S.; Bagwe, R. P.; Dutta, D.; Stanley, J. T.; Walter, G. A.; Tan, W.; Moudgil, B. M.; Mericle, R. A. Adv. Mater. 2005, 17, 2165–2169. 73. Wu, C.; Hong, J.; Guo, X.; Huang, C.; Lai, J.; Zheng, J.; Chen, J.; Mu, X.; Zhao, Y. Chem. Commun. (Camb.) 2008, 750–752. 74. Delgado-Pinar, E.; Frias, J. C.; Jimenez-Borreguero, L. J.; Albelda, M. T.; Alarcon, J.; GarciaEspana, E. Chem. Commun. (Camb.) 2007, 3392–3394.
P1: OTA/XYZ P2: ABC c15 JWBS037-Chen January 25, 2011
23:25
Printer Name: Yet to Come
REFERENCES
371
75. Voisin, P.; Ribot, E. J.; Miraux, S.; Bouzier-Sore, A.-K.; Lahitte, J.-F.; Bouchaud, V.; Mornet, S.; Thiaudiere, E.; Franconi, J.-M.; Raison, L.; Labrugere, C.; Delville, M.-H. Bioconjug. Chem. 2007, 18, 1053–1063. 76. Endres, P. J.; Paunesku, T.; Vogt, S.; Meade, T. J.; Woloschak, G. E. J. Am. Chem. Soc. 2007, 129, 15760–15761. 77. Rehor, I.; Kubicek, V.; Kotek, J.; Hermann, P.; Lukes, I.; Szakova, J.; Elst, L. V.; Muller, R. N.; Peters, J. A. J. Mater. Chem. 2009, 19, 1494–1500. 78. Bruchez, M., Jr.; oronne, M.; Gin, P.; Weiss, S.; Alivisatos, A. P. Science 1998, 281, 2013–2016. 79. Chan, W. C.; Nie, S. Science 1998, 281, 2016–2018. 80. Maxwell, D. J.; Taylor, J. R.; Nie, S. J. Am. Chem. Soc. 2002, 124, 9606–9612. 81. Derfus, A. M.; Chan, W. C. W.; Bhatia, S. N. Nano Lett. 2004, 4, 11–18. 82. Lewinski, N.; Colvin, V.; Drezek, R. Small 2008, 4, 26–49. 83. Kobayashi, H.; Hama, Y.; Koyama, Y.; Barrett, T.; Regino, C. A.; Urano, Y.; Choyke, P. L. Nano Lett. 2007, 7, 1711–1716. 84. Choi, H. S.; Liu, W.; Misra, P.; Tanaka, E.; Zimmer, J. P.; Itty Ipe, B.; Bawendi, M. G.; Frangioni, J. V. Nat. Biotechnol. 2007, 25, 1165–1170. 85. Roullier, V.; Clarke, S.; You, C.; Pinaud, F.; Gouzer, G. G.; Schaible, D.; Marchi-Artzner, V.; Piehler, J.; Dahan, M. Nano Lett. 2009, 9, 1228–1234. 86. Medintz, I. L.; Uyeda, H. T.; Goldman, E. R.; Mattoussi, H. Nat. Mater. 2005, 4, 435–446. 87. Jin, T.; Yoshioka, Y.; Fujii, F.; Komai, Y.; Seki, J.; Seiyama, A. Chem. Commun. (Camb.) 2008, 5764–5766. 88. Gerion, D.; Herberg, J.; Bok, R.; Gjersing, E.; Ramon, E.; Maxwell, R.; Kurhanewicz, J.; Budinger, T. F.; Gray, J. W.; Shuman, M. A.; Chen, F. F. J. Phys. Chem. C 2007, 111, 12542–12551. 89. Bakalova, R.; Zhelev, Z.; Aoki, I.; Masamoto, K.; Mileva, M.; Obata, T.; Higuchi, M.; Gadjeva, V.; Kanno, I. Bioconjug. Chem. 2008, 19, 1135–1142. 90. Koole, R.; van Schooneveld, M. M.; Hilhorst, J.; Castermans, K.; Cormode, D. P.; Strijkers, G. J.; de Mello Donega, C.; Vanmaekelbergh, D.; Griffioen, A. W.; Nicolay, K.; Fayad, Z. A.; Meijerink, A.; Mulder, W. J. Bioconjug. Chem. 2008, 19, 2471–2479. 91. Mulder, W. J.; Koole, R.; Brandwijk, R. J.; Storm, G.; Chin, P. T.; Strijkers, G. J.; de Mello Donega, C.; Nicolay, K.; Griffioen, A. W. Nano Lett. 2006, 6, 1–6. 92. van Tilborg, G. A.; Mulder, W. J.; Chin, P. T.; Storm, G.; Reutelingsperger, C. P.; Nicolay, K.; Strijkers, G. J. Bioconjug. Chem. 2006, 17, 865–868. 93. Mulder, W. J.; Strijkers, G. J.; van Tilborg, G. A.; Cormode, D. P.; Fayad, Z. A.; Nicolay, K. Acc. Chem. Res. 2009, 42, 904–914. 94. van Schooneveld, M. M.; Vucic, E.; Koole, R.; Zhou, Y.; Stocks, J.; Cormode, D. P.; Tang, C. Y.; Gordon, R. E.; Nicolay, K.; Meijerink, A.; Fayad, Z. A.; Mulder, W. J. Nano Lett. 2008, 8, 2517–2525. 95. Boisselier, E.; Astruc, D. Chem. Soc. Rev. 2009, 38, 1759–1782. 96. Ghosh, S. K.; Pal, T. Chem. Rev. 2007, 107, 4797–4862. 97. Lal, S.; Clare, S. E.; Halas, N. J. Acc. Chem. Res. 2008, 41, 1842–1851. 98. Jain, P. K.; El-Sayed, I. H.; El-Sayed, M. A. Nanotoday 2007, 2, 18–29. 99. Hainfeld, J. F.; Slatkin, D. N.; Focella, T. M.; Smilowitz, H. M. Br. J. Radiol. 2006, 79, 248–253. 100. Hainfeld, J. F.; Slatkin, D. N.; Smilowitz, H. M. Phys. Med. Biol. 2004, 49, N309–315. 101. Deboutti`ere, P. J.; Roux, S.; Vocanson, F.; Billotey, C.; Beuf, O.; Favre-R´eguillon, A.; Lin, Y.; Pellet-Rostaing, S.; Lamartine, R.; Perriat, P.; Tillement, O. Adv. Funct. Mater. 2006, 16, 2330–2339.
P1: OTA/XYZ P2: ABC c15 JWBS037-Chen January 25, 2011
372
23:25
Printer Name: Yet to Come
NANOPARTICLES CONTAINING RARE EARTH IONS: A TUNABLE TOOL FOR MRI
102. Alric, C.; Taleb, J.; Le Duc, G.; Mandon, C.; Billotey, C.; Le Meur-Herland, A.; Brochard, T.; Vocanson, F.; Janier, M.; Perriat, P.; Roux, S.; Tillement, O. J. Am. Chem. Soc. 2008, 130, 5908–5915. 103. Alric, C.; Serduc, R.; Mandon, C.; Taleb, J.; Le Duc, G.; A., L. M.-H.; Billotey, C.; Perriat, P.; Roux, S. O., T. Gold Bull. 2008, 41, 90–97. 104. Massue, J.; Quinn, S. J.; Gunnlaugsson, T. J. Am. Chem. Soc. 2008, 130, 6900–6901. 105. Lim, Y. T.; Cho, M. Y.; Choi, B. S.; Lee, J. M.; Chung, B. H. Chem. Commun. (Camb.) 2008, 4930–4932. 106. Huber, M. M.; Staubli, A. B.; Kustedjo, K.; Gray, M. H.; Shih, J.; Fraser, S. E.; Jacobs, R. E.; Meade, T. J. Bioconjug. Chem. 1998, 9, 242–249. 107. Yan, G. P.; Liu, M. L.; Li, L. Y. Bioconjug. Chem. 2005, 16, 967–971. 108. Bull, S. R.; Guler, M. O.; Bras, R. E.; Venkatasubramanian, P. N.; Stupp, S. I.; Meade, T. J. Bioconjug. Chem. 2005, 16, 1343–1348. 109. Bull, S. R.; Guler, M. O.; Bras, R. E.; Meade, T. J.; Stupp, S. I. Nano Lett. 2005, 5, 1–4. 110. Anderson, E. A.; Isaacman, S.; Peabody, D. S.; Wang, E. Y.; Canary, J. W.; Kirshenbaum, K. Nano Lett. 2006, 6, 1160–1164. 111. Langereis, S.; de Lussanet, Q. G.; van Genderen, M. H. P.; Backes, W. H.; Meijer, E. W. Macromolecules 2004, 37, 3084–3091. 112. Talanov, V. S.; Regino, C. A.; Kobayashi, H.; Bernardo, M.; Choyke, P. L.; Brechbiel, M. W. Nano Lett. 2006, 6, 1459–1463. 113. Mulder, W. J.; Strijkers, G. J.; Griffioen, A. W.; van Bloois, L.; Molema, G.; Storm, G.; Koning, G. A.; Nicolay, K. Bioconjug. Chem. 2004, 15, 799–806. 114. Accardo, A.; Tesauro, D.; Roscigno, P.; Gianolio, E.; Paduano, L.; D’Errico, G.; Pedone, C.; Morelli, G. J. Am. Chem. Soc. 2004, 126, 3097–3107. 115. Parac-Vogt, T. N.; Kimpe, K.; Laurent, S.; Pi´erart, C.; Vander Elst, L.; Muller, R. N.; Binnemans, K. Eur. J. Inorg. Chem. 2004, 17, 3538–3543. 116. Reynolds, C. H.; Annan, N.; Beshah, K.; Huber, J. H.; Shaber, S. H.; Lenkinski, R. E.; Wortman, J. A. J. Am. Chem. Soc. 2000, 122, 8940–8945. 117. Turner, J. L.; Pan, D.; Plummer, R.; Chen, Z.; Whittaker, A. K.; Wooley, K. L. Adv. Funct. Mater. 2005, 15, 1248–1254. 118. Platas-Iglesias, C.; Vander Elst, L.; Zhou, W.; Muller, R. N.; Geraldes, C. F.; Maschmeyer, T.; Peters, J. A. Chemistry 2002, 8, 5121–5131. 119. Balkus, K. J.; Shi, J. Langmuir 1996, 12, 6277–6281. 120. Vuu, K.; Xie, J.; McDonald, M. A.; Bernardo, M.; Hunter, F.; Zhang, Y.; Li, K.; Bednarski, M.; Guccione, S. Bioconjug. Chem. 2005, 16, 995–999. 121. Engstrom, M.; Klasson, A.; Pedersen, H.; Vahlberg, C.; Kall, P. O.; Uvdal, K. Magn. Reson. Mater. Phy. 2006, 19, 180–186. 122. McDonald, M. A.; Watkin, K. L. Invest. Radiol. 2003, 38, 305–310. 123. Roberts, D.; Zhu, W. L.; Frommen, C. M.; Rosenzweig, Z. J. Appl. Phys. 2000, 87, 6208–6210. 124. Evanics, F.; Diamente, P. R.; van Veggel, F.; Stanisz, G. J.; Prosser, R. S. Chem. Mater. 2006, 18, 2499–2505. 125. Hu, K.-W.; Jhang, F.-Y.; Su, C.-H.; Yeh, C.-S. J. Mater. Chem. 2009, 19, 2147–2153. 126. Li, I. F.; Su, C. H.; Sheu, H. S.; Chiu, H. C.; Lo, Y. W.; Lin, W. T.; Chen, J. H.; Yeh, C. S. Adv. Funct. Mater. 2008, 18, 766–776. 127. Bazzi, R.; Flores, M. A.; Louis, C.; Lebbou, K.; Zhang, W.; Dujardin, C.; Roux, S.; Mercier, B.; Ledoux, G.; Bernstein, E.; Perriat, P.; Tillement, O. J. Colloid Interface Sci. 2004, 273, 191–197.
P1: OTA/XYZ P2: ABC c15 JWBS037-Chen January 25, 2011
23:25
Printer Name: Yet to Come
REFERENCES
373
128. Bazzi, R.; Flores-Gonzalez, M. A.; Louis, C.; Lebbou, K.; Dujardin, C.; Brenier, A.; Zhang, W.; Tillement, O.; Bernstein, E.; Perriat, P. J. Lumin. 2003, 102–103, 445–450. 129. Louis, C.; Bazzi, R.; Marquette, C. A.; Bridot, J.-L.; Roux, S.; Ledoux, G.; Mercier, B.; Blum, L.; Perriat, P.; Tillement, O. Chem. Mater. 2005, 17, 1673–1682. 130. Louis, C.; Roux, S.; Ledoux, G.; Dujardin, C.; Tillement, O.; Cheng, B. L.; Perriat, P. Chem. Phys. Lett. 2006, 429, 157–160. 131. Bridot, J. L.; Faure, A. C.; Laurent, S.; Riviere, C.; Billotey, C.; Hiba, B.; Janier, M.; Josserand, V.; Coll, J. L.; Elst, L. V.; Muller, R.; Roux, S.; Perriat, P.; Tillement, O. J. Am. Chem. Soc. 2007, 129, 5076–5084. 132. Faure, A. C.; Dufort, S.; Josserand, V.; Perriat, P.; Coll, J. L.; Roux, S.; Tillement, O. Small 2009, 5, 2565–2575. 133. Faure, A. C.; Hoffmann, C.; Bazzi, R.; Goubard, F.; Pauthe, E.; Marquette, C. A.; Blum, L. J.; Perriat, P.; Roux, S.; Tillement, O. ACS Nano. 2008, 2, 2273–2282. 134. Le Duc, G.; Roux, S.; Tillement, O.; Perriat, P.; Billotey, C.; Mandon, C.; Mutelet, B.; Alric, C.; Janier, M.; Louis, C. In Brevet franc¸ais; 07.58348: France, 16/10/2007. 135. Roux, S.; Tillement, O.; Billotey, C.; Coll, J. L.; Le Duc, G.; Marquette, C. A.; Perriat, P. Int. J. Nanotechnol. 2010, 7, 781–801. 136. Bridot, J.-L.; Dayde, D.; Riviere, C.; Mandon, C.; Billotey, C.; Lerondel, S.; Sabattier, R.; Cartron, G.; Pape, A. L.; Blondiaux, G.; Janier, M.; Perriat, P.; Roux, S.; Tillement, O. J. Mater. Chem. 2009, 19, 2328–2335. 137. Soderlind, F.; Pedersen, H.; Petoral, R. M., Jr.; Kall, P. O.; Uvdal, K. J. Colloid Interface Sci. 2005, 288, 140–148. 138. Pedersen, H.; Soderlind, F.; Petoral, R. M.; Uvdal, K.; Kall, P. O.; Ojamae, L. Surf. Sci. 2005, 592, 124–140. 139. Fortin, M. A.; Petoral, R. M.; Soderlind, F.; Klasson, A.; Engstrom, M.; Veres, T.; Kall, P. O.; Uvdal, K. Nanotechnology 2007, 18, 395501. 140. Petoral, R. M.; Soderlind, F.; Klasson, A.; Suska, A.; Fortin, M. A.; Abrikossova, N.; Selegard, L.; Kall, P.-O.; Engstrom, M.; Uvdal, K. J. Phys. Chem. C 2009, 113, 6913–6920. 141. Fizet, J.; Riviere, C.; Bridot, J. L.; Charvet, N.; Louis, C.; Billotey, C.; Raccurt, M.; Morel, G.; Roux, S.; Perriat, P.; Tillement, O. J. Nanosci. Nanotechnol. 2009, 9, 1–9. 142. Norek, M.; Kampert, E.; Zeitler, U.; Peters, J. A. J. Am. Chem. Soc. 2008, 130, 5335–5340. 143. Miyawaki, J.; Yudasaka, M.; Imai, H.; Yorimitsu, H.; Isobe, H.; Nakamura, E.; Iijima, S. J. Phys. Chem. B 2006, 110, 5179–5181. 144. Huang, C. C.; Liu, T. Y.; Su, C. H.; Lo, Y. W.; Chen, J. H.; Yeh, C. S. Chem. Mat. 2008, 20, 3840–3848. 145. Diamente, P. R.; Burke, R. D.; van Veggel, F. C. Langmuir 2006, 22, 1782–1788. 146. Chatterjee, D. K.; Rufaihah, A. J.; Zhang, Y. Biomaterials 2008, 29, 937–943. 147. Kumar, R.; Nyk, M.; Ohulchanskyy, T. Y.; Flask, C. A.; Prasad, P. N. Adv. Funct. Mater. 2009, 19, 853–859. 148. Winter, P. M.; Neubauer, A. M.; Caruthers, S. D.; Harris, T. D.; Robertson, J. D.; Williams, T. A.; Schmieder, A. H.; Hu, G.; Allen, J. S.; Lacy, E. K.; Zhang, H.; Wickline, S. A.; Lanza, G. M. Arterioscler. Thromb. Vasc. Biol. 2006, 26, 2103–2109. 149. Zielhuis, S. W.; Seppenwoolde, J. H.; Mateus, V. A.; Bakker, C. J.; Krijger, G. C.; Storm, G.; Zonnenberg, B. A.; van het Schip, A. D.; Koning, G. A.; Nijsen, J. F. Cancer Biother. Radiopharm. 2006, 21, 520–527. 150. Huang, C. C.; Su, C. H.; Li, W. M.; Liu, T. Y.; Chen, J. H.; Yeh, C. S. Adv. Funct. Mater. 2009, 19, 249–258.
P1: OTA/XYZ P2: ABC c15 JWBS037-Chen January 25, 2011
374
23:25
Printer Name: Yet to Come
NANOPARTICLES CONTAINING RARE EARTH IONS: A TUNABLE TOOL FOR MRI
151. Barth, R. F.; Coderre, J. A.; Vicente, M. G. H.; Blue, T. E. Clin. Cancer Res. 2005, 11, 3987–4002. 152. Le, U. M.; Cui, Z. Int. J. Pharm. 2006, 320, 96–103. 153. Enger, S. A.; Rezaei, A.; af Rosenschold, P. M.; Lundqvist, H. Med. Phys. 2006, 33, 46–51. 154. Matsumura, A.; Zhang, T.; Nakai, K.; Endo, K.; Kumada, H.; Yamamoto, T.; Yoshida, F.; Sakurai, Y.; Yamamoto, K.; Nose, T. J. Exp. Clin. Cancer Res. 2005, 24, 93–98. 155. Wangerin, K.; Culbertson, C. N.; Jevremovic, T. Health Phys. 2005, 89, 135–144. 156. Weishaupt, D.; Kochli, V. D.; Marincek, B. How Does MRI Work? An Introduction to the Physics and Function of Magnetic Resonance Imaging. Springer: New York, 2003. 157. http://www.mr-tip.com (accessed in 2009).
P1: OTA/XYZ P2: ABC c16 JWBS037-Chen January 25, 2011
13:5
Printer Name: Yet to Come
CHAPTER 16
Microfabricated Multispectral MRI Contrast Agents GARY ZABOW and ALAN KORETSKY Laboratory of Functional and Molecular Imaging, National Institute of Neurological Disorders and Stroke, National Institutes of Health, Bethesda, Maryland, USA
16.1 INTRODUCTION Since its inception in 1973 [1, 2], magnetic resonance imaging (MRI) has rapidly become a widely used medical diagnostic imaging technique. Primary advantages of MRI include its nonionizing nature and its excellent soft tissue contrast and high spatial resolution compared to other radiological imaging methods [3]. The relatively low inherent sensitivity of MRI is offset by being able to use generally abundant water as the main molecule for detection. The ability of water to report on a wide range of normal and pathophysiological conditions due to alterations in MRI detectable properties has proved remarkable. These include changes in the relaxation times T1 and T2 that are modified by many processes such as changes in local water content and in endogenous paramagnetic metal content such as iron. Techniques that sensitize MRI to the diffusion of water [4–6], to blood flow [7], and to oxygenation of hemoglobin [8] have dramatically expanded MRI detection of tissue function. In addition to detecting such intrinsic properties, MRI contrast can be modulated by addition of exogenous contrast agents. Most important among the MRI contrast agents are the small molecule chelates of the paramagnetic ion Gd3+ . They are used in a wide variety of applications, such as in detecting leakage of the blood–brain barrier that enables, for example, ready detection of brain tumors [9] and of plaques associated with multiple sclerosis [10]. Another class of MRI contrast agent is made from nanosized dextran-coated iron-oxide particles that are potent relaxation agents and that have been used to measure pathology in liver [11]. Manipulation of endogenous contrast and use of added contrast agents has perhaps made MRI the most versatile radiological imaging tool. The success of MRI for imaging anatomical and physiological/functional changes has led to growing interest in extending MRI for cellular and molecular imaging. Although the low sensitivity of magnetic resonance would initially make MRI seem an unlikely candidate for molecular imaging, the ability of contrast agents to alter the MRI properties of large numbers of surrounding water molecules enables indirect detection via the abundant Nanoplatform-Based Molecular Imaging Edited by Xiaoyuan Chen C 2011 John Wiley & Sons, Inc. Copyright
375
P1: OTA/XYZ P2: ABC c16 JWBS037-Chen January 25, 2011
376
13:5
Printer Name: Yet to Come
MICROFABRICATED MULTISPECTRAL MRI CONTRAST AGENTS
water signal. Because the surrounding water is continually undergoing random diffusion, and because the magnetic field disturbances caused by the contrast agents can often extend appreciable distances from the contrast agent itself, the agents can interact with surrounding volumes of water many times greater than their own physical sizes, resulting in large signal amplifications. Even relatively small quantities of contrast agent are therefore detectable, rendering MRI well suited to molecular and cellular level imaging despite any inherent low NMR sensitivity. For example, concentrations of the paramagnetic ion Mn2+ as low as 2 M from voxels as small as 100 m has been detected via shortening of T1 of water in MRI [12]. This represents a total amount of Mn2+ of 2 femtomoles. MRI detection of individual micrometer-sized superparamagnetic particles has also been demonstrated [13–15] even at imaging resolutions that were 50–100 times larger than the sizes of the particles used. This is due to the large magnetic field disturbances that these superparamagnetic particles generate in the surrounding water. Such superparamagnetic particles have been used to detect single cells in vitro [16–18] and in vivo [19] and to monitor the movement of such cells [20, 21]. Over the past decade, the use of MRI for cell labeling and tracking has enjoyed considerable success [22, 23], as have other MRI-based molecular and cellular imaging approaches. Examples include the possibility of early stage disease detection through particular cell-type or epitope recognition [24–26] and real-time in vivo imaging of gene expression and enzyme activity [27–31]. Also possible with MR techniques is the detection of various in vivo molecular binding events and interactions [32], enabling, for example, real-time analyte concentration measurement [33] thanks to changes in magnetic relaxivities induced by the spatial clustering of magnetic nanoparticle contrast agents [34]. All of this development of molecular- and cellular-based MRI depends on altering MRI contrast with appropriate agents. Thanks to the great success of the Gd chelates for clinical MRI, much research has gone into the development of numerous contrast agents, be they small molecules, macromolecules, or nanoparticles, and many excellent reviews have already been written on their synthesis, characterization, and use [22, 35–45]. All MRI contrast agents developed to date, however, have been chemically synthesized. Recently, it has been demonstrated that top–down microfabrication techniques can be used to generate novel properties for MRI contrast agents [46–48]. This work relied on fabricating specifically shaped microstructures from magnetic materials to control the NMR resonance of water in the structures in ways that make them uniquely identifiable and that open the prospect of large-scale multiplexing with MRI. The goal of this chapter is to describe this new approach to MRI contrast agents. First, however, a brief description of the main classes of chemically synthesized MRI agents is presented to lend perspective to the microfabricated agents.
16.2 BOTTOM–UP CHEMICALLY SYNTHESIZED CONTRAST AGENTS Broadly classified, existing chemically synthesized MRI contrast agents can be thought of as either molecular or particulate agents. The molecular agents are generally based on chelates of some paramagnetic metal ion, often one from the lanthanide series with a large number of unpaired electrons including, most notably, the above-mentioned gadolinium, Gd3+ . These chelates are typically used as T1 agents, requiring direct contact with water in the first or second coordination sphere of the metal. At high concentrations, however, they can also cause appreciable T2 * effects, forming the basis of using Gd chelates to monitor blood flow in the brain [49]. Over the past decade, clever chemistry has been used to control
P1: OTA/XYZ P2: ABC c16 JWBS037-Chen January 25, 2011
13:5
Printer Name: Yet to Come
BOTTOM–UP CHEMICALLY SYNTHESIZED CONTRAST AGENTS
377
water access to the Gd3+ ion by relevant biological processes such as changes in calcium concentration [50] or specific enzyme activity [51]. This work has opened the possibility of using Gd3+ chelates to report on molecular and cellular events. There has also been an increased interest in Mn2+ as an MRI agent. This interest has arisen because Mn2+ goes intracellular and can be used to make unique images of anatomy as well as to monitor a number of interesting cellular properties such as calcium influx and connectivity between neurons [52, 53]. The particulate agents are generally based on various magnetic materials including, most commonly, superparamagnetic iron oxide (SPIO). Unlike the molecular-sized lanthanide chelates, the particulate agents’ linear sizes span over three orders of magnitude. At the smaller end are individual nanocrystals a few to several nanometers in size that form the core of so-called monocrystalline iron-oxide nanoparticles (MIONs) [54] and ultrasmall SPIO (USPIO) particles [55]. These cores are coated, most commonly with dextran, to ensure solubility and biocompatibility. The total hydrodynamic diameters of core and coat are typically in the tens of nanometer range. Various agglomerations of these individual iron oxide nanocrystals have been produced. For example, they have been encapsulated within dendrimer structures that are again a few tens of nanometer in size and known as magnetodendrimers [56], or collected into larger composite SPIO nanoparticles that commonly range from about 50 nm to 300 nm in diameter [57, 58]. At the larger end of the size scale are micrometer-sized particles of iron oxide (MPIOs) [15] with diameters of around 1 m to as large as about 10 m. Although these iron-oxide-based nano- and microparticles can appreciably reduce both T1 and T2 relaxation times, they are most commonly used as potent T2 * contrast agents, which tend to lead to substantial localized image darkening or reduced signal intensity. T2 * effects are caused by dephasing of magnetization due to the particles’ inhomogeneous magnetic fields. Therefore the potency of these agents is based on the size of the magnetic moment generated, which depends on the size and composition of the particles. Since MRI can detect very small perturbations in magnetic field (fractions of a part per million) these particles affect MRI signal over ranges much larger than the size of the particles themselves. Because of this amplification effect, only on the order of a million superparamagnetic nanoparticles are required per voxel for MRI detection, while at the microscale, as mentioned above, even individual microparticles can be detected. This high sensitivity to such particles has enabled MRI to detect targeting of specific receptors on specific cells [59] and is behind the already mentioned large range of work using these particles to label cells for monitoring cell migration with MRI. This success of iron-oxide particles has also encouraged interest in generating nanoparticles of other compositions including manganese oxide [60] and gadolinium oxide [61]. The flexibility of MRI has also led to a number of interesting techniques that let the T2 * properties of particles give signal enhancement rather than signal loss [62, 63]. And the distinction between paramagnetic T1 agents and particulate T2 * agents is blurring even further with proposals to let particulate contrast agents degrade under cellular conditions to give rise to paramagnetic agents [64]. Besides T1 and T2 agents, an exciting new class of MRI contrast agents has been developed in recent years known collectively as chemical-exchange saturation transfer (CEST) agents [65, 66]. These agents are based on off-resonance frequency shifts that indirectly generate contrast through the exchange of free unshifted protons of the surrounding bulk water with various bound—and thus chemically shifted—water, amide, amine, or hydroxyl protons of the CEST molecules. The beauty of such off-resonant agents is that they may be selectively addressable or detectable through their different chemical shifts, allowing a form of multispectral or “color” contrast rather than the traditional “monochrome” brightening or
P1: OTA/XYZ P2: ABC c16 JWBS037-Chen January 25, 2011
378
13:5
Printer Name: Yet to Come
MICROFABRICATED MULTISPECTRAL MRI CONTRAST AGENTS
darkening of T1 and T2 agents. Recently, it has been demonstrated that certain paramagnetic chelates, so called PARACEST agents, can also be used, with the paramagnetic ion leading to larger chemical shifts of the exchangeable protons, increasing the spectral bandwidth achievable [41, 67]. CEST principles have recently also been applied to hyperpolarized xenon agents [68] and to lipid-based nanoparticles [40], including liposome and micelle structures that each aggregate together many CEST molecules [69] to effectively increase contrast per composite particle.
16.3 TOP–DOWN FABRICATED MULTISPECTRAL CONTRAST AGENTS Compared to the above chemically synthesized forms of contrast agents, top–down microfabricated contrast agents [47] represent a departure, being designed around techniques more commonly associated with the microchip industry—see, for example, Madou [70]. Borrowing more from the microengineer’s, than from the traditional chemist’s toolkit, they are most obviously different in terms of the physical forms they can assume and the fabrication methods used in their production. The question then is whether microfabrication, or perhaps even some combination of microfabrication together with bottom–up chemical synthesis (e.g., such as in the design of nucleic acid chips), can offer new and useful properties for MRI contrast agents. Among the many differences between top–down microfabrication and bottom–up chemical synthesis approaches, two in particular bear mention here. First, having been relentlessly honed and enhanced over decades by the microelectronics industry, the micro- and nanoscale spatial patterning capabilities of top–down fabrication approaches allow greater control and flexibility in the design of structure geometries than do purely bottom–up syntheses. Second, being often more physical than chemical in nature, certain microfabrication processes can be relatively independent of material choice, allowing structure compositions that can be more accurately controlled and more easily interchanged. Consequently, top–down fabrication has the potential for creating structures boasting substantially greater degrees of consistency in size, shape, and material composition. Additionally, the increased flexibility in material choices may allow fabricated contrast agents to be more biologically compatible than, for example, those based on lanthanide ions whose toxicity demands chelation and associated loss in agent efficiency [71, 72]. Traded off against these advantages, however, are also relative disadvantages related to minimum structure sizes and fabrication efficiencies and costs. Although direct write serial patterning techniques such an electron beam and scanned probe writing overcome most size limitations by offering true nanoscale definition, they are less efficient than parallel photolithographic processes. And while the electronics industry continues to find new ways to overcome inherent optical diffraction limitations, such parallel top–down fabrication processes are, at least for the time being, still not well suited to structure fabrication at the very smallest of nanoscale sizes that are accessible to bulk phase synthesis. Additionally, even given a massively parallel patterning technique and the economies of scale that accompany such wafer-level processing, the two-dimensional planar nature of top–down fabrication means that such fabrication generally remains less efficient and offers lower throughput than bottom–up chemical synthesis routes. Finally, top–down fabrication may be more demanding in terms of process complexity and cost, often requiring specialized equipment that renders it less accessible to traditional biology/chemistry lab settings. To justify this added complexity and expense, therefore, the top–down fabricated
P1: OTA/XYZ P2: ABC c16 JWBS037-Chen January 25, 2011
13:5
Printer Name: Yet to Come
TOP–DOWN FABRICATED MULTISPECTRAL CONTRAST AGENTS
379
micro- and nanostructures should offer some extra functionality not readily achievable through traditional chemical synthesis methods. The main subject of this chapter presents one example of that, namely, exploiting microfabrication’s control over structure geometry to encode distinct spectral signatures into MRI agents. Note that these multispectral agents are almost certainly not the only form of novel contrast agent that could potentially be microfabricated; they just happen to be the first demonstrated example of one potential new direction opened up by top–down micro- and nanofabricated structures. 16.3.1 Contrast Agent Geometry While the use of ferromagnetic particles as MRI contrast agents was first proposed more than two decades ago [73, 74], interest has always focused primarily on particle composition, size, and clustering, with little attention paid to the geometry of the individual micro- or nanoparticles. For two reasons this is not surprising. First, from a chemical synthesis/selfassembly/magnetic clustering point of view, full control over the structure geometry is generally harder to realize than is control simply over overall structure size. Second, with the exception of a few positive-contrast schemes [62, 63], such agents have thus far primarily been used as negative contrast T2 * agents, designed simply to dephase the surrounding transverse proton magnetization. This makes sense since, with typical MRI voxels measuring hundreds of micrometers or more on a side, virtually all of the dephased hypointense signal results from water lying within the contrast agent nanoparticles’ farfields, that is, at distances away from the particle that are large compared to the particle’s size. With these far-fields being always dipolar in nature, the spatial contrast signal then naturally depends only on overall particle magnetization (i.e., on particle volume and composition) and is essentially independent of structure geometry. Although particle geometry may not directly affect spatial contrast, it has recently been shown [46] that, under certain conditions, the geometry can provide distinct spectrally shifted NMR frequency signals. As such, top–down fabricated ferro- or superparamagnetic agents can be microengineered to simultaneously provide both regular SPIO-type T2 * spatial contrast, as well as distinguishing spectral contrast. Viewed another way, the implication of the above is that by suitably controlling the local distribution of exactly the same quantity of magnetic material that would otherwise have been indiscriminately aggregated together into a traditional chemically synthesized nanoparticle, top–down microfabricated agents’ near-fields can be exploited to provide additional differentiating spectral contrast while their far-fields continue to provide exactly the same spatial contrast as before. As an introductory example, the scanning electron micrograph in Figure 16.1 compares a typical chemically synthesized magnetic microparticle to that of a microparticle whose geometrical structure has been engineered using surface micromachining techniques to produce magnetic field profiles suitable for the above-mentioned spectrally shifted NMR signals. Positioned in front of a standard, commercially available 4.5-m diameter magnetic bead (commonly used in biotechnology research as a separating agent, and in MRI as an MPIO T2 * agent), the image shows a sample microfabricated particle, approximately 2 m in size, formed from two slabs of magnetic material separated by a nonmagnetic post. This particular microstructure is, of course, not intended to represent the only geometry that can be microfabricated, nor the only geometry that can produce distinct NMR spectral shifts; it is intended only as a suggestive visual introduction to some of the differences between typical chemically synthesized structures and those that can be produced with the extra geometrical structuring that top–down fabrication techniques enable.
P1: OTA/XYZ P2: ABC c16 JWBS037-Chen January 25, 2011
380
13:5
Printer Name: Yet to Come
MICROFABRICATED MULTISPECTRAL MRI CONTRAST AGENTS
FIGURE 16.1 Top–down microfabrication versus bottom–up chemical synthesis. Scanning electron micrograph (SEM) showing a sample r = 1-m microfabricated structure in front of a commercial chemically synthesized 4.5-m diameter magnetic particle to show relative size. The top and bottom material slabs of the microfabricated structure are made from magnetic material (nickel). The central support post is nonmagnetic (copper). (Reproduced with permission from Zabow et al. [46].)
Viewed as specialized geometrical variants of regular chemically synthesized magnetic nanoparticles, the microfabricated multispectral agents’ added frequency-shifting ability can be understood by first reconsidering how regular magnetic nanoparticles generate T2 * contrast. As with any coherence phenomenon, SPIO agents’ reduction of the transverse coherence time is directly coupled to an associated frequency broadening; in this case, a broadening of the water NMR line. Proton Larmor precession frequencies are directly proportional to magnetic field magnitude. Therefore the spatially changing magnetic fields surrounding all magnetic particles lead to spatially varying proton precession frequencies that (1) lead to rapid dephasing of the neighboring transverse magnetization and the agents characteristic resulting negative image contrast, and (2) yield NMR frequency signatures that integrate out into broadening of the water line. To achieve instead a distinct frequency shift of the water NMR signal, rather than solely a broadening, the water protons need to be exposed to a field magnitude that is homogeneous over some extended spatial region and that is of a different magnitude than any surrounding field. Given the continuous spatial field decay away from any magnetic particle, this is not possible in any particle’s far-field region, but for appropriate geometries the requisite homogeneous field profiles can be engineered within the particle near-fields. For example, it has been shown [46] that when suitably aligned and magnetized to saturation by the background MRI field, an appropriately spaced pair of magnetizable disks can give rise to a homogeneous field region that can uniformly shift the NMR frequency of the water passing between those disks. A schematic of this double-disk geometry is shown in Figure 16.2 together with the spatial profile of the total field magnitude resulting from the magnetized structure’s field added to the background MRI field. Although not shown in the figure, the magnetic disks are held at a predetermined fixed distance apart by nonmagnetic posts, either internal or external to the structure (see Fig. 16.4(a) and Fig. 16.4(b)). By varying the disk spacings, thicknesses, and radii, or by changing their material composition to increase magnetic moment, the homogeneity and the magnitude of the surrounding magnetic fields can be engineered to yield the desired spectral signature. This enables water NMR frequencies to be shifted almost arbitrarily, certainly over ranges that exceed any that chemical synthesis has yet to achieve. For the shift in NMR frequencies to be well defined, it is critical that the structures
P1: OTA/XYZ P2: ABC c16 JWBS037-Chen January 25, 2011
13:5
Printer Name: Yet to Come
TOP–DOWN FABRICATED MULTISPECTRAL CONTRAST AGENTS
(b)
– Field magnitud
e (a.u.)
(a)
381
–2R
–R
0
R
2R
FIGURE 16.2 Magnetic structure and field diagrams. (a) Diagram of the field (small black arrows) from two parallel disks magnetized to saturation by the background MRI field (large gray arrows). (b) Calculated (negative) field magnitude in the midplane through a typical magnetized disk set, contrasting the homogeneous nature between the disks with the rapid external decay. (Reproduced with permission from Zabow et al. [46].)
be aligned parallel to the applied magnetic field. The structures’ built-in magnetic shape anisotropy automatically ensures that this is the case, causing the particles to self-align to the applied magnetic field. The double-disk structures are by no means the only self-aligning structures that can give a well-controlled frequency shift. As another example, it has recently been demonstrated [48] that hollow magnetic cylinders of suitable aspect ratios can likewise generate shifted NMR signals from the water within the cylindrical shells. These hollow magnetizable cylindrical structures operate, and self-align, according to the same principles as the doubledisk structures and, at least qualitatively, can be understood as each being comprised simply of a series of rotationally superposed double-disk structures [48]. Accordingly, in this case, instead of disk thickness, spacing, and radii, it is the cylinder wall thicknesses, radii, and lengths that control the resulting spectral shifts. For the simplified cases of “thin” disks or “thin-walled” cylinders made from a material with saturation magnetic polarization J S , the resulting NMR frequency shifts ␦ can be analytically approximated as Double disk:
≈ −4␥ JS
Hollow cylinder:
≈ −4␥ JS
hr 2 (4r 2 + 4s 2 )3/2 L t (L 2 + 4 2 )3/2
where ␥ is the gyromagnetic ratio, and the geometrical parameters r, 2s, h, L, , and t denote, respectively, the disk radii, center-to-center separation and thicknesses, and
FIGURE 16.3 Schematic of double-disk and hollow cylindrical contrast agent microstructure geometries. The double-disk structure (left) has disks of radii r, center-to-center separation 2s, and thickness h. The hollow cylinder (right) has radius , length L, and wall thickness t.
P1: OTA/XYZ P2: ABC c16 JWBS037-Chen January 25, 2011
382
13:5
Printer Name: Yet to Come
MICROFABRICATED MULTISPECTRAL MRI CONTRAST AGENTS
the hollow cylinder length, radius, and wall thickness as illustrated schematically in Figure 16.3. In the above, the designation “thin” assumes that h r ≈ 2s and t L ≈ 2 . The above equations have been arranged in similar form and the geometrical correspondence between the two systems becomes even more apparent by noting the natural associations between disk diameter and cylinder length 2r ∼ L, between disk separation and cylinder diameter 2s ∼ 2 , and between disk thicknesses and cylinder wall thickness h ∼ t. SEMs of arrays of different double-disk and hollow cylindrical magnetic structures of various configurations and sizes, all of which can discretely shift NMR spectral peaks, are collected together in Figure 16.4. Note that it is neither the disk nor the cylindrical shell families of geometrical structures that should necessarily be viewed as the critical component for generating distinct spectrally shifted NMR signals. In fact, these particular geometries were chosen primarily on the basis of their being, among others, compatible with microfabrication processing technology. Rather, the critical requirement is only that the total resulting field possess a water-accessible spatially extended region over which its magnitude is homogeneously shifted away from that of all surrounding fields. The ability to encode different spectral signatures into magnetic particles, that would otherwise appear identical in an MRI scan, offers the potential to track and differentiate between multiple different particles and the different biological markers that may be specifically bound to them. While the underlying physical principles behind their operation are different, in spectral terms the engineered magnetic structures can be thought of as radiofrequency (rf) analogs to optically probed quantum dots [75–77]. Just like quantum dots, they therefore have the potential to allow similar multiplexed studies except, being rf rather than optical probes, the obscuring optical scattering properties of biological tissues no longer matter and the probes can potentially be tracked in vivo. A fundamental difference between these magnetic structures and quantum dots, however, is that there is nothing quantized about the magnetic structure sizes. Instead, it is the magnetic structure’s geometrical aspect ratios, rather than quantum dot’s semiconductor crystal size, that is engineered to provide the desired “color” or rf spectral response. Indeed, as implied by the dimensionless groupings of geometrical variables in the above frequency-shift equations, provided all dimensions
(a)
(b)
(c)
FIGURE 16.4 Scanning electron micrographs of various microfabricated contrast agent microstructures. Panels (a) and (b) show arrays of microfabricated double-disk structures with single internal, or three external, support posts, respectively. Panel (c) shows an array of microfabricated hollow cylindrical structures.
P1: OTA/XYZ P2: ABC c16 JWBS037-Chen January 25, 2011
13:5
Printer Name: Yet to Come
TOP–DOWN FABRICATED MULTISPECTRAL CONTRAST AGENTS
383
are scaled proportionally to one another, the shifts in the surrounding water NMR signals are independent of overall structure size, allowing a wide range of particle sizes. Apart from permitting multiple different “colors,” their spectral signals also enable confirmation as to whether hypointense regions observed within an MR image indicate the locations of administered T2 * contrast agent or are instead spurious natural image darkenings or signal voids arising from various sources of magnetic field inhomogeneities. Figure 16.5 illustrates these concepts, showing regular T2 * -weighted gradient echo MR imaging of a series of double-disk structures together with chemical shift MR imaging of those same structures. The unique spectral signatures were shown to enable different particle types (in this case
(a)
(c)
(b)
(d)
FIGURE 16.5 Multispectral MRI. (a–d) Chemical shift imaging of demonstration 1.25-mm diameter particles magnetized by the background MRI field. Particle frequency was varied by changing the thickness of electroplated nickel layers that formed the magnetizable disk pairs. As with normal SPIO detection, magnetic dephasing due to the particles’ external fields enables the spatial imaging shown in the gradient-echo MRI (a). However, comparison between (a) and the chemical shift images (b) shows that the additional spectral information both differentiates between particle types and improves particle localization. The particles are shown schematically (not to scale) in (c). With particle spectra ((d), to the right of the corresponding chemical shift images in (b)) shifted well clear of the water proton line, different planes in the chemical shift imaging map isolate different particle types for unambiguous color coding with minimal background interference ((b), bottom panel). Although still visible in the gradient-echo image, the top corner particle of the letter “B” was damaged, causing its shifted frequency peak to vanish. (Reproduced with permission from Zabow et al. [46].)
P1: OTA/XYZ P2: ABC c16 JWBS037-Chen January 25, 2011
384
13:5
Printer Name: Yet to Come
MICROFABRICATED MULTISPECTRAL MRI CONTRAST AGENTS
all double-disk structures but made with disks of different thicknesses) to be distinguished from one another and from the background. Additionally, because the spectral signals came from water between the structures’ two disks rather than from external surrounding water, as do the dephased magnetization signals responsible for the SPIO-like spatial contrast, the spectral signals afforded improved spatial resolution in the chemical shift images. 16.3.2 Chemical Shifting Versus Magnetic Microstructure Shifting Spectral shifting of NMR frequencies, per se, is nothing new; after all, it underpins all of NMR chemical spectroscopy where the magnetic shielding effect of the surrounding electron clouds locally alters the fields seen by that molecule’s protons, giving rise to different chemical shifts or characteristic NMR spectral “fingerprints” for different molecules. Although chemical shifts may appear unrelated to the shifts produced by the microfabricated agents, noting the parallels that do exist between the two helps illuminate some of the microfabricated agents’ relative advantages. Just as chemical shifts can be understood in terms of the molecules’ surrounding chemical electronic current loops, the microfabricated structures’ field shifting can be regarded analogously as arising from the effective surface Amperian electronic current loops that surround all magnetic objects. The difference is that since the composition, geometrical structure, and relative positionings of these magnetic objects can be arbitrarily engineered, so too can the strengths, locations, and paths of those Amperian surface current loops, providing a degree of explicit field control that is harder to achieve through chemical molecular engineering. Also, since ferromagnetism is fundamentally a multibody phenomenon relying on mutual cooperation between multiple spins, the fabricated microstructures, which represent extended material objects, can readily be ferromagnetic in nature; by contrast, molecular complexes are typically paramagnetic or diamagnetic in nature. When magnetized by the background MRI field, therefore, the microstructures’ greater magnetic permeabilities induce larger equivalent electronic currents than those of dia- or paramagnetic molecules, and therefore enable larger frequency shifts that, as explained below, allow for greater signal amplification. Indeed, since the microfabricated magnetic structures are magnetically saturated already in applied fields well below a tesla, their field shifts can be large and essentially independent of background field for typical MRI fields. This can be seen in the sample NMR spectra acquired from a set of double-disk structures shown in Figure 16.6. The curves, so-called z-spectra [78], show the spectrally shifted NMR water peaks for three different fields strengths, 11.7 T, 7.0 T, and 4.7 T. Although signal amplitudes differed (due to differing transmit/receive coil geometries and field-dependent differences in T1 ), the frequency shifts were independent of MRI field strength. By comparison, given the much lower permeabilities of para- or diamagnetic species, chemical shifts are normally proportional to the background MRI field strength and large molecular shifts therefore require large MRI fields. For the microfabricated structures this then leads to the somewhat unusual situation where conventional chemical shift nomenclature, typically involving measurements represented in relative parts per million (ppm), is no longer helpful. The shifts in Figure 16.6, for example, are therefore given in absolute units, not in ppm. 16.3.3 Magnetization Transfer for Amplification of the Detection of Microfabricated Structures Chemical shifts can be exploited not only for NMR chemical spectroscopy or associated chemical shift MR imaging as seen in Figures 16.5 and 16.6, but also for generating
P1: OTA/XYZ P2: ABC c16 JWBS037-Chen January 25, 2011
13:5
Printer Name: Yet to Come
TOP–DOWN FABRICATED MULTISPECTRAL CONTRAST AGENTS
385
1– Ms/M0
1.00 0.95 0.90
11.7 T 7.0 T 4.7 T
0.85 0.80 –500
–400
–300
–200
–100
Frequency shift (kHz)
FIGURE 16.6 Field-independent spectral shifting: H2 O z-spectra for an array of double-disk structures (r = 2.5 m) taken at three different MRI field strengths all exhibit the same 275-kHz spectral shift, indicating that structures have been fully magnetized to saturation by background MRI fields. (Reproduced with permission from Zabow et al. [46].)
frequency-dependent MRI spatial contrast through off-resonance proton exchange magnetization transfer [79, 80]. For example, such is the mechanism behind CEST and PARACEST contrast agents [41, 65, 67], where off-resonance pulses saturate out the magnetization of particular exchangeable protons within the molecular complexes comprising the CEST/PARACEST agents. Once these off-resonance protons subsequently undergo chemical exchange with the surrounding bulk water protons, the total on-resonance magnetization, and hence the signal intensity acquired from the surrounding bulk water pool, is diminished. Compared to chemical shift imaging, this magnetization transfer scheme represents a more indirect method of imaging, but one that can afford substantial signal enhancement. Specifically, if the free and bound protons exchange at a rate that is faster than the longitudinal relaxation time, T1 , but that is not too fast (explained below), then prior to each on-resonance signal acquisition, several off-resonance saturating pulses can be applied to multiply the total magnetization saturation and effectively increase the contrast agents’ signal-to-noise ratio. Likewise, the microfabricated structures reviewed here can also benefit from similar signal amplification strategies. The difference is that for the microfabricated agents the magnetization transfer does not rely on chemical exchange processes and is therefore not limited by chemical exchange rates; instead, exchange is limited only by the rate at which water molecules naturally self-diffuse around the structures and in and out of the structures’ internal homogeneous field regions. Because this diffusion continually replenishes the water within the structures’ internal homogeneous field regions with “fresh” water that has diffused in from outside, the spectral signals generated can appear to come from volumes of water that are orders of magnitude larger than the internal signal-generating regions of the particles themselves. Since the time period required to diffuse any given distance scales with the square of that distance, as structure sizes shrink, the water exchange rates, and hence potential NMR signal amplification, increase quadratically. This renders the microfabricated structures well suited to miniaturization down to the small sizes required for cellular and molecular level imaging applications since for any given total volume of contrast agent material, the total signal from an ensemble of contrast agent micro- or nanostructures is greater the smaller are the structures comprising that ensemble.
P1: OTA/XYZ P2: ABC c16 JWBS037-Chen January 25, 2011
386
13:5
Printer Name: Yet to Come
MICROFABRICATED MULTISPECTRAL MRI CONTRAST AGENTS
An important caveat to continually increasing the exchange rate, however, is that it should not become so fast that it frequency-broadens the spectral peak by more than its shift. In other words, ensuring that the shifted frequency peak remains clearly resolvable from the unshifted background water requires that the frequency of water exchange remains less than the magnitude of the water’s NMR frequency shift or, equivalently, is slow relative to the NMR time scales involved. A more in-depth discussion of this so-called slowexchange condition and its significance in contrast agent design can be found in Woods’ review of PARACEST agents [41]. Although presented in terms of chemical-exchange limited molecular agents, the arguments in that review carry over equally to the diffusionexchange limited microfabricated agents discussed here. To summarize those arguments, maximizing an exchange-based agent’s potential for signal amplification requires maximizing the exchange rate, which in turn requires maximizing the magnitude of that agent’s frequency shift. Fortunately, given their ferromagnetic nature, even at low MRI background fields the microfabricated structures can produce large field shifts, exceeding those achievable by molecular analogs. For example, a wide range of engineered frequency shifts from tens of kilohertz to substantially beyond a megahertz have already been experimentally demonstrated [46, 48]. In principle, shifts still an order of magnitude larger should also be achievable, potentially allowing structures to be scaled down to the 100-nm range [46], well within typical SPIO nanoparticle size ranges. Apart from enabling large signal amplification, large shifts also bring other benefits. By shifting the signal of interest as far away from the bulk water signal as possible, much of the main competing background noise can be eliminated. Constraints on the power and envelope shapes of the applied rf pulses are significantly reduced since for large offset shifts, there is little risk of any residual rf power overlapping with the bulk water line. Similarly, multiple sets of different frequency shifts can also be clearly distinguished from one another. Also, detection becomes more forgiving of any small system field inhomogeneities, whose resulting frequency shift errors become negligible in comparison to microstructure frequency shifts. Since the amplitude of the agents’ spectrally shifted signals depends on the total magnetization exchange, it depends not only on the chemical- or diffusion-driven exchange rate for each proton, but also on the total number of those protons whose magnetization can be simultaneously saturated at any one time. Not being dependent on any molecule’s particular individual exchangeable proton sites but instead being designed to generate uniform field offsets over spatially extended regions, the microfabricated magnetic structures can interact with larger numbers of protons simultaneously to give large signals per particle. Combined with the signal enhancement due to diffusional exchange, it has been shown [46] that such microfabricated structures should be detectable at low concentrations that may exceed those of certain molecular-based agents and that compare favorably to chemically synthesized SPIO nanoparticle contrast agent concentration requirements. 16.3.4 Combining Imaging and Sensing The microfabricated agents’ spectral signals are derived from the water that flows through the particle structures. This enables several possibilities for transforming the multispectral contrast agents into rf probes that could potentially report on the presence of various biomolecules or enzymes, or monitor various in vivo physiological conditions. By selectively blocking or unblocking water access to the particles’ interior regions, the spectrally
P1: OTA/XYZ P2: ABC c16 JWBS037-Chen January 25, 2011
13:5
Printer Name: Yet to Come
TOP–DOWN FABRICATED MULTISPECTRAL CONTRAST AGENTS
387
FIGURE 16.7 Controlling diffusion to turn tags on or off. (Main panel) High tilt-angle SEM image showing a square array of r = 2.5-m double-disk particles. Except for a defined circular region, all particles have their interiors filled, blocking water diffusion. (Top left inset) A higher magnification SEM image of the boundary between open and filled particles. (Top right inset) The resulting background-subtracted chemical shift MRI showing transferred magnetization saturation from the particles’ shifted resonance. Signal is visible from those particles that have water diffusing through their open interior region (labeled “On”) but not from those particles that have their interiors filled (labeled “Off”). The bottom of the image shows a region that contains no particles (labeled “No tags”), providing a null background signal comparison. A scratch (seen at the lower right corner) removed ∼100 particles (about 10–20 per voxel). Its visibility in the magnetic resonance image suggests the potential for high-resolution imaging to spectrally distinguish individual such particles. (Reproduced with permission from Zabow et al. [46].)
shifted NMR signals can disappear or reappear, respectively, allowing the structures to act as switches. The idea is illustrated in Figure 16.7, showing an array of double-disk structures that are identical in every respect except that some of the structures have their internal regions blocked. Instead of a standard gradient-echo MRI that spatially locates structures based on their resulting T2 * hypointense contrast, the MRI shown is a difference image formed by subtracting from a first on-resonant image, a second image that was generated after first applying a preceding set of off-resonant pulses at the particles’ shift frequencies, as described above. In this way the locations of those double-disk structures with geometries that are resonant with the preceding off-resonant pulses appear highlighted in the difference image while all else disappears. However, only those double-disk structures that have their interior regions accessible generate any off-resonance contrast signal; those structures with their interior regions blocked do not yield any spectrally shifted signal and hence are not visible. The potential for the structures’ contrast signal switching to report on local conditions is then clear: if the material that fills the particles’ internal regions or that coats the particles
P1: OTA/XYZ P2: ABC c16 JWBS037-Chen January 25, 2011
388
13:5
Printer Name: Yet to Come
MICROFABRICATED MULTISPECTRAL MRI CONTRAST AGENTS
was designed to dissolve under certain conditions, such as temperature or pH level, or to break down in the presence of specific enzymes, the presence or absence of the particles’ spectral signals could provide real-time information about in vivo physiological events. This idea is, for example, not unlike one first demonstrated with gadolinium complexes, where water access to the inner Gd3+ ion was compromised by a galactopyranose blocking group sensitive to cleavage by -galactosidase, a common marker gene [27]. However, the difference is that because the particles’ spatial T2 * contrast is independent of whether the particle interior admits water or not, the spatial locations of the particles remain always visible, allowing them to be simultaneously continuously tracked. Additionally, because each particle’s signal can be engineered to occur at a different frequency, multiple different physiological conditions can potentially be simultaneously tested for and differentiated between, even when occurring within the same MRI voxel. The strategy of admitting or denying water access to the particle interior is applicable to both the double-disk and the hollow cylinder structures. However, with their additional separating nonmagnetic posts, the double disks also have the potential to monitor local conditions in a more continuous manner through condition-dependent variations in the magnetic resonant frequencies of the structures themselves. For example, if the separating posts were made of a material responsive to local conditions, then the resulting change in disk spacing could continuously shift the NMR resonances of the water passing between the disks. This change would be instantaneously reflected in the shift in the water resonance, opening the possibility of fast MRI sensors. 16.3.5 Contrast Agent Microfabrication From a scientific point of view, the ability to arbitrarily control NMR resonances increases the magnetic particles’ imaging and sensing potential; from an engineering point of view, however, fully realizing that potential is not without its challenges. With their multispectral nature the magnetic structures epitomize the proverbial double-edged sword: while the structures’ resonances can be intentionally manipulated through controllably changing their geometry, any unintentional geometrical variations from one structure to the next will lead to unintended variation in their resonance frequencies. For NMR signals acquired from individual structures, small fabrication inaccuracies do not necessarily matter: at worst, they diminish that individual structure’s field homogeneity, reducing the sharpness of the structure’s spectral peak. However, for NMR signals acquired from large ensembles of nanostructures, interparticle frequency shifts integrate out into a further broadening of the structures’ combined NMR spectral peak that can significantly reduce signal intensity. For example, the double-disk structures can potentially yield shifted spectral peaks with linewidths on the order of a hundred times narrower than their shifts; accordingly, anything more than a 1% variation in any of the structures’ geometrical parameters or in their saturation magnetization (or material composition) across the ensemble of structures can noticeably reduce their signal intensity [47]. Figure 16.8 shows an example of this rapid signal reduction as a function of the ratio of cross-structure frequency variation to average frequency shift. The plot shows calculated histograms that approximate the double-disk’s NMR signatures by measuring, throughout the space surrounding an ensemble of structures, the total volumes over which each field magnitude (or, equivalently, Larmor precession frequency) occurs. Therefore it is not only the structures’ geometrical forms but also the requirement of minimizing cross-structure variation that favors top–down fabrication techniques over
P1: OTA/XYZ P2: ABC c16 JWBS037-Chen January 25, 2011
13:5
Printer Name: Yet to Come
Field volume (a.u.)
3 +/– 0% +/– 0.5%
2
+/– 1.0% +/– 2.0%
1
Peak height (a.u.)
TOP–DOWN FABRICATED MULTISPECTRAL CONTRAST AGENTS
389
3 2 1 0 0 2 4 6 8 10 % variation (+/–)
+/– 5.0% +/– 10%
0 –650 –600 –550 –500 –450 –400 –350 Frequency shift (kHz)
FIGURE 16.8 Impact of cross-wafer manufacturing variation on spectral linewidths. The main plot shows numerical calculation of the broadening in peak widths when integrating expected signal from an array of double-disk structures that suffer from interparticle cross-wafer processing variation. Curve labels show the integration range of resulting structure frequency shifts about the central frequency. Inset plots the associated relative signal loss as cross-wafer variation increases. Calculations are for structures with r = 1 m, 2s = 0.85 m, h = 50 nm, and J S = 0.6 T (nickel). (Reproduced with permission from Zabow et al. [47].)
traditional chemical syntheses. For example, producing sharp spectral peaks with the double-disk system requires that for each structure the top and bottom disks be well centered on top of each other, the separation gap between those disks be well defined, the radii, thicknesses, and magnetic moments of the top and bottom disks be well controlled and equal to each other, and from one structure to the next there be minimal cross-wafer variation between any of the structures’ magnetic or geometric parameters. Even with the more direct control that top–down fabrication provides, attention must still be paid to various systematic errors and cross-wafer variations particular to each microfabrication processing step. Full discussion of such considerations is beyond the scope of this chapter but can be found detailed elsewhere [47]. Also to be found elsewhere [48] are microfabrication details of the hollow cylindrical structures whose geometry dictates less conventional microfabrication techniques. The double-disk structures, by contrast, are better suited to conventional planar surface micromachining processes and therefore provide a better introduction to some of those processes. Outlined here are illustrative examples of two different approaches for microfabricating double-disk structures: one based on electrochemical deposition and the other on thermal evaporation. The higher deposition rates of chemical electroplating are indispensible for the fabrication of relatively large microscale structures, while the reduced surface roughness and cross-wafer variation possible with thermal evaporation techniques are better suited to more accurate, smaller microscale or nanoscale structures. Whether electroplated or evaporated, a precursor to all double-disk structures is a deposited planar metal trilayer, photolithographically patterned into arrays of trilayer cylindrical stacks, that each consist of a base magnetic layer, a middle sacrificial nonmagnetic layer, and a final top magnetic layer. The sacrificial middle layer is used to define the spacing between what becomes the top and bottom magnetic disks and is either wholly or partially removed depending on the nature of the desired final spacer post(s) linking top and bottom disks. In addition to these three layers, extra layers may be intermingled that may
P1: OTA/XYZ P2: ABC c16 JWBS037-Chen January 25, 2011
390
13:5
Printer Name: Yet to Come
MICROFABRICATED MULTISPECTRAL MRI CONTRAST AGENTS
serve as various protective coatings for the final magnetic layers. For example, interleaving thin-film titanium or gold layers can make the final structures more biologically inert or more amenable to certain bioconjugation protocols that may be required for their specific targeting. In the vertical direction, perpendicular to the processing substrate, accurate definition of the thicknesses of the deposited metal layers can be achieved relatively easily through control of metal deposition rates and times. In this way the stacked layer design allows the thicknesses of top and bottom disks as well as their separation gap to be accurately defined. Disk radii meanwhile, being dependent on lateral patterning of the metal trilayer, are subject to various spatial resolution limits. However, because the disks are circular, their patterning scales relatively well; when encroaching on optical diffraction limits, other shapes would scale less well, since preferential blurring of higher spatial frequencies would lead to shape distortion. Scalability is beneficial because (1) structure frequency shifts are independent of overall structure size and the range of potential applications is therefore increased if structures can be fabricated over a large range of sizes, (2) for any given total amount of material, smaller structures can interact with a larger total volume of water, and (3) being able to fabricate, in particular, very small micro- or nanostructures makes for a less intrusive agent that may afford increased biological compatibility. Spatial patterning of the circular stacks may be performed either before or after the metal deposition steps. In the former approach a masking layer of photoresist is first patterned with arrays of circular openings defining those locations on the substrate wafer, where, after subsequent metal film deposition and photoresist removal, metal will remain in the form of the required cylindrical stacks; in the latter, the metal trilayer is first uniformly deposited everywhere on the substrate and covered by an inverse photoresist mask consisting of arrays of circular disks (instead of circular holes) that protect the metal beneath those circles as the surrounding metal is etched away. Which approach is preferred depends on considerations of desired final structure sizes and associated metal deposition and removal processes, on considerations of the resulting structure accuracy and cross-wafer uniformity that is demanded, and on considerations of the overall process complexity and throughput. The last consideration of processing efficiency is particularly important when considering that the top–down fabrication approach is here replacing more traditional bottom–up chemical syntheses that can offer substantially higher throughput. Indeed, it may often be desirable to trade off some degree of structure accuracy in favor of increased top–down process simplicity and throughput [47]. Turning to specific fabrication details, the diagrams in Figure 16.9 briefly outline the main steps in a sample process flow for microfabricating electroplated double-disk structures. A 10-nm thick adhesion layer of titanium is thermally evaporated onto the wafer substrate followed by a 100-nm layer of gold that serves as the electroplating seed layer and provides the necessary electrical contact. A first layer of magnetic material (e.g., nickel), followed by a nonmagnetic sacrificial layer of copper, and then a second layer of magnetic material are sequentially electroplated on top of this gold-coated substrate. Electroplating times are determined by the desired final disk thicknesses and separations, which are governed by double-disk field homogeneity concerns and desired NMR frequency shifts. A layer of photoresist is then spin-coated on top of this metal stack, photolithographically patterned and developed, and then hardbaked to leave arrays of resilient protective photoresist disks as shown in the schematic cross section of Figure 16.9a. An argon ion beam directed normal to the surface is then used to ion-mill through the exposed top magnetic layer and partway through the underlying copper (Fig. 16.9b). A selective copper wet etch then exposes the
P1: OTA/XYZ P2: ABC c16 JWBS037-Chen January 25, 2011
13:5
Printer Name: Yet to Come
TOP–DOWN FABRICATED MULTISPECTRAL CONTRAST AGENTS
391
FIGURE 16.9 Process flow diagram for electroplated double-disk structures: (a) electroplated nickel–copper–nickel trilayer atop gold–titanium seed layer with patterned photoresist disks on top, (b) ion milling through top nickel layer and part way through copper, (c) partial copper wet etch, (d) additional photoresist coating after flood exposure and development, (e) nickel wet etch and additional photoresist removal, and (f) left: timed copper wet etch, right: SU-8 photopatterned external posts and complete copper removal. (Reproduced with permission from Zabow et al. [47].)
base magnetic layer, but this etch is stopped short of laterally etching completely through the copper (Fig. 16.9c). A second layer of photoresist is then spin-coated over the wafer, flood exposed, and developed (Fig. 16.9d). This results in the removal of this second layer of resist everywhere except for directly beneath the circular disks of the top magnetic layer, which effectively double as an optical mask during the exposure. A selective wet etch of the base magnetic layer and an acetone removal of the second photoresist layer (Fig. 16.9e) therefore leaves the base magnetic disks patterned identically to, and automatically aligned directly beneath, the top magnetic disks. Finally, either the remaining copper is subjected to a timed wet etch to leave single central support posts, or external posts are first patterned using, for example, a biologically compatible photoepoxy [81, 82], before the remaining copper is completely wet etched away (Fig. 16.9f). As another example, the process flow in Figure 16.10 sketches out an alternative fabrication approach based on thermal evaporation. Beginning similarly with a gold–titaniumcoated substrate, a double layer of resist is spin-coated over the wafer. The top layer of this bilayer resist is photosensitive and lithographically patterned to yield an array of circular holes once developed. Meanwhile, the base layer of the resist, a so-called lift-off resist (LOR), is not photosensitive, but instead develops isotropically. Appropriately timed, the resist development then exposes the underlying gold layer and leaves the reentrant undercut profile shown in the cross-section schematic of Figure 16.10a. Sequential evaporation of a first base magnetic layer, a middle sacrificial copper layer, and a final magnetic layer results in the metal deposition profile shown in Figure 16.10b with evaporation times being used to determine layer thicknesses. Because of the undercut bilayer resist profile, the metal deposit on the base gold layer and on the resist layers is discontinuous, allowing clean
P1: OTA/XYZ P2: ABC c16 JWBS037-Chen January 25, 2011
392
13:5
Printer Name: Yet to Come
MICROFABRICATED MULTISPECTRAL MRI CONTRAST AGENTS
FIGURE 16.10 Process flow diagram for evaporated double-disk structures using lift-off patterning: (a) patterned bilayer lift-off resist (undercut profile) atop a gold–titanium base layer, (b) nickel–copper–nickel trilayer evaporation, (c) lift-off resist stack removal and partial copper wet etch, and (d) left: timed copper wet etch, right: SU-8 photopatterned external posts and complete copper removal. (Reproduced with permission from Zabow et al. [47].)
removal of the resist bilayer together with its metal coatings. A subsequent partial copper wet etch then leaves the metal profiles seen in Figure 16.10c. Because of the manner in which the evaporated material accumulates on the resist mask, the exposed circular regions in this mask continually shrink during the metal evaporation leaving tapered metal profiles and top disks that are somewhat smaller than their underlying base disks. This does affect resulting double-disk field homogeneity but can be compensated for by making the top disk layers thinner than the base layers [47]. Finally, as before with the electroplating scheme, the remaining copper may either be partially etched away or, if external posts are used, completely removed (Fig. 16.10d). Note that while each of the above steps show just two representative structures, through the parallel nature of photolithographic processing, each one of those steps is compatible with the simultaneous patterning of large numbers of copies across the substrate wafer. For example, a standard 6-inch wafer can accommodate simultaneous fabrication of on the order of 1010 m-sized structures. The above process flows represent only two possible protocols; many other variants are possible. Ultimately, which processing protocol and materials are used will depend on the contrast agents’ intended applications. Likewise, process protocols will depend on the acceptable levels of compromise between required geometrical accuracy and microfabrication throughput that the final structures’ uses allow.
16.4 FUTURE OUTLOOK While the multispectral contrast agents described in this chapter have been successfully microfabricated in various shapes and sizes and successfully tested in various MRI scanners, they still represent a new approach to MRI contrast agent development. At this point the top–down microfabricated multispectral structures may be regarded as demonstrations only of a new physical platform, a precursor only to any proven new biological imaging agent. At the time of this writing, much remains to be done in terms of biocompatibility studies, the development of efficient strategies for the administering or biological uptake of these
P1: OTA/XYZ P2: ABC c16 JWBS037-Chen January 25, 2011
13:5
Printer Name: Yet to Come
REFERENCES
393
structures, and the structure microengineering itself. For example, structure compositions must still be optimized for biological compatibility, and methods and functionalization strategies developed for the specific biological targeting of the agents. Work continues on further miniaturization of the structures, on the exploration of alternative fabrication materials, strategies, and geometries, and on the improvement of the efficiencies and yields of those microfabrication processes. With the wealth of microfabrication strategies available, the relative ease of switching between materials that those strategies permit, and the many already well-developed proven methods for the bioconjugation of other nanoparticle agents discussed elsewhere in this book, the outlook for microfabricated cellular and molecular imaging agents is an optimistic one. In particular, the range of fabrication techniques that have been developed by the microelectronics industry is substantial. So too is that community’s ability to continually improve upon existing techniques or to invent new ones to overcome new fabrication challenges. Leveraging off such a strong existing technology base, it seems likely that various additional fabrication strategies and even completely alternative types of agents may emerge that will enable additional new, and as yet unenvisioned, functionalities and novel forms of MRI contrast and molecular imaging agents.
ACKNOWLEDGMENTS This work was supported in part by the NINDS NIH Intramural Research Program.
REFERENCES 1. Lauterbur, P. C. Image formation by induced local interactions: examples employing nuclear magnetic resonance. Nature 1973, 242, 190–191. 2. Mansfield, P.; Grannell, P. K. NMR “diffraction” in solids? J. Phys. C 1973, 6, L422–L426. 3. Callaghan, P. T. Principles of Nuclear Magnetic Resonance Microscopy. Oxford University Press: New York, 1991. 4. Moseley, M. E.; Cohen, Y.; Kucharczyk, J.; Mintorovitch, J.; Asgari, H. S.; Wendland, M. F.; Tsuruda, J.; Norman, D. Diffusion-weighted MR imaging of anisotropic water diffusion in cat central-nervous-system. Radiology 1990, 176, 439–445. 5. Basser, P. J. Inferring microstructural features and the physiological state of tissues from diffusion-weighted images. NMR Biomed. 1995, 8, 333–344. 6. LeBihan, D.; Mangin, J. F.; Poupon, C.; Clark, C. A.; Pappata, S.; Molko, N.; Chabriat, H. Diffusion tensor imaging: concepts and applications. J. Magn. Reson. Imaging 2001, 13, 534– 546. 7. Detre, J. A.; Leigh, J. S.; Williams, D. S.; Koretsky, A. P. Perfusion imaging. Magn. Reson. Med. 1992, 23, 37–45. 8. Ogawa, S.; Lee, T. M.; Nayak, A. S.; Glynn, P. Oxygenation-sensitive contrast in magneticresonance image of rodent brain at high magnetic fields. Magn. Reson. Med. 1990, 14, 68–78. 9. Watanabe, M.; Tanaka, R.; Takeda, N. Magnetic-resonance-imaging and histopathology of cerebral gliomas. Neuroradiology 1992, 34, 463–469. 10. McFarland, H. F.; Frank, J. A.; Albert, P. S.; Smith, M. E.; Martin, R.; Harris, J. O.; Patronas, N.; Maloni, H.; McFarlin, D. E. Using gadolinium-enhanced magnetic-resonance-imaging lesions to monitor disease-activity in multiple-sclerosis. Annl. Neurol. 1992, 32, 758–766.
P1: OTA/XYZ P2: ABC c16 JWBS037-Chen January 25, 2011
394
13:5
Printer Name: Yet to Come
MICROFABRICATED MULTISPECTRAL MRI CONTRAST AGENTS
11. Bellin, M. F.; Zaim, S.; Auberton, E.; Sarfati, G.; Duron, J. J.; Khayat, D.; Grellet, J. Liver metastase—safety and efficacy of detection with superparamagnetic iron-oxide in MR-imaging. Radiology 1994, 193, 657–663. 12. Chuang, K. H.; Koretsky, A. P. Improved neuronal tract tracing using manganese enhanced magnetic resonance imaging with fast T1 mapping. Magn. Reson. Med. 2006, 55, 604–611. 13. Lauterbur, P. C.; Bernardo, M. L. Jr.; Mendonca Dias, M. H.; Heldman, A. W. Microscopic NMR imaging of the magnetic fields around magnetite particles. Proc. 5th SMRM 1986, 229–230. 14. Shapiro, E. M.; Skrtic, S.; Sharer, K.; Hill, J. M.; Dunbar, C. E.; Koretsky, A. P. MRI detection of single particles for cellular imaging. Proc. Natl. Acad. Sci. U.S.A. 2004, 101, 10901–10906. 15. Shapiro, E. M.; Skrtic, S.; Koretsky, A. P. Sizing it up: cellular MRI using micron-sized iron oxide particles. Magn. Reson. Med. 2005, 53, 329–338. 16. Dodd, S. J.; Williams, M.; Suhan, J. P.; Williams, D. S.; Koretsky, A. P.; Ho, C. Detection of single mammalian cells by high-resolution magnetic resonance imaging. Biophys. J. 1999, 76, 103–109. 17. Hinds, K. A.; Hill, J. M.; Shapiro, E. M.; Laukkanen, M. O.; Silva, A. C.; Combs, C. A.; Varney, T. R.; Balaban, R. S.; Koretsky, A. P.; Dunbar, C. E. Highly efficient endosomal labeling of progenitor stem cells with large magnetic particles allows magnetic resonance imaging of single cells. Blood 2003, 102, 867–872. 18. Foster-Gareau, P.; Heyn, C.; Alejski, A.; Rutt, B. K. Imaging single mammalian cells with a 1.5 T clinical MRI scanner. Magn. Reson. Med. 2003, 49, 968–971. 19. Shapiro, E. M.; Sharer, K.; Skrtic, S.; Koretsky, A. P. In vivo detection of single cells by MRI. Magn. Reson. Med. 2006, 55, 242–249. 20. Shapiro, E. M.; Gonzalez-Perez, O.; Garcia-Verdugo, J. M.; Alvarez-Buylla, A.; Koretsky, A. P. Magnetic resonance imaging of the migration of neuronal precursors generated in the adult rodent brain. Neuroimage 2006, 32, 1150–1157. 21. Wu, Y. L.; Ye, Q.; Foley, L. M.; Hitchens, T. K.; Sato, K.; Williams, J. B.; Ho, C. In situ labeling of immune cells with iron oxide particles: an approach to detect organ rejection by cellular MRI. Proc. Natl. Acad. Sci. U.S.A. 2006, 103, 1852–1857. 22. Bulte, J. W. M.; Kraitchman, D. L. Iron oxide MR contrast agents for molecular and cellular imaging. NMR Biomed. 2004, 17, 484–499. 23. Frank, J. A.; Miller, B. R.; Arbab, A. S.; Zywicke, H. A.; Jordan, E. K.; Lewis, B. K.; Bryant, L. H.; Bulte, J. W. M. Clinically applicable labeling of mammalian and stem cells by combining superparamagnetic iron oxides and transfection agents. Radiology 2003, 228, 480–487. 24. Sipkins, D. A.; Cheresh, D. A.; Kazemi, M. R.; Nevin, L. M.; Bednarski, M. D.; Li, K. C. Detection of tumor angiogenesis in vivo by ␣V3-targeted magnetic resonance imaging. Nat. Med. 1998, 4, 623–626. 25. Yu, X.; Song, S.-K.; Chen, J.; Scott, M. J.; Fuhrhop, R. J.; Hall, C.-S.; Gaffney, P.-J.; Wickline, S. A.; Lanza, G. M. High-resolution MRI characterization of human thrombus using a novel fibrin-targeted paramagnetic nanoparticle contrast agent. Magn. Reson. Med. 2000, 44, 867–872. 26. Flacke, S.; Fischer, S.; Scott, M. J.; Fuhrhop, R. J.; Allen, J. S.; McLean, M.; Winter, P.; Sicard, G. A.; Gaffney, P. J.; Wickline, S. A.; Lanza, G. M. Novel MRI contrast agent for molecular imaging of fibrin: implications for detecting vulnerable plaques. Circulation 2001, 104, 1280–1285. 27. Louie, A. Y.; H¨uber, M. M.; Ahrens, E. T.; Rothb¨acher, U.; Moats, R.; Jacobs, R. E.; Fraser, S. E.; Meade, T. J. In vivo visualization of gene expression using magnetic resonance imaging. Nat. Biotechnol. 2000, 18, 321–325. 28. Weissleder, R.; Moore, A.; Mahmood, U.; Bhorade, R.; Benveniste, H.; Chiocca, E. A.; Basilion, J. P. In vivo magnetic resonance imagine of transgene expression. Nat. Med. 2000, 6, 351–354.
P1: OTA/XYZ P2: ABC c16 JWBS037-Chen January 25, 2011
13:5
Printer Name: Yet to Come
REFERENCES
395
29. Genove, G.; DeMarco, U.; Xu, H.; Goins, W. F.; Ahrens, E. T. A new transgene reporter for in vivo magnetic resonance imaging. Nat. Med. 2005, 11, 450–454. 30. Gilad, A. A.; McMahon, M. T.; Walczak, P.; Winnard, P. T., Jr.; Raman, V.; van Laarhoven, W. M.; Skoglund, C. M.; Bulte, J. W. M.; van Zijl, P. C. M. Artificial reporter gene providing MRI contrast based on proton exchange. Nat. Biotechnol. 2007, 25, 217–219. 31. Gilad, A. A.; Ziv, K.; McMahon, M. T.; van Zijl, P. C. M.; Neeman, M.; Bulte, J. W. M. MRI reporter genes. J. Nucl. Med. 2008, 49, 1905–1908. 32. Perez, J. M.; Josephson, L.; O’Loughlin, T.; H¨ogemann, D.; Weissleder, R. Magnetic relaxation switches capable of sensing molecular interactions. Nat. Biotechnol. 2002, 20, 816–820. 33. Sun, E. Y.; Weissleder, R.; Josephson, L. Continuous analyte sensing with magnetic nanoswitches. Small 2006, 2, 1144–1147. 34. Tanimoto, A.; Pouliquen, D.; Kreft, B. P.; Stark, D. D. Effects of spatial distribution on proton relaxation enhancement by particulate iron oxide. J. Magn. Reson. Imaging 1994, 4, 653–657. 35. Hyeon, T. Chemical synthesis of magnetic nanoparticles. Chem. Commun. 2003, 927–934. 36. Gupta, A. K.; Gupta, M. Synthesis and surface engineering of iron oxide nanoparticles for biomedical applications. Biomaterials 2005, 26, 3995–4021. 37. Nelson, K. L.; Runge, V. M. Basic principles of MR contrast. Topics Magn. Reson. Imaging 1995, 7(3), 124–136. 38. Merbach, A. E., T´oth, E., Eds. The Chemistry of Contrast Agents in Medical Magnetic Resonance Imaging. Wiley: Hoboken, NJ, 2001. 39. Bottrill, M.; Kwok, L.; Long, N. J. Lanthanides in magnetic resonance imaging. Chem. Soc. Rev. 2006, 35, 557–571. 40. Mulder, W. J. M.; Strijkers, G. J.; van Tilborg, G. A. F.; Griffioen, A. W.; Nicolay, K. Lipidbased nanoparticles for contrast-enhanced MRI and molecular imaging. NMR Biomed. 2006, 19, 142–164. 41. Woods, M.; Woessner, D. E.; Sherry, A. D. Paramagnetic lanthanide complexes as PARACEST agents for medical imaging. Chem. Soc. Rev. 2006, 35, 500–511. 42. Bulte, J. W. M., Modo, M. M. J., Eds. Nanoparticles in Biomedical Imaging. Springer: New York, 2007. 43. Sosnovik, D. E.; Weissleder, R. Emerging concepts in molecular MRI. Curr. Opin. Biotechnol. 2006, 18, 4–10. 44. Lanza, G. M.; Winter, P. M.; Neubauer, A. M.; Caruthers, S. D.; Hockett, F. D.; Wickline, S. A. 1 H/19 F magnetic resonance molecular imaging with perfluorocarbon nanoparticles. Curr. Top. Dev. Biol. 2005, 70, 57–76. 45. Jaschke, W.; Debbage, P. Molecular imaging with nanoparticles: giant roles for dwarf actors. Histochem. Cell Biol. 2008, 130, 845–875. 46. Zabow, G.; Dodd, S.; Moreland, J.; Koretsky, A. Micro-engineered local field control for highsensitivity multispectral MRI. Nature 2008, 453, 1058–1063. 47. Zabow, G.; Koretsky, A. P.; Moreland, J. Design and fabrication of a micromachined multispectral magnetic resonance imaging agent. J. Micromech. Microeng. 2009, 19, 025020. 48. Zabow, G.; Dodd, S.; Moreland, J.; Koretsky, A. Fabrication of uniform cylindrical nanoshells and their use as spectrally tunable MRI contrast agents. Nanotechnology, 2009, 20, 385301. 49. Calamante, F.; Thomas, D. L.; Pell, G. S.; Wiersma, J.; Turner, R. Measuring cerebral blood flow using magnetic resonance imaging techniques. J. Cereb. Blood Flow Metab. 1999, 19, 701–735. 50. Li, W. H.; Fraser, S. E.; Meade, T. J. A calcium sensitive magnetic resonance imaging contrast agent. J. Am. Chem. Soc. 1999, 121, 1413–1414. 51. Moats, R. A.; Fraser, S. E.; Meade, T. J. A “smart” magnetic resonance imaging agent that reports on specific enzymatic activity. Angew. Chem. Int. Ed. 1997, 36, 726–728.
P1: OTA/XYZ P2: ABC c16 JWBS037-Chen January 25, 2011
396
13:5
Printer Name: Yet to Come
MICROFABRICATED MULTISPECTRAL MRI CONTRAST AGENTS
52. Koretsky, A. P.; Silva, A. C. Manganese-enhance magnetic resonance imaging. NMR Biomed. 2004, 17, 527–531. 53. Silva, A. C.; Lee, J. H.; Aoki, I.; Koretsky, A. P. Manganese-enhanced magnetic resonance imaging (MEMRI): methodological and practical considerations. NMR Biomed. 2004, 17, 532–543. 54. Shen, T.; Weissleder, R.; Papisov, M.; Bogdanov, A. Jr.; Brady, T. J. Monocrystalline iron oxide nanocompounds (MION): physicochemical properties. Magn. Reson. Med. 1993, 29, 599–604. 55. Weissleder, R.; Elizondo, G.; Wittenberg, J.; Rabito, C. A.; Bengele, H. H.; Josephson, L. Ultrasmall superparamagnetic iron oxide: characterization of a new class of contrast agents for MR imaging. Radiology 1990, 175, 489–493. 56. Bulte, J. W. M.; Douglas, T.; Witwer, B.; Zhang, S.-C.; Strable, E.; Lewis, B. K.; Zywicke, H.; Miller, B.; van Gelderen, P.; Moskowitz, B. M.; Duncan, I. D.; Frank, J. A. Magnetodendrimers allow endosomal magnetic labeling and in vivo tracking of stem cells. Nat. Biotechnol. 2001, 19, 1141–1147. 57. Jung, C. W.; Jacobs, P. Physical and chemical properties of superparamagnetic iron oxide MR contrast agents: ferumoxides, ferumoxtran, ferumoxsil. Magn. Reson. Imaging 1995, 13, 661–674. 58. Wang, Y. X.; Hussain, S. M.; Krestin, G. P. Superparamagnetic iron oxide contrast agents: physicochemical characteristics and applications in MR imaging. Eur. Radiol. 2001, 11, 2319–2331. 59. Weissleder, R.; Reimer, R.; Lee, A. S.; Wittenberg, J.; Brady, T. J. MR receptor imaging— ultrasmall iron-oxide particles targeted to asialoglycoprotein receptors. Am. J. Roentgenol. 1990, 155, 1161–1167. 60. Gilad, A. A.; Walczak, P.; McMahon, M. T.; Na, H. B.; Lee, J. H.; An, K.; Hyeon, T.; van Zijl, P. C. M.; Bulte, J. W. M. MRI tracking of transplanted cells with “positive contrast” using manganese oxide nanoparticles. Magn. Reson. Med. 2008, 60, 1–7. 61. Bridot, J.-L.; Faure, A.-C.; Laurent, S.; Rivi`ere, C.; Billotey, C.; Hiba, B.; Janier, M.; Josserand, V.; Coll, J-L.; Vander Elst, L.; Muller, R.; Roux, S.; Perriat, P.; Tillement, O. Hybrid gadolinium oxide nanoparticles: multimodal contrast agents for in vivo imaging. J. Am. Chem. Soc. 2007, 129, 5076–5084. 62. Seppenwoolde, J.-H.; Viergever, M. A.; Bakker, C. J. G. Passive tracking exploiting local signal conservation: the white marker phenomenon. Magn. Reson. Med. 2003, 50, 784–790. 63. Cunningham, C. H.; Arai, T.; Yang, P. C.; McConnell, M. V.; Pauly, J. M.; Conolly, S. M. Positive contrast magnetic resonance imaging of cells labeled with magnetic nanoparticles. Magn. Reson. Med. 2005, 53, 999–1005. 64. Shapiro, E. M.; Koretsky, A. P. Convertible manganese contrast for molecular and cellular MRI. Magn. Reson. Med. 2008, 60, 265–269. 65. Ward, K. M.; Aletras, A. H.; Balaban, R. S. A new class of contrast agents for MRI based on proton chemical exchange dependent saturation transfer (CEST). J. Magn. Reson. 2000, 143, 79–87. 66. McMahon, M. T.; Gilad, A. A.; DeLiso, M. A.; Cromer Berman, S. M.; Bulte, J. W. M.; van Zijl, P. C. M. New “multicolor” polypeptide diamagnetic chemical exchange saturation transfer (DIACEST) contrast agents for MRI. Magn. Reson. Med. 2008, 60, 803–812. 67. Zhang, S.; Merritt, M.; Woessner, D. E.; Lenkinski, R. E.; Sherry, A. D. PARACEST agents: modulating MRI contrast via water proton exchange. Acc. Chem. Res. 2003, 36, 783–790. 68. Schr¨oder, L.; Lowery, T. J.; Hilty, C.; Wemmer, D. E.; Pines, A. Molecular imaging using a targeted magnetic resonance hyperpolarized biosensor. Science 2006, 314, 446–449. 69. Aime, S.; Delli Castelli, D.; Terreno, E. Highly sensitive MRI chemical exchange saturation transfer agents using liposomes. Angew. Chem. Int. Ed. 2005, 44, 5513–5515. 70. Madou, M. J. Fundamentals of Microfabrication: The Science of Miniaturization, 2nd ed. CRC Press: Boca Raton, FL, 2002.
P1: OTA/XYZ P2: ABC c16 JWBS037-Chen January 25, 2011
13:5
Printer Name: Yet to Come
REFERENCES
397
71. Runge, V. M.; Wells, J. W. Update: safety, new applications, new MR agents. Topics Magn. Reson. Imaging 1995, 7(3), 181–195. 72. Shellock, F. G.; Kanal, E. Safety of magnetic resonance imaging contrast agents. J. Magn. Reson. Imaging 1999, 10, 477–484. 73. Mendonca Dias, M. H.; Lauterbur, P. C. Ferromagnetic particles as contrast agents for magnetic resonance imaging of liver and spleen. Magn. Reson. Med. 1986, 3, 328–330. 74. Renshaw, P. F.; Owen, C. S.; McLaughlin, A. C.; Frey, T. G.; Leigh, J. S., Jr. Ferromagnetic contrast agents: a new approach. Magn. Reson. Med. 1986, 3, 217–225. 75. Chan, W. C. W.; Nie, S. Quantum dot bioconjugates for ultrasensitive nonisotopic detection. Science 1998, 281, 2016–2018. 76. Bruchez, M. Jr.; Moronne, M.; Gin, P.; Weiss, S.; Alivisatos, A. P. Semiconductor nanocrystals as fluorescent biological labels. Science 1998, 281, 2013–2016. 77. Alivisatos, P. The use of nanocrystals in biological detection. Nat. Biotechnol. 2004, 22, 47–52. 78. Grad, J.; Bryant, R. G. Nuclear magnetic cross-relaxation spectroscopy. J. Magn. Reson. 1990, 90, 1–8. 79. Henkelman, R. M.; Stanisz, G. J.; Graham, S. J. Magnetization transfer in MRI: a review. NMR Biomed. 2001, 14, 57–64. 80. Zurkiya, O.; Hu, X. Off-resonance saturation as a means of generating contrast with superparamagnetic nanoparticles. Magn. Reson. Med. 2006, 56, 726–732. 81. Kotzar, G.; Freas, M.; Abel, P.; Fleischman, A.; Roy, S.; Zorman, C.; Moran, J. M.; Melzak, J. Evaluation of MEMS materials of construction for implantable medical devices. Biomaterials 2002, 23, 2737–2750. 82. Voskerician, G.; Shive, M. S.; Shawgo, R. S.; von Recum, H.; Anderson, J. M.; Cima, M. J.; Langer, R. Biocompatibility and biofouling of MEMS drug delivery devices. Biomaterials 2003, 24, 1959–1967.
P1: OTA/XYZ P2: ABC c17 JWBS037-Chen January 25, 2011
13:16
Printer Name: Yet to Come
CHAPTER 17
Radiolabeled Nanoplatforms: Imaging Hot Bullets Hitting Their Target RAFFAELLA ROSSIN Department of Biomolecular Engineering, Philips Research Europe, Eindhoven, The Netherlands
Nanomedicine is a rapidly evolving field. Every year, new and more sophisticated nanoplatforms (NPs) designed to deliver drugs to specific targets or for molecular imaging of disease states make their appearance on the scientific scene. The assessment of the in vivo distribution of these NPs is the first step in order to fully understand their potential as candidate nanomedicines and to predict and prevent toxic side effects. To this aim, the incorporation of a radionuclide in the NP is of great advantage, as radiometric assays are rapid and quantitative. Furthermore, noninvasive nuclear imaging techniques are highly sensitive, achieve high tissue penetration and good spatial resolution, and can accelerate the translation of new NPs into the clinic. This chapter gives an overview of the strategies to label NPs with a variety of radionuclides (radiometals and radiohalogens), focusing on some of the hurdles that scientists must overcome (or take into consideration) when administering these radioconjugates in vivo. Furthermore, the latest achievements in molecular imaging of angiogenesis with radiolabeled NPs are reviewed. The most recent publications on nanoplatforms labeled with ␥ - and + -emitting radionuclides for SPECT and PET imaging are reviewed. Different strategies to label nanoplatforms are described, particularly those involving radiometals (64 Cu, 111 In, 67/68 Ga, 86 Y, 99m Tc, and 186/188 Re) and radiohalogens (123/124/125/131 I, 76 B, and 18 F). Some of the problems occurring when evaluating radiolabeled nanoplatforms in vivo are also covered. Finally, the use of radiolabeled nanoplatforms for angiogenesis imaging with PET and SPECT is discussed.
ABBREVIATIONS %ID/g, percent injected dose per gram tissue; 111 In-acac, 111 In-acetylacetonate; FpyME, 1-[3-(2-[18 F]fluoropyridin-3-yloxy)propyl]pyrrole-2,5-dione; 18 FDG, 2-deoxy2-[18 F]fluoro-d-glucose; BFC, bifunctional chelating agent; BMEDA, N,N-bis(2-mercaptoethyl)-N ,N -diethyl-ethylenediamine; c(RGDyK), cyclo(Arg-Gly-Asp-d-Tyr-Lys); 18
Nanoplatform-Based Molecular Imaging Edited by Xiaoyuan Chen C 2011 John Wiley & Sons, Inc. Copyright
399
P1: OTA/XYZ P2: ABC c17 JWBS037-Chen January 25, 2011
400
13:16
Printer Name: Yet to Come
RADIOLABELED NANOPLATFORMS: IMAGING HOT BULLETS HITTING THEIR TARGET
CAT, Chloramine-T; CB-DO2A, 4,10-bis(carboxymethyl)-1,4,7,10-tetraazabicyclo[5.5.2] tetradecane; CB-TE2A, 4,11-bis(carboxymethyl)-1,4,8,11-tetraazabicyclo[6.6.2]hexadecane; CNT, carbon nanotube; CT, computed tomography; DF, deferoxamine; DOTA, 1,4,7,10-tetraazacyclododecane-1,4,7,10-tetraacetic acid; Doxil, PEGylated liposomal doxorubicin; DTPA, diethylenetriaminepentaacetic acid; EC, electron capture; ECM, extracellular matrix; EPR, enhanced permeability and retention; HMPAO, hexamethyl propyleneamine oxime; HPMA, 2-(hydroxypropyl)methacrylamide; HYNIC, hydrazino nicotinamide; IO, iron oxide particle; IT, isomeric transition; MB, microbubble; MPS, mononuclear phagocyte system; MRI, magnetic resonance imaging; MWNT, multiwall carbon nanotube; NIRF, near-infrared fluorescence; NP, nanoplatform; PAMAM, poly(amido amino); PE, phosphatidylethanolamine; PEG, polyethylene glycol; PET, positron emission tomography; PPI, poly(propylene imine); QD, quantum dot; RGD, Arg-Gly-Asp; SarAr, 1-N-(4-aminobenzyl)-3,6,10,13,16,19-hexaazabicyclo[6.6.6]eicosane-1,8-diamine; siRNA, small interfering RNA; SOD, superoxide dismutase; SPECT, single photon emission computed tomography; SPIO, superparamagnetic iron oxide particle; SWNT, single-wall carbon nanotube; SWNTol, hydoxylated single-wall carbon nanotube; TETA, 1,4,8,11-tetraazacyclotetradecane-1,4,8,11-tetraacetic acid; US, ultrasound; VCAM-1, vascular cell adhesion molecule-1; VEGF, vascular endothelial growth factor; VEGFR-2, vascular endothelial growth factor receptor-2; + , positron.
17.1 INTRODUCTION Recent breakthroughs revealing the mechanisms of disease that have emerged from genomics and proteomics prompted great advances in the design of new therapy strategies. It was Paul Ehrlich who, at the beginning of the twentieth century, conceived the idea of a “magic bullet” capable of selectively hitting the diseased cells [1]. In the clinic, however, this concept is hampered by the fact that most small molecule drugs distribute throughout the entire body and few are perfectly selective in terms of action. This results in intrinsic toxic side effects when the drug has a cytotoxic action and/or a suboptimal efficacy. Therefore the design of “smart” drug carriers has become an integral part of drug development [2–5]. Such vehicles should be biocompatible, encapsulate a therapeutic cargo, and release it only at the site of action in a controlled and timely fashion and, possibly, provide a feedback to the clinician. Composite NPs are ideal candidates for this multitask assignment as they can carry multiple functionalities (targeting ligands and detectable labels among others). In addition, some NPs have intrinsic cytotoxic properties [6,7] or can be chemically engineered to degrade and release a therapeutic payload when triggered by the microenvironment (e.g., pH, enzymes) or by external stimuli (e.g., temperature, pressure, light) [8–10]. Although the first liposomes were described in 1965 [11] and the first controlled release of a protein from a polymer was described in 1976 [12], the term “nanomedicine”—the application of nanotechnology to healthcare—was introduced only recently [13]. The properties of drug carriers need constant tuning to meet healthcare needs. For this reason, nowadays the development of new “magic nanobullets” entails the combined multidisciplinary efforts of chemists, material scientists, biologists, engineers, mathematicians, and clinicians. Some examples of advanced carrier NPs include late-generation dendrimers,
P1: OTA/XYZ P2: ABC c17 JWBS037-Chen January 25, 2011
13:16
Printer Name: Yet to Come
INTRODUCTION
401
core–shell nanoparticles, CNTs, polymersomes, bubbles, multilayered polyelectrolyte assemblies, and mesoporous silica nanoparticles. The literature relating to new and more sophisticated NPs increases year by year. Biological data is provided in many of these studies, but the vast majority of experiments are carried out in vitro while these candidate drug vectors must perform their multiple tasks (reaching a target tissue or organ, delivering their intact cargo, and clear) in a living organism. The body is well protected from external agents by layers of epithelial cells that line the skin, the airways, the gastrointestinal tract, and the genital mucosa. Low molecular weight agents (<500 Da) can penetrate these barriers but, in general, macromolecular drugs and NPs cannot pass through the intercellular junctions [14]. To bypass these hurdles, most drug delivery systems are injected parenterally (intravenously, intraperitoneally, or subcutaneously). However, the macrophages of the MPS are very efficient in capturing circulating “foreign objects” (i.e., senescent cells, macromolecules, and particulates) and often NPs are eliminated from the blood circulation within minutes from injection and end up in the liver, spleen, and lymph nodes. The blood clearance was shown to depend on charge, size, and surface chemistry of the NP [15]. Significant improvements in biological half-life were obtained by tampering with the opsonization and complement activation processes driving macrophage recognition, for instance, by covering the NP surface with PEG [15–19]. However, eventually all injected NPs are eliminated from the circulation and clear mainly through the liver and intestine or through the kidneys (the size limits for glomerular filtration are 70 kDa for linear polymers [20] and 3–5 nm in diameter for globular assemblies [21]). Knowing the bioavailability of a candidate drug carrier is crucial, especially for cancer therapy, as the active (e.g., receptor mediated) or passive targeting (driven by the EPR effect [22]) of a tumor may require time. Also, the NP distribution and residence time in nontarget tissues need to be evaluated to predict and minimize toxic side effects in sensitive organs. Finally, the evaluation of candidate NPs in living organisms gives the scientist vital information on the effects of different administration routes on the kinetics and specific pathways of NP distribution upon administration in vivo [23–26]. The localization of a NP at the site of interest is often accomplished by conjugating multiple targeting moieties (peptides, antibody fragments, receptor binding small molecules, etc.) to its surface. Information on the NP localization within a living system (from a living cell to the entire body) can be obtained with multiple imaging modalities by including detectable functions in the nanostructure or by exploiting intrinsic imageable characteristics of the NP. This adds diagnostic capability to the therapeutic action of the NP, leading to the emerging field of “theranostic” (therapeutic + diagnostic) agents [27]. To date, the specific imaging capabilities of targeting NPs have been utilized mainly to visualize vascular markers of disease (e.g., integrins and VCAM-1 on activated endothelial cells, macrophages in atherosclerotic plaques), which are easily accessible to blood-borne bulky structures without the need of extravasation from blood vessels. In fact, the accumulation of macromolecular structures within the interstitial space of solid tumors is mostly driven by passive diffusion (the EPR effect) and not by specific interaction with tumor cell targets [28–30]. The application of IOs for MRI, QDs for optical imaging micro- and nanobubbles for US imaging, perfluorocarbon nanoparticles for 19 F nuclear magnetic resonance spectroscopy, and more are discussed in detail in other chapters of this book. This chapter focuses on the achievements in the use of radioactive labels to track NPs in vivo.
P1: OTA/XYZ P2: ABC c17 JWBS037-Chen January 25, 2011
402
13:16
Printer Name: Yet to Come
RADIOLABELED NANOPLATFORMS: IMAGING HOT BULLETS HITTING THEIR TARGET
17.2 NUCLEAR TECHNIQUES AND RADIONUCLIDES FOR NANOPLATFORM IN VIVO TRACKING Classic “cut and count” radiometric assays have been used for decades to assess the biodistribution of polymers and NPs labeled with long-lived radionuclides such as 3 H, 14 C, and 125 I [25, 31, 32]. More recently, high-resolution (<100 m) autoradiography equipment and preclinical PET and SPECT have entered research laboratories worldwide, making it easier to evaluate the distribution of radiolabeled compounds in normal animals and disease models. Nuclear imaging modalities hold a key position among the noninvasive imaging techniques because of their exquisite sensitivity [33]. Microdosing combined with good spatial resolution (from 0.5–1 mm in small animal scanners to 5–10 mm in clinical scanners [34]) and the ability to provide quantitative information on tracer accumulation in deep tissues allow a rapid translation of investigational tracers from mouse to human. Furthermore, the use of noninvasive nuclear imaging techniques for longitudinal studies speeds up the drug development process while keeping the costs to a reasonable level, especially when pricey disease models and transgenic animals are used. The choice of radionuclide for NP labeling depends on the imaging technique and on the timing for in vivo evaluation. The radionuclides used in PET imaging decay by emitting + particles, which travel in tissue until they collide with one electron generating two back-to-back 511-keV photons nearly 180◦ apart (annihilation). The detection of all pairs of photons hitting opposite detectors in coincidence allows the construction of an image defining the position of the annihilation event. On the contrary, SPECT uses single photons emitted by ␥ -emitting radionuclides. These photons travel through the tissue and hit the scanner detector. A collimator (usually a lead adsorbing plate with parallel holes or a pinhole collimator) selects the photons along defined axes and stops the others, therefore allowing the construction of a projection image defining the tracer distribution in the scanned object. Multiple projections, acquired sequentially, are then reconstructed in a three-dimensional (3D) tomographic image. The different physics underlying PET and SPECT imaging lead to intrinsic strengths and weaknesses of these two techniques. For instance, the availability of + -emitting radionuclides for PET imaging can be a limiting factor, especially for short-lived cyclotronproduced radionuclides, which must be produced in-house. On the contrary, most ␥ emitters are relatively long-lived and commercially available. Furthermore, the detection of 511-keV annihilation photons does not allow imaging different + -emitting radiotracers at the same time with PET while the simultaneous detection of photons with different energies can be performed with SPECT by choosing the appropriate energy window in the scanner. However, the need for a collimator reduces significantly (up to two orders of magnitude [34]) the sensitivity of SPECT scanners compared to PET scanners. Overall, the high detection sensitivity, deep tissue penetration, and quantitative accuracy of PET and SPECT make them powerful techniques for elucidating physiology, metabolic pathways, and molecular targets in vivo. Therefore great effort has been dedicated to integrating these techniques with those providing detailed structural information, such as CT and MRI [35]. As a result, SPECT/CT and PET/CT systems have become available over the past years and, recently, the scientific community has welcomed the introduction of the first integrated PET/MR scanners [36]. Functionalized nanocarriers are ideal platforms for multimodality imaging, as they can produce enough contrast at the target to effect a signal change detectable with CT, MRI, or US imaging while carrying also labels for nuclear or optical imaging. Consequently, the high resolution anatomical information delivered by CT,
P1: OTA/XYZ P2: ABC c17 JWBS037-Chen January 25, 2011
13:16
Printer Name: Yet to Come
RADIONUCLIDES AND STRATEGIES FOR NANOPLATFORM RADIOLABELING
403
MRI, and US imaging can be used to localize the target, while the functional information provided by more sensitive techniques such as PET, SPECT, and optical imaging, can be used to detect the release of the therapeutic payload or to monitor the therapeutic effects.
17.3 RADIONUCLIDES AND STRATEGIES FOR NANOPLATFORM RADIOLABELING The radionuclides most commonly used to track NPs in living systems are listed in Table 17.1 (␥ -emitters) and Table 17.2 (+ -emitters). The selection of the right radionuclide for an in vivo assay depends on several factors such as the technique of choice, the radionuclide availability, the experimental time window, and the radiochemistry. The need to detect enough counts to differentiate the target tissue from the background noise, especially for imaging, limits the use of ␥ - and + -emitting radionuclides to approximately four decay half-lives. Therefore the physical life of the radionuclide must match the biological life of the NP and the time window of the biological process under scrutiny. Short-lived radionuclides such as 18 F and 68 Ga find an application with short circulating NPs or in dynamic imaging studies early after NP injection. On the contrary, the prolonged evaluation of long-circulating NPs is usually carried out with medium- to long-lived radionuclides such as 64 Cu, 111 In, and others. The energy of the radionuclide emission is also a key issue. High-energy + emitted by many “nonstandard” PET radionuclides (i.e., + -emitters other than 18 F, 11 C, 15 O, and 13 N) travel for a relatively long distance (a few millimeters) in the target tissue before annihilating with an electron. Some of these radionuclides (e.g., 66 Ga) also emit highenergy ␥ rays in cascade during the decay. These two factors reduce the quality of the PET images and complicated algorithms must be used for image reconstruction [37–39]. The emission energy is a limitation also for SPECT imaging. In fact, low-energy photons (<100 keV) produce too much scatter in tissues while ␥ energies greater than 200 keV are difficult to collimate, therefore posing a challenge to image resolution and quantitation. TABLE 17.1 Physical Dataa for Useful ␥-Emitting Radionuclides for NP Radiolabeling Decay Mode (%)
Main ␥ keV (%)
max MeV (ave )
140 (89.1) 137 (9.5)
1.069 (0.347)
17.0 h 13.2 h 59.4 days
IT − (93) EC (7) − EC EC
I
8.0 days
−
Ga
3.3 days
EC
2.8 days
EC
Radionuclide
Half-life
99m
Tc 186 Re
6.0 h 3.7 days
188
Re I 125 I 131
123
67
111
In
155 (15.6) 159 (83.3) 27 (74.0); 31 (13.2); 35 (6.7) 284 (6.1); 364 (81.5); 636 (7.2) 93 (39.2); 185 (21.4); 300 (16.6) 171 (90.7); 245 (94.1)
Production 99
2.120 (0.763)
0.807 (0.182)
Mo/99m Tc generator Re(n,␥ )186 Re 186 W(p,n)186 Re 188 W/188 Re generator 124 Te(p,2n)123 I 124 Xe(n,␥ )125 Xe → 125 I 185
130
68
Te(n,␥ )131 Te → 131 I
Zn(p,2n)67 Ga
111 112
Cd(p,n)111 In Cd(p,2n)111 In
a Some data are obtained from http://www.nndc.bnl.gov (searchable National Nuclear Data Center, Brookhaven National Laboratory) (accessed October 1, 2009).
P1: OTA/XYZ P2: ABC c17 JWBS037-Chen January 25, 2011
404
13:16
Printer Name: Yet to Come
RADIOLABELED NANOPLATFORMS: IMAGING HOT BULLETS HITTING THEIR TARGET
TABLE 17.2 Physical Dataa for Useful + -Emitting Radionuclides for NP Radiolabeling
Radionuclide Half-life 64
Cu
12.7 h
68
Ga
67.7 min
86
Y
14.7 h
124
I
76
Br
18
F
Decay Mode (%)
Main ␥ keV (%)
+ (18) 511 (35.2) EC (44) − (38) + (89) 511 (178.3)
EC (11) + (33) EC (67) 4.2 days + (23) EC (77) 16.2 h + (55) EC (45) 109.8 min + (97) EC (3)
511 (64); 628 (32.6); 1077 (85.2); 1153 (30.5) 511 (45); 603 (62.9); 723 (10.4) 511 (109); 559 (74.0); 657 (15.9); 1854 (14.7) 511 (193.5)
max MeV (ave )
Mean + Range (mm)
Production
0.653 (0.278) 0.70
64
Ni(p,n)64 Cu
1.899 (0.829) 2.40
68
Ge/68 Ga generator
3.141 (0.660) 2.46
86
Sr(p,n)86 Y
2.138 (0.820) 3.24
124
Te(p,n)124 I Te(d,2n)124 I 76 Se(p,n)76 Br 76 Se(d,2n)76 Br 18 O(p,n)18 F 20 Ne(d,␣)18 F 124
3.941 (1.180) 5.07 0.633 (0.250) 0.69
a Some
data are obtained from http://www.nndc.bnl.gov (searchable National Nuclear Data Center, Brookhaven National Laboratory) (accessed on October 1, 2009). See also Nayak and Brechbiel [85].
The stability of the radiolabeled species is yet another aspect of paramount importance when evaluating its behavior in a living system. For instance, some radiometals bind with relatively high affinity to serum proteins (e.g., transferrin) and therefore they need to be stably anchored to the molecule under evaluation. The metabolic breakdown of radiolabeled species may also produce radiocatabolites with a different distribution with respect to that of the parent molecule. For these reasons, assessing the integrity of a radiolabeled species is crucial when using nuclear techniques to track its distribution in a living system, as these techniques do not discriminate between the intact species and the degradation products. For the same reason, extra care needs to be taken in order to administer in vivo NP solutions free of radioactive impurities such as free radionuclides, unbound radiometal chelates, and radioactive by-products. A comprehensive review of all the different methods to radiolabel NPs is beyond the scope of this chapter due to the variety of nanostructures in preclinical evaluation and the diversity of radiolabeling approaches used to this aim. Detailed methods to radiolabel liposomes [40] and general procedures for specific radionuclides (64 Cu and 76 Br) [41] were published recently. The following sections contain an overview of the most recent publications on NPs labeled with radionuclides that form coordinative bonds (radiometals) and covalent bonds (radiohalogens). The properties of each radionuclide and the radiochemistry are discussed briefly. The discussion focuses also on some specific problems that have occurred during radiolabeling and/or in vivo evaluation. 17.3.1 Copper Radiolabeling Among the + -emitting radionuclides, 64 Cu is becoming the radionuclide of choice for assessing the in vivo distribution and targeting capabilities of NPs with small animal PET imaging. Copper-64 is produced with high specific activity on medical cyclotrons [42] and decays by emitting + (18%) and − particles (38%). Hence 64 Cu has potential for both molecular imaging with PET and radiotherapy. From the imaging point of view, the
P1: OTA/XYZ P2: ABC c17 JWBS037-Chen January 25, 2011
13:16
Printer Name: Yet to Come
RADIONUCLIDES AND STRATEGIES FOR NANOPLATFORM RADIOLABELING
COOH
HOOC N
N N HOOC
405
N COOH
N
COOH
COOH
NOTA
DTPA
HOOC
HOOC
COOH N
N
N
N
N
HOOC
COOH
HOOC
COOH
COOH N
N
N
N
HOOC
COOH
TETA
DOTA
COOH
COOH N
N
N
N
N
N
N
N
HOOC
HOOC
CB-TE2A
CB-DO2A
NH H2N
HN
NH
HN
NH
HN
NH2
diamSar
FIGURE 17.1 Representative chelators for radiometal complexation.
low energy of the + emitted by 64 Cu produces images with superb spatial resolution (0.70-mm mean + range). In addition, due to the convenient half-life (12.7 hours), 64 Culabeled tracers can be tracked in vivo for 24–48 hours. Besides 64 Cu, the family of Cu radionuclides includes other isotopes currently under investigation for PET (60 Cu, 61 Cu, and 62 Cu) and radiotherapy (67 Cu) applications. For this reason, the chemistry of radio-Cu applied to the labeling of small molecules, peptides, and antibodies has been thoroughly investigated [43–45]. A variety of NPs were labeled with 64 Cu by using BFCs, such as derivatives of the tetraazamacrocyclic TETA [29, 46, 47] and DOTA [48–61] or the poly-azacarboxylate DTPA [62] (Figs. 17.1 and 17.2), that bind the metal with high affinity. Usually, the BFC is conjugated to the preformed NP before radiolabeling by using reactive functions available on its surface. When using negatively charged nanosystems, the length of the spacer between the chelator and the surface was shown to influence significantly the conjugation yield of the BFC. In fact, the use of short spacers affected the specific activity of the 64 Cu-labeled NP [59]. In this scenario, direct conjugation of the BFCs to the NP building blocks before
P1: OTA/XYZ P2: ABC c17 JWBS037-Chen January 25, 2011
406
13:16
Printer Name: Yet to Come
RADIOLABELED NANOPLATFORMS: IMAGING HOT BULLETS HITTING THEIR TARGET
HOOC
NCS
HOOC HOOC
NCS N N HOOC
N COOH
N
H N
COOH
COOH
HOOC
p-SCN-Bz-DTPA O
N
N
N
Br O
COOH
COOH
HOOC
COOH
BAT COOH
O COOH N
O N O
N
HOOC
N
H N
N COOH
HOOC N
N
CHX-A"-DTPA
HOOC N
COOH
N
COOH
HOOC
DOTA-NHS
N
SCN
N N
N
HOOC
COOH
COOH N COOH
p-SCN-Bz-DOTA
NODAGA
O NH
O
N OH HN O
H2N HO N
O
NH
HN
NH
HN
NH
HN
NH2 N H
O N
Deferoxamine
OH
SarAr
NH2
FIGURE 17.2 Representative BFCs for radiometal complexation.
NP assembly was shown to improve markedly the radiolabeling yields [57]. Nonetheless, with this strategy some chelators end up buried within the core of the NP, where they are not available to bind the radiometal [50, 54, 57]. Jarret and co-workers reported also the conjugation of preformed 64 Cu-DOTA groups onto NPs [51]. This approach, however, is not ideal for the in vivo administration of high specific activity 64 Cu-NPs due to the length of the labeling, conjugation, and purification procedures. Despite the attractive features of 64 Cu, one drawback for using 64 Cu labeling for NP evaluation is the partial instability of Cu complexes with tetraazamacrocyclic chelators. In fact, in vivo transchelation of Cu-DOTA and Cu-TETA has been reported both in rodents [63, 64] and in humans [65, 66]. The precise location and mechanism of copper release from the chelators are still under debate. However, it is known that the release is caused by the reduction of Cu(II) to Cu(I). In liver, this is followed by Cu binding to SOD [63, 64], while in blood the radiometal binds to ceruloplasmin, metallothioneins, and other serum proteins [64, 66]. To overcome this problem, a series of new macrocyclic ligands (the cross-bridged CB-DO2A and CB-TE2A, and diamSar, Fig. 17.1), that form Cu complexes with improved stability, were developed [45]. To date, there are no published data suggesting metabolism of 64 Cu-DOTA-NPs and formation of 64 Cu-SOD in liver. The in vitro stability of a series of 64 Cu-NPs was confirmed for as long as 48 hours in serum [47, 50, 53, 54]. Moreover, Liu and co-workers used the intrinsic optical properties of 64 Cu-DOTA labeled SWNTs and found a correlation between the amount of radioactivity and the amount of SWNTs in mouse liver by using Raman spectroscopy [53]. A thorough evaluation of the in vivo metabolism of a series of polymeric NPs is under way at Washington University in St. Louis. The results of this study will help us understand the fate of 64 Cu-DOTA-NPs in liver.
P1: OTA/XYZ P2: ABC c17 JWBS037-Chen January 25, 2011
13:16
Printer Name: Yet to Come
RADIONUCLIDES AND STRATEGIES FOR NANOPLATFORM RADIOLABELING
407
Mechanisms other than Cu transchelation may also cause degradation of 64 Cu-NPs in vivo. For instance, when using small animal PET to evaluate the distribution of 64 CuDOTA-siRNA particles in mice, Bartlett and co-workers observed a high accumulation of radioactivity in kidneys and bladder shortly after injection [48]. This result was not expected due to the large size of the siRNA NPs (80–120 nm). Therefore the authors investigated the possible release of 64 Cu-DOTA-siRNA from the NP in vitro and observed approximately 50% NP dissociation in 0.4-M NaCl. Given these findings, they hypothesized a salt-mediated disruption of the NP within the kidney as the possible cause for the observed in vivo behavior. Recently, we investigated the possible use of PET to image the accumulation of 64 Cu-labeled immunobeads in the lung endothelium in mice [55]. We investigated the metabolism of the beads in this organ and observed significant dissociation of radioactivity (approximately 50% of the total) from the beads 1 hour after injection. The radioactivity elimination was not due to Cu transchelation; it was more likely due to the release of the noncovalently bound 64 Cu-DOTA-IgG from the surface of the beads and to degradation of the protein. The formation of a small amount of a hydrophilic radiocatabolite was observed also by Liu et al. [53] following the injection of 64 Cu-DOTA-CNTs in mice. In fact, virtually no CNT kidney uptake was observed based on Raman measurements, while some kidney activity was visible in the PET images.
17.3.2 Indium Radiolabeling The cyclotron-produced 111 In (2.8-day half-life) is another commonly used radiometal for NP evaluation in living systems, as it allows one to detect NP in vivo trafficking for several days. Indium-111 decays by EC with two major ␥ emissions (171 and 245 keV) and is widely available through commercial sources. Like 64 Cu, 111 In binds avidly to poly-azacarboxylate chelators such as DTPA and DOTA. For NP radiolabeling, BFCs have been chemically conjugated to the surface [67–70] or appended to hydrophobic domains constituting the nanostructure (i.e., the lipid bilayer of liposomes or the core of lipid NPs [19, 71]). Non-BFC-based methods to label NPs with 111 In were also developed. For instance, Oyewumi et al. [72] labeled Gd NPs by entrapping 111 In-acac (Fig. 17.3), a small lipophilic complex, in the particle core during synthesis by an oil-in-water microemulsion procedure. A similar approach is used to label liposomes after manufacturing (after-loading approach). In fact, 111 In-acac and 111 In-oxine (Fig. 17.3) freely cross the lipid bilayer of liposomes [40, 73]. The premature release of the label in vivo is avoided by including DTPA within the liposome aqueous core, which transchelates the radiometal forming a hydrophilic moiety [40, 74]. Recently, Cartier and co-workers [75] reported also the direct labeling of NPs with 111 InCl3 , 68 GaCl3 , and GdCl3 . Surprisingly, the trivalent metal cations bound quickly to latex nanoparticles in water without the need for a BFC, although with low specific activity (220 kBq/mg NP for both radiometals and 10 g/mg NP for Gd) [75]. This finding was explained with the formation of strong coordination bonds between the metal ions and multiple carboxylic groups on the NP surface. However, the in vivo data reported by Cartier et al. [75] were limited to 1 hour post NP injection. In another interesting study, H¨afeli et al. [76] labeled magnetic microspheres prepared from metallic iron and activated carbon by directly mixing the particles with 111 InCl3 , 111 In-oxine, 111 In-DOTA, and 111 In-Abz-DOTA (an amino benzyl derivative of DOTA). Indium-111-oxine and 111 In-Abz-DOTA adsorbed onto the surface of the microspheres with high efficiency but only 111 In-Abz-DOTA was
P1: OTA/XYZ P2: ABC c17 JWBS037-Chen January 25, 2011
408
13:16
Printer Name: Yet to Come
RADIOLABELED NANOPLATFORMS: IMAGING HOT BULLETS HITTING THEIR TARGET
O
O
O N
O
O
M
O O
N
N O N
O
N
N
O
O
M
O
Tc
N
H
In/Ga-Acac
In/Ga-Oxine
S S
S
Tc-HMPAO
SH
N
M S
S
N O
S
S
M S
N
S N
Tc/Re-SSS
Tc/Re-BMEDA
FIGURE 17.3 Representative complexes used to radiolabel NPs comprising a lipophilic domain.
stably retained on the particle for at least 7 days, both in vitro and in vivo. The authors did not make hypotheses of the interaction mechanisms and did not study further the different release of radiolabels from the particles.
17.3.3 Gallium Radiolabeling Gallium radionuclides are also used to label NPs. The ␥ -emitting 67 Ga is one of the most widespread radiometals in nuclear imaging because of the 185-keV photon emission and relatively long half-life (3.3 days), which allows marketing. Gallium-67 is often used to radiolabel liposomes with convenient after-loading methods [40]. Like 111 In, 67 Ga can pass through lipid bilayers as 67 Ga-oxine (Fig. 17.3) and then it is trapped within the aqueous core of the NP preloaded with a chelator. To this aim, strong chelators such as DF [77] (Fig. 17.2) or DTPA [30, 78] are preferred to weaker ones such as nitriloacetic acid. In fact, in case of release of the liposome content in vivo, Ga-DF and Ga-DTPA undergo fast renal elimination while the nitriloacetate complex exchanges the radiometal with serum proteins [79, 80]. The + -emitter 66 Ga could be a good candidate for in vivo evaluation of NPs with PET because of the 9.5-hour half-life [81]. Unfortunately, 66 Ga has a suboptimal decay for imaging applications, as it emits high-energy + (4.153 MeV) and ␥ rays (1.039 and 2.752 MeV) that degrade the image quality (8.06-mm + mean range) and increase the patient dose [37]. For this reason, 68 Ga is the only gallium radionuclide used for PET imaging nowadays. This is due to its superb imaging characteristics and to the development of a user-friendly
P1: OTA/XYZ P2: ABC c17 JWBS037-Chen January 25, 2011
13:16
Printer Name: Yet to Come
RADIONUCLIDES AND STRATEGIES FOR NANOPLATFORM RADIOLABELING
409
and long-lasting generator that delivers 68 Ga solutions on demand without the need for an in-house cyclotron [82]. However, the 67.7-min half-life of 68 Ga is a drawback for in vivo NP applications because it gives only a few hours of imageable signal. Few studies on 68 Ga-labeled NPs have appeared in the literature. In 1979, Wagner and Welch [83] used 68 Ga-oxine to label DTPA-conjugated human serum albumin microspheres. These radiolabeled particles were found to be very stable toward transchelation in vitro. After intravenous injection in dogs with one surgically ligated lung, the microspheres quickly accumulated in the perfused lung, allowing PET imaging at 45 min postinjection. Recently, as previously mentioned, Cartier et al. [75] directly labeled the surface of latex NPs with 68 Ga and other trivalent (radio)metals. The in vivo distribution of such NPs was studied in normal rats with PET (68 Ga), MRI (Gd), and cut-and-count techniques (111 In). In this study, the only time point used for PET imaging was 15 min postinjection. At the same time point, no significant MRI contrast was visible in any organ despite the administration of a double amount of Gd-NPs (20 mg per rat) compared to 68 Ga-NPs [75]. Reasonably, this was due to the low specific activity of the injected (radio)labeled NPs. Therefore in spite of the obvious convenience of direct labeling methods, BFC-based approaches appear to be more suited to in vivo molecular imaging with NPs.
17.3.4 Yttrium Radiolabeling The medium-lived + -emitter 86 Y (14.7-hours half-life) can be produced with high specific activity on a medical cyclotron [84] and is being considered as an imaging analog of the therapeutic radionuclide 90 Y (100% − ) [85]. Unfortunately, the widespread use of 86 Y is held back by its limited availability and complicated decay scheme (emission of high-energy + and prompt ␥ rays in cascade). Despite these drawbacks, McDevitt and colleagues recently used 86 Y to study the distribution of CNTs in normal mice injected intravenously and intraperitoneally [24]. The DOTA-conjugated CNTs were 86 Y-labeled with high yield and specific activity (555 MBq/mg) in mild conditions (30-min incubation at 60 ◦ C). The prolonged retention of radioactivity in the peritoneal cavity suggested in vivo stability of these 86 Y-labeled NPs, confirmed also by the presence of intact 86 Y-CNTs in the urine 1 hour postinjection. Both small animal PET imaging and cut-and-count techniques showed fast blood clearance of CNTs and high uptake in kidney, liver, and spleen up to 24 hours postinjection [24].
17.3.5 Technetium and Rhenium Radiolabeling The establishment and evolution of diagnostic nuclear medicine can probably be attributed to the existence of 99m Tc [86]. Technetium-99m has been the most used ␥ -emitter since the introduction of ␥ -cameras in hospitals and today approximately 80% of the diagnostic imaging procedures are still carried out using 99m Tc radiotracers. The prominent role of 99m Tc is due to its optimal nuclear properties (>99% decay by IT with the emission of a 140-keV ␥ photon) and convenient half-life (6.0 hours) and to the existence of an affordable 99 Mo/99m Tc generator. In the hospital pharmacy, the radiometal is eluted daily from the generator in saline solution and added to commercial kits containing all the chemicals necessary for the tracer synthesis, thus making the production of 99m Tc tracers widely accessible and cost effective. Another Tc isotope with potential application in nuclear imaging is 94m Tc. Technetium-94m is a cyclotron-produced radionuclide with a 52.0-min
P1: OTA/XYZ P2: ABC c17 JWBS037-Chen January 25, 2011
410
13:16
Printer Name: Yet to Come
RADIOLABELED NANOPLATFORMS: IMAGING HOT BULLETS HITTING THEIR TARGET
half-life. It decays 72% by emission of 2.438-MeV + (2.74-mm mean + range), and it is a candidate PET tracer [87]. Based on the similarity of the chemical properties to those of Tc, in the past two decades considerable effort was dedicated also to the development of tracers containing Re radionuclides. Rhenium-186 and 188 Re decay by emitting − particles and imageable ␥ photons. As a result, Re radionuclides can be tracked in vivo during internal radiotherapy procedures and can be used for dosimetry studies. Rhenium-186 can be produced in nuclear reactors (low specific activity) and cyclotrons (high specific activity) [88] while high specific activity 188 Re is available from the 188 W/188 Re generator [89]. For this reason, 188 Re is becoming the most popular Re radioisotope for labeling of a variety of molecules, including NPs. For labeling purposes, the radiometal contained in the generator eluate (pertechnetate or perrhenate) is reduced from the nonreactive heptavalent state to a lower oxidation state. Usually, this reaction is carried out in the presence of mono- or polydentate ligands or BFCs containing heterofunctions (e.g., thiol, amino, amido, hydrazino, or phosphino groups [90]) that stabilize the radiometal in its final oxidation state and avoid the formation of over reduced, inert Tc/Re-oxide species (usually referred to as Tc/Re radiocolloids). Despite this general approach, a variety of NPs were labeled with a direct method, that is, without the use of a metal chelator. In theory, this is possible when the NP contains multiple heteroatoms able to bind 99m Tc in one of its reduced states. However, the in vivo stability of these 99m Tc-NPs varies from case to case. For instance, Banerjee and co-workers studied the labeling of native chitosan NPs with 99m Tc in the presence of SnCl2 or NaBH4 as reducing agents [91]. In this study, the use of SnCl2 in the absence of a metal chelator led to the formation of radiocolloids, which could not be separated from the NP mixture. The formation of colloids was not observed with NaBH4 , but the reduced Tc was weakly bound on the NPs. For these reasons, the in vivo administration of the first batch (SnCl2 ) of 99m Tc-labeled chitosan NPs in normal rabbits resulted in high radioactivity uptake in lungs and liver (colloids) while the administration of the second batch (NaBH4 ) produced high radioactivity uptake in the stomach (99m Tc-pertechnetate) [91]. Presumably, partial radiolabel release in vivo occurred also when injecting glucosamine modified MWNTs directly labeled with 99m Tc and SnCl2 . In fact, in the biodistribution study published by Guo et al. [92], the uptake and release of radioactivity in the stomach in the first few hours after MWNT administration mirrors that of free 99m Tc-pertechnetate [93]. Such high stomach uptake was not observed with SWNTs carrying a variety of functionalities and radiolabels [24, 53, 68]. The release of pertechnetate from directly labeled NPs is not a general phenomenon. In fact, Agashe et al. [94] directly labeled PPI dendrimers with 99m Tc-pertechnetate and SnCl2 and proved that the label was stably attached to the NPs both in vitro and in vivo. Similarly, Park and co-workers did not observe any release of 99m Tc-pertechnetate after intratumor injection of directly 99m Tc-labeled pullulan NPs [95]. Solid lipid nanoparticles were also labeled with 99m Tc in aqueous solution though a direct method [96, 97]. Interestingly, despite the paucity of heteroatoms in the structure, these NPs were able to retain 99m Tc in vitro when challenged with excess DTPA and Cys, and did not release the label in vivo. A more traditional approach to label the lipid core of lipid nanocapsules with 99m Tc was reported by Cahouet et al. [98]. In an early study, 99m Tcoxine was encapsulated in the NP lipid core while 125 I was used to label PE in the shell of the capsule with the Bolton–Hunter method [98]. Unfortunately, shortly after intravenous injection the 125 I activity was found in plasma while the 99m Tc activity was associated with
P1: OTA/XYZ P2: ABC c17 JWBS037-Chen January 25, 2011
13:16
Printer Name: Yet to Come
RADIONUCLIDES AND STRATEGIES FOR NANOPLATFORM RADIOLABELING
411
red blood cells, reasonably due to 99m Tc-oxine exchange with red blood cell membranes. Markedly different results were obtained when including a lipophilic 99m Tc/188 Re-SSS complex (Fig. 17.3) in the NP core [99]. This formulation proved to be very stable in vivo and no release of radiolabel was observed in normal rodents 24 hours after intravenous administration. In addition, NPs loaded with 188 Re with the same strategy were injected intratumor in a glioma model and only 10% of the injected dose was released as perrhenate from the injection site in 3 days [100]. Various direct labeling approaches were used also with magnetic NPs. Silica-coated magnetite and SPIOs were functionalized with an antibody targeting human hepatocarcinoma cells and directly labeled with 188 Re-perrhenate after reduction of the protein disulfide bridges to sulfhydryl groups [101, 102]. With this approach, the radiometal incorporation was efficient but approximately 25% of the NP-bound radioactivity was released in vitro in 24 hours. On the contrary, when the surface of the magnetite/silica NPs was functionalized with a chelator (His) and reacted with the [188 Re(CO)3 (H2 O)3 ]+ synthon, less than 20% of the bound radioactivity was released in vitro over a 3-day period [103]. To the best of the author’s knowledge, these 188 Re-NPs were not tested in vivo. Similar to 111 In and 67 Ga, liposomes can be labeled with 99m Tc with an after-loading approach by using a lipophilic 99m Tc complex. In the method introduced by Phillips and co-workers, 99m Tc-HMPAO (Fig. 17.3) is used to shuttle the radiometal through the bilayer of the liposome [104]. Preloaded glutathione reduces the complex to a more hydrophilic species that is trapped in the aqueous core and retained in vivo. This labeling approach is highly efficient (85–90% 99m Tc incorporation) and it is not influenced by the size of the liposomes [105], by the lipid composition [106, 107], or by the presence of PEG on the liposome surface [108]. Another lipophilic moiety used for after-loading liposome labeling is the SNS/S complex 99m Tc-BMEDA (Fig. 17.3) [109]. This complex contains tertiary amino functions that become protonated at low pH, and therefore it can be trapped inside the liposome core by using a pH gradient. Bao and co-workers used this approach to label Doxil with good yields (40–80% 99m Tc encapsulation) and proved that this labeling procedure does not alter the in vivo distribution of the NP [110]. However, slow release of the 99m Tc complex from the liposome in vivo caused some radioactivity accumulation in the kidney in rats. Surprisingly, when doxorubicin-free liposomes were labeled with 186 Re-BMEDA with the same procedure, a higher stability was observed both in vitro and in vivo [111]. BFC-based strategies to label liposomes with 99m Tc have also been developed [40]. At first, the use of single-chain fatty acid to anchor the DTPA to the lipid bilayer gave relatively poor in vivo stability, reasonably due to lipid exchange with blood components [112]. The stability of 99m Tc-liposomes was significantly improved by attaching the BFC to a double-chain fatty acid [113]. More recently, Laverman and co-workers developed a method to 99m Tc-label liposomes conjugated to the HYNIC chelator, using tricine as a coligand (Fig. 17.4) [114]. When compared with the after-loading method based on 99m Tc-HMPAO, the HYNIC method showed higher incorporation efficiency, with no need for postlabeling purification, and less in vitro transchelation in the presence of excess DTPA and Cys. Due to the high stability, higher radioactivity uptake in bacterial infection loci and lower radioactivity uptake in kidneys were observed in rats administered 99m Tc-HYNIC liposomes compared to rats injected with a 99m TcHMPAO liposomal preparation. Significant differences were also observed in liver and spleen uptakes, reasonably due to the different surface chemistries of these two NP preparations [114].
P1: OTA/XYZ P2: ABC c17 JWBS037-Chen January 25, 2011
412
13:16
Printer Name: Yet to Come
RADIOLABELED NANOPLATFORMS: IMAGING HOT BULLETS HITTING THEIR TARGET
H N
O
N N O
O
N
O
O
M HO HO
N
O
N
OH OH
OH
FIGURE 17.4 Structure of the 99m Tc-HYNIC(tricine)2 complex.
17.3.6 Iodine Radiolabeling Iodine has a variety of radioisotopes that can be used in radioactive assays and nuclear medicine procedures, with a wide range of half-lives and emitted radiations. The long-lived ␥ -emitter 125 I (59.4-day half-life) emits low-energy photons and is mostly used for in vitro assays, ex vivo biodistribution, and autoradiography studies. The ␥ - and − -emitter 131 I (8day half-life) was used for both radiotherapy of thyroid pathologies and diagnostic imaging for over four decades. Nowadays, 123 I (13.2-hour half-life) has replaced this “classic” iodine isotope in SPECT imaging because of the lower radiation dose and because of the 159-keV ␥ emission, ideal for ␥ -cameras. Finally, in recent years the + -emitter 124 I (4.2-day half-life) has received much consideration for PET imaging [85]. For decades, antibodies and proteins have been radioiodinated using radioiodide in the presence of mild oxidizing agents such as Iodogen and CAT. These reagents oxidize iodide to iodine, which then reacts at the position ortho of Tyr residues in situ. The same strategy is applied to polymers containing Tyr residues or phenol groups. For instance, the Iodogen method [115, 116] and CAT [117] were used to radioiodinate HPMA polymers obtained by radical polymerization in the presence of small amounts (0.3–1 mol %) of methacryloyltyrosinamide. [125 I]Iodide and CAT were also used to label several polyester dendrimers containing phenol rings or tyramine groups in their backbone [118–120]. Interestingly, the radioiodide/CAT approach was applied also to NP lacking phenol groups. For instance, Wang and co-workers used this method to covalently bind 125 I to SWNTols, presumably at defect sites on the wall and at the open ends of the tubes [26]. As an alternative to labeling phenol groups, NPs containing amino functions are reacted with the Bolton–Hunter reagent [121, 122] or with succinimidyl 3-[125 I]iodobenzoate [123, 124]. Cell internalization followed by lysosomal degradation is a major concern when using compounds radioiodinated on Tyr residues. In fact, the resulting lipophilic radiocatabolite (iodotyrosine) diffuses back through cellular membranes and is poorly retained in tissues [125]. Furthermore, if dehalogenation occurs in vivo, the released radioiodide concentrates in organs expressing the sodium–iodide symporter, such as thyroid, stomach, and salivary gland [93]. The degradation processes, however, may happen slowly when the radiolabel is inserted within the core of a macromolecular assembly. In addition, the degradation mechanisms may change depending on the structure and composition of the NP.
P1: OTA/XYZ P2: ABC c17 JWBS037-Chen January 25, 2011
13:16
Printer Name: Yet to Come
RADIONUCLIDES AND STRATEGIES FOR NANOPLATFORM RADIOLABELING
413
Interestingly, when evaluating the influence of molecular weight on the biodistribution and tumor accumulation of radioiodinated poly(HPMA) in rats, small polymers (20–30 kDa) showed a higher release of radioiodide compared to a ∼60-kDa polymer, as suggested by the different radioactivity uptake in the thyroids in the scintigraphic images [115, 116]. This may reflect the higher chance of the radioiodinated phenols to be situated close to the surface, exposed to enzymatic degradation, in smaller NPs compared to larger ones. Unfortunately, the authors of these studies did not determine the content of radioactivity in the rat stomach and did not analyze the urine to quantify the amount of radioiodide that was released in vivo. On the contrary, no significant release of radioiodide from the Tyr residues in a series of PEGylated polyester dendrimers was detected in vivo [119]. However, 48 hours after NP injection in mice, Guillaudeu and co-workers [119] observed the presence of a high amount of small molecular weight radiometabolites in the urine (65% of the total activity), suggesting dendrimer degradation. The nature of these degradation products was not investigated further. A yet different behavior was observed by Wilbur et al. [124] and Sato et al. [123] after the injection in mice of biotinylated PAMAM dendrimers labeled with succinimidyl 3-[125 I]iodobenzoate. In these studies, very low radioactivity accumulation was observed in thyroid and stomach 4 hours after the NP administration [124] and only intact dendrimer was detected in urine, suggesting in vivo NP stability [123]. High in vivo stability was observed also for the 125 I-SWNTols. In the study by Wang et al. [26], these radioiodinated NPs were administered in mice though different routes (intravenously, subcutaneously, intraperitoneally, and orally). In all cases, the SWNTols exhibited a rapid distribution throughout the entire body and a fast urinary excretion. Noticeably, the amount of radioactivity in the stomach was higher than that observed after administration of a comparable amount of free radioiodide [26] and similar to that reported by Liu et al. [53] for PEGylated SWNT labeled with 64 Cu. Therefore, although the metabolism of 125 I-SWNTols was not investigated, the authors concluded that the presence of radioactivity in the stomach was probably due to accumulation of 125 I-SWNTols in this organ and not to dehalogenation. 17.3.7 Bromine Radiolabeling In recent years, there has been considerable interest in using 76 Br for PET evaluation of long-circulating molecules, especially antibodies [85]. The high abundance + emission and high production yields in medical cyclotrons [126] combined with the convenient halflife (16.2 hours) make this PET radionuclide a valid alternative to 124 I. As with iodine, NPs containing phenol groups can readily be radiolabeled using electrophilic [76 Br]bromine. This reactive species is obtained in situ from bromide with oxidizing agents (CAT, peracetic acid, etc.) or enzymes (bromoperoxidase/H2 O2 ) [127–129]. One advantage of bromine over iodine is that 76 Br tracers are less prone to dehalogenate in vivo compared to radioiodinated analogs due to the stronger C Br bond. Unfortunately, if in vivo debromination occurs, the elimination of [76 Br]bromide from the system is very slow (10–12 days) [130]. Almutairi et al. [131] used 76 Br and small animal PET imaging to evaluate the in vivo distribution of PEGylated dendrimers targeting the ␣v 3 integrin. In this study, the radiohalogen was efficiently bound to the Tyr functionalized core of the dendrimer using CAT. The nanoprobe cleared effectively from the blood (less than 1 % ID/g 48 hours postinjection in rats) and a small amount of a small molecular weight radiometabolite was detected in plasma (approximately 30% of the radioactivity circulating at 48 hours). No
P1: OTA/XYZ P2: ABC c17 JWBS037-Chen January 25, 2011
414
13:16
Printer Name: Yet to Come
RADIOLABELED NANOPLATFORMS: IMAGING HOT BULLETS HITTING THEIR TARGET
significant accumulation of radioactivity was detected in brain and stomach [131], target organs for free radiobromide [132], confirming the in vivo stability of this 76 Br-NP. 17.3.8 Fluorine Radiolabeling Fluorine-18 is probably the most important PET radionuclide, as it is used to produce the widely used 18 FDG and many other tracers. Fluorine-18 decays 97% by + emission and 3% by EC. The maximum energy of the emitted + (0.633 MeV) is ideal for imaging with PET cameras because of the short mean + range (0.69 mm) providing the highest resolution of all PET radionuclides. The relatively long half-life of 18 F (109.8 min) allows the distribution of [18 F]fluoride within a few hours drive from commercial sources and a variety of synthetic strategies can be carried out. However, considering the length of the syntheses and purification steps, the use of PET scanners to assess the distribution of 18 F tracers in living systems is limited to a few hours. NP radiolabeling with 18 F can be accomplished by using small reactive 18 F-labeled synthons that bind to chemical functionalities on the NP surface. For instance, Ducong´e and co-workers functionalized the surface of PEGylated phospholipid micelles encapsulating QDs with thiol groups and then used a 18 F-maleimido agent (18 FPyME) for radiolabeling [133]. The complete 18 F-NP radiosynthesis was accomplished within 145 min from the 18 F production with a high specific activity. The stability of the 18 F-QDs was confirmed both in vitro and in vivo. Dynamic PET imaging in mice (90 min) showed a long blood circulation (approximately 2-hour half-life) and delayed MPS uptake for these PEGylated QDs. The QD uptake in liver and spleen macrophages was also monitored at the cellular level by in vivo fiber confocal fluorescence imaging and then confirmed ex vivo by fluorescence microscopy [133]. Willmann et al. [134] used a different 18 F-synthon (succinimidyl [18 F]fluorobenzoate) to label the surface of targeted MBs designed for US imaging of angiogenesis. In this case, 18 F labeling was performed on the targeting ligand (a biotinylated antibody) to retain the integrity of the gas-filled NPs. After radiosynthesis, the 18 F-antibody was bound to the NP surface using biotin–streptavidin interaction. Dynamic PET imaging (90 min) confirmed very short MB circulation in blood (few minutes) and rapid uptake in MPS organs in mice. However, incomplete blood clearance and slow accumulation of radioactivity in the kidney suggested partial degradation of 18 F-MBs in vivo. The direct nucleophilic radiofluorination of NPs was also reported. In a recent study, Matson and co-workers synthesized crosslinked polymeric micelles carrying mesylate groups in the outer shell [135]. These robust NPs were efficiently and safely labeled with [18 F]fluoride in acetonitrile in the presence of Kriptofix-222 (1-hour reaction at 120 ◦ C). The biologic studies on these 18 F-NPs are ongoing. Fluorine-18 is used also for liposome radiolabeling. At first, 18 FDG was encapsulated in the liposome core during manufacturing [136]. With this straightforward labeling approach, Oku and co-workers performed a real-time analysis of liposomal in vivo trafficking with PET and thoroughly evaluated the influence of size, composition, and charge on the early biodistribution of long-circulating stealth liposomes encapsulating 18 FDG [137]. Recently, a more general approach to 18 F-label NPs comprising a lipophilic domain was reported by Marik and co-workers [138]. In this study, the authors synthesized a dipalmitin derivative carrying a tosylate group, which was 18 F-labeled with high specific activity (greater than 110 GBq/mmol) and incorporated in the membrane of long-circulating liposomes during manufacturing (approximately 2 hours for the complete radiosynthesis). As expected, dynamic PET showed a near-constant level of radioactivity in blood and major organs during
P1: OTA/XYZ P2: ABC c17 JWBS037-Chen January 25, 2011
13:16
Printer Name: Yet to Come
ANGIOGENESIS IMAGING WITH RADIOLABELED NANOPLATFORMS
415
FIGURE 17.5 Representative full body PET projection (90-min continuous bed motion scan) of a normal rat injected with long-circulating PEGylated liposomes radiolabeled with [18 F]fluorodipalmitin. (Adapted from Marik et al. [138].)
the 90-min scan (Fig. 17.5) [138]. The short half-life of 18 F did not allow extending the imaging study beyond this point. The same 18 F-dipalmitin derivative was used also to radiolabel the shell of lipid MBs [139]. Compared to stealth liposomes, 18 F-MBs exhibited a much shorter blood circulation and a higher spleen uptake. Dynamic PET imaging was also capable of quantifying lipid material deposition in the kidney as the result of MB disruption following US pulses.
17.4 ANGIOGENESIS IMAGING WITH RADIOLABELED NANOPLATFORMS Several formulations containing NPs chemically bound to or entrapping a chemotherapeutic agent have reached the market or are in clinical trials for the treatment of cancer [140]. These agents accumulate in cancer tissues mainly through passive targeting (EPR effect) [22]. By attaching specific ligands to the NP surface, active targeting of tumor cells can be accomplished. For instance, NPs conjugated to folic acid [28, 72, 141] and transferrin [48, 142, 143] are known to specifically enter cancer cells through endocytic pathways triggered by the interaction with overexpressed membrane receptors both in vitro and in vivo. However, in solid tumors the rate-determining step for NP accumulation is the extravasation from the blood vessels. For this reason, the use of targeting ligands has
P1: OTA/XYZ P2: ABC c17 JWBS037-Chen January 25, 2011
416
13:16
Printer Name: Yet to Come
RADIOLABELED NANOPLATFORMS: IMAGING HOT BULLETS HITTING THEIR TARGET
generally failed in increasing the overall accumulation of NPs in solid tumors and specific accumulation of targeted NPs was seen only in disseminated lymphoma models [68] and ascitic tumor cells [144]. Obviously, this apparent lack of specificity is a drawback when aiming at using NPs for molecular imaging of cancer, especially with nonactivatable labels such as radionuclides. Furthermore, many candidate imaging NPs are too large or leave the bloodstream too rapidly to diffuse out of the blood vessels into the tumor interstitial space. Therefore the evaluation of targeted NPs for molecular imaging of cancer has been mostly limited to vascular markers of disease (e.g., markers of angiogenesis) that are readily accessible and do not require extravasation. Angiogenesis—the formation of new capillaries by cellular outgrowth from existing microvessels—is a fundamental process in tumor development. Small tumors (1–2 mm) rely on oxygen and nutrient diffusion from the surrounding tissues to survive (avascular phase). Tumor progression and metastasis propagation occur only after induction of a tumor vasculature (angiogenic switch) [145]. Besides tumors, angiogenesis takes place also in other tissues as an adaptation to hypoxic stress. For instance, neovascularization occurs as part of the natural healing process after ischemic injury to restore tissue oxygenation [146]. The presence of hypoxia within thickened atherosclerotic lesions causes the formation of neovessels, which may mark plaque vulnerability [147]. Therefore molecular imaging of angiogenesis could be critical for defining the pathophysiology of a variety of diseases including cancer and atherosclerosis, and for assessing the efficacy of anti- and proangiogenic therapies. An altered expression of integrins and growth factor receptors on endothelial cells are known molecular events associated with the angiogenic process. The next sections summarize the progress in molecular imaging of angiogenesis with radiolabeled NPs targeting the ␣v 3 integrin and VEGFR-2. 17.4.1 Imaging of ␣ v 3 Integrins Integrins are a family of ␣ heterodimeric transmembrane glycoproteins that mediate cell–cell and cell–matrix adhesion through interactions with specific ligands [148]. Integrin signaling plays a key role in immune response, hemostasis, wound healing, and tissue remodeling. Integrin binding supports the development of tumor vasculature, tumor growth, and metastasis. In vascular, inflammatory, autoimmune, or hyperproliferative diseases, integrin function can also become dysregulated and contribute to pathogenesis [148]. It is currently accepted that the vitronectin receptor ␣v 3 is absent on quiescent endothelial cells in established vessels and that its expression is induced on endothelial cells undergoing the process of active angiogenesis. However, the exact role of ␣v 3 as a proangiogenic or antiangiogenic factor is still a matter of debate [149]. A subset of integrins including the ␣v 3 recognizes a common motif in their ligands: the RGD sequence. An arsenal of contrast agents containing single or multiple RGD peptides has been developed for angiogenesis imaging and some of them have made the transition from bench to bedside [150]. To date, there has been only one report on SPECT imaging of angiogenesis with radiolabeled NPs. Hu and co-workers [71] tested 111 In-labeled perfluorocarbon particles functionalized with a quinolone-based vitronectin receptor antagonist [151] in a nascent Vx-2 lung carcinoma tumor implanted in rabbit hindlimbs. The presence of neovessels at the capsular interface between the tumor tissue and the thigh muscle 2 weeks after tumor implant was confirmed with a stain for vascular endothelium. In this animal model, the
P1: OTA/XYZ P2: ABC c17 JWBS037-Chen January 25, 2011
13:16
Printer Name: Yet to Come
ANGIOGENESIS IMAGING WITH RADIOLABELED NANOPLATFORMS
417
tumor-to-muscle ratio obtained with targeted 111 In-NPs was approximately twofold higher than that observed with nontargeted NPs. In addition, the administration of a threefold excess of nonradioactive targeted NPs competitively blocked the uptake in the neovascularized tumor region, suggesting specific interaction with the ␣v 3 integrin. In this study, the uptake in the tumor was visible by ␥ imaging as early as 15 min postinjection and persisted for at least 18 hours [71]. This was a significant improvement with respect to a previous imaging experiment performed with a paramagnetic version of the NP [152]. In fact, the MRI contrast in the tumor periphery was markedly lower than that obtained by ␥ imaging and visible only at later time points. In recent years, the group at Stanford University dedicated major efforts to the development of NP-based probes for multimodal imaging of ␣v 3 integrins in tumor neovasculature. Copper-64-labeled QDs and IOs carrying RGD moieties were evaluated for PET/NIRF imaging [49] and PET/MRI imaging [52], respectively. The specific targeting capabilities of these NPs were investigated in mice bearing human glioblastoma xenografts (U87MG), a highly malignant tumor hallmarked by extensive invasion and angiogenesis [153, 154]. Following the intravenous administration of 64 Cu-labeled RGD-QDs in mice, small animal PET confirmed the rapid and extensive QD accumulation in MPS organs. Despite the high MPS uptake, the U87MG xenografts were clearly visible as early as 5 hours postinjection and the contrast increased over time (4.0 ± 1.0% ID/g at 18 hours postinjection) [49]. The specific interaction between RGD-QDs and ␣v 3 integrins was confirmed both in vitro (competitive blocking experiment) and in vivo (less than 1% ID/g in tumors in mice injected with nontargeted QDs). The confinement of the vast majority of RGD-QDs into the tumor vasculature expressing ␣v 3 was assessed ex vivo by immunohistochemistry and fluorescence microscopy. Unfortunately, due to the higher sensitivity and deeper tissue penetration of PET with respect to optical imaging, the small amount of 64 Cu-QDs used in this PET/NIRF study (20 pmol QDs per mouse) did not allow noninvasive fluorescence imaging. However, a good correlation between the PET and NIRF signals was observed ex vivo in isolated organs [49]. Yet, the findings of these experiments agree with the semiquantitative results of a previous NIRF imaging study carried out with nonradiolabeled RGD-QD in U87MG xenografted mice (200 pmol QDs injected per mouse) [155]. Real dual-modality imaging of tumor angiogenesis was achieved with 64 Cu-labeled RGD-IOs and PET/MRI [52]. In this experiment, small amounts of 64 Cu-labeled NPs were diluted with nonradioactive IOs and injected in mice bearing U87MG xenografts (3.7 MBq 64 Cu and 300 g iron per mouse). The tumors were clearly visualized with small animal PET in mice administered IO-RGD (10.1 ± 2.1% ID/g 4 hours postinjection). On the contrary, almost no contrast was observed in mice injected with nontargeted IOs or coinjected with IO-RGD and c(RGDyK) (10 mg/kg mouse body weight). In vivo T2-weighted MRI was also performed on a clinical 3-T scanner 4 hours post-IO injection. In tumors, a decrease in signal intensity was observed after the injection of DOTA-IO-RGD compared to DOTA-IO and DOTA-IO-RGD with blocking agent (qualitative measurements) [52]. In another study by the same group, SWNTs were functionalized with DOTA and c(RGDyK) functionalities appended at the end of phospholipid-PEG chains of two different lengths (2 and 5.4 kDa) [53]. Quantitative analysis of the PET images in mice confirmed a significantly longer blood circulation of 64 Cu-labeled PEG5400 SWNTs (2hour half-life) with respect to the PEG2000 derivatives (30-min half-life). Surprisingly, the SWNTs functionalized with the shorter RGD-PEG chain exhibited only a slightly higher retention in U87MG tumors (∼3–4 % ID/g over a 24-hour period) with respect to the nontargeted analogs (∼2.5% ID/g). On the contrary, the tumor uptake of the RGD-PEG5400
P1: OTA/XYZ P2: ABC c17 JWBS037-Chen January 25, 2011
418
13:16
Printer Name: Yet to Come
RADIOLABELED NANOPLATFORMS: IMAGING HOT BULLETS HITTING THEIR TARGET
functionalized SWNTs was significantly higher than that of the nontargeted analogs at each considered time point (∼13% ID/g and ∼4% ID/g, respectively, over a 24-hour period). Furthermore, this tumor uptake was competitively blocked by coinjecting c(RGDyK) (15 mg/kg mouse body weight). The quantitative measurements from PET image analysis were confirmed by Raman spectroscopy [53]. The reason why long-circulating SWNTs accumulated more in tumor compared to the short-circulating analogs in a ␣v 3 -dependent fashion was not investigated further by the authors. Previous work carried out with different CNTs revealed that these NPs easily pass through a number of compartments to reach targets in the whole mouse [92]. Therefore, possibly, the long-circulating RGD-SWNTs extravasated from the tumor blood vessels with a mechanism that was initiated or facilitated by ␣v 3 -mediated adhesion to the endothelial wall. Alternatively, the high tumor uptake of long-circulating RGD-SWNTs could have been the result of ␣v 3 -mediated internalization in endothelial cells followed by integrin recycling to the cell surface. Recently, Almutairi et al. [131] designed a dendritic imaging NP carrying multiple RGD copies appended to PEG chains and labeled with 76 Br for PET imaging of angiogenesis. This nanoprobe was tested in a mouse model of hindlimb ischemia producing slow revascularization of the thigh muscle [156, 157]. In normal rodents, the dendritic NP exhibited sustained blood circulation and did not accumulate significantly in nontarget organs (including the MPS organs). Seven days after the surgical intervention in the mouse hindlimbs, intravenous injection of the ␣v 3 -targeted nanoprobe resulted in radioactivity accumulation in the entire ischemic muscle compared to the nonischemic limb (sham surgery), particularly in the area distal to the femoral artery excision (Fig. 17.6). On the contrary, the
(a)
(b)
(c)
FIGURE 17.6 (A) Structure of a 76 Br-labeled dendritic nanoprobe targeting the ␣v 3 integrin designed for PET images of angiogenesis. (B) Representative small animal PET/CT coronal and sagittal slices obtained 24 hours after the injection of the RGD nanoprobe (left mouse) and nontargeted nanoprobe (right mouse) in a mouse model of hindlimb ischemia. The images show high RGD nanoprobe uptake in the ischemic limb with respect to the control limb (sham injury) and with respect to the nontargeted nanoprobe. (C) Quantitative analysis of PET images (right-to-left hindlimb ratios) showing an approximate twofold higher uptake of the RGD nanoprobe in the distal and proximal areas of the ischemic thigh respect to the non-targeted nanoprobe. (Adapted from Almutairi et al. [131].)
P1: OTA/XYZ P2: ABC c17 JWBS037-Chen January 25, 2011
13:16
Printer Name: Yet to Come
ANGIOGENESIS IMAGING WITH RADIOLABELED NANOPLATFORMS
419
administration of nontargeted nanoprobe did not produce significant contrast [131]. This behavior was similar to that observed with a small 99m Tc-RGD derivative in a similar animal model [157], thus suggesting RGD-dendrimer uptake in angiogenic vessels.
17.4.2 Imaging of VEGFR-2 Among the seven structurally related VEGF molecules, VEGF-A (usually referred to as VEGF) is the major mediator of tumor angiogenesis [158]. VEGF is a homodimeric glycoprotein existing in several splice variants differing in the number of amino acids and heparin-binding properties. VEGF121 is a freely diffusible protein. The dominant isoform VEGF165 is also secreted, but a significant fraction is bound to heparin-like moieties on the cell surface and ECM. The longer VEGF189 and VEGF206 are almost completely sequestered by the ECM and can be released in diffusible form by the action of heparin or heparinase [159]. Most types of human cancer cells express VEGF as a response to environmental factors (hypoxia, pH, inflammatory cytokines, sex hormones, etc.) or genetic changes (activation of oncogenes or inactivation of tumor-suppressor genes) and increased expression is often associated with poor prognosis [159]. VEGF signals mainly through the tyrosine kinase VEGFR-2, which is expressed at elevated levels by endothelial cells. VEGF diffusion and binding to VEGFR-2 leads to a cascade of signaling events resulting in proliferation, migration, and survival of endothelial cells and increased vascular permeability, ultimately leading to the formation of tumor neovessels [158, 159]. The pivotal role of VEGF/VEGFR signaling in tumor angiogenesis has prompted significant efforts in the search for antiVEGF treatments to “starve” solid tumors [160]. Besides tumors, VEGF/VEGFR signaling is involved also in several hematologic malignancies, intraocular neovascularization syndromes, inflammatory disorders, and other pathologies and plays an important role in bone healing and in the angiogenesis associated with myocardial infarction and ischemia [146, 159]. The development of reliable VEGF/VEGFR-targeted imaging tracers has also been the subject of extensive research. Several VEGF121 and VEGF165 derivatives, anti-VEGF antibodies, and VEGFR-2-specific fusion proteins were functionalized with radionuclides and optical labels or conjugated to detectable NPs for angiogenesis imaging and therapy monitoring [161]. Chen and co-workers conjugated VEGF and DOTA to QDs to perform dual-modality imaging of tumor angiogenesis and evaluated these nanoconjugates in U87MG xenografts with PET (64 Cu) and NIRF (Fig. 17.7) [61]. Small animal PET showed a significantly higher tumor uptake of 64 Cu-labeled VEGF-QDs (up to 4.5 ± 0.5% ID/g 24 hours postinjection) compared to nontargeted QDs (<1% ID/g). Also in this case, as for 64 Cu-labeled RGD-QDs, NIRF imaging was carried out on a different group of mice, administered a 10-fold higher amount of QDs, and showed good agreement with PET (Fig. 17.7). Immunofluorescence staining on tumor slices treated with an anti-VEGFR-2 antibody showed a precise overlay between the receptor expression and QD fluorescent signal. Interestingly, prominent receptor expression was visible only on small and immature tumor vessels and not on bigger ones [61]. Nonhomogeneous distribution of VEGFR-2 on the U87MG tumor vasculature and a correlation between tumor size, VEGFR-2 expression, and 64 Cu-DOTAVEGF121 uptake were the findings of another study published by Chen and co-workers [162]. These results further stress the importance of imaging VEGFR expression before
P1: OTA/XYZ P2: ABC c17 JWBS037-Chen January 25, 2011
420
13:16
Printer Name: Yet to Come
RADIOLABELED NANOPLATFORMS: IMAGING HOT BULLETS HITTING THEIR TARGET
(a)
(b)
(c)
O O N H
N
NH—VEGF
S O
O
ZnS HN O N
HO C N
64Cu N
CO H
N
HO C
FIGURE 17.7 (A) Representative structure of a 64 Cu-labeled DOTA-QD-VEGF conjugate designed for PET/NIRF imaging of angiogenesis. (B) Representative small animal PET coronal slices obtained 1 hour and 24 hours after the injection of 10 pmol nontargeted QDs (top) and VEGF-QDs (bottom) in mice bearing U87MG xenografts. (C) in vivo NIRF images of mice bearing U87MG xenografts 1 hour after the injection of 200 pmol nontargeted QDs (top) and VEGF-QDs (bottom). White arrows indicate the tumors. The images show high radioactivity uptake in liver and spleen. The tumors of the mice administered QD-VEGF are clearly visible with both modalities. (Adapted from Chen et al. [61].)
and during anti-VEGFR cancer treatments, due to the dynamic nature of VEGFR expression through tumor progression and to the high variability among different tumor types. Recently, the feasibility of anti-VEGFR-2 therapy monitoring with targeted MBs and contrast-enhanced ultrasonography was proved in a murine model of human pancreatic adenocarcinoma [163]. Hence Willmann and co-workers used a 18 F-labeled anti-VEGFR-2 to functionalize MBs and performed dynamic small animal PET imaging to assess the whole body distribution and absolute tumor targeting capabilities of the targeted MBs [134]. As expected, the MBs disappeared from the blood circulation within a few minutes from intra-venous injection and accumulated in MPS organs. However, the MB retention in subcutaneous angiosarcomas was significantly higher than that in the adjacent skeletal muscle during the whole imaging session (1.14% ID/g in tumor and 0.84% ID/g in muscle, averaged over 60 min) [134]. The tumor targeting capabilities of these MBs were further enhanced by using a dual-targeting approach [164]. In fact, the administration of MBs targeting both VEGFR-2 and ␣v 3 produced significantly higher US contrast in ovarian adenocarcinoma xenografts compared to that obtained with MBs targeting only VEGFR-2 or ␣v 3 , injected separately or in combination.
17.5 CONCLUSION Radionuclide-based assays and noninvasive imaging techniques are powerful tools in the hands of scientists for assessing the distribution and targeting capabilities of new molecules. These versatile and highly sensitive techniques allow a more comprehensive understanding of the behavior of candidate drugs at the cell and tissue level and can expedite enormously the path from small to large animals and, finally, to humans. For this reason, the production of new clinical and research PET, SPECT, and hybrid imaging devices has flourished over
P1: OTA/XYZ P2: ABC c17 JWBS037-Chen January 25, 2011
13:16
Printer Name: Yet to Come
REFERENCES
421
the last decade. Nanomedicine also benefits greatly from the use of these techniques. In fact, more and more material scientists adopt radionuclide-based methods to study the characteristics of new nanomaterials in vitro, in vivo, and ex vivo. It is expected that this trend will continue with the increased availability of nonstandard medium- to long-lived radionuclides such as 64 Cu, 89 Zr, 124 I, and 76 Br and with the distribution of radionuclide generators in medical centers and research institutions worldwide. The thrilling achievements in the field of angiogenesis imaging reported in the previous paragraphs, obtained with the support of PET, SPECT, and other imaging modalities, are only one example of the great potential of NPs for molecular imaging of disease. In the near future, the close partnership among material scientists, chemists, biologists, clinicians, engineers, physicists, and mathematicians will produce more sophisticated nanobullets, better targets, higher sensitivity, and higher resolution imaging devices and better reconstruction algorithms for the needs of the healthcare of tomorrow.
ACKNOWLEDGMENTS Special thanks to Dr. Sander Langereis, Dr. Marc Robillard, and Dr. Holger Gr¨ull for stimulating discussions and assistance in writing this chapter.
REFERENCES 1. Schwartz, R. S. Paul Ehrlich’s magic bullets. N. Engl. J. Med. 2004, 350(11), 1079–1080. 2. Farokhzad, O. C.; Langer, R. Nanomedicine: developing smarter therapeutic and diagnostic modalities. Adv. Drug Deliv. Rev. 2006, 58(14), 1456–1459. 3. Marcato, P. D.; Dur´an, N. New aspects of nanopharmaceutical delivery systems. J. Nanosci. Nanotechnol. 2008, 8, 2216–2229. 4. Moghimi, S. M.; Kissel, T. Particulate nanomedicine. Adv. Drug Deliv. Rev. 2006, 58(14), 1451–1455. 5. Sahoo, S. K.; Parveen, S.; Panda, J. J. The present and future of nanotechnology in human health care. Nanomedicine 2007, 3(1), 20–31. 6. Bhattacharya, R.; Mukherjee, P. Biological properties of “naked” metal nanoparticles. Adv. Drug Deliv. Rev. 2008, 60(11), 1289–1306. 7. Hirsch, L.; Gobin, A.; Lowery, A.; et al. Metal nanoshells. Ann. Biomed. Eng. 2006, 34(1), 15–22. 8. Ganta, S.; Devalapally, H.; Shahiwala, A.; Amiji, M. A review of stimuli-responsive nanocarriers for drug and gene delivery. J. Control Release 2008, 126(3), 187–204. 9. Hernot, S.; Klibanov, A. L. Microbubbles in ultrasound-triggered drug and gene delivery. Adv. Drug Deliv. Rev. 2008, 60(10), 1153–1166. 10. Husseini, G. A.; Pitt, W. G. Micelles and nanoparticles for ultrasonic drug and gene delivery. Adv. Drug Deliv. Rev. 2008, 60(10), 1137–1152. 11. Bangham, A. D.; Standish, M. M.; Watkins, J. C. Diffusion of univalent ions across the lamellae of swollen phospholipids. J. Mol. Biol. 1965, 13(1), 238–252. 12. Langer, R.; Folkman, J. Polymers for the sustained release of proteins and other macromolecules. Nature 1976, 263(5580), 797–800. 13. Freitas, R. A., Jr. Nanomedicine, Volume 1: Basic Capabilities. Landes Bioscience: Georgetown, TX, 1999.
P1: OTA/XYZ P2: ABC c17 JWBS037-Chen January 25, 2011
422
13:16
Printer Name: Yet to Come
RADIOLABELED NANOPLATFORMS: IMAGING HOT BULLETS HITTING THEIR TARGET
14. Duncan, R. Targeting and intracellular delivery of drugs. In Encyclopedia of Molecular Cell Biology and Molecular Medicine, 2nd ed.; Meyers, R. A., Ed.; Wiley-VCH: Weinheim, Germany, 2005; Vol. 14, pp. 163–203. 15. Moghimi, S. M.; Hunter, A. C.; Murray, J. C. Long-circulating and target-specific nanoparticles: theory to practice. Pharmacol. Rev. 2001, 53(2), 283–318. 16. Drummond, D. C.; Meyer, O.; Hong, K.; Kirpotin, D. B.; Papahadjopoulos, D. Optimizing liposomes for delivery of chemotherapeutic agents to solid tumors. Pharmacol. Rev. 1999, 51(4), 691–744. 17. Moghimi, S. M.; Szebeni, J. Stealth liposomes and long circulating nanoparticles: critical issues in pharmacokinetics, opsonization and protein-binding properties. Progr. Lipid Res. 2003, 42(6), 463–478. 18. Owens, D. E. III; Peppas, N. A. Opsonization, biodistribution, and pharmacokinetics of polymeric nanoparticles. Int. J. Pharm. 2006, 307(1), 93–102. 19. Gref, T.; Minamitake, Y.; Peracchia, M. T.; Trubetskoy, V.; Torchilin, V.; Langer, R. Biodegradable long-circulating polymeric nanospheres. Science 1994, 263(5153), 1600–1603. 20. Caliceti, P.; Veronese, F. M. Pharmacokinetic and biodistribution properties of poly(ethylene glycol)-protein conjugates. Adv. Drug Deliv. Rev. 2003, 55(10), 1261–1277. 21. Ota, Z.; Makin, H.; Takaya, Y.; Ofuji, T. Molecular sieve in renal glomerular and tubular basement membranes as revealed by electron microscopy. Renal. Physiol. 1980, 3, 317–323. 22. Maeda, H. The enhanced permeability and retention (EPR) effect in tumor vasculature: the key role of tumor-selective macromolecular drug targeting. Adv. Enzyme Regul. 2001, 41(1), 189–207. 23. Cartlidge, S. A.; Duncan, R.; Lloyd, J. B.; Kopeckova-Rejmanova, P.; Kopecek, J. Soluble, crosslinked N-(2-hydroxypropyl)methacrylamide copolymers as potential drug carriers: 2. Effect of molecular weight on blood clearance and body distribution in the rat after intravenous administration. Distribution of unfractionated copolymer after intraperitoneal, subcutaneous or oral administration. J. Control Release 1987, 4(4), 253–264. 24. McDevitt, M. R.; Chattopadhyay, D.; Jaggi, J. S.; et al. PET imaging of soluble yttrium-86labeled carbon nanotubes in mice. PLoS ONE 2007, 2(9), e907. 25. Seymour, L. W.; Duncan, R.; Strohalm, J.; Kopecek, J. Effect of molecular weight (Mw) of N-(2-hydroxypropyl)methacrylamide copolymers on body distribution and rate of excretion after subcutaneous, intraperitoneal, and intravenous administration to rats. J. Biomed. Mater. Res. 1987, 21(11), 1341–1358. 26. Wang, H.; Wang, J.; Deng, X.; et al. Biodistribution of carbon single-wall carbon nanotubes in mice. J. Nanosci. Nanotechnol. 2004, 4(8), 1019–1024. 27. McCarthy, J. R.; Jaffer, F. A.; Weissleder, R. A macrophage-targeted theranostic nanoparticle for biomedical applications. Small 2006, 2(8-9), 983–987. 28. Gabizon, A.; Shmeeda, H.; Horowitz, A. T.; Zalipsky, S. Tumor cell targeting of liposomeentrapped drugs with phospholipid-anchored folic acid-PEG conjugates. Adv. Drug Deliv. Rev. 2004, 56(8), 1177–1192. 29. Rossin, R.; Pan, D.; Qi, K.; et al. 64 Cu-labeled folate-conjugated shell cross-linked nanoparticles for tumor imaging and radiotherapy: synthesis, radiolabeling, and biologic evaluation. J. Nucl. Med. 2005, 46(7), 1210–1218. 30. Kirpotin, D. B.; Drummond, D. C.; Shao, Y.; et al. Antibody targeting of long-circulating lipidic nanoparticles does not increase tumor localization but does increase internalization in animal models. Cancer Res. 2006, 66(13), 6732–6740. 31. Kimelberg, H. K.; Tracy, T. F., Jr.; Biddlecome, S. M.; Bourke, R. S. The effect of entrapment in liposomes on the in vivo distribution of [3 H]methotrexate in a primate. Cancer Res. 1976, 36(8), 2949–2957.
P1: OTA/XYZ P2: ABC c17 JWBS037-Chen January 25, 2011
13:16
Printer Name: Yet to Come
REFERENCES
423
32. Sjoholm, I.; Edman, P. Acrylic microspheres in vivo. I. Distribution and elimination of polyacrylamide microparticles after intravenous and intraperitoneal injection in mouse and rat. J. Pharmacol. Exp. Ther. 1979, 211(3), 656–662. 33. Cherry, S. R. In vivo molecular and genomic imaging: new challenges for imaging physics. Phys. Med. Biol. 2004, 49(3), R13-R49. 34. Rowland, D. J.; Cherry, S. R. Small-animal preclinical nuclear medicine instrumentation and methodology. Semin. Nucl. Med. 2008, 38(3), 209–222. 35. Cherry, S. R. Multimodality in vivo imaging systems: twice the power or double the trouble? Annu. Rev. Biomed. Eng. 2006, 8(1), 35–62. 36. Pichler, B. J.; Wehrl, H. F.; Kolb, A.; Judenhofer, M. S. Positron emission tomography/magnetic resonance imaging: the next generation of multimodality imaging? Semin. Nucl. Med. 2008, 38(3), 199–208. 37. Laforest, R.; Rowland, D. J.; Welch, M. J. MicroPET imaging with nonconventional isotopes. IEEE Trans. Nucl. Sci. 2002, 49(5), 2119–2126. 38. Vandenberghe, S. Three-dimensional positron emission tomography imaging with 124 I and 86 Y. Nucl. Med. Commun. 2006, 27, 237–245. 39. Walrand, S. P.; Jamar, F. O.; Mathieu, I.; et al. Quantitation in PET using isotopes emitting prompt single gammas: application to yttrium-86. Eur. J. Nucl. Med. Mol. Imaging 2003, 30(3), 354– 361. 40. Laverman, P.; Boerman, O. C.; Storm, G. Radiolabeling of liposomes for scintigraphic imaging. Methods Enzymol. 2003, 373, 234–248. 41. Shokeen, M.; Fettig, N. M.; Rossin, R. Symthesis, in vitro and in vivo evaluation of radiolabeled nanoparticles. Q. J. Nucl. Med. Mol. Imaging 2008, 52(3), 267–277. 42. McCarthy, D. W.; Shefer, R. E.; Klinkowstein, R. E.; et al. Efficient production of high specific activity 64 Cu using a biomedical cyclotron. Nucl. Med. Biol. 1997, 24(1), 35–43. 43. Blower, P. J.; Lewis, J. S.; Zweit, J. Copper radionuclides and radiopharmaceuticals in nuclear medicine. Nucl. Med. Biol. 1996, 23(8), 957–980. 44. Smith, S. V. Molecular imaging with copper-64. J. Inorg. Biochem. 2004, 98(11), 1874–1901. 45. Wadas, T. J.; Wong, E. H.; Weisman, G. R.; Anderson, C. J. Copper chelation chemistry and its role in copper radiopharmaceuticals. Curr. Pharm. Des. 2007, 13(1), 3–16. 46. Sun, X.; Rossin, R.; Turner, J. L.; et al. An assessment of the effects of shell cross-linked nanoparticle size, core composition, and surface PEGylation on in vivo biodistribution. Biomacromolecules 2005, 6(5), 2541–2554. 47. Seo, J. W.; Zhang, H.; Kukis, D. L.; Meares, C. F.; Ferrara, K. W. A novel method to label preformed liposomes with 64 Cu for positron emission tomography (PET) imaging. Bioconjug. Chem. 2008, 19(12), 2577–2584. 48. Bartlett, D. W.; Su, H.; Hildebrandt, I. J.; Weber, W. A.; Davis, M. E. Impact of tumor-specific targeting on the biodistribution and efficacy of siRNA nanoparticles measured by multimodality in vivo imaging. Proc. Natl. Acad. Sci. U.S.A. 2007, 104(39), 15549–15554. 49. Cai, W.; Chen, K.; Li, Z-B; Gambhir, S. S.; Chen, X. Dual-function probe for PET and nearinfrared fluorescence imaging of tumor vasculature. J. Nucl. Med. 2007, 48(11), 1862–1870. 50. Fukukawa, K.; Rossin, R.; Hagooly, A.; et al. Synthesis and characterization of core-shell star copolymers for in vivo PET imaging applications. Biomacromolecules 2008, 9(4), 1329–1339. 51. Jarrett, B. R.; Gustafsson, B.; Kukis, D. L.; Louie, A. Y. Synthesis of 64 Cu-labeled magnetic nanoparticles for multimodal imaging. Bioconjug. Chem. 2008, 19(7), 1496–1504. 52. Lee, H.-Y.; Li, Z.; Chen, K.; et al. PET/MRI dual-modality tumor imaging using arginine–glycine–aspartic (RGD)-conjugated radiolabeled iron oxide nanoparticles. J. Nucl. Med. 2008, 49(8), 1371–1379.
P1: OTA/XYZ P2: ABC c17 JWBS037-Chen January 25, 2011
424
13:16
Printer Name: Yet to Come
RADIOLABELED NANOPLATFORMS: IMAGING HOT BULLETS HITTING THEIR TARGET
53. Liu, Z.; Cai, W.; He, L.; et al. In vivo biodistribution and highly efficient tumour targeting of carbon nanotubes in mice. Nat. Nano. 2007, 2(1), 47–52. 54. Pressly, E. D.; Rossin, R.; Hagooly, A.; et al. Structural effects on the biodistribution and positron emission tomography (PET) imaging of well-defined 64 Cu-labeled nanoparticles comprised of amphiphilic block graft copolymers. Biomacromolecules 2007, 8(10), 3126–3134. 55. Rossin, R.; Muro, S.; Welch, M. J.; Muzykantov, V. R.; Schuster, D. P. In vivo imaging of 64 Culabeled polymer nanoparticles targeted to the lung endothelium. J. Nucl. Med. 2008, 49(1), 103–111. 56. Schipper, M. L.; Cheng, Z.; Lee, S-W; et al. MicroPET-based biodistribution of quantum dots in living mice. J. Nucl. Med. 2007, 48(9), 1511–1518. 57. Sun, G.; Hagooly, A.; Xu, J.; et al. Facile, efficient approach to accomplish tunable chemistries and variable biodistributions for shell cross-linked nanoparticles. Biomacromolecules 2008, 9(7), 1997–2006. 58. Sun, G.; Xu, J.; Hagooly, A.; et al. Strategies for optimized radiolabeling of nanoparticles for in vivo PET imaging. Adv. Mater. 2007, 19(20), 3157–3162. 59. Xu, J.; Sun, G.; Rossin, R.; et al. Labeling of polymer nanostructures for medical imaging: importance of cross-linking extent, spacer length, and charge density. Macromolecules 2007, 40(9), 2971–2973. 60. Yuan, J.; You, Y.; Lu, X.; Muzik, O.; Oupicky, D.; Peng, F. Synthesis of nano-sized poly(APMA)DOTA-64 Cu conjugates for interventional radionuclide therapy of prostate cancer: assessment of intra-tumoral retention by microPET imaging Mol. Imaging 2007, 6(1), 10–17. 61. Chen, K.; Li, Z-B; Wang, H.; Cai, W.; Chen, X. Dual-modality optical and positron emission tomography imaging of vascular endothelial growth factor receptor on tumor vasculature using quantum dots. Eur. J. Nucl. Med. Mol. Imaging 2008, 35(12), 2235–2244. 62. Nahrendorf, M.; Zhang, H.; Hembrador, S.; et al. Nanoparticle PET-CT imaging of macrophages in inflammatory atherosclerosis. Circulation 2008, 117(3), 379–387. 63. Bass, L. A.; Wang, M.; Welch, M. J.; Anderson, C. J. In vivo transchelation of copper-64 from TETA-octreotide to superoxide dismutase in rat liver. Bioconjug. Chem. 2000, 11(4), 527–532. 64. Boswell, C. A.; Sun, X.; Niu, W.; et al. Comparative in vivo stability of copper-64-labeled cross-bridged and conventional tetraazamacrocyclic complexes. J. Med. Chem. 2004, 47(6), 1465–1474. 65. Anderson, C. J.; Dehdashti, F.; Cutler, P. D.; et al. 64 Cu-TETA-octreotide as a PET imaging agent for patients with neuroendocrine tumors. J. Nucl. Med. 2001, 42(2), 213–221. 66. DeNardo, G. L.; DeNardo, S. J.; Kukis, D.; et al. Maximum tolerated dose of 67 Cu-2ITBAT-LYM-1 for fractionated radioimmunotherapy of non-Hodgkin’s lymphoma: a pilot study. Anticancer Res. 1998, 18(4B), 2779–2788. 67. Kommareddy, S.; Amiji, M. Biodistribution and pharmacokinetic analysis of long-circulating thiolated gelatin nanoparticles following systemic administration in breast cancer-bearing mice. J. Pharm. Sci. 2007, 96(2), 397–407. 68. McDevitt, M. R.; Chattopadhyay, D.; Kappel, B. J.; et al. Tumor targeting with antibodyfunctionalized, radiolabeled carbon nanotubes. J. Nucl. Med. 2007, 48(7), 1180–1189. 69. Singh, R.; Pantarotto, D.; Lacerda, L.; et al. Tissue biodistribution and blood clearance rates of intravenously administered carbon nanotube radiotracers. Proc. Natl. Acad. Sci. U.S.A. 2006, 103(9), 3357–3362. 70. Okuda, T.; Kawakami, S.; Akimoto, N.; Niidome, T.; Yamashita, F.; Hashida, M. PEGylated lysine dendrimers for tumor-selective targeting after intravenous injection in tumor-bearing mice. J. Control. Release 2006, 116(3), 330–336. 71. Hu, G.; Lijowski, M.; Zhang, H.; et al. Imaging of Vx-2 rabbit tumors with ␣v 3 -integrintargeted 111 In nanoparticles. Int. J. Cancer 2007, 120(9), 1951–1957.
P1: OTA/XYZ P2: ABC c17 JWBS037-Chen January 25, 2011
13:16
Printer Name: Yet to Come
REFERENCES
425
72. Oyewumi, M. O.; Yokel, R. A.; Jay, M.; Coakley, T.; Mumper, R. J. Comparison of cell uptake, biodistribution and tumor retention of folate-coated and PEG-coated gadolinium nanoparticles in tumor-bearing mice. J. Control. Release 2004, 95(3), 613–626. 73. Chow, T.-H.; Lin, Y.-Y.; Hwang, J.-J.; et al. Diagnostic and therapeutic evaluation of 111 Invinorelbine-liposomes in a human colorectal carcinoma HT-29/luc-bearing animal model. Nucl. Med. Biol. 2008, 35(5), 623–634. 74. Harrington, K. J.; Mohammadtaghi, S.; Uster, P. S.; et al. Effective targeting of solid tumors in patients with locally advanced cancers by radiolabeled pegylated liposomes. Clin. Cancer Res. 2001, 7(2), 243–254. 75. Cartier, R.; Kaufner, L.; Paulke, B. R.; et al. Latex nanoparticles for multimodal imaging and detection in vivo. Nanotechnology. 2007, 18, 195102. 76. H¨afeli, U. O.; Yu, J.; Farudi, F.; Li, Y.; Tapolsky, G. Radiolabeling of magnetic targeted carriers (MTC) with indium-111. Nucl. Med. Biol. 2003, 30(7), 761–769. 77. Gabizon, A.; Hliberty, J.; Straubinger, R. M.; Price, D. C.; Papahadjopoulos, D. An improved method for in vivo tracing and imaging of liposomes using a gallium-67–deferoxamine complex. J. Liposome Res. 1988, 1(1), 123–135. 78. Woodle, M. C. 67 Gallium-labeled liposomes with prolonged circulation: preparation and potential as nuclear imaging agents. Nucl. Med. Biol. 1993, 20(2), 149–155. 79. Hwang, K. J.; Merriam, J. E.; Beaumier, P. L.; Luk, K.-F. S. Encapsulation, with high efficiency, of radioactive metal ions in liposomes. Biochim. Biophys. Acta 1982, 716(1), 101–109. 80. Ogihara, I.; Kojima, S.; Jay, M. Differential uptake of gallium-67-labeled liposomes between tumors and inflammatory lesions in rats. J. Nucl. Med. 1986, 27(8), 1300–1307. 81. Lewis, M. R.; Reichert, D. E.; Laforest, R.; et al. Production and purification of gallium-66 for preparation of tumor-targeting radiopharmaceuticals. Nucl. Med. Biol. 2002, 29(6), 701–706. 82. Fani, M.; Andr´e, J. P.; Maecke, H. R. 68 Ga-PET: a powerful generator-based alternative to cyclotron-based PET radiopharmaceuticals. Contrast Media Mol. Imaging 2008, 3(2), 67–77. 83. Wagner, S. J.; Welch, M. J. Gallium-68 labeling of albumin and albumin microspheres. J. Nucl. Med. 1979, 20(5), 428–433. 84. Yoo, J.; Tang, L.; Perkins, T. A.; et al. Preparation of high specific activity 86 Y using a small biomedical cyclotron. Nucl. Med. Biol. 2005, 32(8), 891–897. 85. Nayak, T. K.; Brechbiel, M. W. Radioimmunoimaging with longer-lived positron-emitting radionuclides: potentials and challenges. Bioconjug. Chem. 2009, 20(5), 825–841. 86. Banerjee, S.; Pillai, M. R.; Ramamoorthy, N. Evolution of Tc-99m in diagnostic radiopharmaceuticals. Semin. Nucl. Med. 2001, 31(4), 260–277. 87. Bigott, H. M.; Prior, J. L.; Piwnica-Worms, D. R.; Welch, M. J. Imaging multidrug resistance Pglycoprotein transport function using microPET with technetium-94m-sestamibi. Mol. Imaging 2005, 4(1), 30–39. 88. Moustapha, M. E.; Ehrhardt, G. J.; Smith, C. J.; Szajek, L. P.; Eckelman, W. C.; Jurisson, S. S. Preparation of cyclotron-produced 186 Re and comparison with reactor-produced 186 Re and generator-produced 188Re for the labeling of bombesin. Nucl. Med. Biol. 2006, 33(1), 81–89. 89. Knapp, F. F.; Beets, A. L.; Guhlke, S.; et al. Availability of rhenium-188 from the alumina-based tungsten-188/rhenium-188 generator for preparation of rhenium-188-labeled radiopharmaceuticals for cancer treatment. Anticancer Res. 1997, 17(3B), 1783–1795. 90. Liu, S. Bifunctional coupling agents for radiolabeling of biomolecules and target-specific delivery of metallic radionuclides. Adv. Drug. Deliv. Rev. 2008, 60(12), 1347–1370. 91. Banerjee, T.; Singh, A. K.; Sharma, R. K.; Maitra, A. N. Labeling efficiency and biodistribution of technetium-99m labeled nanoparticles: interference by colloidal tin oxide particles. Int. J. Pharm. 2005, 289(1-2), 189–195.
P1: OTA/XYZ P2: ABC c17 JWBS037-Chen January 25, 2011
426
13:16
Printer Name: Yet to Come
RADIOLABELED NANOPLATFORMS: IMAGING HOT BULLETS HITTING THEIR TARGET
92. Guo, J.; Zhang, X.; Li, Q.; Li, W. Biodistribution of functionalized multiwall carbon nanotubes in mice. Nucl. Med. Biol. 2007, 34(5), 579–583. 93. Zuckier, L. S.; Dohan, O.; Li, Y.; Chang, C. J.; Carrasco, N.; Dadachova, E. Kinetics of perrhenate uptake and comparative biodistribution of perrhenate, pertechnetate, and iodide by NaI symporter-expressing tissues in vivo. J. Nucl. Med. 2004, 45(3), 500–507. 94. Agashe, H. B.; Babbar, A. K.; Jain, S.; et al. Investigations on biodistribution of technetium99m-labeled carbohydrate-coated poly(propylene imine) dendrimers. Nanomedicine 2007, 3(2), 120–127. 95. Park, K. H.; Song, H. C.; Na, K.; et al. Ionic strength-sensitive pullulan acetate nanoparticles (PAN) for intratumoral administration of radioisotope: ionic strength-dependent aggregation behavior and 99 mtechnetium retention property. Colloids Surf. B. Biointerfaces 2007, 59(1), 16–23. 96. Reddy, L. H.; Sharma, R. K.; Chuttani, K.; Mishra, A. K. Etoposide-incorporated tripalmitin nanoparticles with different surface charge: formulation, characterization, radiolabeling and biodistribution studies. AAPS J. 2004, 6(3), 23. 97. Thakkar, H.; Kumar Sharma, R.; Murthy, R. S. Enhanced retention of celecoxib-loaded solid lipid nanoparticles after intra-articular administration. Drugs R&D 2007, 8(5), 275–285. 98. Cahouet, A.; Denizot, B.; Hindr´e, F.; et al. Biodistribution of dual radiolabeled lipidic nanocapsules in the rat using scintigraphy and ? counting. Int. J. Pharm. 2002, 242(1-2), 367–371. 99. Ballot, S.; Noiret, N.; Hindr´e, F.; et al. 99m Tc/188 Re-labelled lipid nanocapsules as promising radiotracers for imaging and therapy: formulation and biodistribution. Eur. J. Nucl. Med. Mol. Imaging 2006, 33(5), 602–607. 100. Allard, E.; Hindre, F.; Passirani, C.; et al. 188 Re-loaded lipid nanocapsules as a promising radiopharmaceutical carrier for internal radiotherapy of malignant gliomas. Eur. J. Nucl. Med. Mol. Imaging 2008, 35(10), 1838–1846. 101. Liang, S.; Wang, Y.; Yu, J.; Zhang, C.; Xia, J.; Yin, D. Surface modified superparamagnetic iron oxide nanoparticles: as a new carrier for bio-magnetically targeted therapy. J. Mater. Sci. Mater. Med. 2007, 18(12), 2297–2302. 102. Liang, S.; Wang, Y.; Zhang, C.; Liu, X. Synthesis of amino-modified magnetite nanoparticles coated with Hepama-1 and radiolabeled with 188 Re for bio-magnetically targeted radiotherapy. J. Radioanal. Nucl. Chem. 2006, 269(1), 3–7. 103. Cao, J.; Wang, Y.; Yu, J.; et al. Preparation and radiolabeling of surface-modified magnetic nanoparticles with rhenium-188 for magnetic targeted radiotherapy. J. Magn. Magn. Mater. 2004, 277(1-2), 165–174. 104. Phillips, W. T.; Rudolph, A. S.; Goins, B.; Timmons, J. H.; Klipper, R.; Blumhardt, R. A simple method for producing a technetium-99m-labeled liposome which is stable in vivo. Int. J. Rad. Appl. Instrum. B 1992, 19(5), 539–547. 105. Awasthi, V. D.; Garcia, D.; Goins, B. A.; Phillips, W. T. Circulation and biodistribution profiles of long-circulating PEG-liposomes of various sizes in rabbits. Int. J. Pharm. 2003, 253(1-2), 121–132. 106. Mirahmadi, N.; Babaei, M. H.; Vali, A. M.; Johari Daha, F.; Kobarfard, F.; Dadashzadeh, S. 99m Tc-HMPAO-labeled liposomes: an investigation into the effects of some formulation factors on labeling efficiency and in vitro stability. Nucl. Med. Biol. 2008, 35(3), 387–392. 107. Carmo, V. A.; Ferrari, C. S.; Reis, E. C.; et al. Biodistribution study and identification of inflammation sites using 99m Tc-labelled stealth pH-sensitive liposomes. Nucl. Med. Commun. 2008, 29(1), 33–38. 108. Lee, C.-M.; Choi, Y.; Huh, E. J.; et al. Polyethylene glycol (PEG) modified 99m Tc-HMPAOliposome for improving blood circulation and biodistribution: the effect of the extent of PEGylation. Cancer Biother. Radiopharm. 2005, 20(6), 620–628.
P1: OTA/XYZ P2: ABC c17 JWBS037-Chen January 25, 2011
13:16
Printer Name: Yet to Come
REFERENCES
427
109. Bao, A.; Goins, B.; Klipper, R.; Negrete, G.; Phillips, W. T. Direct 99m Tc labeling of pegylated liposomal doxorubicin (Doxil) for pharmacokinetic and non-invasive imaging studies. J. Pharmacol. Exp. Ther. 2004, 308(2), 419–425. 110. Gabizon, A.; Shmeeda, H.; Barenholz, Y. Pharmacokinetics of pegylated liposomal doxorubicin: review of animal and human studies. Clin. Pharmacokinet. 2003, 42, 419–436. 111. Bao, A.; Goins, B.; Klipper, R.; Negrete, G.; Phillips, W. T. 186 Re-liposome labeling using 186Re-SNS/S complexes: in vitro stability, imaging, and biodistribution in rats. J. Nucl. Med. 2003, 44(12), 1992–1999. 112. Hnatowich, D. J.; Friedman, B.; Clancy, B.; Novak, M. Labeling of preformed liposomes with Ga-67 and 99m Tc by chelation. J. Nucl. Med. 1981, 22(9), 810–814. 113. Ahkong, Q. F.; Tilcock, C. Attachment of 99m Tc to lipid vesicles containing the lipophilic chelate dipalmitoylphosphatidylethanolamine-DTTA. Int. J. Rad. Appl. Instrum. B. 1992, 19(8), 831–840. 114. Laverman, P.; Dams, E.; Oyen, W.; et al. A novel method to label liposomes with 99m Tc by the hydrazino nicotinyl derivative. J. Nucl. Med. 1999, 40(1), 192–197. 115. Kissel, M.; Peschke, P.; Subr, V.; et al. Synthetic macromolecular drug carriers: biodistribution of poly[(N-2-hydroxypropyl)methacrylamide] copolymers and their accumulation in solid rat tumors. J. Pharm. Sci. Technol. 2001, 55(3), 191–201. 116. Lammers, T.; K¨uhnlein, R.; Kissel, M.; et al. Effect of physicochemical modification on the biodistribution and tumor accumulation of HPMA copolymers. J. Control. Release 2005, 110(1), 103–118. 117. Seymour, L. W.; Miyamoto, Y.; Maeda, H.; et al. Influence of molecular weight on passive tumour accumulation of a soluble macromolecular drug carrier. Eur. J. Cancer 1995, 31(5), 766–770. 118. Gillies, E. R.; Dy, E.; Frechet, J. M.; Szoka, F. C. Biological evaluation of polyester dendrimer: poly(ethylene oxide) “bow-tie” hybrids with tunable molecular weight and architecture. Mol. Pharm. 2005, 2(2), 129–138. 119. Guillaudeu, S. J.; Fox, M. E.; Haidar, Y. M.; Dy, E. E.; Szoka, F. C.; Frechet, J. M. J. PEGylated dendrimers with core functionality for biological applications. Bioconjug. Chem. 2008, 19(2), 461–469. 120. Padilla De Jesus, O. L.; Ihre, H. R.; Gagne, L.; Frechet, J. M. J.; Szoka, F. C. Polyester dendritic systems for drug delivery applications: in vitro and in vivo evaluation. Bioconjug. Chem. 2002, 13(3), 453–461. 121. Huang, R. Q.; Qu, Y. H.; Ke, W. L.; Zhu, J. H.; Pei, Y. Y.; Jiang, C. Efficient gene delivery targeted to the brain using a transferrin-conjugated polyethyleneglycol-modified polyamidoamine dendrimer. FASEB J. 2007, 21(4), 1117–1125. 122. Malik, N.; Wiwattanapatapee, R.; Klopsch, R.; et al. Dendrimers: relationship between structure and biocompatibility in vitro, and preliminary studies on the biodistribution of 125 I-labelled polyamidoamine dendrimers in vivo. J. Control. Release 2000, 65(1-2), 133–148. 123. Sato, N.; Park, C. W.; Kim, H.-S.; et al. Synthesis of dendrimer-based biotin radiopharmaceuticals to enhance whole-body clearance. Nucl. Med. Biol. 2003, 30(6), 617–625. 124. Wilbur, D. S.; Pathare, P. M.; Hamlin, D. K.; Buhler, K. R.; Vessella, R. L. Biotin reagents for antibody pretargeting. 3. synthesis, radioiodination, and evaluation of biotinylated starburst dendrimers. Bioconjug. Chem. 1998, 9(6), 813–825. 125. Tolmachev, V.; Orlova, A.; Lundqvist, H. Approaches to improve cellular retention of radiohalogen labels delivered by internalising tumour-targeting proteins and peptides. Curr. Med. Chem. 2003, 10(22), 2447–2460. 126. Tolmachev, V.; Lovqvist, A.; Einarsson, L.; Schultz, J.; Lundqvist, H. Production of 76 Br by a low-energy cyclotron. Appl. Radiat. Isot. 1998, 49(12), 1537–1540.
P1: OTA/XYZ P2: ABC c17 JWBS037-Chen January 25, 2011
428
13:16
Printer Name: Yet to Come
RADIOLABELED NANOPLATFORMS: IMAGING HOT BULLETS HITTING THEIR TARGET
127. Lovqvist, A.; Sundin, A.; Ahlstrom, H.; Carlsson, J.; Lundqvist, H. 76 Br-labeled monoclonal anti-CEA antibodies for radioimmuno positron emission tomography. Nucl. Med. Biol. 1995, 22(1), 125–131. 128. Lovqvist, A.; Sundin, A.; Ahlstrom, H.; Carlsson, J.; Lundqvist, H. Pharmacokinetics and experimental PET imaging of a bromine-76-labeled monoclonal anti-CEA antibody. J. Nucl. Med. 1997, 38(3), 395–401. 129. McElvany, K. D.; Welch, M. J. Characterization of bromine-77-labeled proteins prepared by using bromoperoxidase. J. Nucl. Med. 1980, 21(10), 953–960. 130. Gudjonsson, O.; Bergstr¨om, M.; Kristjansson, S.; et al. Analysis of 76 Br-BrdU in DNA of brain tumors after a PET study does not support its use as a proliferation marker. Nucl. Med. Biol. 2001, 28(1), 59–65. 131. Almutairi, A.; Rossin, R.; Shokeen, M.; et al. Biodegradable dendritic positron emitting nanoprobes for the non invasive imaging of angiogenesis. Proc. Natl. Acad. Sci. U.S.A. 2009, 106(3), 685–690. 132. Soremark, R.; Ullberg, S. Distribution of bromide in mice. An autoradiography study with Br-82. Int. J. Appl. Radiat. Isot. 1960, 8, 192–197. 133. Duconge, F.; Pons, T.; Pestourie, C.; et al. Fluorine-18-labeled phospholipid quantum dot micelles for in vivo multimodal imaging from whole body to cellular scales. Bioconjug. Chem. 2008, 19(9), 1921–1926. 134. Willmann, J. K.; Cheng, Z.; Davis, C.; et al. Targeted microbubbles for imaging tumor angiogenesis: assessment of whole-body biodistribution with dynamic micro-PET in mice. Radiology. 2008, 249(1), 212–219. 135. Matson, J. B.; Grubbs, R. H. Synthesis of fluorine-18 functionalized nanoparticles for use as in vivo molecular imaging agents. J. Am. Chem. Soc. 2008, 130(21), 6731–6733. 136. Oku, N.; Tokudome, Y.; Tsukada, H.; Okada, S. Real-time analysis of liposomal trafficking in tumor-bearing mice by use of positron emission tomography. Biochim. Biophys. Acta 1995, 1238(1), 86–90. 137. Oku, N. Delivery of contrast agents for positron emission tomography imaging by liposomes. Adv. Drug Deliv. Rev. 1999, 37(1-3), 53–61. 138. Marik, J.; Tartis, M. S.; Zhang, H.; et al. Long-circulating liposomes radiolabeled with [18 F]fluorodipalmitin ([18 F]FDP). Nucl. Med. Biol. 2007, 34(2), 165–171. 139. Tartis, M. S.; Kruse, D. E.; Zheng, H.; et al. Dynamic microPET imaging of ultrasound contrast agents and lipid delivery. J. Control. Release 2008, 131(3), 160–166. 140. Cho, K.; Wang, X.; Nie, S.; Chen, Z.; Shin, D. M. Therapeutic nanoparticles for drug delivery in cancer. Clin. Cancer Res. 2008, 14(5), 1310–1316. 141. Leamon, C. P.; Reddy, J. A. Folate-targeted chemotherapy. Adv. Drug Deliv. Rev. 2004, 56(8), 1127–1141. 142. Qian, Z. M.; Li, H.; Sun, H.; Ho, K. Targeted drug delivery via the transferrin receptor-mediated endocytosis pathway. Pharmacol. Rev. 2002, 54(4), 561–587. 143. Sahoo, S. K.; Labhasetwar, V. Enhanced antiproliferative activity of transferrin-conjugated paclitaxel-loaded nanoparticles is mediated via sustained intracellular drug retention. Mol. Pharm. 2005, 2(5), 373–383. 144. Gabizon, A.; Horowitz, A. T.; Goren, D.; Tzemach, D.; Shmeeda, H.; Zalipsky, S. In vivo fate of folate-targeted polyethylene-glycol liposomes in tumor-bearing mice. Clin. Cancer Res. 2003, 9(17), 6551–6559. 145. Folkman, J. Angiogenesis in cancer, vascular, rheumatoid and other disease. Nat. Med. 1995, 1, 27–30. 146. Sinusas, A. J. Imaging of angiogenesis. J. Nucl. Cardiol. 2004, 11(5), 617–633.
P1: OTA/XYZ P2: ABC c17 JWBS037-Chen January 25, 2011
13:16
Printer Name: Yet to Come
REFERENCES
429
147. Doyle, B.; Caplice, N. Plaque neovascularization and antiangiogenic therapy for atherosclerosis. J. Am. Coll. Cardiol. 2007, 49(21), 2073–2080. 148. Staunton, D. E.; Lupher, M. L.; Liddington, R.; Gallatin, W. M.; Frederick, W. A. Targeting integrin structure and function in disease. Adv. Immunol. 2006, 91, 111–157. 149. Hodivala-Dilke, K. ␣v 3 Integrin and angiogenesis: a moody integrin in a changing environment. Curr. Opin. Cell. Biol. 2008, 20(5), 514–519. 150. Beer, A.; Schwaiger, M. Imaging of integrin ␣v 3 expression. Cancer Metastasis Rev. 2008, 27(4), 631–644. 151. Harris, T. D.; Kalogeropoulos, S.; Nguyen, T.; et al. Design, synthesis, and evaluation of radiolabeled integrin ␣v 3 receptor antagonists for tumor imaging and radiotherapy. Cancer Biother. Radiopharm. 2003, 18(4), 627–641. 152. Winter, P. M.; Caruthers, S. D.; Kassner, A.; et al. Molecular imaging of angiogenesis in nascent Vx-2 rabbit tumors using a novel ␣v 3 -targeted nanoparticle and 1.5 tesla magnetic resonance imaging. Cancer Res. 2003, 63(18), 5838–5843. 153. Millauer, B.; Shawver, L. K.; Plate, K. H.; Risaui, W.; Ullrich, A. Glioblastoma growth inhibited in vivo by a dominant-negative Flk-1 mutant. Nature 1994, 367(6463), 576–579. 154. Sallinen, S-L; Sallinen, P. K.; Haapasalo, H. K.; et al. Identification of differentially expressed genes in human gliomas by dna microarray and tissue chip techniques. Cancer Res. 2000, 60(23), 6617–6622. 155. Cai, W.; Shin, D.-W.; Chen, K.; et al. Peptide-labeled near-infrared quantum dots for imaging tumor vasculature in living subjects. Nano. Lett. 2006, 6(4), 669–676. 156. Couffinhal, T.; Silver, M.; Zheng, L. P.; Kearney, M.; Witzenbichler, B.; Isner, J. M. Mouse model of angiogenesis. Am. J. Pathol. 1998, 152(6), 1667–1679. 157. Hua, J.; Dobrucki, L. W.; Sadeghi, M. M.; et al. Noninvasive imaging of angiogenesis with a 99m Tc-labeled peptide targeted at ␣v 3 integrin after murine hindlimb ischemia. Circulation 2005, 111(24), 3255–3260. 158. Kerbel, R. S. Tumor angiogenesis. N. Engl. J. Med. 2008, 358(19), 2039–2049. 159. Ferrara, N. Vascular endothelial growth factor: basic science and clinical progress. Endocr. Rev. 2004, 25(4), 581–611. 160. Tonra, J. R.; Hicklin, D. J. Targeting the vascular endothelial growth factor pathway in the treatment of human malignancy. Immunol. Invest. 2007, 36(1), 3–23. 161. Cai, W.; Chen, X. Multimodality imaging of vascular endothelial growth factor and vascular endothelial growth factor receptor expression. Front. Biosci. 2007, 12, 4267–4279. 162. Chen, K.; Cai, W.; Li, Z.-B.; Wang, H.; Chen, X. Quantitative PET imaging of VEGF receptor expression. Mol. Imaging Biol. 2009, 11(1), 15–22. 163. Korpanty, G.; Carbon, J. G.; Grayburn, P. A.; Fleming, J. B.; Brekken, R. A. Monitoring response to anticancer therapy by targeting microbubbles to tumor vasculature. Clin. Cancer Res. 2007, 13(1), 323–330. 164. Willmann, J. K.; Lutz, A. M.; Paulmurugan, R.; et al. Dual-targeted contrast agent for US assessment of tumor angiogenesis in vivo. Radiology 2008, 248(3), 936–944.
P1: OTA/XYZ P2: ABC c18 JWBS037-Chen February 2, 2011
21:51
Printer Name: Yet to Come
PART III
NANOPARTICLE PLATFORMS AS MULTIMODALITY IMAGING AND THERAPY AGENTS
P1: OTA/XYZ P2: ABC c18 JWBS037-Chen February 2, 2011
21:51
Printer Name: Yet to Come
CHAPTER 18
Lipoprotein-Based Nanoplatforms for Cancer Molecular Imaging IAN R. CORBIN Department of Medical Biophysics, University of Toronto, and Division of Biophysics and Bioimaging, Ontario Cancer Institute, Toronto, Ontario, Canada
KENNETH NG Institute of Biomaterials and Biomedical Engineering, University of Toronto, Toronto, Ontario, Canada
GANG ZHENG Department of Medical Biophysics and Institute of Biomaterials and Biomedical Engineering, University of Toronto, and Division of Biophysics and Bioimaging, Ontario Cancer Institute, Toronto, Ontario, Canada
18.1 LIPOPROTEINS: NATURE’S ENDOGENOUS NANOCARRIER Efficient transport of key lipids, such as triacylglycerols (TAGs) and cholesterol, is an essential process in mammalian systems. These molecules respectively serve as a high-energy fuel source and as a structural building block for all cells. Like other essential nutrients and substrates these molecules must be delivered to cells through the blood. The hydrophobicity of these molecules, however, poses a daunting problem for their transport through the aqueous channels of the vascular system. Nature has uniquely addressed this dilemma through the lipoprotein carrier system. Plasma lipoproteins are macromolecular vehicles that transport neutral lipids (fat and cholesterol) through the circulatory system and extracellular fluid compartments of the body. These spherical emulsion–like complexes display a range of physicochemical properties; however, they have a common structural organization (Fig. 18.1.) consisting of an apolar core of TAG and cholesterol esters surrounded by a monolayer of phospholipids and free cholesterol. Interspersed across the phospholipid monolayer are specific amphipathic proteins (apolipoproteins) whose hydrophilic domains extend toward the aqueous environment. These apolipoproteins act to stabilize and confer structural framework to the lipoproteins, modulate enzyme activity, as well as serve as receptor ligands for the lipoprotein delivery system [1].
Nanoplatform-Based Molecular Imaging Edited by Xiaoyuan Chen C 2011 John Wiley & Sons, Inc. Copyright
433
P1: OTA/XYZ P2: ABC c18 JWBS037-Chen February 2, 2011
434
21:51
Printer Name: Yet to Come
LIPOPROTEIN-BASED NANOPLATFORMS FOR CANCER MOLECULAR IMAGING
FIGURE 18.1 General structure of a lipoprotein.
18.1.1 Classification and Composition of Plasma Lipoproteins Plasma lipoproteins can be designated into various classes based on numerous physical parameters (e.g., electrophoretic mobility, diameter). The most commonly accepted classification is based on the density of the different lipoprotein species (see Table 18.1). According to this classification scheme the major density categories include the following. (1) Chylomicrons (d < 0.95 g/mL) are TAG-rich emulsions particles (80–88% by weight) that are synthesized by the intestine after a fatty meal. Chylomicrons are the largest particles in the lipoprotein family (80 nm to 1 m in diameter) and have the highest lipid-to-protein ratio. (2) Very low density lipoproteins (d = 0.95–1.006 g/mL) are also TAG-rich particles, however, they are synthesized by the liver. They are smaller than chylomicrons (30–80 nm in diameter) and contain relatively less TAG but more cholesterol and protein. (3) Low density lipoproteins (d = 1.020–1.063 g/mL) are particles formed from the intravascular catabolism of VLDLs. The LDL core is predominantly cholesterol ester molecules. Finally, (4) high density lipoproteins (d = 1.063–1.210 g/mL) These carriers are the smallest (6–12 nm in diameter) member of the lipoprotein family. Their core is mainly composed of cholesterol esters and they are composed of a relatively high proportion of protein (35–56% by weight). TABLE 18.1 Density, Size, Electrophoretic Mobility, and Chemical Composition of Lipoproteins [2] Classa
Density (g/mL)
Chylomicrons <0.95 VLDL 0.95–1.006 LDL 1.020–1.063 HDL 1.063–1.210 a b
Diameter Electrophoretic Cholesteryl (nm) Mobility Proteinb Phospholipidb Estersb Cholesterolb TAGa,b 75–100 30–80 18–25 6–12
— Pre-  ␣
1–2 6–10 18–22 35–56
3–6 15–20 18–24 26–32
2–4 6–10 45–50 15–20
1–3 4–8 6–8 3–5
VLDL, very low density lipoprotein; LDL, low density lipoprotein; HDL, high density lipoprotein; TAG, triacylglycerol. Expressed in % dry weight.
80–95 45–65 4–8 2–7
P1: OTA/XYZ P2: ABC c18 JWBS037-Chen February 2, 2011
21:51
Printer Name: Yet to Come
LIPOPROTEINS: NATURE’S ENDOGENOUS NANOCARRIER
435
18.1.2 Plasma Apoproteins Apoproteins play an important role in imparting structure and function to the lipoproteins [1]. To date at least ten different species of apoproteins have been identified in human plasma. These include apo A-I, apo A-II, apo A-IV, apo B-100, apo B-48, apo C-I, apo C-II, apo C-III, apo D, and apo E. Based on structural homologies and solution properties, apolipoproteins can be divided into three groups. Group 1 apoproteins are A-I, A-II, A-IV, C-I, C-II, C-III, and E. These apoproteins, also termed the exchangeable apoproteins, are all water-soluble proteins that transfer between the different classes of lipoproteins within the circulation. With the exception of apo A-II, which is a dimer, all of the group 1 apoproteins consist of a single polypeptide chain. Another unifying feature of this group of proteins is their high degree of amphipathic alpha helix secondary structure [3, 4]. The distribution of the amino acid residues in this structural motif is such that they form a lipid binding domain with polar and nonpolar surfaces. The polar side faces and protrudes into the surrounding aqueous environment, while the nonpolar face penetrates into the hydrophobic region of the phospholipid monolayer that surrounds the neutral lipid core [3, 4]. Apo A-I and A-II are the major protein components on HDL. Both of these apoproteins interact and bind to the scavenger receptor type I (SR-BI) (HDL receptor) [5]. Apo A-I also acts as a cofactor for the enzyme lecithin-cholesterol acyltransferase (LCAT) which catalyzes the conversion of cholesterol and phosphatidylcholine to esterified cholesterol and lysophosphatidylcholine [6]. Furthermore, apo A-I when combined with phospholipids can facilitate the process of reverse cholesterol transport (the transfer of cholesterol from cell membranes back to the liver) [7]. The C apoproteins are present on TAG-rich lipoproteins and HDLs. They work in a coordinated manner to regulate the activity of lipoprotein lipase at the capillary bed. This enzyme hydrolyzes TAGs present in the core of lipoproteins. Apo C-I and C-II activate lipoprotein lipase [8, 9], while apo C-III is postulated to inhibit the activation of this enzyme [10]. Apo E is a component of several lipoprotein species. It serves as an important ligand for LDL receptor-mediated catabolism of lipoproteins [11]. Group 2 apoproteins include the B apoproteins. Apo B proteins are large hydrophobic proteins that contain a large portion of beta sheet and relatively little alpha helix secondary structure [12]. As such, these apoproteins do not exchange between the various lipoproteins. Apo B is present in the plasma in two isoforms: apo B-100 and apo B-48. The apo B gene encodes for a large 549-kDa protein product called apo B-100. This apoprotein is synthesized by the liver and is found on the surface of LDLs and VLDLs, where it acts as a ligand for LDLR. The second isoform of apo B, apo B-48, is found exclusively on chylomicrons. This apolipoprotein is synthesized by the intestine and is a truncated version of apo B-100 consisting of only 48% of the normal gene product [13]. The presence of a translational stop codon in the intestine only translates approximately half of the N-terminal region of apo B-100 [13]. Given that the LDLR binding domain of apo B-100 is located in the carboxy terminal region of the sequence, apo B-48 does not contain any receptor binding properties and thus does not interact with LDLR [14]. Finally, apo D belongs to group 3 apoproteins. These proteins bind and associate with monomeric lipids. Apo D has been found to associate with HDL, cholesterol, progesterone, pregnenolone, bilirubin, and arachidonic acid. The function of apo D has yet to be elucidated [15].
P1: OTA/XYZ P2: ABC c18 JWBS037-Chen February 2, 2011
436
21:51
Printer Name: Yet to Come
LIPOPROTEIN-BASED NANOPLATFORMS FOR CANCER MOLECULAR IMAGING
FIGURE 18.2 Schematic representation of lipoprotein metabolism. See text for details.
18.1.3 Lipoprotein Metabolism Circulating lipoproteins operate in a complex and highly coordinated manner to enable the efficient transport of neutral lipids throughout the body (Fig. 18.2). The lipoprotein carrier system transports up to 200–300 g of TAG and cholesterol between the intestine/liver and peripheral tissues each day [16]. This load of neutral lipids arises from both exogenous (dietary) and endogenous (hepatic synthesis) sources. Dietary lipids are packaged into chylomicrons within the enterocytes of the intestine. These fat-laden particles are released into the circulation, where they are acted upon by extrahepatic capillary bound lipoprotein lipases to liberate fatty acids to peripheral cells. En route within the circulation the chylomicron acquires exchangeable apoproteins (apo A, E, and C) from other lipoproteins. As the chylomicron continues to lose its fatty acid cargo, a resulting smaller chylomicron remnant is eventually recognized (via apo E) and internalized within hepatocytes by LDLR and non-LDLR uptake pathways. Endogenous sources of neutral lipid arise from the liver in the form of VLDL. Like the chylomicron, VLDL circulates to the periphery to release its fatty acid cargo by capillary bound lipoprotein lipases. The resulting TAG depleted VLDL transition through an intermediate density lipoprotein stage, which is partly extracted by the liver via apo E mediated uptake, finally giving rise to the cholesterol ester rich LDL. LDL is metabolized by peripheral and hepatic tissue via the apo B-100 and LDLR pathway. The HDL class of lipoproteins first enters the circulation as lipid-free or lipid-poor apo A-I (secreted by the liver and intestine). In the extracellular medium of peripheral tissues the ATP-binding cassette (ABC)-A1 mediates the efflux of phospholipids and free cholesterol from cells to apo A-I to form nascent HDL particles (initial step of reverse
P1: OTA/XYZ P2: ABC c18 JWBS037-Chen February 2, 2011
21:51
Printer Name: Yet to Come
LIPOPROTEINS: NATURE’S ENDOGENOUS NANOCARRIER
437
cholesterol transport). These newly formed HDL particles are discoidal in shape, consisting only of a phospholipid bilayer (interdigitated with free cholesterol) surrounded by a “belt” of apo A-I protein [17]. Acquisition of core molecules (CE) and its transformation into mature spherical HDL is accomplished through the activity of LCAT, which esterifies the 3-hydroxyl position of cholesterol to form CE. The CE molecule then moves into the center of the nascent HDL to form a hydrophobic core, thereby making the particle spherical. Within the circulation, HDLs are constantly remodeled as LCAT and phospholipid transfer protein (PLTP) introduce CE and phospholipids to HDLs, while cholestryl ester transfer protein (CETP) and hepatic lipases mediate the exchange of CE and TAGs between HDLs and other lipoprotein species. In addition to apo A-I, other exchangeable apoproteins such as apo A-II, C, and E may be incorporated onto the HDL surface from other lipoproteins. HDLs are finally catabolized through interactions with scavenger receptor B1 (SR-BI) mainly in the liver, where core CE are selectively transferred into SRB1 expressing cells without whole particle uptake [18]. The remaining apo A-I on the remnant lipid-free (poor) HDL is removed from the circulation by renal filtration. 18.1.4 Drug Interactions with Plasma Lipoproteins Circulating lipoproteins are highly dynamic macromolecules that are under a constant flux of interchanging lipid and apoprotein components. During their transit time in the vascular system other hydrophobic molecules may also associate with the lipophilic compartments of lipoproteins to assume a more favorable thermodynamic state. This is clearly demonstrated by the fact that several hydrophobic/basic drugs and xenobiotics (e.g., vitamins, cyclosporin A, amiodarone, and amphotericin B) passively associate with plasma lipoproteins [19–21]. While many drugs simply associate with lipoprotein through thermodynamic mechanisms, there has been some evidence to suggest that cholesterol ester transfer protein may facilitate the transfer of drugs between the lipoprotein classes [22, 23]. Regardless of the mechanism of uptake, the compartmentalization of the drug within lipoproteins can occur by absorption either within the apolar core or on the external polar milieu of the particle. The relative preference of the drug to reside in the core or surface of the lipoprotein is likely due to the log P value (n-octanol–water partition coefficient) of the drug. Conceivably, a drug might bind to a site within the apolipoprotein; however, no specific examples of such interactions have been reported in the literature. Once associated with lipoproteins the natural course and disposition of the drug/xenobiotic can become significantly altere [24]. Thus binding or encapsulation of drugs with lipoproteins may facilitate receptor-mediated uptake and sequestration of the drug into tissues expressing high levels of lipoprotein receptors, thereby altering the pharmacokinetic, pharmacodynamic, and toxicological profile of the drug. Additionally, the impact that lipoproteins can have on drug disposition is further exaggerated in conditions of dyslipidemia (disruption in normal distribution of lipoprotein classes within the plasma) [24]. 18.1.5 Cholesterol and Cancer To date, numerous studies have been performed documenting the central role of cholesterol in the development of cardiovascular disease. Substratified analyses of cancer patients within these large clinical studies yielded unexpected results. An association between low plasma cholesterol levels and an increased incidence of cancer was observed [25]. This correlation was found to hold for a number of cancers of different origin [26].
P1: OTA/XYZ P2: ABC c18 JWBS037-Chen February 2, 2011
438
21:51
Printer Name: Yet to Come
LIPOPROTEIN-BASED NANOPLATFORMS FOR CANCER MOLECULAR IMAGING
Hypocholesterolemia, resulting from lower total plasma, LDL, and HDL cholesterol is believed to be a consequence rather than a cause of cancer [27, 28]. This is further supported by the observation that cholesterol levels decrease with disease progression. As well, blood cholesterol levels have been shown to normalize relative to pretreatment levels with successful chemotherapy or remission of the disease [29]. The expression of lipoprotein receptors is markedly elevated on many cancer cells. Early reports of Ho et al. [30] cited that acute myelogenous leukemia has 3–100 times higher LDL receptor activity over normal white blood cells. Similarly, Gal et al. [31] reported that numerous human gynecological cancer cell lines exhibit higher uptake of LDL than corresponding normal tissue. More recently, numerous investigators have documented high levels of HDL uptake in breast cancer cells. It is generally believed that upregulated expression of lipoprotein receptor among cancer cells functions to provide high amounts of cholesterol needed for active membrane turnover [32]. This sequestration of cholesterol to an increasing burden of malignant cells may explain the corresponding low levels of plasma cholesterol found in cancer patients.
18.2 CANCER IMAGING WITH LIPOPROTEIN VEHICLES The previous sections have highlighted (1) the tendency of exogenous lipophilic agents (drugs) to associate with circulating lipoproteins and (2) the strong proclivity that cancer cells have toward acquiring lipoproteins. Given these findings, the next logical step would be to preload or formulate lipoproteins to carry exogenous agents for cancer treatment or detection. In fact, many investigators in the fields of drug delivery and oncology have attempted this dating as far back as 1981 [31]. A number of excellent reviews have been written on the utility of lipoproteins for the delivery of anticancer drugs [33–36]. In this chapter we focus solely on the use of lipoprotein (primarily LDL and HDL) for cancer imaging. 18.2.1 Methods of Probe Incorporation In general, there are three main strategies employed to incorporate probes into lipoproteins. These strategies include covalent modification of the apoprotein component, intercalation of probes within the phospholipid monolayer, and lastly packaging the probe into the lipoprotein core using the “reconstitution” procedure (Fig. 18.3).
Covalent Modification The first strategy in which probes can be incorporated in lipoproteins is through covalent modification of the amino acid residues of the apoproteins. Various residues have been selected as sites for modification; the most common residues include the ε-amino groups of lysine and arginine, the phenolic groups of tyrosine, and the thiol group of cysteine. The number of amino acids that make up the different apoprotein species varies greatly. Apo C-I consist of only 57 amino acids whereas over 4500 amino acids make up apo B-100 [1, 12]. Given the complexity of these multifunctional proteins to conduct lipid binding, receptor recognition, and an array of enzyme catalyzed reactions, these apoproteins should be subject to minimal modification in order to retain normal lipoprotein activity. In the case of apo B-100, the lysine residues that are a common site for chemical modification also play a central role in LDLR recognition. Of the 357 lysine residues within the apo B-100 sequence, 225 are exposed and available for
P1: OTA/XYZ P2: ABC c18 JWBS037-Chen February 2, 2011
21:51
Printer Name: Yet to Come
CANCER IMAGING WITH LIPOPROTEIN VEHICLES
439
FIGURE 18.3 Sites for probe incorporation on or into lipoprotein particles. (1) Covalent modification with amino acid residues in apoprotein component. (2) Surface labeling through intercalation of probes within the phospholipid monolayer. (3) Loading of probe within the lipoprotein core through reconstitution methods.
lipid–protein or protein–protein interactions [37]. However, modification to only 20% of these exposed lysines will abolish apo B-100 capacity to recognize LDLR [37]. Given that apo E is also a ligand for LDLR, it is anticipated that a similar degree of modification to apo E would also result in a similar disruption of receptor recognition. For apo A-I (HDL) its interaction with SR-BI is markedly different from that of the LDL–LDLR system. The apo A-I/SR-BI interaction is complex in that there is not a specific amino acid residue sequence or unique recognition site in apo A-I that mediates binding. Instead, it is the structure motif of the amphipathic ␣-helix present in this apoprotein that is recognized by the SR-BI receptor [38, 39]. As such, several apoproteins (apo A-II, apo-E and apo-C) and multiple sites in the apoprotein may be recognized by SR-BI [38, 40]. Therefore when labeling HDL through covalent modification, the integrity of the amphipathic ␣-helix (clusters of basic amino acid residues located at the polar–nonpolar interfaces of the helix and acidic residues located at the center of the polar face) must be preserved in order to retain normal SR-BI binding. Although covalent modification provides a simple and stable means of labeling lipoproteins, relatively few alterations of the apoprotein are permitted if regular lipoprotein receptor interactions are to be retained. As a result of this stringency, covalent modification methods are limited to probes with high sensitivity, since only a few probe molecules can provide the necessary signal contrast for imaging. Contrast agents such as radiolabels or fluorescent probes lend themselves to this procedure [41, 42]. Numerous reports in the literature have used these methods to study lipoprotein metabolism. Interestingly, excessive modification of apoprotein components has also served as a means to intentionally
P1: OTA/XYZ P2: ABC c18 JWBS037-Chen February 2, 2011
440
21:51
Printer Name: Yet to Come
LIPOPROTEIN-BASED NANOPLATFORMS FOR CANCER MOLECULAR IMAGING
disrupt lipoprotein/receptor interactions (e.g., preparation of acetylated LDL or rerouting of lipoproteins—see Section 18.4) [43, 44].
Intercalation The next strategy in which imaging probes can be incorporated into lipoproteins is through noncovalent association with the phospholipid monolayer. More specifically, the imaging compound nestles or “intercalates” between the phospholipid molecules at the lipoprotein surface. To facilitate this process, the diagnostic agent should be amphiphilic in nature like the phospholipids intended to surround it. Hence for labeling via intercalation, the imaging molecule should contain polar and nonpolar termini. The majority of conventional imaging agents are hydrophilic; thus amphiphilic analogs of these compounds are often prepared by covalently attaching a long hydrocarbon chain or sterol moiety to the parent molecule. Depending on the polarity of the lipophilic conjugate, the analog may partition closer to the hydrophilic surface or the apolar core. This method of labeling is rather attractive as it is relatively simple: usually involving a single incubation step followed by sample purification to separate the free probe from that complexed to the lipoprotein. In addition, very high yields can be achieved with this method. The capacity to ferry intercalated agents correlates with the amount of surface area available on the lipoprotein. Whereas LDL has been shown to carry 200 paramagnetic probes [45] HDL has been cited to only transport 10 [46]. The main caveat to this strategy is that intercalated compounds have a high propensity of non-specific exchange with other membrane surfaces within their immediate vicinity. Such events likely occur in the narrow spacing of distal capillaries, where the lipoproteins are in close apposition with the endothelial lining of the vessel. The mechanism of lipid transfer is a general phenomenon that operates through contact-mediated pathways of intermembrane lipid exchange or mixing that is free of a cellular energy source or complex transport processes. The high free energy stored within the lipoprotein membrane itself (i.e., monolayer bending modulus) likely provides sufficient energy for the thermodynamically favorable exchange of lipids [47]. Lipid exchange may also occur between different lipoprotein species through enzymemediated processes. These reactions are catalyzed by phospholipid transfer protein or cholesterol ester transfer protein [22, 48]. Although direct evidence of enzyme-mediated transfer of exogenous agents has not been documented in the literature, speculative observations do warrant consideration of this process. Regardless of the mechanism of lipid exchange, serum stability tests should be performed to ensure integrity of the surface labeled lipoprotein complex. Core-Loading “Reconstitution” Procedure The previous two strategies for incorporating probes into lipoproteins can be viewed as surface-loading methods as they involve attaching the probes to the protein or phospholipid components of lipoproteins. The third strategy that will be covered is the core-loading strategy. This method of inserting exogenous agents into the interior of lipoproteins, called “reconstitution,” has been most extensively described for LDL particles. This procedure was pioneered by Gustafson and colleague [49] and was later popularized by Krieger in the late 1970s [50, 51]. The method involves the selective extraction of the core lipids using highly nonpolar solvents (e.g., heptane) in the presence of starch. Throughout this procedure the apoprotein is protected and stabilized by the starch; as well, the phospholipid shell remains “intact” in the apolar solvent system. Exogenous compounds solubilized in apolar solvents are then added to the starch–LDL residue, followed by evaporation of the solvent with an inert gas. During
P1: OTA/XYZ P2: ABC c18 JWBS037-Chen February 2, 2011
21:51
Printer Name: Yet to Come
CANCER IMAGING WITH LIPOPROTEIN VEHICLES
441
the course of the solvent evaporation the LDL particle spontaneously reassembles as the exogenous agents are concentrated into the LDL core. This method of labeling lipoproteins is rather attractive as the diagnostic agent is shielded from the external environment until it reaches its designated cellular target, where it is released by LDL catabolism. Another merit of the reconstitution procedure is the carrying capacity of the lipoprotein core. The number of exogenous compounds that can potentially be packaged into the lipoprotein interior can be estimated from the number of endogenous lipid molecules it can normally transport in its core. For example, VLDL can carry an estimated 15,000 molecules of cholesterol esters and triglycerides, LDL 1500 molecules, and HDL2 (10-nm diameter) 109 molecules [52]. In light of these values a study by Li et al. [53] showed that 400 molecules of the fluorophore/photosensitizer derivative tetra-t-butyl silicon phthalocyanine bisoleate could be reconstituted into the core of LDL. The versatility of this method does have some restrictions as reconstituted agents must have sufficient hydrophobicity to be loaded into the LDL core. It should be noted that the reconstitution process is facilitated if the compound contains a long unsaturated hydrocarbon chain (e.g., oleic or linoleic acid) or branched side chains (e.g., polyisoprenes such as phytol) [54]. The sterol ester moiety is not required as modified triglycerides and fatty acids have been loaded into LDL. Furthermore, numerous contrast agent analogs containing either unsaturated fatty acyl chains or sterol ester moieties have also been successfully reconstituted into the LDL core [41, 55, 56]. To date, the core extraction/reconstitution procedure has only been reported for LDL in the literature. The seminal work of Pitman and colleagues demonstrated that synthetic HDL particles can be made from apo A-I and commercial lipids [57]. This process of building HDL from individual constituents has also been cited as a reconstitution method; hence in the literature these synthetic constructs are referred to as “reconstituted HDL particles.” This method of preparing HDL provides the flexibility of introducing exogenous agents into the particle formulation. Similar to LDL, compounds intended for HDL core-loading must be hydrophobic and contain an unsaturated hydrocarbon chain (>10 carbons) or a sterol ester moiety. Several examples of synthetic HDL reconstituted with drug or imaging probes have been described [58–60].
18.2.2 Radionuclide Imaging Investigators in the field of nuclear medicine were the first to exploit the natural cancer targeting capability of lipoproteins for tumor detection and characterization. The small amounts of radiotracers needed for detection and the facile chemistry used for labeling lipoprotein make this modality highly amenable for lipoprotein-based cancer imaging. The two most common radiopharmaceuticals used for labeling lipoproteins have been iodine and technetium. Radioiodination and technetium labeling can be performed by fixing the radionucleotide to the lipoprotein surface. With this method the majority of the radioactive probe will be conjugated to the amino acid residues of the apoprotein component. However, despite this fact, as much as 20% of the radioactivity has been known to become associated with the surface lipids [61, 62]. Alternatively, analogs of lipoprotein core components can be synthesized labeled with a radiopharmaceutical and incorporated into the lipoprotein core through reconstitution or detergent solubilization methods (Fig. 18.4) [55]. Radioiodinated cholesteryl esters and hexaiodinated diglycerides are two such examples [63, 64]. In addition, intercalation of amphiphilic metal chelating complexes, gallium-68
P1: OTA/XYZ P2: ABC c18 JWBS037-Chen February 2, 2011
442
21:51
Printer Name: Yet to Come
LIPOPROTEIN-BASED NANOPLATFORMS FOR CANCER MOLECULAR IMAGING
FIGURE 18.4 Radiolabeled probes for lipoprotein incorporation. (a) Radioiodinated cholesteryl ester. (b) Hexaiodinated diglyceride. (c) Indium-111 DTPA(stearylamide).
or indium-111 DTPA(stearylamide) (see Fig. 18.4), into the phospholipid monolayer can be performed to construct radiolabeled lipoprotein complexes [65]. A great number of studies have been performed using radiolabeled lipoproteins to monitor and characterize lipoprotein metabolism in normal organs, particularly for the liver, adrenal gland, and reproductive tissues [41]. While the number of papers focused on tumor detection has not been as numerous as that of the former, the findings from the tumor studies have been promising. Gal and co-workers were one of the first groups to use radiolabeled lipoproteins to screen a number of different cancer cells of gynecological origin [66]. In these studies they demonstrate that neoplastic cells possess higher binding capacities (15– 30×) and metabolized LDL at an enhanced rate (up to 20×) compared to nonneoplastic cells [66]. In vivo studies by Norata’s group and Hynd’s group went on to support Gal’s initial findings. Norata and co-workers found that murine MS-2 fibrosarcomas took up high levels of iodinated and cyclohexanedione-modified iodinated LDL (via receptormediated and independent pathways) exceeding even that of the host liver (∼2.5-fold)(Fig. 18.5) [67]. Similarly, Hynds and colleagues used scintillation measurements to show that moderate radioactivity was associated with MAC 13 tumors following intravenous (IV) injection of iodinated LDL into tumor bearing mice. The tumor radioactivity was second only to that found in the liver. Furthermore, the investigators showed that they could enhance the tumor targeting selectivity of radiolabeled LDL by pretreating animals with sodium taurocholate and hydrocortisone sodium succinate, which selectively downregulate LDL receptor-mediated uptake in the liver and adrenal glands without affecting the tumor activity [68]. All of the aforementioned studies have relied on ex vivo measurements to demonstrate the assimilation of radiolabeled lipoproteins into tumors. To date, few studies have gone on to assess the utility of these probes for in vivo tumor detection. One such study performed by Ponty and co-workers utilized 99m Tc-LDL to visualize B16-melanoma xenographs in mice [69]. Scintigraphic images demonstrated strong accumulation of the labeled probe in the upper abdomen and in the region of the tumor xenograft (Fig. 18.6). Subsequent
P1: OTA/XYZ P2: ABC c18 JWBS037-Chen February 2, 2011
21:51
Printer Name: Yet to Come
CANCER IMAGING WITH LIPOPROTEIN VEHICLES
443
Relative LDL Uptake (cpm.g-1/plasma cpm)
7 6 5
Control Mouse
Tumor Bearing Mouse
4 3 2 1 0
Liver
Spleen
Muscle
Tumor
Liver
Spleen
Muscle
125
FIGURE 18.5 Relative uptake of I-LDL in tissues from BALB/c mice bearing a fibrosarcoma tumor. The radiolabeled LDL was injected 8–10 days after the injection of 106 cells in the hind leg and the mice were killed 24 h later. Data represents means ±SEM for 10 animals. C = Controls; T = tumor bearing mice. (Reproduced with permission from Norata et al. [67].)
biodistribution analysis of excised tissues revealed that the gamma counts in the tumor were 1.4-fold greater than that of host liver. Interestingly, the authors found that the magnitude of LDL uptake by the tumor was associated with the tumor growth rate, where higher levels of LDL uptake occurred during the early period of tumor growth [69]. One caveat associated with this nuclear imaging method is the poor spatial resolution of this modality. Clear demarcation of tumor and host tissue boundaries is often critical during radiographic tumor detection. Alternate modalities may be enlisted to help address this issue.
UA
RL
UA
LL
T
(a)
LL (b)
FIGURE 18.6 Scintigraphic images of a (a) normal mouse and (b) B16 tumor-bearing mouse taken 18 h after the injection of 99m Tc-LDL[2 2.2 MBq (600 Ci) per mouse]. (Reprinted with permission of Ponty et al. [69].)
P1: OTA/XYZ P2: ABC c18 JWBS037-Chen February 2, 2011
444
21:51
Printer Name: Yet to Come
LIPOPROTEIN-BASED NANOPLATFORMS FOR CANCER MOLECULAR IMAGING
18.2.3 Near-Infrared Fluorescent Optical Imaging Near-infrared (NIR) optical imaging is a sensitive, safe, and inexpensive imaging modality that can be applied for cancer detection [70]. The spectral domain of NIR probes provides a substantial optical window where hemoglobin and water absorption are minimal (650– 900 nm). Differentiation of tumor and normal tissue is based on differences in tissue absorption and fluorescence. Given that tissues are relatively transparent to NIR light, selective delivery of NIR probes to malignant lesions would facilitate the high sensitivity and specificity needed for tumor detection. Packaging these fluorophores into lipoproteins could provide a means whereby NIR probes could be selectively delivered to tumors. Krieger and colleagues were the first to demonstrate that fluorescent probes could be stably formulated into lipoproteins. In these early studies the authors showed that oleoylconjugated fluorophores (3-pyrenemethyl-23, 24-dinor-5-cholen-22-oate-3 beta-yl oleate and dioleyl fluorescein) and a fluorescent derivative of cholesteryl ester (cholesteryl 12-O[methyl 3-O-methyl-5 (6 )-carboxyfluorescein]ricinoleyl carbonate) could be reconstituted into LDL to monitor the LDLR activity in normal and LDLR deficient cells [51, 71]. Several years later, Zheng et al. [72] utilized a similar strategy where they conjugated a tricarbocyanine NIR dye (IRD41; Li-Cor Biotechnology Division, Lincoln, NE) to cholesterol laurate through a thiourea linkage (Fig. 18.7a). NIR carbocyanine dyes are excellent fluorophores due to their long wavelength absorption (750–850 nm), high extinction coefficients (>100,000), and fluorescent quantum yields, making them an ideal probe for cancer detection [73]. These investigators also demonstrated that the visible carbocyanine dye, 1,1 -dioctadecyl-3,3,3 ,3 -tetramethylindocarbocyanine perchlorate (DiI), could also be used to label LDL [74]. Both carbocyanine probes were attached to LDL through surfaceloading methods. They went on to show that these fluorophore-labeled LDL particles could be selectively delivered to various tumors through LDLR-mediated processes [72, 74]. Low-temperature three-dimensional (3D) redox scanning showed that hepatoma, HepG2 xenograft, localized the labeled LDL complexes into the tumor core. Conversely, the distribution of these probes in B16 melanoma xenografts was limited to the tumor periphery due the large central necrosis typically found in these tumors [72, 74]. These findings were also confirmed with a neutral NIR fluorescent bioconjugate, pyropheophorbide cholesterol ester (Pyro-CE) (Fig. 18.7b), which can be loaded within the core of LDL [56]. The tumor distribution of Pyro-CE-LDL appears to be dictated by the perfusion properties of the xenograft, where the B16 melanoma was only well perfused at its periphery while HepG2 tumor was uniformly perfused (Fig. 18.8). Pyropheophorbide is an attractive agent as it can serve both as an NIR fluorescent probe and as a PDT agent [75]. Subsequent studies by this group have shown that several hydrophobic PDT analogs can be loaded in LDL for therapeutic purposes [53, 76]. Complementary work from the lab of Therien showed that NIR fluorescent supermolecular meso-to-meso ethyne-bridged tris[(porphinato)zinc (II)] (PZn3 ) fluorophores can be successfully reconstituted into the core of LDL [77]. The potent fluorescence emission of this agent combined with the minimal cytotoxicity profile of the LDL(PZn3 ) complex make this a promising targeting contrast agent. Indeed, studies from this lab went on to show that LDL(PZn3 ) is actively taken up in B16 melanoma cells and can be imaged at very low doses in the nanomolar range [77]. In vivo fluorescent imaging of tumors using fluorophore-labeled LDL was accomplished with the carbocyanine derivative DiR-BOA (1,1 -dioctadecyl-3,3,3 ,3 tetramethylindotricarbocyanine iodide bis-oleate). Fluorescent LDL was prepared by
P1: OTA/XYZ P2: ABC c18 JWBS037-Chen February 2, 2011
21:51
Printer Name: Yet to Come
445
CANCER IMAGING WITH LIPOPROTEIN VEHICLES
(a)
Me (b) Me
NH N N HN
H Me
H C HN
O C
Me O
O
O
FIGURE 18.7 Structures of (a) tricarbocyanine dye–cholesteryl laurate conjugate and (b) pyropheophorbide cholesterol oleate.
reconstituting DiR-BOA into the lipoprotein core. Figure 18.9 displays serial in vivo optical images of a HepG2 tumor bearing mouse following an IV injection of r-DiRBOA-LDL (molar ratio 90:1). In the immediate postinjection period, low fluorescence can be seen throughout the body. At 1 hour, high-intensity signal can be seen in the abdomen (liver) and by 6 hours high fluorescence can also be seen in the tumor. This pattern persists through to 24 hours postinjection. Similar to the earlier described scintigraphic images, the in vivo fluorescent images show that the majority of the diagnostic signal emanates from the liver and the tumor. 18.2.4 Magnetic Resonance Imaging The high-resolution tomographic capacity that magnetic resonance imaging offers makes it an attractive modality for cancer imaging. The use of amphiphilic chelates has provided a means whereby gadolinium (Gd) contrast agents can be tethered to LDL. The prototype
P1: OTA/XYZ P2: ABC c18 JWBS037-Chen February 2, 2011
446
21:51
Printer Name: Yet to Come
LIPOPROTEIN-BASED NANOPLATFORMS FOR CANCER MOLECULAR IMAGING
r-pyropheophorbide-LDL
250
B-16 R-(pyro-CE)-LDL signal
200
200u m
120 100
150 80
300u m
60
100
40
400u m
20
50
20
40
60
80
100
120
0
R-pyroCE signal
HepG2 R-pyroCE-LDL signal
R-pyroCE-LDL signal
250 120
120
100
100
80
80
60
60
40
40
20
20
200 150 100 50
20
40
60
80
100
120
20
40
60
80
100 120
0
FIGURE 18.8 Redox imaging of the reconstituted pyropheophorbide cholesterol oleate labeled low-density lipoprotein in the HepG2 and B16 tumors.
amphiphilic contrast agent (PTIR267) was the first in this series; it contained both a fluorophore and a Gd(III) moiety along with two palmitate hydrocarbon chains to facilitate intercalation into the LDL phospholipid monolayer (Fig. 18.10) [78]. Approximately 50 molecules of PTIR267 can be attached to each LDL macromolecule. Intravenous administration of this agent into a B16 tumor bearing mouse demonstrated that the labeled LDL accumulated within the liver and tumor (Fig. 18.10) [78]. However, the bulky nature of the PTIR267 complex limits the number of probes that can be loaded onto the LDL surface (upper limit of 50). This becomes a critical issue for MR molecular imaging as high paramagnetic payloads are required to overcome the low inherent sensitivity of MRI. To address this limitation a study by Corbin et al. [45] synthesized a simple and less bulky amphiphilic Gd chelate, DTPA-Bis(stearylamide). The authors cited that it was possible to incorporate much higher payloads (150–500) of the Gd(III) contrast agent into LDL than was previously described. Modest relaxivities were achieved on a per Gd(III) basis, but more compelling relaxation enhancements were identified when viewed with respect to the entire LDL nanoparticle. Coronal MR images of tumor bearing mice display clear enhancement of the tumor xenografts 24 and 36 hours following IV injection of Gd-DTPA-SA-LDL (Fig. 18.11) [45]. These studies represent initial attempts at using LDL as a delivery vehicle for Gd(III). Follow-up investigations by Crich and co-workers further improved upon this system by using novel lipophilic Gd complexes possessing two coordinated water molecule (q = 2)
P1: OTA/XYZ P2: ABC c18 JWBS037-Chen February 2, 2011
21:51
Printer Name: Yet to Come
CANCER IMAGING WITH LIPOPROTEIN VEHICLES
447
FIGURE 18.9 Uptake of core-loaded DiR-LDL in HepG2 tumor-bearing mouse. Intravenous injection consists of 73.9 nmol of DiR and 0.82 nmol of LDL carrier.
coordination sites [79]. The Gd complex, Gd-AAZTAC17, displayed markedly improved relaxivity when inserted into LDL over the initial Gd- DTPA-SA-LDL construct (22 versus 6.5 mM-1 ·s−1 ). In vivo experiments displayed in Figure 18.12 showed that the highest levels of tissue enhancement were achieved in the early postinjection period (first 4 hours) and steadily decreased thereafter over the remaining 48-hour study period.
FIGURE 18.10 (a) Structure of PTIR267, a prototype of dual MRI and fluorescent contrast agent. (b) T1 maps of B16 melanoma in vivo before and 16 h after injection of PTIR267-labeled LDL. The tumors were outlined by dashed contours. The gray level of each pixel represents its T1 value as indicated by the gray scale bar, which presents the T1 value in milliseconds (T1 value greater than 2000 ms is saturated). (Reprinted with permission of Li et al. [74].)
P1: OTA/XYZ P2: ABC c18 JWBS037-Chen February 2, 2011
448
21:51
Printer Name: Yet to Come
LIPOPROTEIN-BASED NANOPLATFORMS FOR CANCER MOLECULAR IMAGING
Pre-contrast
5h
24h
36h
FIGURE 18.11 T1-weighted coronal spin-echo images of a nude mouse with a subcutaneously implanted HepG2 tumor. Images are from the mouse prior to administration of Gd-DTPA-SA-LDL (precontrast) and at various times following the intravenous administration of MR contrast agent (5, 24, and 36 h). Arrow indicates tumor. (Reprinted with permission from Corbin et al. [45].)
To date, only gadolinium-labeled lipoproteins have been used for tumor targeting. For a number of years there has been considerable interest in incorporating superparamagnetic agents (iron oxide) into lipoproteins for this purpose, but until recently this approach remained elusive. In a recent publication by Cormode et al. [80], oleate-coated iron oxide nanocrystals were successfully formulated into the core of HDL particles. This new technology may now open opportunities for targeted superparamagnetic–lipoprotein-based cancer imaging.
FIGURE 18.12 Structure of Gd-AAZTAC17. Fat-suppressed T1-weighted multislice multiecho MR images of C57BL/6 mice grafted subcutaneously with B16 melanoma cells. Images were obtained (a) before and 8 h after the administration of (b) Gd-AAZTAC17 and (c) Gd-AAZTAC17/LDL. Arrows show the enhanced tumor region (outlined in blank). (d) Course of SI enhancement measured in the ROI corresponding to the whole tumors and muscle from mice injected with Gd-AAZTAC17 (full and empty blue circles, respectively) and Gd-AAZTAC17/LDL (full and empty red squares, respectively). Graph shows the standard error of the means (SEM). (Reprinted with permission from Crich et al. [79].)
P1: OTA/XYZ P2: ABC c18 JWBS037-Chen February 2, 2011
21:51
Printer Name: Yet to Come
CANCER IMAGING WITH LIPOPROTEIN VEHICLES
449
18.2.5 Computed Tomography Similar to MRI, X-ray computed tomography (CT) offers excellent spatial resolution and multiplane imaging capacity. However, the high content of contrast media required to elicit significant Housfield unit enhancement precludes the use of X-ray contrast agents for many nanoparticle/molecular imaging approaches [81]. To date, neither LDL nor HDL particles have been used as a tumor targeting vehicles for X-ray contrast media. While the finite loading capacity of LDL and HDL limits their use for such applications, this may not be the case for the much larger VLDL or chylomicron-like particles. Weichart and colleagues have shown that polyiodinated triglyceride analogs (50% by weight iodine) can be loaded into synthetic chylomicron-like lipid emulsions for liver imaging [82, 83]. These particles proved to be highly hepatoselective, as over 70% of the administered dose of these particles localized in the liver within 30 minutes after injection (Fig. 18.13) [82]. Although chylomicrons do not naturally home to tumors, it may prove interesting to reroute
FIGURE 18.13 (a) Chemical structure of the polyiodinated triglyceride compound, 2oleoylglycerol-1,3-bis[7-(3-amino-2,4,6-triiodophenyl)heptanoate] (DHOG). The aromatic iodine in the compound makes up approximately 50% of the weight of the molecule. (b) Volume-weighted particle size distribution profile for a typical 125 I-DHOG-LE formulation. The mean diameter of the formulation was 219 nm with a standard deviation about the mean of 68 nm (31.2%). (c) Representative CT images after injection of DHOG-LE formulations in normal rats. Images were taken before and 30 min, 2 h, or 24 h after injection of (i) 50 mg I/kg or (ii) 90 mg I/kg. (Reprinted with permission from Bakan et al. [82].)
P1: OTA/XYZ P2: ABC c18 JWBS037-Chen February 2, 2011
450
21:51
Printer Name: Yet to Come
LIPOPROTEIN-BASED NANOPLATFORMS FOR CANCER MOLECULAR IMAGING
this iodinated chylomicron particles to tumors/tumor vessels for targeted CT imaging of cancer (see Section 18.4).
18.2.6 Ultrasound To date, there have been no attempts to use lipoproteins to ferry ultrasound contrast agents for cancer detection.
18.3 HDL FOR TUMOR TARGETING Of all the lipoprotein species, LDL has been most extensively documented as a drug delivery vehicle for targeting cancer cells. More recently, there has been an increasing interest in the association of HDL with cancer. In a large clinical study consisting of 530 cancer patients, serum HDL cholesterol, like LDL cholesterol, was seen to be consistently low compared to noncancer subjects [84]. Within this study group, patients with hematological malignancies displayed the lowest HDL cholesterol while those with breast tumors showed the highest [84]. In addition to supplying cholesterol for active membrane turnover, there is also evidence to suggest that HDL may have a mitogenic effect on various cell types [85]. Early work by Favre et al. [86] showed that HDL or HDL components were able to stimulate proliferation of adenocarcinoma cell lines. Since then, others have also concurred with these findings and have reported on the stimulating effects of HDL on various other cancer cell lines. Furthermore, studies by Cao and co-workers showed that by disrupting the integrity of the HDL receptor (mutation of residues at the carboxy terminus) the proliferation of breast cancer cells can be markedly inhibited [87]. Interestingly, HDL stimulation of cellular proliferation is not the only mechanism for augmenting cell growth, as other studies have suggested that HDL can induce antiapoptotic effects in various cell types [88]. Finally, high levels of HDL receptor have also been reported among cancer cells over surrounding normal cells [89], thus providing malignant cells a selective advantage for acquiring this potent growth stimulant. These collective findings have led several investigators to explore the utility of HDL as a tumor targeting vehicle for drug delivery. Unlike LDL, HDL-like particles (reconstituted HDL) can be prepared from apoproteins (apo A-I with or without apo A-II) and commercial lipids [57]. This affords the opportunity to easily introduce foreign constituents to the HDL particles. Lacko and colleagues have formulated the microtubule toxin paclitaxel into HDL and have shown enhanced efficacy of this complex against breast cancer cells compared to the conventional paclitaxel formulation (Fig. 18.14) [90]. Similar results were also observed by Lou et al. [91], who incorporated aclacinomycin into HDL particles to treat hepatoma cells. To date, there have not been any reports in the literature of labeling HDL particles with diagnostic probes for cancer detection. However, the laboratory of Fayad has published a number of papers utilizing paramagnetic and fluorescent labeled HDL particles for MRI and optical detection of atherosclerotic plaques (Fig. 18.15) [46, 92]. More recently, this group has demonstrated that gold, iron oxide, or quantum dot nanocrystals may also be incorporated into HDL particles for CT, MRI, and fluorescent imaging, respectively (Fig. 18.16) [80]. The convergence of the HDL delivery system with nanocrystal-based probes will undoubtedly open new opportunities for imaging studies in biomedical and cancer research [93].
P1: OTA/XYZ P2: ABC c18 JWBS037-Chen February 2, 2011
21:51
Printer Name: Yet to Come
REROUTING LIPOPROTEINS TO CANCER SPECIFIC TARGETS
451
FIGURE 18.14 Cytotoxicity of the reconstituted high-density lipoprotein (rHDL)/paclitaxel (PTX) nanoparticles versus free PTX estimated by the MTT assay. (Reproduced with permission from McConathy et al. [90].)
18.4 REROUTING LIPOPROTEINS TO CANCER SPECIFIC TARGETS Although contrast agent-bearing lipoprotein nanoparticles are able to accumulate in cancer cells, this platform suffers from limited tumor selectivity over normal tissues. The liver, adrenal glands, and reproductive organs all express high levels of LDLR [94, 95] and thus would effectively compete and be exposed to high levels of the administered imaging agent. Even though the lipoprotein receptor expression levels in these tissues may be modulated by the concomitant use of steroids, bile acids, or dietary fats [96, 97], the downregulation of lipoprotein receptor is not absolute and the extent of this protective effect is uncertain. Moreover, the range for potential lipoprotein-based applications in cancer medicine is
apo A-1 NDB-DPPE
Phospholipid Cholesterol Ester Triglyceride
Gd-DTPA-DMPE
Pre
Unesterified Cholesterol
24 h
FIGURE 18.15 Gd (III) labeled HDL-like nanoparticle contrast agent displays selective enhancement for atherosclerotic plaques 24 h following intravenous injection of the agent. Schematic of the HDL-like MRI contrast agent is also shown. (Reprinted with permission of Frias et al. [46].)
P1: OTA/XYZ P2: ABC c18 JWBS037-Chen February 2, 2011
452
21:51
Printer Name: Yet to Come
LIPOPROTEIN-BASED NANOPLATFORMS FOR CANCER MOLECULAR IMAGING
FIGURE 18.16 Nanocrystal core high-density lipoprotein. (a) Schematic depiction of the different agents in this study. (b) Negative stain TEM images of FeO-HDL. (c) Photograph of the QD-HDL in (left) normal light and (right) under UV illumination. (d) Phantoms of Au-HDL imaged using CT. (Reprinted with permission of Cormode et al. [80].)
limited by the narrow purview of lipoprotein receptor-positive tumors. Currently, there are a number of “cancer signatures” that bind to surface receptor/epitopes that are much more tumor specific than LDLR/HDLR; these include, for example, Her2/neu [98], EGF [99], somatostatin [100], folate (FA [101], and ␣v 3 integrins [102]. Incorporating such targeting ligands into lipoprotein carriers would greatly improve both the targeting specificity and range of cancers reached by this system. In fact, the laboratory of Zheng and co-workers has recently reported that lipoprotein particles can be redirected from their native receptors to alternate surface receptors or epitopes of choice [43]. The redirecting strategy is based on the chemistry of the exposed active lysine residues in the apoprotein component. These residues play a central role in recognition and binding of lipoprotein to lipoprotein receptor; thus by conjugating homing ligands to the ε-amino groups of these lysine residues one effectively abolishes the affinity of the lipoprotein particle for its lipoprotein receptor and simultaneously redirects the particles to alternate receptors. Proof of the redirection strategy was first demonstrated with fluorescent DiI labeled and photosensitizer reconstituted LDL, using folic acid (FA) modification of apo B-100 to redirect the LDL particle to the folate receptor (FR) [43]. Apo B-100 contains 357 Lys residues; of these, 225 are exposed and are available for ligand conjugation [37]. Fifty-three of the exposed Lys ε-amino groups are considered “active” with a low pK a value of 8.9, while the remaining 172 Lys are “normal” with a pK a
P1: OTA/XYZ P2: ABC c18 JWBS037-Chen February 2, 2011
21:51
Printer Name: Yet to Come
REROUTING LIPOPROTEINS TO CANCER SPECIFIC TARGETS
453
FIGURE 18.17 (a) Schematic of DiR-LDL-FA. (b) Confocal images of (a) KB (FR+) alone, (b) KB + DiR-LDL-FA, (c) KB+DiR-LDL-FA + 200-fold FA, (d) HT1080 (FR-) alone, and (e) HT1080 + DiR-LDL-FA. (c) Fluroescence images of tissues and tumors excised from tumor-bearing mouse 24 h following the intravenous injection of DiR-LDL-FA (5.8 M). (Reprinted with permission of Chen et al. [103].)
value of 10.5 [37]. Thus by adjusting the pH of the conjugation reaction, the majority of the active lysines can be conjugated with FA. FA-conjugated LDL was shown to be selectively taken up in FR-positive cancer cells (KB cells) over FR-deficient cells (HT1080 or CHO) or cells expressing high levels of LDLR (HepG2) (Fig. 18.17) [43]. Similar results were also generated in vivo in nude mice bearing dual tumor xenografts of KB and HT1080. Although these findings demonstrated that FR targeting can be achieved with the LDL nanoparticle, competitive uptake processes (RES, opsonization, and possible absorption of exchangeable apoproteins) by the liver and spleen compromise the tumor uptake of FA-LDL (Fig. 18.17) [103]. The rerouting technology was next investigated with HDL particles due to their small size and intriguing pharmacokinetic properties. HDL was first reconstituted with the lipophilic NIR optical dye (DiR-BOA); this served as an optical tracer for the activity and distribution of the nanoparticle. In a similar manner as with LDL, folic acid moieties were conjugated to the lysine residues of HDL’s apoprotein component (apo A-1) [60]. The Lys residues in apo A-I display various pK a values ranging from 8.3 to 10.5. Thus careful titration of the reaction pH is critical to saturate all the Lys residues in apo A-I. Following successful preparation of this FA-conjugated fluorescent HDL, rerouting from SR-BI to the FR was demonstrated in in vitro and in vivo experiments. The in vivo data showed particularly impressive results as exceptional targeting activity was displayed by the FA-conjugated HDL nanoparticle (Fig. 18.18). High levels of nanoparticle accumulation were detected within the target FR-expressing tumor while minimal uptake was evident in the FR-negative tumor or in tissues of the RES [60]. Ongoing experiments are being conducted to explore other ligands such as EGF and PSMA for studies targeting HDL to lung and prostate cancer, respectively.
P1: OTA/XYZ P2: ABC c18 JWBS037-Chen February 2, 2011
454
21:51
Printer Name: Yet to Come
LIPOPROTEIN-BASED NANOPLATFORMS FOR CANCER MOLECULAR IMAGING
FIGURE 18.18 (a) In vivo fluorescent images of dual tumor (KB/HT-1080) bearing mouse before and at several times following intravenous injection of (DiR-BOA)rHDL-FA (15 M). (b) In vivo inhibition study was performed with 30-fold excess free folic acid.
18.5 FUTURE DIRECTION FOR LIPOPROTEIN-BASED CARRIERS One of the major drawbacks of utilizing lipoproteins as a delivery vehicle for exogenous agents is the need to isolate particles or their respective apoproteins from fresh donor plasma. Relying on donor plasma to acquire lipoprotein is problematic because lipoprotein samples may vary from batch to batch, methods for isolating the various lipoprotein species can be lengthy (up to 72 hours), large quantities of lipoproteins can be difficult to attain, and safety issues surrounding the use of blood products remains a matter of concern. Furthermore, isolated lipoproteins can only be stored for finite periods before aggregation and degradative processes compromise the integrity of the sample. As a result of these limitations, attempts have been made to prepare synthetic lipoprotein-like particles. In the case of LDL, Nikanjam et al. [104] have demonstrated that synthetic LDL-like particles can be prepared from a lipid emulsion and a synthetic bifunctional 29 amino acid peptide, which contains a lipid binding motif and the LDLR binding domain of apo B-100. These LDL-like particles were shown to mimic the behavior of native LDL and target the LDLR on carcinoma cells. The authors also went on to show that a lipid derivative prodrug of
P1: OTA/XYZ P2: ABC c18 JWBS037-Chen February 2, 2011
21:51
Printer Name: Yet to Come
FUTURE DIRECTION FOR LIPOPROTEIN-BASED CARRIERS
455
Nonpolar face (a) Ala 17
Leu 3
Val 10
Leu 14
Tyr 7
Lys 13
Phe 18 Ala 11 Lys 4
Trp 2 Lys + 9
Lys 15 Asp 8 –
Phe 6
+
–
Glu 16
Ala 5
Glu 12
Asp 1
+ +
–
–
(b)
Polar face 18A 40
A. 1
30 20 10 80
120
160
200
240
280
320
FIGURE 18.19 (a) Axial projection of 18A peptide showing the relative location of the amino acid residues with segregation of the hydrophobic and hydrophilic amino acids. (b) Size distribution and transmission electron microscopy of discoidal complexes. (Reprinted with permission from Anantharamaiah et al. [106].)
paclitaxel, paclitaxel oleate, could be loaded in these particles and effectively delivered and activated in glioblastoma multiforme cancer cells via the LDLR-mediated pathway [105]. For a number of years the pharmaceutical industry has invested a considerable amount of effort into investigating apo A-I mimetic peptides as a potential cardiovascular therapeutic. This investigation led to the synthesis of an 18 amino acid peptide known as 18A [106]. The sequence of 18A (D-W-L-K-A-F-Y-D-K-V-A-E-K-L-K-E-A-F) does not have any sequence homology to apo A-I. However, the peptide displays characteristic features of apo A-I, a Class A amphipathic ␣-helix protein. The peptide contains opposing polar and nonpolar faces orientated along the long axis of the helix with positively charged lysine residues residing at the polar–nonpolar interface and clusters of negatively charged residues at the apex of the polar face (Fig. 18.19a) [107]. This unique amphipathic ␣ helix orientation imparts strong lipid binding properties to the peptide. Early work by Anantharamaiah’s group showed that these peptides could not only facilitate reverse cholesterol transport but when mixed with phospholipids they also possessed capacity to spontaneously form discoidal complexes similar to nascent HDL particles (Fig. 18.19b) [106, 107]. A few groups have recently explored the utility of these peptide stabilized disks as a delivery platform for drugs or imaging agents. Tufteland and co-workers showed that peptide–phospholipid disks could be formulated with the antifungal amphotericin B [108]. These complexes
P1: OTA/XYZ P2: ABC c18 JWBS037-Chen February 2, 2011
456
21:51
Printer Name: Yet to Come
LIPOPROTEIN-BASED NANOPLATFORMS FOR CANCER MOLECULAR IMAGING
displayed potent antifungal activity. Thereafter, Cormode and colleagues used a dimer of 18A to make a paramagnetic labeled peptide–phospholipid disk to image atherosclerotic plaques [109]. Investigations utilizing fully synthetic lipoprotein-like particles for drug and contrast agent will undoubtedly continue, and over the next few years it is anticipated that many more studies in this area will be published. As progress continues in this field, issues of peptide immunocompatibility and particle/payload stability of these synthetic lipoproteinlike particles should be addressed. Although several issues remain to be resolved, this technology shows great promise and will certainly advance the drug delivery field. Providing synthetic lipoprotein-like particles that are free of plasma-derived components will help overcome the many obstacles that have hampered the progress of lipoprotein-based delivery systems in both basic and clinical research. 18.6 CONCLUSION Lipoproteins are the main transport system for important hydrophobic molecules such as cholesterol and fatty acids. The compartmentalized organization of these carriers, which allows for the native transport of these molecules, also makes them amenable to facile incorporation of exogenous compounds. From the perspective of cancer molecular imaging, loading of contrast agents into modified lipoproteins confers several advantages, such as payload delivery of imaging probe, improved pharmacokinetics, and the ability to selectively target these probes to cancer cells. In summary, this chapter provides an introduction to the biology of lipoproteins: from the metabolism of lipoproteins to its relationship with malignant tumors. In addition, it has outlined existing techniques for incorporating imaging agents into lipoproteins, as well as introducing novel methods of rerouting lipoprotein to cancer-specific eptitopes. While lipoproteins remain an intriguing candidate for delivery of diagnostic agents to cancer cells, the future clinical application of lipoproteins remains unclear. This is due in large part to issues involving the utilization of human blood products as the starting material. However, recent improvements in protein synthesis and recombinant technology have provided an alternative means for starting materials. Development of recombinant apoproteins and synthetic apoprotein mimetics provides a plausible strategy for future application of this technology and may herald a new age for cancer-targeted diagnostic imaging. ACKNOWLEDGMENT Work from the lab of G.Z. was supported by the Canadian Institutes of Health Research (82847), the Ontario Institute for Cancer Research through funding provided by the Government of Ontario, the Canadian Cancer Society (018510), the National Institute of Health (N01 CO37119 and R21/R33 CA114463), and the Joey and Toby Tanenbaum/Brazilian Ball Chair in Prostate Cancer Research. REFERENCES 1. Mahley, R. W.; Innerarity, T. L.; Rall, S. C., Jr.; Weisgraber, K. H. Plasma lipoproteins: apolipoprotein structure and function. J. Lipid Res. 1984, 25, 1277–1294.
P1: OTA/XYZ P2: ABC c18 JWBS037-Chen February 2, 2011
21:51
Printer Name: Yet to Come
REFERENCES
457
2. Gotto, A. M.; Pownall, H.; Havel, R. J. Introduction to the plasma lipoproteins. Methods Enzymol. 1986, 128, 3–40. 3. Segrest, J. P.; De Loof, H.; Dohlman, J. G.; Brouillette, C. G.; Anantharamaiah, G. M. Amphipathic helix motif: classes and properties. Proteins 1990, 8, 103–117. 4. Segrest, J. P.; et al. The amphipathic helix in the exchangeable apolipoproteins: a review of secondary structure and function. J. Lipid Res. 1992, 33, 141–166. 5. Duriez, P.; Fruchart, J. C. High-density lipoprotein subclasses and apolipoprotein A-I. Clin. Chim. Acta: Int. J. Clin. Chem. 1999, 286, 97–114. 6. Francone, O. L.; Gurakar, A.; Fielding, C. Distribution and functions of lecithin:cholesterol acyltransferase and cholesteryl ester transfer protein in plasma lipoproteins. Evidence for a functional unit containing these activities together with apolipoproteins A-I and D that catalyzes the esterification and transfer of cell-derived cholesterol. J. Bio. Chem. 1989, 264, 7066–7072. 7. Barter, P. J.; Rye, K. A. Molecular mechanisms of reverse cholesterol transport. Curr. Opin. Lipidol. 1996, 7, 82–87. 8. LaRosa, J. C.; Levy, R. I.; Herbert, P.; Lux, S. E.; Fredrickson, D. S. A specific apoprotein activator for lipoprotein lipase. Biochem. Biophys. Res. Commun. 1970, 41, 57–62. 9. Soutar, A. K.; et al. Effect of the human plasma apolipoproteins and phosphatidylcholine acyl donor on the activity of lecithin: cholesterol acyltransferase. Biochemistry 1975, 14, 3057– 3064. 10. Breckenridge, W. C.; Little, J. A.; Steiner, G.; Chow, A.; Poapst, M. Hypertriglyceridemia associated with deficiency of apolipoprotein C-II. N. Engl. J. Med. 1978, 298, 1265–1273. 11. Innerarity, T. L.; Mahley, R. W. Enhanced binding by cultured human fibroblasts of apo-Econtaining lipoproteins as compared with low density lipoproteins. Biochemistry 1978, 17, 1440–1447. 12. Segrest, J. P.; Jones, M. K.; De Loof, H.; Dashti, N. Structure of apolipoprotein B-100 in low density lipoproteins. J. Lipid Res. 2001, 42, 1346–1367. 13. Powell, L. M.; et al. A novel form of tissue-specific RNA processing produces apolipoproteinB48 in intestine. Cell 1987, 50, 831–840. 14. Knott, T. J.; et al. Complete protein sequence and identification of structural domains of human apolipoprotein B. Nature 1986, 323, 734–738. 15. Rassart, E.; et al. Apolipoprotein D. Biochim. Biophys. Acta 2000, 1482, 185–198. 16. Spady, D. K. Lipoproteins in biological fluids and compartments: synthesis, interconversions, and catabolism. Targeted Diagn. Ther. 1991, 5, 1–44. 17. Koppaka, V.; Silvestro, L.; Engler, J. A.; Brouillette, C. G.; Axelsen, P. H. The structure of human lipoprotein A-I. Evidence for the “belt” model. J. Biol. Chem. 1999, 274, 14541–14544. 18. Acton, S.; et al. Identification of scavenger receptor SR-BI as a high density lipoprotein receptor. Science (N.Y.) 1996, 271, 518–520. 19. Wasan, K. M.; Brazeau, G. A.; Keyhani, A.; Hayman, A. C.; Lopez-Berestein, G. Roles of liposome composition and temperature in distribution of amphotericin B in serum lipoproteins. Antimicrob. Agents Chemother. 1993, 37, 246–250. 20. Yau, J. C.; Brunner, L. J.; Lopez-Berestein, G.; LeMaistre, C. F.; Luke, D. R. Therapeutic drug monitoring of cyclosporine-lipoprotein levels. Pharmacotherapy 1991, 11, 291–295. 21. Shayeganpour, A.; Lee, S. D.; Wasan, K. M.; Brocks, D. R. The influence of hyperlipoproteinemia on in vitro distribution of amiodarone and desethylamiodarone in human and rat plasma. Pharm. Res. 2007, 24, 672–678. 22. Wasan, K. M.; Morton, R. E.; Rosenblum, M. G.; Lopez-Berestein, G. Decreased toxicity of liposomal amphotericin B due to association of amphotericin B with high-density lipoproteins: role of lipid transfer protein. J. Pharm. Sci. 1994, 83, 1006–1010.
P1: OTA/XYZ P2: ABC c18 JWBS037-Chen February 2, 2011
458
21:51
Printer Name: Yet to Come
LIPOPROTEIN-BASED NANOPLATFORMS FOR CANCER MOLECULAR IMAGING
23. Wasan, K. M.; Ramaswamy, M.; Wong, W.; Pritchard, P. H. Lipid transfer protein I facilitated transfer of cyclosporine from low- to high-density lipoproteins is only partially dependent on its cholesteryl ester transfer activity. J. Pharmacol. Exp. Ther. 1998, 284, 599–605. 24. Wasan, K. M.; Brocks, D. R.; Lee, S. D.; Sachs-Barrable, K.; Thornton, S. J. Impact of lipoproteins on the biological activity and disposition of hydrophobic drugs: implications for drug discovery. Nat. Rev. 2008, 7, 84–99. 25. Feinleib, M. Review of the epidemiological evidence for a possible relationship between hypocholesterolemia and cancer. Cancer Res. 1983, 43, 2503s–2507s. 26. Schatzkin, A.; et al. Site-specific analysis of total serum cholesterol and incident cancer in the National Health and Nutrition Examination Survey I Epidemiologic Follow-up Study. Cancer Res. 1988, 48, 452–458. 27. Chao, F. C.; Efron, B.; Wolf, P. The possible prognostic usefulness of assessing serum proteins and cholesterol in malignancy. Cancer 1975, 35, 1223–1229. 28. Rose, G.; Shipley, M. J. Plasma lipids and mortality: a source of error. Lancet 1980, 1, 523–526. 29. Budd, D.; Ginsberg, H. Hypocholesterolemia and acute myelogenous leukemia. Association between disease activity and plasma low-density lipoprotein cholesterol concentrations. Cancer 1986, 58, 1361–1365. 30. Ho, Y. K.; Smith, R. G.; Brown, M. S.; Goldstein, J. L. Low-density lipoprotein (LDL) receptor activity in human acute myelogenous leukemia cells. Blood 1978, 52, 1099–1114. 31. Gal, D.; Ohashi, M.; MacDonald, P. C.; Buchsbaum, H. J.; Simpson, E. R. Low-density lipoprotein as a potential vehicle for chemotherapeutic agents and radionucleotides in the management of gynecologic neoplasms. Am. J. Obstet. Gynecol. 1981, 139, 877–885. 32. Favre, G. [Targeting of tumor cells by low density lipoproteins: principle and use of ellipticin derivatives]. C. R. Soc. Biol. 1992, 186, 73–87. 33. Kader, A.; Pater, A. Loading anticancer drugs into HDL as well as LDL has little effect on properties of complexes and enhances cytotoxicity to human carcinoma cells. J. Control. Release 2002, 80, 29–44. 34. Masquelier, M.; Vitols, S.; Peterson, C. Low-density lipoprotein as a carrier of antitumoral drugs: in vivo fate of drug–human low-density lipoprotein complexes in mice. Cancer Res. 1986, 46, 3842–3847. 35. Rensen, P. C.; et al. Recombinant lipoproteins: lipoprotein-like lipid particles for drug targeting. Adv. Drug Deliv. Rev. 2001, 47, 251–276. 36. Shaw, J. M.; et al. Delivery of lipophilic drugs using lipoproteins. Ann. N.Y. Acad. Sci. 1987, 507, 252–271. 37. Lund-Katz, S.; Ibdah, J. A.; Letizia, J. Y.; Thomas, M. T.; Phillips, M. C. A 13 C NMR characterization of lysine residues in apolipoprotein B and their role in binding to the low density lipoprotein receptor. J. Biol. Chem. 1988, 263, 13831–13838. 38. Thuahnai, S. T.; Lund-Katz, S.; Anantharamaiah, G. M.; Williams, D. L.; Phillips, M. C. A quantitative analysis of apolipoprotein binding to SR-BI: multiple binding sites for lipid-free and lipid-associated apolipoproteins. J. Lipid Res. 2003, 44, 1132–1142. 39. Williams, D. L.; et al. Binding and cross-linking studies show that scavenger receptor BI interacts with multiple sites in apolipoprotein A-I and identify the class A amphipathic alpha-helix as a recognition motif. J. Biol. Chem. 2000, 275, 18897–18904. 40. Liadaki, K. N.; et al. Binding of high density lipoprotein (HDL) and discoidal reconstituted HDL to the HDL receptor scavenger receptor class B type I. Effect of lipid association and APO A-I mutations on receptor binding. J. Biol. Chem. 2000, 275, 21262–21271. 41. Counsell, R. E.; Schwendner, S. W.; DeForge, L. E.; DeGalan, M. R.; Newton, R. S. Lipoproteins as carriers for organ-imaging radiopharmaceuticals. Targeted Diagn. Ther. 1991, 5, 251–314.
P1: OTA/XYZ P2: ABC c18 JWBS037-Chen February 2, 2011
21:51
Printer Name: Yet to Come
REFERENCES
459
42. Hofer, G.; Steyrer, E.; Kostner, G. M.; Hermetter, A. LDL-mediated interaction of Lp[a] with HepG2 cells: a novel fluorescence microscopy approach. J. Lipid Res. 1997, 38, 2411–2421. 43. Zheng, G.; Chen, J.; Li, H.; Glickson, J. D. Rerouting lipoprotein nanoparticles to selected alternate receptors for the targeted delivery of cancer diagnostic and therapeutic agents. Proc. Natl. Acad. Sci. U.S.A. 2005, 102, 17757–17762. 44. Basu, S. K.; Goldstein, J. L.; Anderson, G. W.; Brown, M. S. Degradation of cationized low density lipoprotein and regulation of cholesterol metabolism in homozygous familial hypercholesterolemia fibroblasts. Proc. Nat. Acad. Sci. U.S.A. 1976, 73, 3178–3182. 45. Corbin, I. R.; et al. Low-density lipoprotein nanoparticles as magnetic resonance imaging contrast agents. Neoplasia (N.Y.) 2006, 8, 488–498. 46. Frias, J. C.; Williams, K. J.; Fisher, E. A.; Fayad, Z. A. Recombinant HDL-like nanoparticles: a specific contrast agent for MRI of atherosclerotic plaques. J. Am. Chem. Soc. 2004, 126, 16316–16317. 47. Markin, V. S.; Albanesi, J. P. Membrane fusion: stalk model revisited. Biophy. J. 2002, 82, 693–712. 48. Kwong, M.; Sivak, O.; Kwong, E. H.; Wasan, K. M. Cyclosporine A transfer between highand low-density lipoproteins: independent from lipid transfer protein I-facilitated transfer of lipoprotein-coated phospholipids because of high affinity of cyclosporine a for the protein component of lipoproteins. J. Pharm. Sci. 2001, 90, 1308–1317. 49. Gustafson, A.; Alaupovic, P.; Furman, R. H. Studies of the composition and structure of serum lipoproteins: isolation, purification, and characterization of very low density lipoproteins of human serum. Biochemistry 1965, 4, 596–605. 50. Krieger, M.; Brown, M. S.; Faust, J. R.; Goldstein, J. L. Replacement of endogenous cholesteryl esters of low density lipoprotein with exogenous cholesteryl linoleate. Reconstitution of a biologically active lipoprotein particle. J. Biol. Chem. 1978, 253, 4093–4101. 51. Krieger, M.; et al. Reconstituted low density lipoprotein: a vehicle for the delivery of hydrophobic fluorescent probes to cells. J. Supramol. Structure 1979, 10, 467–478. 52. Shen, B. W.; Scanu, A. M.; Kezdy, F. J. Structure of human serum lipoproteins inferred from compositional analysis. Proc. Natl. Acad. Sci. U.S.A. 1977, 74, 837–841. 53. Li, H.; Marotta, D.; Kim, S.K.; et al. High payload delivery of optical imaging and photodynamic therapy agents to tumors using phthalocyanine-reconstituted LDL nanoparticles. J. Biomed. Opt. 2005, 10, 41203. 54. Krieger, M. Reconstitution of the hydrophobic core of low-density-lipoprotein. Methods Enzymol. 1986, 128, 608–613. 55. Deforge, L. E.; Degalan, M. R.; Ruyan, M. K.; Newton, R. S.; Counsell, R. E. Comparison of methods for incorporating a radioiodinated residualizing cholesteryl ester analog into low density lipoprotein. Int. J. Radiat. Applications Instrum. 1992, 19, 775–782. 56. Zheng, G.; et al. Low-density lipoprotein reconstituted by pyropheophorbide cholesteryl oleate as target-specific photosensitizer. Bioconjug. Chem. 2002, 13, 392–396. 57. Pittman, R. C.; Glass, C. K.; Atkinson, D.; Small, D. M. Synthetic high density lipoprotein particles. Application to studies of the apoprotein specificity for selective uptake of cholesterol esters. J. Biol. Chem. 1987, 262, 2435–2442. 58. Bijsterbosch, M. K.; van de Bilt, H.; van Berkel, T. J. Specific targeting of a lipophilic prodrug of iododeoxyuridine to parenchymal liver cells using lactosylated reconstituted high density lipoprotein particles. Biochem. Pharmacol. 1996, 52, 113–121. 59. Bijsterbosch, M. K.; et al. Carrier-mediated delivery improves the efficacy of 9-(2phosphonylmethoxyethyl)adenine against hepatitis B virus. Mol. Pharmacol. 2001, 60, 521–527.
P1: OTA/XYZ P2: ABC c18 JWBS037-Chen February 2, 2011
460
21:51
Printer Name: Yet to Come
LIPOPROTEIN-BASED NANOPLATFORMS FOR CANCER MOLECULAR IMAGING
60. Corbin, I.; et al. Enhanced cancer-targeted delivery using engineered high-density lipoproteinbased nanocarriers. J. Biomed. Nanotechnol. 2007, 3, 367–376. 61. Eisenberg, S.; Windmueller, H. G.; Levy, R. I. Metabolic fate of rat and human lipoprotein apoproteins in the rat. J. Lipid Res. 1973, 14, 446–458. 62. Sigurdsson, G.; Noel, S. P.; Havel, R. J. Quantification of the hepatic contribution to the catabolism of high density lipoproteins in rats. J. Lipid Res. 1979, 20, 316–324. 63. DeForge, L. E.; Ruyan, M. K.; Schwendner, S. W.; Newton, R. S.; Counsell, R. E. Synthesis and evaluation of radioiodinated cholesteryl ethers as lipoprotein probes. Bioconjug. Chem. 1991, 2, 254–260. 64. Xiao, W.; Wang, L.; Ryan, J. M.; Pater, A.; Liu, H. Incorporation of an (125)I-labeled hexaiodinated diglyceride analog into low-density lipoprotein and high specific uptake by cells of cervical carcinoma cell lines. Radiat. Res. 1999, 152, 250–256. 65. Moerlein, S. M.; Daugherty, A.; Sobel, B. E.; Welch, M. J. Metabolic imaging with gallium-68and indium-111-labeled low-density lipoprotein. J. Nucl. Med. 1991, 32, 300–307. 66. Gal, D.; MacDonald, P. C.; Porter, J. C.; Simpson, E. R. Cholesterol metabolism in cancer cells in monolayer culture. III. Low-density lipoprotein metabolism. Int. J. Cancer 1981, 28, 315–319. 67. Norata, G.; et al. In vivo assimilation of low density lipoproteins by a fibrosarcoma tumour line in mice. Cancer Lett. 1984, 25, 203–208. 68. Hynds, S. A.; et al. Low-density lipoprotein metabolism in mice with soft tissue tumours. Biochim. Biophys. Acta 1984, 795, 589–595. 69. Ponty, E.; et al. Biodistribution study of 99m Tc-labeled LDL in B16-melanoma-bearing mice. Visualization of a preferential uptake by the tumor. Int. J. Cancer 1993, 54, 411–417. 70. Sevick-Muraca, E. M.; Lopez, G.; Reynolds, J. S.; Troy, T. L.; Hutchinson, C. L. Fluorescence and absorption contrast mechanisms for biomedical optical imaging using frequency-domain techniques. Photochem. Photobiol. 1997, 66, 55–64. 71. Krieger, M.; Ho, Y. K.; Falck, J. R. Reconstitution of LDL with lipophilic fluorescein derivatives: quantitative analysis of the receptor activity of human lymphocytes. J. Receptor Res. 1983, 3, 361–375. 72. Zheng, G.; et al. Tricarbocyanine cholesteryl laurates labeled LDL: new near infrared fluorescent probes (NIRFs) for monitoring tumors and gene therapy of familial hypercholesterolemia. Bioorg. Med. Chem. Lett. 2002, 12, 1485–1488. 73. Mujumdar, R. B.; Ernst, L. A.; Mujumdar, S. R.; Waggoner, A. S. Cyanine dye labeling reagents containing isothiocyanate groups. Cytometry 1989, 10, 11–19. 74. Li, H.; et al. Carbocyanine labeled LDL for optical imaging of tumors. Acad. Radiol. 2004, 11, 669–677. 75. Blessington, D.; et al. Detection and imaging of the reconstituted pyropheophorbide-cholesterol oleate labeled low-density lipoprotein in the hepG2 tumor. Proc. SPIE 2003, 4955, 497–504. 76. Song, L.; et al. Naphthalocyanine-reconstituted LDL nanoparticles for in vivo cancer imaging and treatment. Int. J. Nanomed. 2007, 2, 767–774. 77. Wu, S. P.; et al. Near-infrared optical imaging of B16 melanoma cells via. lowdensity lipoprotein-mediated uptake and delivery of high emission dipole strength tris[(porphinato)zinc(II)] fluorophores. Bioconjug. Chem. 2005, 16, 542–550. 78. Li, H.; et al. MR and fluorescent imaging of low-density lipoprotein receptors. Acad. Radiol. 2004, 11, 1251–1259. 79. Crich, S. G.; et al. Magnetic resonance imaging detection of tumor cells by targeting low-density lipoprotein receptors with Gd-loaded low-density lipoprotein particles. Neoplasia (N.Y.) 2007, 9, 1046–1056.
P1: OTA/XYZ P2: ABC c18 JWBS037-Chen February 2, 2011
21:51
Printer Name: Yet to Come
REFERENCES
461
80. Cormode, D. P.; et al. Nanocrystal core high-density lipoproteins: a multimodality contrast agent platform. Nano Lett. 2008, 8, 3715–3723. 81. Speck, U. Contrast agents: X-ray contrast agents and molecular imaging—a contradiction? In Handbook of Experimental Pharmacology, Springer: New York, 2008, pp. 167–175. 82. Bakan, D. A.; Weichert, J. P.; Longino, M. A.; Counsell, R. E. Polyiodinated triglyceride lipid emulsions for use as hepatoselective contrast agents in CT: effects of physicochemical properties on biodistribution and imaging profiles. Invest. Radiol. 2000, 35, 158–169. 83. Weichert, J. P.; et al. Polyiodinated triglyceride analogs as potential computed tomography imaging agents for the liver. J. Med. Chem. 1995, 38, 636–646. 84. Fiorenza, A. M.; Branchi, A.; Sommariva, D. Serum lipoprotein profile in patients with cancer. A comparison with non-cancer subjects. Int. J. Clin. Lab. Res. 2000, 30, 141–145. 85. Pussinen, P. J.; et al. The human breast carcinoma cell line HBL-100 acquires exogenous cholesterol from high-density lipoprotein via CLA-1 (CD-36 and LIMPII analogous 1)-mediated selective cholesteryl ester uptake. Biochem. J. 2000, 349, 559–566. 86. Favre, G.; et al. High density lipoprotein3 binding sites are related to DNA biosynthesis in the adenocarcinoma cell line A549. J. Lipid Res. 1993, 34, 1093–1106. 87. Cao, W. M.; et al. A mutant high-density lipoprotein receptor inhibits proliferation of human breast cancer cells. Cancer Res. 2004, 64, 1515–1521. 88. Suc, I.; Escargueil-Blanc, I.; Troly, M.; Salvayre, R.; Negre-Salvayre, A. HDL and Apo A prevent cell death of endothelial cells induced by oxidized LDL. Arterioscl. Thromb. Vascular Biol. 1997, 17, 2158–2166. 89. Lacko, A. G.; Nair, M.; Paranjape, S.; Johnso, S.; McConathy, W. J. High density lipoprotein complexes as delivery vehicles for anticancer drugs. Anticancer Res. 2002, 22, 2045–2049. 90. McConathy, W. J.; Nair, M. P.; Paranjape, S.; Mooberry, L.; Lacko, A. G. Evaluation of synthetic/reconstituted high-density lipoproteins as delivery vehicles for paclitaxel. Anti-cancer Drugs 2008, 19, 183–188. 91. Lou, B.; et al. High-density lipoprotein as a potential carrier for delivery of a lipophilic antitumoral drug into hepatoma cells. World J. Gastroenterol. 2005, 11, 954–959. 92. Frias, J. C.; Ma, Y.; Williams, K. J.; Fayad, Z. A.; Fisher, E. A. Properties of a versatile nanoparticle platform contrast agent to image and characterize atherosclerotic plaques by magnetic resonance imaging. Nano Lett. 2006, 6, 2220–2224. 93. Corbin, I.; Scott, B.; Zheng, G. Convergence of nanocrystal imaging and lipoprotein delivery. Nanomedicine 2009, 4, 7–10. 94. Carew, T. E.; Pittman, R. C.; Steinberg, D. Tissue sites of degradation of native and reductively methylated [14 C]sucrose-labeled low density lipoprotein in rats. Contribution of receptordependent and receptor-independent pathways. J. Biol. Chem. 1982, 257, 8001–8008. 95. Pittman, R. C.; Steinberg, D. Sites and mechanisms of uptake and degradation of high density and low density lipoproteins. J. Lipid Res. 1984, 25, 1577–1585. 96. de Smidt, P. C.; van Berkel, T. J. LDL-mediated drug targeting. Crit. Rev. Ther. Drug Carrier Syst. 1990, 7, 99–120. 97. Firestone, R. A. Low-density lipoprotein as a vehicle for targeting antitumor compounds to cancer cells. Bioconjug. Chem. 1994, 5, 105–113. 98. Slamon, D. J.; et al. Studies of the HER-2/neu proto-oncogene in human breast and ovarian cancer. Science (N.Y.) 1989, 244, 707–712. 99. Wong, A. J.; et al. Structural alterations of the epidermal growth factor receptor gene in human gliomas. Proc. Natl. Acad. Sci. U.S.A. 1992, 89, 2965–2969. 100. Becker, A.; et al. Receptor-targeted optical imaging of tumors with near-infrared fluorescent ligands. Nat. Biotechnol. 2001, 19, 327–331.
P1: OTA/XYZ P2: ABC c18 JWBS037-Chen February 2, 2011
462
21:51
Printer Name: Yet to Come
LIPOPROTEIN-BASED NANOPLATFORMS FOR CANCER MOLECULAR IMAGING
101. Leamon, C. P.; Low, P. S. Delivery of macromolecules into living cells: a method that exploits folate receptor endocytosis. Proc. Natl. Acad. Sci. U.S.A. 1991, 88, 5572–5576. 102. Ruoslahti, E. Specialization of tumour vasculature. Nat. Rev. Cancer 2002, 2, 83–90. 103. Chen, J.; et al. Ligand conjugated low-density lipoprotein nanoparticles for enhanced optical cancer imaging in vivo. J. Am. Chem. Soc. 2007, 129, 5798–5799. 104. Nikanjam, M.; et al. Synthetic nano-low density lipoprotein as targeted drug delivery vehicle for glioblastoma multiforme. Int. J. Pharm. 2007, 328, 86–94. 105. Nikanjam, M.; Gibbs, A. R.; Hunt, C. A.; Budinger, T. F.; Forte, T. M. Synthetic nano-LDL with paclitaxel oleate as a targeted drug delivery vehicle for glioblastoma multiforme. J. Control. Release 2007, 124, 163–171. 106. Anantharamaiah, G. M.; et al. Studies of synthetic peptide analogs of the amphipathic helix. Structure of complexes with dimyristoyl phosphatidylcholine. J. Biol. Chem. 1985, 260, 10248–10255. 107. Anantharamaiah, G. M. Synthetic peptide analogs of apolipoproteins. Methods Enzymol. 1986, 128, 627–647. 108. Tufteland, M.; Pesavento, J. B.; Bermingham, R. L.; Hoeprich, P. D., Jr.; Ryan, R. O. Peptide stabilized amphotericin B nanodisks. Peptides 2007, 28, 741–746. 109. Cormode, D. P.; et al. An Apo A-I mimetic peptide high-density-lipoprotein-based MRI contrast agent for atherosclerotic plaque composition detection. Small (Weinheim, Germany) 2008, 4, 1437–1444.
P1: OTA/XYZ P2: ABC c19 JWBS037-Chen January 25, 2011
13:19
Printer Name: Yet to Come
CHAPTER 19
Protein Cages as Multimode Imaging Agents MASAKI UCHIDA, LARS LIEPOLD, MARK YOUNG, and TREVOR DOUGLAS Department of Chemistry and Biochemistry and Department of Plant Sciences, Center for Bio-Inspired Nanomaterials, Montana State University, Bozeman, Montana, USA
19.1 INTRODUCTION Protein cage architectures such as those of small virus capsids and ferritins are selfassembled from a limited and defined number of subunit building blocks [1, 2]. These protein cage architectures typically range in size from tens to hundreds of nanometers and are spherical or rod shaped. By combining both chemical and genetic modifications of the subunits, one can impart novel functions to these protein cage architectures that are quite different from their native function in biology. There are three distinct surfaces of the assembled protein cage architecture that can be manipulated to impart function by design [1, 2]: these are the external surface, the interior surface, and the surface that forms the interface between the subunits. Manipulation of the interactions at the subunit interface allows control over disassembly and reassembly [3, 4]. The interior interface can be used for chemical attachment of small molecules and for encapsulation of nanoparticles, sequestered inside the cage, useful for imaging and therapeutics. The exterior surface can also be modified by small molecule attachment or by presentation of cell-specific targeting ligands (peptides, antibodies, carbohydrates) [5–10]. Considerable effort has gone into the development of protein cage architectures as robust systems that combine advanced imaging capacity with highly selective cell targeting within the same protein cage platform [2, 11]. A library of protein cages, which support a range of synthetic chemistries and genetic manipulations useful for creating high performance multimode imaging agents, have been employed (Fig. 19.1) [12, 13]. These include the 28-nm capsid of Cowpea chlorotic mottle virus (CCMV) [14], the 30-nm capsid of Cowpea mosaic virus (CPMV) [15], the bacteriophages MS2 (27 nm) [16] and Q (30 nm) [17], the 12-nm small heat shock protein from Methanococcus jannaschii [18], and the 12-nm human ferritin cage [19]. Many closed shell protein cage architectures define an interior space that can be used to selectively direct synthesis or entrapment of a guest cargo. A significant advantage to the use of protein cages is the ability to produce monodisperse nanoparticles with a precisely Nanoplatform-Based Molecular Imaging Edited by Xiaoyuan Chen C 2011 John Wiley & Sons, Inc. Copyright
463
P1: OTA/XYZ P2: ABC c19 JWBS037-Chen January 25, 2011
464
13:19
Printer Name: Yet to Come
PROTEIN CAGES AS MULTIMODE IMAGING AGENTS
FIGURE 19.1 Protein cages including the viral capsids CCMV (28-nm diameter), CPMV (30-nm), MS2 (27-nm), and Q (30-nm), the heat shock protein Hsp (12-nm), and a ferritin Fn (12-nm). The images of CCMV, CPMV, MS2, and Q were reproduced from VIPER (http://viperdb.scripps.edu/) at the Scripps Research Institute. Those of Hsp and Fn were reproduced using the UCSF Chimera package (http://www.cgl.ucsf.edu/chimera/) from the Resource for Biocomputing, Visualization, and Informatics at the University of California (supported by NIH P41 RR-01081).
defined size and shape and since the properties of nanophase materials are intimately related to their dimensions, any heterogeneity in size is reflected as heterogeneity in their physical properties. The controlled formation of nanomaterials within protein cage architectures thus has broad applications in a variety of emerging technologies including drug delivery and medical diagnostics. Using approaches that mimic the structure–function relationships in virus particles, that is, selective encapsulation, homogeneous size, and targeted cell delivery, creates the potential for protein cage nanoparticles for imaging and therapeutic applications [20]. Key to the success of this approach is an ability to produce and to modify protein cages to impart novel functionality. The highly symmetrical, spatially controlled presentation of functional groups has allowed the attachment of organic species, such as targeting peptides [6, 8], fluorescent labels [11, 21, 22], metal chelating groups [23, 24], and inorganic species such as magnetic nanomaterials [25, 26] and gadolinium ions [27–31], specifically to either the interior or exterior of the protein cages [32]. These agents represent a wide range of potential imaging modalities and, when coupled with targeting capability on the outside of the cages, provide a powerful platform for medial imaging.
19.2 PROTEIN CAGE BASED MAGNETIC MATERIALS AS MRI CONTRAST AGENTS The high resolution of magnetic resonance imaging (MRI) affords excellent anatomical detail; however, the ability to detect low copy number tissues is limited by the technique’s low sensitivity. Targeted contrast agents increase the sensitivity of MR to tissues of interest
P1: OTA/XYZ P2: ABC c19 JWBS037-Chen January 25, 2011
13:19
Printer Name: Yet to Come
PROTEIN CAGES AS T1 AGENTS
465
by accelerating the NMR relaxation rate of water protons in the vicinity of these tissues. Two relaxation processes, either T1 relaxation (spin–lattice or longitudinal) or T2* relaxation (spin–spin or transverse), are probed to produce MR images. The bulk of the image is produced from the nuclear spins of protons that are in water or fatty molecules. Contrast agents are materials that catalyze this realignment of nuclear spins and in doing so they increase the relaxation rate and therefore increase the contrast of and image. This effect is called paramagnetic relaxation enhancement (PRE) and a common class of contrast agents contain gadolinium (Gd) ions or Fe3 O4 nanoparticles as catalysts for this process.
19.3 PROTEIN CAGES AS T1 AGENTS Gd agents typically are used to enhance the T1 relaxation process and “T1” specifically refers to a time in units of seconds equal to a realignment decay constant for the proton’s spin in a given sample. The T1 of water, for example, is approximately 3 seconds. The PRE of a contrast agent is given in units of mM-1 ·seconds-1 (mM-1 ·s-1 ) and is referred to as relaxivity. The Gd concentration (mM) is used for the calculation of ionic relaxivity while the concentration of a particle (mM) containing one or more Gd ions is used to calculate the particle relaxivity. For example, a 1 mM aqueous solution of Magnevist, a clinically used contrast agent, has a relaxivity of r1 ≈ 5 mM-1 ·s-1 with a T1 of 0.19 second based on Eq. (19.1) [33]. T1observed =
1 T1sample
+ r1 CA ConcentrationCA
−1 (19.1)
There are a few parameters that must be optimized with regard to development of Gdbased contrast agents. The first consideration is that Gd is toxic and must therefore be used in the form of a tightly associated metal ion chelate. Since the Gd-based relaxivity is a spin–lattice effect, direct water interaction with the Gd ion is necessary through an inner sphere mechanism. The number of metal-bound water molecules that can occupy a coordination sphere of Gd is referred to as “q” and, in general, as q increases the stability of the chelator–metal interaction decreases. Therefore a balance must be struck between increasing the number of water coordination sites and the overall stability of the metal–chelator interaction. Another parameter that must be optimized is the residence time of the Gd-bound water, referred to as M . It is important for the water to remain bound to Gd long enough for the proton spins to relax. However, when M is too high, and the water molecule remains bound to Gd long after it has been relaxed, the result is an effective decrease in the “duty cycle” of the Gd ion resulting in a decrease of the overall relaxivity. The M value can be modulated by varying the chemical nature of the metal chelator. The final parameter that is important for efficient PRE is the rotational correlation time of the Gd ion ( R ). Larger Gd complexes result in greater R values, which yields more efficient PRE. Figure 19.2 shows the parameters that effect PRE. Protein cages are well suited as vehicles for Gd contrast agents since their large size provides a large exterior and interior surface area as well as an interior volume, all of which can be functionalized to carry high densities of Gd. Also, the large size of these particles
P1: OTA/XYZ P2: ABC c19 JWBS037-Chen January 25, 2011
466
13:19
Printer Name: Yet to Come
PROTEIN CAGES AS MULTIMODE IMAGING AGENTS
q
τM τR
FIGURE 19.2 A schematic of DTPA-Gd and the important physical parameters that affect the relaxivity, including the number of water molecules that directly interact with the Gd ion (q), the exchange lifetime of the metal-bound water molecule ( M ), and the rotational correlation time of the Gd chelate ( R ).
results in ideal rotational properties ( R ) and therefore theoretical maximum relaxivity values can be realized in these systems. To date, several groups have reported the utilization of viral capsids or other protein cage structures as MRI contrast agents. In the first report of this kind, Allen and co-workers utilized the endogenous metal binding site in the plant virus CCMV and bound Gd3+ ions to endogenous Ca-binding sites in the capsid [27]. Since the metal ion in this capsid is situated at the interface of three subunits, it is likely that the R for Gd is similar to R of the entire cage, therefore this Gd–protein cage construct should exhibit near ideal R values. This construct has a relatively weak affinity for the Gd ion (K d = 31 M), which along with analysis of the crystal structure of the metal binding pocket both suggest that q is likely to be greater than one. Recently performed fitting of the CCMV construct’s r1 NMRD profile to the Solomon–Bloembergen–Morgan (SBM) analytical model for relaxivity returned values of q = 4, M = 3 × 10-9 second, and R > 10-7 second, all of which are highly desirable. Not surprisingly, the optimal relaxivity properties of this construct resulted in extremely high ionic relaxivities (r1 = 202 mM-1 ·s-1 @ 1.5 T) [27]. Unfortunately, the low Gd affinity of this construct makes the clinical applicability impractical. Several groups have developed a range of next generation constructs, which can now be divided into three categories: (1) endogenous metal binding sites, [27], (2) genetic insertion of a metal binding peptide [30], and (3) chemical attachment of small molecule chelates [28–32]. Of these three categories the first two are less significant from a clinical standpoint due to their low Gd binding affinities. The third category is the most common approach with three groups reporting similar results. Figure 19.3 outlines the single step coupling of chelated Gd ions to the CCMV protein cage. Anderson et al. [28] attached DTPA-Gd to lysine residues on the surface of the MS2 capsid (MS2 DTPA ITC). Despite the relatively low ionic relaxivity of this construct, the high density of Gd labeling resulted in the highest particle relaxivity for constructs with sufficient Gd binding stability and is approximately three orders of magnitude higher than the small molecule contrast agent. Finn and co-workers produced a similar construct by using click chemistry to attach DOTA-Gd to CPMV and Q-beta and this scheme also produced a construct with a relatively low ionic relaxivity (r1 = 16 mM-1 ·s-1 @1.5 T); however, high levels of Gd ions were attached [31]. A CCMV-DOTAGd construct and a MS2 Bis (HOPO) TAM construct both have higher ionic relaxivities
P1: OTA/XYZ P2: ABC c19 JWBS037-Chen January 25, 2011
13:19
Printer Name: Yet to Come
PROTEIN CAGES AS T1 AGENTS
467
O N O
NH2
O O
O
N
N
N
N
+
N
O-
O
-O
O
O
O
O
N
N
N
N
N
O N
O
-O
+ GdCI3
O
O-
-O
O
O-
O-
O
N
Gd3+ N
N O
O-
FIGURE 19.3 The reaction scheme to attach DOTA-Gd to the CCMV viral capsid. Endogenous lysines on the viral capsid are reacted to a DOTA/NHS conjugation. Next, GdCl3 is added to produce a viral capsid conjugated with Gd3+ ions.
(r1 = 46 mM-1 ·s-1 @1.5 T; r1 = 31 mM-1 ·s-1 @1.5 T, respectively) but lower density of Gd labeling resulted in lower particle relaxivities [29]. To optimize the relaxivity properties of Gd-based contrast agents, it is helpful to have quantitative information of the parameters important to PRE. Since R , M , and q have different dependencies on the external magnetic field, it is possible to extract information about these values from a plot of relaxivities obtained for one contrast agent taken over a range of field strengths. These plots are referred to as nuclear magnetic resonance dispersion (NMRD) plots and they can be fit to the SBM analytical model for PRE and values for R , M , and q can be obtained [34]. Francis and co-workers labeled the MS2 viral capsids on the interior surface and on the exterior surface with a novel Gd chelator (Fig. 19.4). A fit of their NMRD data with a modified SBM model suggests that the M for this construct is near optimal and q = 2 [32]. However, local mobility in the chelator attachment scheme partially negated the ideal rotational properties of this large construct. Furthermore, these authors reported a relaxivity difference between the interior and exterior labeled versions of the MS2 construct that was attributed to a difference in the rigidity of the these two anchoring schemes. The R values
P1: OTA/XYZ P2: ABC c19 JWBS037-Chen January 25, 2011
468
13:19
Printer Name: Yet to Come
PROTEIN CAGES AS MULTIMODE IMAGING AGENTS
FIGURE 19.4 The MS2 viral capsid labeled externally (4) and internally (5) with a Gd chelate. The crystal structure of the MS2 virus capsid is shown, highlighting the amino groups on the exterior: red, K106; yellow, K113; and orange, the N terminus. Y85 is highlighted in green on the interior surface. (Copyright © 2008 American Chemical Society.)
were higher for the interior labeled construct, which has higher relaxivity values, suggesting that local mobility can dampen the ideal rotational effects of protein cages. Recently, a strategy has been developed for synthesizing a branched structure, via azide alkyne cycloaddition reactions, that projects into the interior cavity of the heat shock protein (Hsp) cage from Methanococcus jannaschii (Fig. 19.5) [35]. With this synthetic strategy a high Gd payload was incorporated in the interior cavity of a protein cage by anchoring a Gd chelator to this interiorly grown branched polymer (Fig. 19.5). In doing so, very high Gd densities in the protein cage and reasonably high ionic relaxivity values were achieved. Therefore the resulting particle has exceptional particle relaxivity properties. The most promising aspect of coupling chelated Gd to the polymer is the development of a method to take advantage of the interior volume of the protein cage. To date, only the
FIGURE 19.5 Shown here is the synthesis of the branch polymer in the interior of the Hsp protein cage via the alkyne–azide click coupling reaction. Amines on the polymer are functionalized with DTPA-Gd after polymer growth.
P1: OTA/XYZ P2: ABC c19 JWBS037-Chen January 25, 2011
13:19
Printer Name: Yet to Come
PROTEIN CAGES AS T2 MRI CONTRAST AGENTS
469
surface of protein cages including viral capsids have been modified to carry Gd chelators by single step chemical labeling. Here we expand the chemical modification to fill the interior volume with a DTPA-Gd labeled polymer network. This results in a particle that has a per particle relaxivity of 4200 mM-1 ·s-1 , which, while lower than the MS2-DTPA-Gd value (7200 mM-1 ·s-1 ), exhibits relaxivity properties that are clinically significant. The MS2 construct is significantly larger than the Hsp, occupying 12 times the volume. When the relaxivity behaviors of these materials are compared on a volume basis, the Hsp construct exhibits a relaxivity that is seven times that of the MS2 construct. Until now the relaxivity properties of wtCCMV-Gd seemed unattainable in systems with high-affinity chelators. This has been achieved by filling the interior cavity of the Hsp with DTPA-Gd and the per volume relaxivity or the “relaxivity density” values are equal to that of the wtCCMV-Gd construct. The advantages of a particle with high relaxivity densities become apparent when one considers the relaxivity properties required to image a cell. A previously reported equation returns the amount of contrast agent required to image a mammalian cell as a function of the contrast agent’s particle relaxivity [36]. We used this equation along with values for the surface area of the contrast agent and the surface area of a mammalian cell to calculate the percentage of the cell that would be covered with the contrast agent for successful imaging, assuming that the contrast agent would bind irreversibly and would not be taken into the cell. These calculations resulted in 11% of the mammalian cell surface covered with Hsp-BP-(DTPA-Gd) for successful imaging, while the MS2 construct would cover 29% of the surface and Magnevist would cover 72%. While the details of cellular imaging are likely to be more complex than this simple calculation assumes, the importance of a contrast agent with high relaxivity densities is apparent. There are significant benefits—namely, increased stability and efficient particle relaxivity properties—that result from a synthetic approach that takes full advantage of interior volume of protein cages. This strategy can be extended to larger protein cages and it will be interesting to see if the ionic relaxivity remains constant in larger systems. Also, developing similar protein cage–polymer hybrid constructs with Gd chelators that have more desirable q and M values should result in particles with very attractive relaxivity properties.
19.4 PROTEIN CAGES AS T2 MRI CONTRAST AGENTS The T2 relaxation is a time constant in units of seconds that describes the length of time for the precession of water proton to lose coherence and become randomized after an excitation pulse. This is dependent on the local inhomogeneities in the magnetic field, which can be controlled by the presence of magnetic nanoparticles. By using a T2-weighted pulse sequence, the MR scanner is biased to detect the T2 relaxation process, an extremely useful clinical option for some diagnostic applications. Thus T2 contrast agents are used to enhance contrast in diagnostic imaging and have been manipulated to yield maximal efficiency of T2 relaxivity. Among the variety of protein cage architectures, ferritin has been the most intensively investigated for its application as a T2 MRI contrast agent. Ferritin’s inherent biological function is the sequestration and storage of iron as an amorphous iron oxide nanoparticle of ferrihydrite. Ferritins are assembled from 24 subunits and exhibit cage-like structure 12 nm in diameter with an interior cavity of 8 nm where ferrihydrite is encapsulated [19]. This weakly superparamagnetic ferrihydrite [37] functions as an endogenous contrast agent that
P1: OTA/XYZ P2: ABC c19 JWBS037-Chen January 25, 2011
470
13:19
Printer Name: Yet to Come
PROTEIN CAGES AS MULTIMODE IMAGING AGENTS
significantly shortens the T2 of surrounding water protons. Since iron is well known to play an important role in a number of diseases, endogenous ferritin in humans has been considered as a useful marker for MRI examination to assess the amount of iron in liver, spleen, and brain [38–40]. In addition, ferritin has recently been proposed as a reporter protein to monitor transgene expression using MRI [41, 42]. These studies suggest ferritin could be used like green fluorescence protein and luciferase for in vivo optical imaging of gene expression. However, here we will focus on development of synthetic ferritin–iron oxide nanocomposites for MRI contrast agents rather than exploitation of endogenous ferritin. One disadvantage of endogenous ferritin as an MRI contrast agent is the fact that it exhibits 10–100-fold lower relaxivity per iron than commercially available synthetic iron oxide MRI contrast agents. In order to overcome this drawback, demineralized ferritin (apoferritin) was developed as a size-constrained reaction environment for the synthesis of superparamagnetic iron oxide nanoparticles (magnetoferritin) which exhibit much higher R2 compared to endogenous ferritin [43, 44]. Recently, recombinant ferritin cages have been developed, which have demonstrated advantages for the encapsulation of magnetite or maghemite (Fig. 19.6) [45, 46]. Native mammalian ferritins are composed of a mixture of two different types of subunit, light (L) chain and heavy (H) chain [37]. A catalytic ferroxidase site, which catalyzes the oxidation of Fe(II) to Fe(III) exists in the H chain, but is absent in the L chain. Oxidation of Fe(II) to Fe(III) might be a critical step in the formation of iron oxide because of the low solubility of Fe(III) [19], even in synthetic systems. Using recombinant human H chain ferritin cages, it has been shown that Fe3 O4 cores formed close to the theoretical core diameter, calculated from input Fe loading with narrow size distribution. The average size of the mineralized Fe3 O4 particles in HFn increases from 3.6 to 5.9 nm with increasing Fe loading from 1000 Fe to 5000 Fe per cage. This means that both the Fe3 O4 core of the mineralized HFn and the exterior diameter are comparable to ultrasmall
FIGURE 19.6 (a) Illustration of synthesis of magnetite (Fe3 O4 ) nanoparticle in HFn. (b) TEM image of mineralized particles within HFn with 5000 Fe/cage (right) and size distribution of the particles analyzed from the TEM image (left).
P1: OTA/XYZ P2: ABC c19 JWBS037-Chen January 25, 2011
13:19
Printer Name: Yet to Come
PROTEIN CAGES AS CONTRASTAGENTS
471
TABLE 19.1 Comparison of Size and r1 and r2 Relaxivities of the Mineralized HFn, SPIO, and USPIO HFn1000Fe
HFn3000Fe
HFn5000Fe
Ferumoxides
Ferumoxtran-10
Exterior (nm) Interior particle (nm)
12 3.6 ± 0.7
12 5.1 ± 0.9
12 5.8 ± 0.9
58.5 ± 185.8 15.2 ± 8.9
29.5 ± 23.1 4.9 ± 1.5
r1 (mM·s)−1 r2 (mM·s)−1
2.3 11
4.2 31
8.4 93
7.8 119
12 101
superparamagnetic iron oxide (USPIO) contrast agents such as Ferumoxtran-10 rather than small superparamagnetic iron oxide (SPIO) such as Ferumoxides [47] (Table 19.1). As expected, the r1 and r2 relaxivities of the mineralized HFn increase with increasing Fe per cage (i.e., the particle size) and the r1 and r2 of the HFn loaded with 5000 Fe were comparable to those of commercially available SPIO and USPIO materials [46] (see Table 19.1).
19.5 PROTEIN CAGES AS CONTRAST AGENTS FOR IMAGING ATHEROSCLEROTIC PLAQUES One of the unique features of the mineralized HFn with regard to MRI contrast is that it is readily taken up by macrophage cells and therefore provides a strong T2* effect [46]. Since macrophage cells play a significant role in the progression of inflammatory responses, imaging of macrophages involved in inflammatory diseases such as atherosclerosis is useful to assess and diagnose these diseases [48, 49]. SPIO particles are taken up by macrophages more readily than the mineralized HFn in vitro, likely due to its larger particle size. However, in vivo the SPIO particles are trapped by Kupffer cells in liver tissue during the first pass after injection in vivo [50] and therefore are less useful for imaging macrophages outside the liver. In contrast, USPIO particles have a longer blood circulation time and hence are able to label atherosclerotic plaques in vivo [51, 52], but their uptake efficiency into macrophages is low. Recently, it has been shown that mineralized HFn particles are taken up by macrophages much more efficiently than USPIO in vitro, even though the overall size of the mineralized HFn is similar to that of the USPIO (Fig. 19.7a). These findings hold the promise that mineralized ferritin can be developed as an MRI contrast agent to monitor inflammatory events induced by macrophages such as atherosclerotic plaque progression (Fig. 19.7b). Indeed, fluorescence imaging studies have revealed that fluorescently labeled HFn accumulates significantly more in atherosclerotic plaque lesions formed in the carotid arteries of mice than in normal carotid arteries and colocalizes with macrophage cells in these lesions [53] (Fig. 19.8). Furthermore, in vivo MRI assessment of the atherosclerotic mouse model confirmed reduction of carotid artery lumen caliber size on a T2* enhanced image taken 24 h and 48 h after injection of the mineralized HFn only when the atherosclerotic plaque was induced within the artery [53]. This reduction is likely due to a T2* enhanced effect caused by the mineralized HFn accumulated in the plaque lesion. These results indicate that the mineralized HFn is a promising MRI contrast agent to monitor inflammation in atherosclerosis.
P1: OTA/XYZ P2: ABC c19 JWBS037-Chen January 25, 2011
472
13:19
Printer Name: Yet to Come
PROTEIN CAGES AS MULTIMODE IMAGING AGENTS
FIGURE 19.7 (a) Fe content taken up by macrophage cells after incubation with 165 g/mL of various contrast agents. (b) In vitro MR images of macrophage cells incubated with 165 g/mL of various contrast agents for 24 h. (Copyright © 2008 Wiley-Liss, Inc.)
Another unique feature of protein cage architectures is that cell- and/or tissue-specific targeting moieties can be functionalized on their exterior surface using either chemical or genetic approaches. For example, the amino acid sequence RGD-4C (CDCRGDCFC), known to bind integrins ␣v 3 and ␣v 5 [54], has been genetically introduced as an N-terminal fusion protein into HFn [45]. The HFn mutant (RGD4C-Fn) exhibited a cagelike structure indistinguishable from the wild-type HFn by either size or morphology. In addition, there was no significant difference in the average diameter and size distribution of the mineralized Fe3 O4 formed within the RGD4C-Fn and HFn under the same Fe loading conditions. These results indicate that the introduction of the RGD-4C targeting
FIGURE 19.8 In situ and ex vivo fluorescence images of HFn-Cy5.5 nanoparticles in mouse carotid arteries. The fluorescently labeled HFn is accumulated in atherosclerotic plaques formed in a ligated carotid artery.
P1: OTA/XYZ P2: ABC c19 JWBS037-Chen January 25, 2011
13:19
Printer Name: Yet to Come
PROTEIN CAGES AS VEHICLES FOR TARGETED DELIVERY OF IMAGING AGENTS
473
FIGURE 19.9 FACS analysis of THP-1 cells incubated with fluorescence-labeled protein cages. The data are plotted as histograms with their corresponding geometric (geo) mean fluorescence values. The increased level of fluorescence intensity of the cells incubated with RGD4C-Fn indicates specific binding of the cages to THP-1 cells. (Copyright © 2009 Wiley-VCH Verlag GmbH & Co.)
peptide has little effect on either the protein cage assembly or the mineralization behavior of this protein cage mutant. Analysis using fluorescence-activated cell sorting demonstrated that the RGD4C-Fn clearly exhibited enhanced targeting interaction with C32 amelanotic melanoma cells and THP-1 monocyte cells, which are known to overexpress integrins ␣v 3 [55] (Fig. 19.9). Furthermore, TEM observation of macrophage cells incubated with the Fe3 O4 mineralized HFn or RDG4C-Fn revealed that the RDG4C-Fn particles both bind to the cell surface and were also internalized into the cells more efficiently than the nontargeted HFn controls [55]. These results suggest that introduction of cell and/or tissue targeting moieties onto the exterior of a protein cage architecture shows promise for enhancing and directing the MR imaging capability.
19.6 PROTEIN CAGES AS VEHICLES FOR TARGETED DELIVERY OF IMAGING AGENTS TO BIOFILM INFECTIONS Another promising application of protein cage-based imaging agents is the diagnosis and imaging of bacterial infection in biofilms. Chronic infections associated with device implantation, also known as biomaterial centered infections, are rare but the consequences are dire for the patient. It is now well accepted that biofilms, which are structured communities of microorganisms that develop on tissue or material surfaces, are the etiological agents of biomaterial centered infections. In terms of our current understanding, the most cost-effective approach for treatment of biofilm infections would be early detection followed by early treatment with an appropriate antimicrobial. In principle, there are currently a variety of noninvasive imaging techniques that could be used to diagnose biomaterial centered infections. Among various types of imaging techniques, MRI has some advantages for imaging biofilms. The quality of MR images of soft tissues is superior to those produced by any other technique. In addition, it does not require exposure to potentially harmful radiation or administration of radioisotopes. However, MRI is a relatively insensitive technique compared to radionuclide imaging. In addition, the expected size of biofilms
P1: OTA/XYZ P2: ABC c19 JWBS037-Chen January 25, 2011
474
13:19
Printer Name: Yet to Come
PROTEIN CAGES AS MULTIMODE IMAGING AGENTS
(a)
(b)
(c)
(d)
FIGURE 19.10 Targeting of CCMV to a pathogenic bacterium: (a) targeting scheme—SpA, protein A; Ab, monoclonal antibody against protein A; B, biotin; StAv, streptavidin; IA, imaging agent; (b,c) SEM micrographs showing the high density of targeting; and (d) SEM micrograph of cell exposed to CCMV without Ab targeting.
colonizing implants is small (less than 1 mm in dimension). Therefore developing the capability to image biofilms in vivo by MRI is nontrivial and it will be necessary to develop new methods for targeted delivery of imaging agents to achieve early detection of biofilms by MRI. Protein cage architectures are an ideal platform to accomplish loading of designed targeting moieties, imaging agents, and possibly therapeutic agents simultaneously. Cowpea chlorotic mottle virus (CCMV) has been used as a delivery vehicle and Staphylococcus aureus (S. aureus) as a model organism for some initial studies. Staphylococcus aureus is a primary culprit in prosthetic joint and graft infections [56, 57] and also tissue infections such as endocarditis, osteomyelitis, and native heart valve infections that conform to the biofilm model. In general, bacterial pathogens express cell surface proteins that are distinct from those of the host tissues and can be used for targeting. For example, S. aureus expresses at least 22 different cell wall anchored proteins, many of which are involved in conferring virulence and are likely to be expressed at a high level in infections [58, 59]. Among these is protein A (SpA), which binds to the Fc region of IgG, and thereby may serve to conceal the cells from immune detection. We designed the CCMV capsid to target protein A displayed on S. aureus using streptavidin (StAv) to mediate the interaction between biotinylated monoclonal antibody (mAb) against protein A and CCMV (Fig. 19.10). This provided a clear experimental advantage since mAb against protein A is commercially available. By using the indirect targeting scheme outlined in Figure 19.10a, a relatively high density of CCMV binding to the cell surface of S. aureus could be achieved (Fig. 19.10b,c)
P1: OTA/XYZ P2: ABC c19 JWBS037-Chen January 25, 2011
13:19
Printer Name: Yet to Come
CONCLUSION
475
[60]. Negative controls (e.g., Fig. 19.10d) indicated binding was specifically mediated by the mAb–SpA interaction. The specificity of the binding was corroborated by the results of epifluorescence microscopy, flow cytometry, and TEM [60]. An additional issue to be considered is the penetration of the protein cage into the biofilm since biofilm cells are enclosed in a matrix of extracellular polymeric substances (EPSs) that can impede transport of nanoplatforms. In biofilms of S. aureus treated with CCMV that was dual labeled with biotin and a fluorescent tag, it was shown that the functionalized CCMV penetrated into the biofilm microcolonies to a depth of about 20 m during a 90-min exposure period [60]. Thus cells in relatively thin biofilms (<10 m) such as those expected to be involved in biomaterial centered infections [61–63] can be fully penetrated during an exposure period that is feasible to achieve clinically. More recently, we have successfully prepared a “Janus-like” protein cage nanoparticle using the DNA-binding protein from Listeria innocua (LiDps) [64]. The cage is dual labeled with biotin and a fluorescent tag similar to CCMV, but possesses two spatially distinct faces, one labeled only with biotin and the other with a fluorescent tag. These Janus-like LiDps cages can be used as a “plug and play” nanoplatform for antibody directed cell targeting with the potential for control over the orientation of the targeted nanoplatform [65]. In addition, the protein cages, such as CCMV, can be conjugated with DOTA, a clinically relevant Gd chelate, to load CCMV with the Gd-based MRI contrast agent [60]. Using this approach, the mean number of Gd ions bound to this biotinylated construct was estimated at 166 per cage. Because of the selected targeting of the CCMV cage to S. aureus discussed above, roughly 1.8 × 105 Gd atoms per cell were delivered. A typical S. aureus biofilm is quite dense (approximately 3 × 1010 cells/cm3 ) with a distance between cells of less than 3 m [66]. At this cell density the level of Gd loading onto S. aureus cells would result in about 10 mM Gd chelate. It has been estimated that about 50 mM Gd chelate is required for discrimination of tissues targeted with MRI contrast agent [67]. This estimate is based on a Gd chelate with an ionic relaxivity of about 5 mM-1 ·s-1 . The concentration of Gd chelate required for discrimination of tissues targeted with contrast agent is inversely proportional to R1 . As discussed above, a large effort to increase relaxivity of targeted contrast agents is ongoing and we, and others, have demonstrated that ionic relaxivities can be increased by more than an order of magnitude by coupling Gd to a protein cage [27]. In addition, previous levels of labeling with fluorescence imaging agents suggests that Gd loading onto CCMV can be increased substantially [68], effectively producing large particle relaxivities and increasing the localized Gd concentration above 50 mM. Even more promising in this respect are protein cages in which the Gd is incorporated into a crosslinked polymer that fills the protein cage interior (Fig. 19.5). The possibility of high density targeting combined with the high particle relaxivities of Gd coupled to protein cages suggests that protein cages have great potential as vehicles for targeted delivery of Gd to pathogenic microbes for MR imaging of biofilm infections.
19.7 CONCLUSION In this chapter, we outlined how protein cages can be used as platforms for MRI contrast agents and fluorescent imaging agents. Protein cages have a number of advantages as templates for development of nanoplatforms for targeted delivery of both T1 and T2 contrast agents, fluorescent agents, and potentially other imaging modalities as well. One salient
P1: OTA/XYZ P2: ABC c19 JWBS037-Chen January 25, 2011
476
13:19
Printer Name: Yet to Come
PROTEIN CAGES AS MULTIMODE IMAGING AGENTS
feature in this respect is their unique versatility, which allows many different design strategies for incorporation of both cell-specific targeting ligands and imaging agents (or therapeutic agents) to be implemented. In addition, monodisperse preparations of nanoparticles can be produced with relative ease. We have introduced examples of targeted delivery of protein cages to both atherosclerotic plaques and microbial biofilms. Coupling of T1 contrast agents to nanoplatforms can increase their relaxivity substantially and protein cages can be loaded with contrast agents both on the exterior and interior surfaces and in the interior cavity. T2 contrast agents such as iron oxides can also be loaded into the interior cavity with control over both the size and phase. The synthetic manipulation of the cages to allow incorporation of metal chelating agents, and other small organics, opens up the promise for using these multivalent platforms with other imaging modalities such as PET and CT. The development of protein cages as a biomaterial has progressed only for a decade. There are a wide variety of proteins that assemble into cage-like structures. We anticipate that the chemical and biophysical diversity of the family of protein cages will lead to an expanding range of biomedical applications.
REFERENCES 1. Douglas, T.; Young, M. Viruses: Making friends with old foes. Science 2006, 312(5775), 873– 875. 2. Uchida, M.; Klem, M. T.; Allen, M.; Flenniken, M. L.; Gillitzer, E.; Varpness, Z.; Suci, P.; Young, M. J.; Douglas, T. Protein cage architecture: containers as templates for materials synthesis. Adv. Mater. 2007, 19, 1025–1042. 3. Aime, S.; Frullano, L.; Crich, S. G. Compartmentalization of a gadolinium complex in the apoferritin cavity: a route to obtain high relaxivity contrast agents for magnetic resonance imaging. Angew. Chem. Int. Ed. 2002, 41(6), 1017. 4. Sun, J.; DuFort, C.; Daniel, M. C.; Murali, A.; Chen, C.; Gopinath, K.; Stein, B.; De, M.; Rotello, V. M.; Holzenburg, A.; Kao, C. C.; Dragnea, B. Core-controlled polymorphism in virus-like particles. Proc. Nat. Acad. Sci. U.S.A 2007, 104(4), 1354–1359. 5. Wang, Q.; Raja, K. S.; Janda, K. D.; Lin, T. W.; Finn, M. G. Blue fluorescent antibodies as reporters of steric accessibility in virus conjugates. Bioconjug. Chem. 2003, 14(1), 38–43. 6. Chatterji, A.; Burns, L. L.; Taylor, S. S.; Lomonossoff, G. P.; Johnson, J. E.; Lin, T.; Porta, C. Cowpea mosaic virus: from the presentation of antigenic peptides to the display of active biomaterials. Intervirology 2002, 45, (4-6), 362–370. 7. Taylor, K. M.; Lin, T. W.; Porta, C.; Mosser, A. G.; Giesing, H. A.; Lomonossoff, G. P.; Johnson, J. E. Influence of three-dimensional structure on the immunogenicity of a peptide expressed on the surface of a plant virus. J. Mol. Recognition 2000, 13(2), 71–82. 8. Flenniken, M. L.; Willits, D. A.; Harmsen, A. L.; Liepold, L. O.; Harmsen, A. G.; Young, M. J.; Douglas, T. Melanoma and lymphocyte cell-specific targeting incorporated into a heat shock protein cage architecture. Chem. Biol. 2006, 13(2), 161–170. 9. Miermont, A.; Barnhill, H.; Strable, E.; Lu, X.; Wall, K. A.; Wang, Q.; Finn, M. G.; Huang, X. Cowpea mosaic virus capsid: a promising carrier for the development of carbohydrate based antitumor vaccines. Chemistry 2008, 14(16), 4939–4947. 10. Raja, K. S.; Wang, Q.; Finn, M. G. Icosahedral virus particles as polyvalent carbohydrate display platforms. Chembiochem 2003, 4(12), 1348–1351. 11. Lewis, J. D.; Destito, G.; Zijlstra, A.; Gonzalez, M. J.; Quigley, J. P.; Manchester, M.; Stuhlmann, H. Viral nanoparticles as tools for intravital vascular imaging. Nat. Med. 2006, 12(3), 354–360.
P1: OTA/XYZ P2: ABC c19 JWBS037-Chen January 25, 2011
13:19
Printer Name: Yet to Come
REFERENCES
477
12. Carrillo-Tripp, M.; Shepherd, C. M.; Borelli, I. A.; Venkataraman, S.; Lander, G.; Natarajan, P.; Johnson, J. E.; Brooks, C. L., 3rd; Reddy, V. S. VIPERdb2: an enhanced and web API enabled relational database for structural virology. Nucleic Acids Res. 2009, 37 (Database issue), D436–D442. 13. Pettersen, E. F.; Goddard, T. D.; Huang, C. C.; Couch, G. S.; Greenblatt, D. M.; Meng, E. C.; Ferrin, T. E. UCSF Chimera—a visualization system for exploratory research and analysis. J. Comput. Chem. 2004, 25(13), 1605–1612. 14. Speir, J. A.; Munshi, S.; Wang, G. J.; Baker, T. S.; Johnson, J. E. Structures of the native and swollen forms of cowpea chlorotic mottle virus determined by X-ray crystallography and cryoelectron microscopy. Structure 1995, 3(1), 63–78. 15. Lin, T. W.; Chen, Z. G.; Usha, R.; Stauffacher, C. V.; Dai, J. B.; Schmidt, T.; Johnson, J. E. The refined crystal structure of Cowpea mosaic virus at 2.8 angstrom resolution. Virology 1999, 265(1), 20–34. 16. Valegard, K.; Liljas, L.; Fridborg, K.; Unge, T. The 3-dimensional structure of the bacterial-virus Ms2. Nature 1990, 345(6270), 36–41. 17. Golmohammadi, R.; Fridborg, K.; Bundule, M.; Valegard, K.; Liljas, L. The crystal structure of bacteriophage Q beta at 3.5 angstrom resolution. Structure 1996, 4(5), 543–554. 18. Kim, K. K.; Kim, R.; Kim, S. H. Crystal structure of a small heat-shock protein. Nature 1998, 394(6693), 595–599. 19. Lawson, D. M.; Artymiuk, P. J.; Yewdall, S. J.; Smith, J. M. A.; Livingstone, J. C.; Treffry, A.; Luzzago, A.; Levi, S.; Arosio, P.; Cesareni, G.; Thomas, C. D.; Shaw, W. V.; Harrison, P. M. Solving the structure of human H-ferritin by genetically engineering intermolecular crystal contacts. Nature 1991, 349(6309), 541–544. 20. Manchester, M.; Steinmetz, N. F. Viruses and Nanotechnology; Springer-Verlag: Berlin, 2009; Vol. 327. 21. Wang, Q.; Lin, T. W.; Tang, L.; Johnson, J. E.; Finn, M. G. Icosahedral virus particles as addressable nanoscale building blocks. Angew. Chem. Int. Ed. 2002, 41(3), 459–462. 22. Flenniken, M. L.; Willits, D. A.; Brumfield, S.; Young, M.; Douglas, T. The small heat shock protein cage from Methanococcus jannaschii is a versatile nanoscale platform for genetic and chemical modification. Nano Lett. 2003, 3(11), 1573–1576. 23. Kang, S.; Lucon, J.; Varpness, Z.; Liepold, L.; Uchida, M.; Willits, D.; Young, M.; Douglas, T. Monitoring biomimetic platinum nanocluster formation using mass spectrometry and clusterdependent H2 production. Angew. Chem. Int. Ed. 2008, 47(41), 7845–7848. 24. Niemeyer, J.; Abe, S.; Hikage, T.; Ueno, T.; Erker, G.; Watanabe, Y. Noncovalent insertion of ferrocenes into the protein shell of apo-ferritin. Chem. Comm. 2008, 48, 6519–6521. 25. Klem, M. T.; Willits, D.; Solis, D. J.; Belcher, A. M.; Young, M.; Douglas, T. Bio-inspired synthesis of protein-encapsulated CoPt nanoparticles. Adv. Funct. Mater. 2005, 15(9), 1489–1494. 26. Iwahori, K.; Yoshizawa, K.; Muraoka, M.; Yamashita, I. Fabrication of ZnSe nanoparticles in the apoferritin cavity by designing a slow chemical reaction system. Inorg. Chem. 2005, 44, 6393–6400. 27. Allen, M.; Bulte, J. W. M.; Liepold, L.; Basu, G.; Zywicke, H. A.; Frank, J. A.; Young, M.; Douglas, T. Paramagnetic viral nanoparticles as potential high-relaxivity magnetic resonance contrast agents. Magn. Reson. Med. 2005, 54(4), 807–812. 28. Anderson, E. A.; Isaacman, S.; Peabody, D. S.; Wang, E. Y.; Canary, J. W.; Kirshenbaum, K. Viral nanoparticles donning a paramagnetic coat: conjugation of MRI contrast agents to the MS2 capsid. Nano Lett. 2006, 6(6), 1160–1164. 29. Hooker, J. M.; Datta, A.; Botta, M.; Raymond, K. N.; Francis, M. B. Magnetic resonance contrast agents from viral capsid shells: a comparison of exterior and interior cargo strategies. Nano Lett. 2007, 7(8), 2207–2210.
P1: OTA/XYZ P2: ABC c19 JWBS037-Chen January 25, 2011
478
13:19
Printer Name: Yet to Come
PROTEIN CAGES AS MULTIMODE IMAGING AGENTS
30. Liepold, L.; Anderson, S.; Willits, D.; Oltrogge, L.; Frank, J. A.; Douglas, T.; Young, M. Viral capsids as MRI contrast agents. Magn. Reson. Med. 2007, 58(5), 871–879. 31. Prasuhn, D. E.; Yeh, R. M.; Obenaus, A.; Manchester, M.; Finn, M. G. Viral MRI contrast agents: coordination of Gd by native virions and attachment of Gd complexes by azide-alkyne cycloaddition. Chem. Commun. 2007, 12, 1269–1271. 32. Datta, A.; Hooker, J. M.; Botta, M.; Francis, M. B.; Aime, S.; Raymond, K. N. High relaxivity gadolinium hydroxypyridonate–viral capsid conjugates: nanosized MRI contrast agents. J. Am. Chem. Soc. 2008, 130(8), 2546–2552. 33. Koenig, S. H.; Baglin, C.; Brown, R. D. Magnetic-field dependence of solvent proton relaxation induced by Gd3+ and Mn2+ complexes. Magn. Reson. Med. 1984, 1(4), 496–501. 34. Helm, L. Relaxivity in paramagnetic systems: theory and mechanisms. Prog. Nucl. Magn. Reson. Spectrosc. 2006, 49(1), 45–64. 35. Abedin, M. J.; Liepold, L.; Suci, P.; Young, M.; Douglas, T. Synthesis of a cross-linked branched polymer network in the interior of a protein cage. J. Am. Chem. Soc. 2009, 131(12), 4346– 4354. 36. Digilio, G.; Catanzaro, V.; Fedeli, F.; Gianolio, E.; Menchise, V.; Napolitano, R.; Gringeri, C.; Aime, S. Targeting exofacial protein thiols with Gd(III) complexes. An efficient procedure for MRI cell labelling. Chem. Commun. (Camb.) 2009, 8, 893–895. 37. Harrison, P. M.; Arosio, P. Ferritins: molecular properties, iron storage function and cellular regulation. Biochim. Biophys. Acta—Bioenergetics 1996, 1275(3), 161–203. 38. Stark, D. D.; Bass, N. M.; Moss, A. A.; Bacon, B. R.; McKerrow, J. H.; Cann, C. E.; Brito, A.; Goldberg, H. I. Nuclear magnetic resonance imaging of experimentally induced liver disease. Radiology 1983, 148(3), 743–751. 39. Gossuin, Y.; Muller, R.; Gillis, P.; Bartel, L. Relaxivities of human liver and spleen ferritin. Magn. Reson. Imaging 2005, 23(10), 1001–1004. 40. Vymazal, J.; Righini, A.; Brooks, R. A.; Canesi, M.; Mariani, C.; Leonardi, M.; Pezzoli, G. T1 and T2 in the brain of healthy subjects, patients with Parkinson disease, and patients with multiple system atrophy: relation to iron content. Radiology 1999, 211(2), 489–495. 41. Genove, G.; Demarco, U.; Xu, H.; Goins, W.; Ahrens, E. A new transgene reporter for in vivo magnetic resonance imaging. Nat. Med. 2005, 11(4), 450–454. 42. Cohen, B.; Dafni, H.; Meir, G.; Harmelin, A.; Neeman, M. Ferritin as an endogenous MRI reporter for noninvasive imaging of gene expression in C6 glioma tumors. Neoplasia 2005, 7(2), 109–117. 43. Bulte, J. W. M.; Douglas, T.; Mann, S.; Frankel, R. B.; Moskowitz, B. M.; Brooks, R. A.; Baumgarner, C. D.; Vymazal, J.; Strub, M. P.; Frank, J. A. Magnetoferritin—characterization of a novel superparamagnetic MR contrast agent. J. Magn. Reson. Imaging 1994, 4(3), 497–505. 44. Bulte, J. W.; Douglas, T.; Mann, S.; Vymazal, J.; Laughlin, P. G.; Frank, J. A. Initial assessment of magnetoferritin biokinetics and proton relaxation enhancement in rats. Acad. Radiol. 1995, 2(10), 871–878. 45. Uchida, M.; Flenniken, M. L.; Allen, M.; Willits, D. A.; Crowley, B. E.; Brumfield, S.; Willis, A. F.; Jackiw, L.; Jutila, M.; Young, M. J.; Douglas, T. Targeting of cancer cells with ferrimagnetic ferritin cage nanoparticles. J. Am. Chem. Soc. 2006, 128(51), 16626–16633. 46. Uchida, M.; Terashima, M.; Cunningham, C. H.; Suzuki, Y.; Willits, D. A.; Willis, A. F.; Yang, P. C.; Tsao, P. S.; McConnell, M. V.; Young, M. J.; Douglas, T. A human ferritin iron oxide nano-composite magnetic resonance contrast agent. Magn. Reson. Med. 2008, 60(5), 1073– 1081. 47. Jung, C. W.; Jacobs, P. Physical and chemical-properties of superparamagnetic iron-oxide MR contrast agents—Ferumoxides, Ferumoxtran, Ferumoxsil. Magn. Reson. Imaging 1995, 13(5), 661–674.
P1: OTA/XYZ P2: ABC c19 JWBS037-Chen January 25, 2011
13:19
Printer Name: Yet to Come
REFERENCES
479
48. Li, A. C.; Glass, C. K. The macrophage foam cell as a target for therapeutic intervention. Nat. Med. 2002, 8(11), 1235–1242. 49. Ross, R. Mechanisms of disease—atherosclerosis—an inflammatory disease. N. Engl. J. Med. 1999, 340(2), 115–126. 50. Lim, J. H.; Choi, D.; Cho, S. K.; Kim, S. H.; Lee, W. J.; Lim, H. K.; Park, C. K.; Paik, S. W.; Kim, Y. I. Conspicuity of hepatocellular nodular lesions in cirrhotic livers at ferumoxides-enhanced MR imaging: importance of Kupffer cell number. Radiology 2001, 220(3), 669–676. 51. Kooi, M. E.; Cappendijk, V. C.; Cleutjens, K.; Kessels, A. G. H.; Kitslaar, P.; Borgers, M.; Frederik, P. M.; Daemen, M.; van Engelshoven, J. M. A. Accumulation of ultrasmall superparamagnetic particles of iron oxide in human atherosclerotic plaques can be detected by in vivo magnetic resonance imaging. Circulation 2003, 107(19), 2453–2458. 52. Ruehm, S. G.; Corot, C.; Vogt, P.; Kolb, S.; Debatin, J. F. Magnetic resonance imaging of atherosclerotic plaque with ultrasmall superparamagnetic particles of iron oxide in hyperlipidemic rabbits. Circulation 2001, 103(3), 415–422. 53. Terashima, M.; Uchida, M.; Kosuge, H.; Cunningham, C. H.; Tsao, P.; Young, M.; Conolly, S. M.; Douglas, T.; McConnell, M. V. Human ferritin protein cage nanoparticles as macrophage-targeted contrast agents for detecting inflammation in atherosclerosis. Submitted to Biomaterials. 54. Arap, W.; Pasqualini, R.; Ruoslahti, E. Cancer treatment by targeted drug delivery to tumor vasculature in a mouse model. Science 1998, 279(5349), 377–380. 55. Uchida, M.; Willits, A. D.; Muller, K.; Willis, A. F.; Jackiw, L.; Jutila, M.; Young, M.; Porter, A. E.; Douglas, T. Intracellular distribution of macrophage targeting ferriti–iron oxide nanocomposite. Adv. Mat. 2009, 21(4), 458–462. 56. Zimmerli, W.; Trampuz, A.; Ochsner, P. E. Current concepts: prosthetic-joint infections. N. Engl. J. Med. 2004, 351(16), 1645–1654. 57. Taylor, M. D.; Napolitano, L. M. Methicillin-resistant Staphylococcus aureus infections in vascular surgery: increasing prevalence. Surg. Infect. (Larchmt.) 2004, 5(2), 180–187. 58. Marraffini, L. A.; DeDent, A. C.; Schneewind, O. Sortases and the art of anchoring proteins to the envelopes of gram-positive bacteria. Microbiol. Mol. Biol. Rev. 2006, 70(1), 192– 221. 59. Harris, L. G.; Foster, S. J.; Richards, R. G. An introduction to Staphylococcus aureus, and techniques for identifying and quantifying S. aureus adhesins in relation to adhesion to biomaterials: review. Eur. Cells Mater. 2002, 4, 39–60. 60. Suci, P. A.; Berglund, D. L.; Liepold, L.; Brumfield, S.; Pitts, B.; Davison, W.; Oltrogge, L.; Hoyt, K. O.; Codd, S.; Stewart, P. S.; Young, M.; Douglas, T. High-density targeting of a viral multifunctional nanoplatform to a pathogenic, biofilm-forming bacterium. Chem. Biol. 2007, 14(4), 387–398. 61. Kadurugamuwa, J. L.; Sin, L.; Albert, E.; Yu, J.; Francis, K.; DeBoer, M.; Rubin, M.; BellingerKawahara, C.; Parr, T. R.; Contag, P. R. Direct continuous method for monitoring biofilm infection in a mouse model. Infect. Immun. 2003, 71(2), 882–890. 62. Fluckiger, U.; Ulrich, M.; Steinhuber, A.; Doring, G.; Mack, D.; Landmann, R.; Goerke, C.; Wolz, C. Biofilm formation, icaADBC transcription, and polysaccharide intercellular adhesin synthesis by staphylococci in a device-related infection model. Infect. Immun. 2005, 73(3), 1811– 1819. 63. Monzon, M.; Garcia-Alvarez, F.; Lacleriga, A.; Gracia, E.; Leiva, J.; Oteiza, C.; Amorena, B. A simple infection model using pre-colonized implants to reproduce rat chronic Staphylococcus aureus osteomyelitis and study antibiotic treatment. J. Orthop. Res. 2001, 19(5), 820–826. 64. Kang, S.; Suci, P. A.; Broomell, C. C.; Iwahori, K.; Kobayashi, M.; Yamashita, I.; Young, M.; Douglas, T. Janus-like protein cages. Spatially controlled dual-functional surface modifications of protein cages. Nano Lett. 2009, 9(6), 2360–2366.
P1: OTA/XYZ P2: ABC c19 JWBS037-Chen January 25, 2011
480
13:19
Printer Name: Yet to Come
PROTEIN CAGES AS MULTIMODE IMAGING AGENTS
65. Suci, P. A.; Kang, S.; Young, M.; Douglas, T. A streptavidin–protein cage janus particle for polarized targeting and modular functionalization. J. Am. Chem. Soc. 2009, 131(26), 9164. 66. Leid, J. G.; Shirtliff, M. E.; Costerton, J. W.; Stoodley, P. Human leukocytes adhere to, penetrate, and respond to Staphylococcus aureus biofilms. Infect. Immun. 2002, 70(11), 6339–6345. 67. Ahrens, E. T.; Rothbacher, U.; Jacobs, R. E.; Fraser, S. E. A model for MRI contrast enhancement using T-1 agents. Proc. Nat. Acad. Sci. U.S.A 1998, 95(15), 8443–8448. 68. Gillitzer, E.; Willits, D.; Young, M.; Douglas, T. Chemical modification of a viral cage for multivalent presentation. Chem. Commun. 2002, 20, 2390–2391.
P1: OTA/XYZ P2: ABC c20 JWBS037-Chen January 25, 2011
13:21
Printer Name: Yet to Come
CHAPTER 20
Biomedical Applications of Single-Walled Carbon Nanotubes WEIBO CAI Departments of Radiology and Medical Physics, School of Medicine and Public Health, University of Wisconsin–Madison, and University of Wisconsin Carbone Cancer Center, Madison, Wisconsin, USA
TING GAO Tyco Electronics Corporation, Menlo Park, California, USA
HAO HONG Departments of Radiology and Medical Physics, School of Medicine and Public Health, University of Wisconsin–Madison, Madison, Wisconsin, USA
20.1 INTRODUCTION Nanotechnology, an interdisciplinary research field involving chemistry, engineering, biology, material science, and medicine, has enormous potential for early detection, accurate diagnosis, and personalized treatment of diseases. With the size of many orders of magnitude smaller than human cells, nanoparticles can offer unprecedented interactions with biomolecules both on the surface of and inside the cells, which may revolutionize disease diagnosis and treatment. Over the last decade, there have been numerous nanotechnology centers established worldwide and it is expected that nanotechnology will mature into a clinically useful field in the near future [1, 2]. To date, the most well-studied nanoparticles include quantum dots [3, 4], paramagnetic nanoparticles [5], nanowires [6], carbon nanotubes (CNTs) [7], gold nanoparticles [8], nanoshells [9], and many others [10, 11]. No other nanoparticles have been investigated for such a wide variety of biomedical applications like the single-walled carbon nanotubes (SWNTs). Over the last decade, SWNTs have been explored for sensing, drug/gene/vaccine delivery, thermal therapy, molecular imaging, tissue engineering, biomedical instrumentation, and other applications. With the capacity to provide enormous sensitivity, throughput, and flexibility, SWNTs have the potential to profoundly impact disease diagnosis and patient management in the near future. In this chapter, we summarize the current state-of-the-art of SWNTs for biomedical Nanoplatform-Based Molecular Imaging Edited by Xiaoyuan Chen C 2011 John Wiley & Sons, Inc. Copyright
481
P1: OTA/XYZ P2: ABC c20 JWBS037-Chen January 25, 2011
482
13:21
Printer Name: Yet to Come
BIOMEDICAL APPLICATIONS OF SINGLE-WALLED CARBON NANOTUBES
applications. Big strides have been made and many proof-of-principle studies have been successfully performed. The future of SWNTs looks brighter than ever yet many hurdles remain to be conquered. Discovery of the multiwalled carbon nanotube (MWNT) and synthesis of SWNTs in the early 1990s opened up a new arena for research on the nanoscale [12–14]. A search in PubMed for “carbon nanotube” returned more than 5000 publications. CNTs are generally produced by three major techniques: electric arc discharge, laser ablation, and thermal or plasma-enhanced chemical vapor deposition (CVD) [15]. A number of purification techniques have been developed for CNTs such as physical separation and gas/liquid phase oxidation in combination with chemical treatment [16]. Structural characterization of CNTs generally includes scanning electron microscopy (SEM), transmission electron microscopy (TEM), atomic force microscopy (AFM), scanning tunneling microscopy (STM), and many other techniques [17]. Among these, TEM and SEM are the most popular choices for imaging CNTs [18–20]. Intriguingly, the confined space inside SWNTs has been utilized for high-resolution TEM imaging of the dynamic behavior of individual small molecules [21, 22]. These findings opened up the possibility of investigating the biological activities of these molecules on an individual basis, which may potentially be applied to a wide range of systems. A number of methods have been explored for the chemical functionalization of CNTs, a prerequisite for their potential biomedical applications [23]. In addition, the unique properties of CNTs, including but not limited to electrical conductance, high mechanical stiffness, light weight, transistor behavior, piezoresistance, thermal conductivity, luminescence, and electrochemical bond expansion, have made them superb materials for a broad spectrum of applications ranging from energy storage (e.g., H2 , Li) to sensing to imaging and therapy.
20.2 SWNT-BASED SENSORS Besides the superb physical properties such as high strength and excellent electrical/thermal conductivity, SWNT also has a large surface area (estimated to be as high as 1600 m2 /g) which provides numerous sites at which chemicals can react [24]. Furthermore, SWNTs can be chemically modified or permeated with other materials to modify their physical, chemical, and electronic properties, which make them very versatile materials for many applications such as sensing. Compared to MWNTs, SWNTs are more durable, have less structural defects, are much easier to purify, and therefore are more suitable for sensing applications. Since the diameters of SWNTs are similar to or smaller than those of proteins, they can serve as high-performance electrical conduits for interfacing with biological systems, which is also highly desirable for sensing applications [25]. Several review articles focusing on CNT-based chemical and biochemical sensors have been published previously [26–28]. Chemical and biochemical sensors based on SWNTs vary by their transduction mechanisms, which can be electrical, electrochemical, optical, or mass resonant in nature [29, 30]. SWNT-based pressure and flow sensors, which can be used in a wide variety of scenarios in the clinic, are also briefly discussed. 20.2.1 Field-Effect Transistor (FET) Sensors SWNTs can be either metallic or semiconducting depending on the tube diameter and the chirality (i.e., the direction in which the graphite sheet is rolled to form a tube) [31, 32].
P1: OTA/XYZ P2: ABC c20 JWBS037-Chen January 25, 2011
13:21
Printer Name: Yet to Come
SWNT-BASED SENSORS
483
FIGURE 20.1 SWNT-based FET sensor: (a) a bottom gate FET, (b) a liquid gate FET, and (c) typical SWNT FET transfer characteristics in air (solid line), using the bottom gate, and in water (dotted line), using the liquid gate. S, source; D, drain. (Adapted from Bradley et al. [47].)
While metallic SWNTs can be used as coulomb islands in single-electron transistors [33, 34], semiconducting SWNTs can be used to build FETs [35, 36]. The electrical properties of SWNTs can be significantly affected by the interaction between the SWNTs and foreign molecular species, which makes them outstanding candidates for chemical nanosensors [37, 38]. To minimize the nonspecific binding (NSB) of SWNTs to chemicals or biomolecules, SWNTs can be functionalized to be biocompatible for the recognition of specific chemical species [26, 39–41]. SWNT FETs are typically fabricated on an SiO2 /Si chip, composed of multiple or single SWNT which connects two closely spaced metal electrodes that serve as the source and drain for transistors [42]. The conductance of the semiconductor between the source and drain is switched on and off by a third gate electrode that is capacitively coupled through a thin dielectric layer. The gate voltage is applied either on the Si chip (bottom gate) or on the liquid through a Pt electrode in which the entire device is immersed in (liquid gate) (Fig. 20.1). The bottom gate and liquid gate configurations are similar in direct current characterization yet they differ in scale because the capacitance between the SWNT and the two gates is different, with one being the oxide insulating layer (bottom gate) and the other being a hydration layer (liquid gate) (Fig. 20.1C). In an SWNT FET, the sensing signal could be monitored as the conductance values determined by the gate configurations (the current–voltage characteristics at various gate voltages) or as the current changes at a fixed gate voltage. The observed changes in conductance at a given gate voltage can be attributed to two phenomena, the Schottky barrier
P1: OTA/XYZ P2: ABC c20 JWBS037-Chen January 25, 2011
484
13:21
Printer Name: Yet to Come
BIOMEDICAL APPLICATIONS OF SINGLE-WALLED CARBON NANOTUBES
modulation at the SWNT/metal contact or the chemical gating at the SWNT channel [43, 44]. Several studies claimed that the majority of the source–drain current and conductance changes originated from the former [45, 46] while a few other reports suggested the latter [47, 48]. Certain small molecule (e.g., NO2 ) adsorption on SWNT can lead to p-type hole doping, which can increase positive hole carriers, thereby increasing the conductance of the SWNT [37]. On the other hand, n-type electron doping of the SWNT, which is caused by molecules such as NH3 , can result in a decrease in conductance. Many groups have investigated the use of SWNT FETs for sensing a wide variety of biolmolecules, such as streptavidin [39, 49], human immunoglobulin G (IgG) [40], carcinoembryonic antigen [50], hemagglutinin [51], Chromogranin A [52, 53], single-stranded DNA (ssDNA) [41], and ethanol vapor [54]. In one report, the enzymatic degradation of starch was monitored electronically using an SWNT-based FET sensor [55]. Incubation of the device in aqueous buffer solutions of amyloglucosidase resulted in the removal of starch from both the silicon surfaces and the side walls of the SWNTs in the FETs, as evidenced by both direct imaging and electronic measurements. In many of these studies, the removal of NSB to SWNTs was found to be critical and could dramatically affect the experimental results. One report suggested that reversible adsorption sites can be created on the SWNT via noncovalent functionalization with amine-terminated molecules and the analyte adsorption is largely affected by the basicity of the surface groups on the SWNT [56]. 20.2.2 Resistance Chemical Sensors Due to the high resistance of SWNT at the gate voltage of 0 V, direct detection of the change in SWNT resistance, when exposed to chemicals and biomolecules, is rarely used. This method has mostly been employed for gas detection with surface functionalized SWNTs. For example, palladium functionalized SWNTs showed good sensitivity and selectivity to H2 at room temperature [57], while palladium-doped SWNTs on an interdigitated electrode resistance sensor exhibited unique adsorption and electronic properties to methane [58]. Polyaniline functionalized SWNTs were demonstrated to have excellent sensitivity to NH3 , more than 60 times higher than that of nonfunctionalized SWNT sensors, as well as good reproducibility upon repeated exposure to high concentrations of NH3 [59]. In one report, network films of SWNT bundles on polyethylene terephthalate were used to detect nerve agent simulants such as diisopropyl methylphosphonate [60]. 20.2.3 Electrochemical Sensors Electrochemical sensors can facilitate direct electron exchange between the electrodes and a biomolecule [61, 62]. SWNTs can be incorporated into the electrode in many ways: random deposition; vertically “standing” on the electrode surface with one end of the SWNT exposed to the electrolyte solution and the other end in contact with the underlying electrode (termed “SWNT forest”); or integration into the electrode material [26, 27, 63–65]. Among these, SWNT forest-assembled electrodes have attracted the most attention, especially after they can be made to be stable for weeks by either electrostatic interactions [66, 67] or carbodiimide-assisted covalent coupling of amine-functionalized substrates with carboxylterminated SWNTs [68–70]. Cyclic amperometry/voltammetry and/or electrical impedance measurements are typically used for the monitoring of these SWNT-based electrochemical sensors.
P1: OTA/XYZ P2: ABC c20 JWBS037-Chen January 25, 2011
13:21
Printer Name: Yet to Come
SWNT-BASED SENSORS
485
SWNT modified electrodes have been used directly for sensing applications. For example, a carbon fiber electrode with SWNT deposited on the surface was used to detect nitric oxide release from single isolated human umbilical vein endothelial cells [71]. SWNT modification of carbon fiber microelectrodes was reported to increase the sensitivity and reduce fouling in the detection of dopamine and serotonin in vivo [72, 73]. SWNT modification of glass carbon electrode also proved useful for continuous and sensitive monitoring of the ascorbate level and its depletion in the rat striatum when induced by global ischemia [74]. In another study, SWNT forest self-assembled on graphite electrodes was demonstrated to be highly sensitive to mediated immunosensing [75]. In one interesting report, a freestanding SWNT film electrode has been studied as a nonenzymatic glucose sensor [76]. Lastly, a number of different polymer–SWNT composites have also been investigated for electrode modification [77–82]. The sensitivity and dynamic range of these sensors vary to a great extent depending on a number of experimental parameters; therefore they cannot be directly compared with each other. In many studies, the SWNTs that were used to modify the electrodes were either functionalized or immobilized with enzymes such as bilirubin oxidase [83], alcohol dehydrogenase [83], glucose dehydrogenase [83], glucose oxidase [84–87], and horseradish peroxidase [88]. These enzyme-containing electrodes exhibited good catalytic activities and some were suggested to also have a promising future in areas beyond sensing, such as biofuel cell applications. ssDNA is another popular choice for SWNT modification and a few examples have been given in the previous discussion regarding SWNT FET sensors [41]. In electrochemical sensing, electrodes coated with ssDNA-modified SWNTs have also been investigated [89, 90]. In one report, hydrogen peroxide reduction was detected with amplified electrochemical response using DNA immobilized on a ferrocene-filled SWNT electrode [91]. Several other studies have also focused on hydrogen peroxide detection. SWNT-protein dopant modified electrode [92] and a ferrocene-filled SWNT-based sensor [93] both exhibited enhanced sensor performance to hydrogen peroxide with good stability and reproducibility. In one intriguing report, the aligned reconstitution of glucose oxidase on the edge of SWNTs linked to an electrode surface was demonstrated for the first time [94]. The SWNT acted as a connector that electrically connected the active site of the enzyme and the electrode (Fig. 20.2). The electrons were transported along distances of greater than 150 nm and the rate of electron transport could be controlled by the length of the SWNTs [94]. These findings clearly demonstrated the compatibility of SWNTs with novel biomaterial hybrid systems that may have fascinating properties.
20.2.4 Capacitance Chemical Sensors Unlike chemiresistor sensors, which typically suffer from slow recovery after chemical exposure [95], because they detect charge transfer from the analytes, SWNT-based chemicapacitors detect the polarization of surface adsorbates. Therefore they are much more sensitive, recover faster, and are completely reversible. In the presence of a dilute chemical vapor, molecular adsorbates are polarized by the fringing electric fields radiating from the surface of an SWNT electrode, which causes an increase in its capacitance. By coating thin chemoselective materials on the SWNTs, the sensor can provide a large gain to the capacitance response, hence making it capable of detecting both volatile organic compounds (VOCs) and low vapor pressure explosives [96].
P1: OTA/XYZ P2: ABC c20 JWBS037-Chen January 25, 2011
486
13:21
Printer Name: Yet to Come
BIOMEDICAL APPLICATIONS OF SINGLE-WALLED CARBON NANOTUBES
(a)
(b)
nm 5
nm 800 600 400 200
nm
0 0
(c)
200
400
600
800
60 50
I / μA
40 30 20 10 0 –10 –0.2 –0.1 0 0.1 0.2 0.3 0.4 0.5 E / V (vs SCE)
FIGURE 20.2 Long-range electrical contact of enzymes by SWNT connectors. (a) Assembly of the electrode. FAD: flavin adenine dinucleotide, cofactor of the enzyme. (b) An AFM image of the glucose oxidase reconstituted on the FAD-functionalized SWNT monolayer associated with the Au surface. (c) Cyclic voltammograms corresponding to the catalyzed oxidation of different concentrations of glucose by the electrode. (Adapted from Patolsky et al. [94].)
The electronic response of SWNTs to trace levels of chemical vapors was explored [97]. It was found that adsorption at the defect sites produced a large electronic response that dominated the SWNT capacitance and conductance sensitivity. Furthermore, it was demonstrated that controlled introduction of oxidation defects can be used to enhance the sensitivity of an SWNT network sensor to a variety of chemical vapors. These initial results indicated that optimization of the chemoselective monolayers (to better cover the SWNTs), in combination with improved sensor design, may result in a new class of sorption-based sensors with high sensitivity and fast response time. To date, this sensing mechanism has not been widely used for biomolecule sensing yet.
20.2.5 Resonant Chemical Sensors Quartz crystal microbalance (QCM), which can measure a mass change by detecting the change in frequency of a quartz crystal resonator, has been employed for sensing applications [98]. In one study, SWNTs functionalized with a rheumatoid arthritis specific peptide were deposited on a quartz crystal [99]. Specific antibody binding was then detected in the serum from rheumatoid arthritis patients with sensitivity (in the femtomole range) significantly higher than that of established enzyme-linked immunosorbent assay (ELISA) and microarray systems. SWNT-based sensors for streptavidin and IgG detection were also reported to have sensitive changes in the electronic conductance and mass uptake [100]. By depositing SWNTs onto SiO2 film coated, ST-X quartz substrate as surface acoustic wave (SAW) sensors, VOCs could be detected [101, 102]. In subsequent studies, SWNTbased QCM, SAW, and silica optical fiber sensors were compared in the detection of
P1: OTA/XYZ P2: ABC c20 JWBS037-Chen January 25, 2011
13:21
Printer Name: Yet to Come
SWNT-BASED SENSORS
487
VOCs [103–105]. It was found that the frequency changes in SWNT-based SAW sensors, due to the conductivity and permittivity of the composite changes arising from chemical adsorption, could be tracked with a loop antenna, which makes them more advantageous for remote monitoring. 20.2.6 Pressure Sensors and Flow Sensors The electronic properties of SWNTs, strongly depending on their chirality and diameter, are affected by distortion (e.g., bending and twisting), which makes them excellent candidates for piezoelectric pressure sensors [28]. SWNT-based pressure sensors can be used in many diagnostic and therapeutic procedures/devices, such as in eye surgery, hospital beds, respiratory devices, patient monitors, inhalers, and kidney dialysis instruments. SWNT films on suspended polysilicon membrane were found to exhibit a change in resistance under uniform air pressure, which can be restored when the pressure is released [106]. Furthermore, randomly oriented SWNT bundle films showed a linear response in the measured change in voltage when subjected to tensile and compressive stress tests. Using first principle quantum transport calculations, molecular dynamics simulation, and continuum mechanics analysis, it was demonstrated that hydrostatic pressure can induce radial deformation, which can lead to electrical transition of SWNTs [107]. In this study, a pressure induced metal-to-semiconductor transition in armchair SWNTs was observed, which provided a basis for the design of novel nanoscale tunable pressure sensors. The fabrication and characterization of pressure sensors based on individual SWNTs as the active electromechanical transducer elements was reported [108]. Electromechanical measurements on strained metallic SWNTs adhering to a membrane revealed a piezoresistive gauge factor of approximately 210 for metallic SWNTs. SWNT bundles, which can produce an electrical signal in response to fluid flow when packed between two metal electrodes, have also been investigated [109]. 20.2.7 Sensors Based on Optical Techniques Several reports have focused on optical techniques for SWNT-based sensing. DNA–SWNT hybrids in aqueous solution, in the presence of glucose oxidase, were optically sensitive to hydrogen peroxide and glucose [110]. The DNA–SWNT hybrid surface could be regenerated by dialytically removing the hydrogen peroxide, which makes it potentially useful in immunoassays and glucose sensing. The dynamics of an SWNT-based optical sensor were compared to a flux-measuring electrochemical sensor for in vivo fluorescence sensing of glucose [111]. Both sensors demonstrated an approximately linear response to blood glucose levels. However, the SWNT-based optical sensor, which transduces glucose concentration, not flux, directly was found to be significantly more stable to membrane biofouling. In another report, hollow-core optical fibers integrated with SWNTs were used for VOC detection [112]. Good sensitivity and fast response time were achieved by characterizing the reflectance of the sensing probe. A different approach for the immobilization of glucose oxidase on SWNTs was also explored [113]. In this report, the near-infrared (NIR; 700–1200 nm) luminescence properties of the aqueous solution of DNA wrapped SWNTs did not change after addition of glucose oxidase. However, with surface bound potassium ferricyanide (PFC) as an electron transfer mediator, quenched emission of the SWNTs (after addition of PFC) was easily restored by glucose injection, as a result of the reaction between PFC and hydrogen peroxide, which
P1: OTA/XYZ P2: ABC c20 JWBS037-Chen January 25, 2011
488
13:21
Printer Name: Yet to Come
BIOMEDICAL APPLICATIONS OF SINGLE-WALLED CARBON NANOTUBES
FIGURE 20.3 SWNT-based protein microarray chip. (a) A schematic view of protein microarray formation on an SWNT substrate. (b) An AFM image of the SWNT film grown by the CVD method. (c) Fluorescence images after reaction of the substrates (containing biotin-BSA and SpA) with Cy5SA, Cy3-IgG, and a mixture of both. SA, streptavidin; SpA, a specific antigen recognized by the IgG. (Adapted from Byon et al. [114].)
was generated during glucose oxidation. The increased SWNT luminescence with rising glucose concentration can therefore be detected using this method. High-yield CVD-grown SWNT films were reported to be efficient substrates for microarray protein chip applications (Fig. 20.3) [114]. A simple one-step treatment of the SWNT films allows both the immobilization of probe proteins and efficient suppression of NSB without the need for additional bovine serum albumin (BSA). Since BSA treatment frequently leads to ambiguous results in various assays, mostly arising from the size and intrinsic NSB characteristics of BSA itself, it was suggested that such a BSA-free system should provide new opportunities for extensive and detailed studies of diverse biomolecular interactions. 20.2.8 Challenges and Future Directions for SWNT-Based Sensors SWNT-based sensors can have enormous impact on both disease diagnosis and treatment monitoring. However, clinical use of SWNT-based sensors is not yet mature. Separation of
P1: OTA/XYZ P2: ABC c20 JWBS037-Chen January 25, 2011
13:21
Printer Name: Yet to Come
GENE DELIVERY WITH SWNTs
489
metallic and semiconducting SWNTs to fabricate homogeneous sensing arrays, as well as the purification of SWNTs, is a major obstacle, unlike nanowires made from semiconducting materials [6]. Thus whether SWNT is the ideal material for nanosensing applications is still debatable. More importantly, whether SWNT-based sensing can compete well with DNA microarrays and mass spectroscopy-based proteomics remains unclear. As nanoscale sensing evolves, incorporation of SWNT-based sensors for multiplexed sensing will need to be applied to more complicated proteins and sample systems. Site-specific modification to improve selectivity/reproducibility, protective coatings to minimize NSB of biomolecules, and arraying sensor elements are current approaches to achieving reliable and consistent performance. SWNT arrays for conductivity-based sensing are probably more ideal than other sensing mechanisms in that they do not require labeling, external probes, or optical excitation. In many cases, literature reports have mixed findings. It will be ideal if different sensors (SWNT-based as well as those based on other nanoparticles) can be compared side-by-side using the same model system, which may significantly help in deciding which sensors are the best candidates for potential clinical testing. The National Cancer Institute (NCI) has required each of its funded nanotechnology centers to test their newly developed nanosensors using the same standard samples, which is expected to readily identify which new sensors truly stand out from the large pool of candidates. Extending a similar standard to a much broader range of research laboratories across the country would be highly beneficial to patients. Choosing the right candidate(s) at an early stage not only saves precious research time, but also significantly reduces the cost for new nanosensor development. Breakthrough in several research areas may lead to significant advances in the development of SWNT-based sensors. One area is the development of responsive biosensors that can change their configuration (e.g., size, gap, alignment, or stiffness) to improve sensitivity, selectivity, and/or performance based on the measurement of a specific analyte. The idea of this biosensor design is to use the intrinsic actuation properties of SWNTs to create a device that can be adjusted to improve its sensing capabilities or to add other functionalities such as delivering a drug or controlling a process that is related to an in-body device. Another promising research direction is hybrid biosensors that combine engineering and biological materials to form a more sensitive device, where the key is to find proteins that gate the process based on specific biomolecules. Lastly, the development of nanosensors that could be injected intravenously or implanted into the body, in order to monitor and record key physiological variables continuously, is another vibrant research area. Three major problems must be overcome to develop an in vivo biosensor: communication between the sensor and the outside world, elimination of biofouling, and potential toxicity of the sensor. Synthesis, processing, and device fabrication techniques for SWNTs have greatly improved after intensive research over the last decade. The future of SWNT-based sensors looks brighter than ever, yet many hurdles remain to be overcome.
20.3 GENE DELIVERY WITH SWNTs Gene therapy holds great potential for the treatment of a wide variety of diseases. However, the safety of gene therapy has always been a major concern for scientists, clinicians, and the general public [115]. Currently, gene therapy is generally performed through local administration. In addition, various strategies have been investigated to improve the gene
P1: OTA/XYZ P2: ABC c20 JWBS037-Chen January 25, 2011
490
13:21
Printer Name: Yet to Come
BIOMEDICAL APPLICATIONS OF SINGLE-WALLED CARBON NANOTUBES
expression efficiency and the safety profile of gene therapy for clinical applications [116]. SWNTs have been explored as a novel class of gene delivery vehicles because of their unique chemical/physical properties. In the first example of an SWNT-based gene delivery system, SWNTs were functionalized with ammonium to complex plasmid DNA through ionic interactions [117]. These complexes were proposed to enter cells by an endosome-independent mechanism. It was shown that these complexes gave higher DNA uptake and gene expression in vitro than DNA alone. However, such comparison is of no practical significance. Comparison with other DNA delivery agents, such as viral vectors and cationic polymers [118, 119], should have been carried out. After this proof-of-concept study, which demonstrated the potential of SWNTs for gene delivery applications, the interactions of different types of functionalized SWNTs with plasmid DNA were analyzed by various methods [120]. It was concluded that both the SWNT surface area and charge density were critical for the interaction between the SWNTs and the DNA. Systematic investigations of the cellular uptake mechanism for SWNTs were carried out [121]. Unlike the above-mentioned reports, it was found that intracellular transportation of proteins and DNA by SWNTs is through clathrin-dependent endocytosis, rather than caveolae or lipid rafts. The use of SWNTs as molecular transporters was demonstrated for various cargoes such as small molecules and proteins [122]. Recently, SWNTs have been employed in a number of studies for the delivery of genes such as antisense nucleic acids and small interfering RNAs (siRNAs). The major limitations of antisense therapy in the clinical setting are the rapid degradation of antisense nucleic acids and poor diffusion across the cell membrane [123]. To explore the potential of SWNTs in antisense therapy, SWNTs were used to deliver antisense myc (a oncogene [124]) into cells [125]. When compared with either antisense myc or SWNTs, the SWNT–antisense myc conjugate performed much better in the inhibition of cell proliferation, induction of apoptosis, and downregulation of myc at both the mRNA and protein levels. Recently, it was found that SWNTs can cause the polyA tail of single-stranded mRNA to form a duplex structure [126]. Since all mRNAs in eukaryotic cells have a polyA tail at the 3 end, this finding may potentially enable the application of SWNTs in targeting specific gene sequences. Gene silencing with siRNA, an attractive approach to probe gene function in mammalian cells, is also expected to be highly efficacious for therapeutic applications in the clinic [127–129]. The major limitation of siRNA-based therapy is the efficiency of siRNA delivery, where cytoplasmic delivery is a prerequisite for silencing mRNA expression through RNA interference (RNAi). Because of their ability to cross cell membranes, SWNTs were used to deliver siRNA and DNA through a cleavable disulfide bond linkage (Fig. 20.4) [130]. Subsequently, SWNTs were explored for siRNA delivery into human T cells and primary cells (Fig. 20.4) [131]. It was shown that the delivery capability and RNAi efficiency of SWNTs exceeded that of nonviral transfection agents such as liposomes. Further, cellular uptake was dependent on the functionalization and the polyethylene glycol (PEG) chain length on the SWNTs. It was proposed that hydrophobic interaction is one of the major driving forces for cellular delivery of siRNA. In another study, delivery of telomerase reverse transcriptase siRNA with positively charged SWNTs was found to suppresses tumor growth both in vitro and in vivo [132]. SWNT-based siRNA delivery, targeting various genes, has also been achieved in primary cardiomyocytes and chronic myelogenous leukemia cells [133, 134].
P1: OTA/XYZ P2: ABC c20 JWBS037-Chen January 25, 2011
13:21
Printer Name: Yet to Come
DRUG DELIVERY WITH SWNTs
491
FIGURE 20.4 SWNT-based gene (DNA and siRNA) delivery into cells. (a) Functionalized SWNTs were conjugated to the cargo (DNA or siRNA) through cleavable disulfide linkages. (b) Three confocal images recorded with different focal planes (top, middle, and bottom of cells) along the viewing direction for a cell after incubation with SWNT–DNA conjugate. The red color surrounding the nucleus corresponds to Cy3-labeled DNA. (c) Confocal images of untreated human T cells (left), cells treated with siRNA in complex with a commercial transfection agent (middle), and cells treated with SWNT–siRNA complexes (right), respectively, after staining for the cell surface marker targeted by the siRNA. Scale bar: 40 m. (Adapted from Kam et al. [130] and Liu et al. [131].)
These reports clearly demonstrated the potential of SWNTs as gene delivery systems. However, much future research and optimization will be needed before any clinical investigation since systemic delivery of genes (either DNA or siRNA) with SWNTs has not been achieved in living subjects. The barriers for gene delivery in vivo far supersede the barriers posed by the cell membrane in culture. Whether these SWNT conjugates can efficiently extravasate and deliver sufficient quantity of genes into the targeted cells in live animals remains unclear. In particular, delivering foreign DNAs into the cell nucleus is very challenging. The large surface area of SWNTs, which offers ample sites for the incorporation of nuclear localization sequences [135, 136], may be advantageous for DNA delivery and should be explored in future studies.
20.4 DRUG DELIVERY WITH SWNTs The same properties that make SWNTs attractive as gene delivery vehicles also make them desirable for drug delivery applications. In one report, it was found that various proteins adsorb spontaneously on the side walls of SWNTs through NSB [137]. More importantly, the adsorbed proteins could be delivered into the cytoplasm of various mammalian cells, presumably via the endocytosis pathway, which can lead to various biological sequences depending on the proteins used. A recent study showed that dispersion of SWNTs by ultrasonication with phospholipid-PEG can fragment the polymer, thus interfering with its ability to block nonspecific uptake by cells [138]. Since ultrasonication is a technique commonly used to disperse SWNTs [139], this may be a concern for drug delivery applications. One report showed that large surface areas exist for supramolecular chemistry on functionalized SWNTs, either noncovalently or covalently [140]. Water-soluble SWNTs with PEG functionalization can allow for very high degrees of -stacking of aromatic molecules such as anticancer drugs and/or certain fluorescent molecules, which can be
P1: OTA/XYZ P2: ABC c20 JWBS037-Chen January 25, 2011
492
13:21
Printer Name: Yet to Come
BIOMEDICAL APPLICATIONS OF SINGLE-WALLED CARBON NANOTUBES
subsequently released in a controllable fashion by varying the pH value. The dependence of the -stacking strength on the diameter of SWNTs could also be exploited for controlled release of the loaded molecules. The fact that multiple species such as PEG, drugs, and fluorescent tags can be attached onto the surface of the same SWNT makes it highly useful for a wide variety of future applications. Covalent linkages have been explored for SWNT-based drug delivery. One of the major disadvantages of low molecular weight (MW) anticancer drugs is their short circulation halflives, which typically lead to low tumor uptake and poor therapeutic efficacy [141]. Aminefunctionalized soluble SWNTs have been derivatized with cisplatin prodrug conjugates, where multiple prodrug centers were incorporated onto each SWNT [142]. Through pHmediated release following endocytosis in cancer cells, the cytotoxicity of the SWNT– prodrug conjugate was >100-fold more potent than the prodrug itself. In a follow-up study, targeted SWNT-mediated cisplatin prodrug delivery was reported using folate as the targeting ligand to cells that overexpress the folate receptor [143]. A recent report from another group also showed that an SWNT-based, tumor-targeted prodrug delivery system exhibited high potency toward cancer cells [144]. Recently, SWNT-based drug delivery was shown to suppress tumor growth in living mice [145]. Paclitaxel (PTX, Taxol), a widely used anticancer drug [146], was attached to PEG chains on the surface of SWNTs via a cleavable ester bond. The SWNT–PTX conjugate afforded higher efficacy in suppressing tumor growth than PTX in a murine breast cancer model, primarily due to prolonged blood circulation, which led to higher PTX uptake in the tumor through the enhanced permeability and retention (EPR) effect (Fig. 20.5). Such EPR effect is due to the nonspecific accumulation of SWNT–PTX in the tumor, since the tumor vasculature is more leaky than normal blood vessels and there is no lymphatic drainage in the tumor [147]. In future studies, tumor-specific targeting may further enhance the therapeutic index of anticancer drugs delivered by SWNTs. Various antibodies have been attached to SWNTs, either noncovalently or covalently. Due to the large surface area of SWNTs, it was shown that antibodies adsorbed onto SWNT bundles could stimulate T cells more effectively than equivalent concentrations of the antibody itself, potentially useful for applications such as immunotherapy [148]. In another study, a multivalent system based on SWNTs was constructed to target T cells both in vitro and in living animals [149]. Using a biotin–neutravidin linkage, the functionalized SWNTs were linked with antibodies for T cell receptor postsignaling endocytosis, as well as a synthetic fusogenic polymer for lysosome disruption. Although cytoplasmic delivery of SWNTs was accomplished in live animals, whether this system can enable therapeutically efficacious drug/gene delivery in vivo remains to be demonstrated in the future. In summary, SWNTs have been explored for drug delivery applications in cancer cells and animals through NSB, targeted and nontargeted approaches, as well as controlled release. Several studies in animals have been reported, which makes SWNT-based drug delivery one step closer to potential clinical applications than SWNT-based gene delivery. More importantly, the biological barriers faced by drug delivery are much less challenging than those confronted by gene delivery. In many cases, EPR-based drug delivery is already therapeutically efficacious and many anticancer drug formulations have been approved for clinical use without any targeting moiety [150, 151]. Once the potential toxicity of SWNTs is fully addressed, they may find widespread use in the clinic because of their versatile chemistry.
P1: OTA/XYZ P2: ABC c20 JWBS037-Chen January 25, 2011
13:21
Printer Name: Yet to Come
VACCINE DELIVERY WITH SWNTs
493
FIGURE 20.5 Drug delivery with SWNTs for anticancer therapy in live animals. (a) A schematic illustration of PTX conjugation to phospholipid-PEG functionalized SWNT, which allows for PTX release in vivo by cleavage of the ester bond. (b) A photo of representative tumors harvested from untreated mice, Taxol-treated mice, and SWNT–PTX treated mice at the end of the study. (c) TUNEL (indicating apoptosis status) and Ki67 (indicating cell proliferation) staining of tumor slices from mice after different treatment, which confirmed the therapeutic efficacy of SWNT–PTX. The fluorescence signal is shown in red and blue staining indicates the nuclei. Scale bar: 100 m. (Adapted from Liu et al. [145].)
20.5 VACCINE DELIVERY WITH SWNTs A few studies have demonstrated that SWNTs can also be useful for vaccine delivery. For example, a peptide from the foot-and-mouth disease virus, when linked to SWNTs covalently, was able to retain the structural integrity and immunogenicity hence could be recognized by monoclonal and polyclonal antibodies [152]. In another report, a neutralizing and protective B cell epitope was linked to SWNTs via selective chemical ligation [153]. The
P1: OTA/XYZ P2: ABC c20 JWBS037-Chen January 25, 2011
494
13:21
Printer Name: Yet to Come
BIOMEDICAL APPLICATIONS OF SINGLE-WALLED CARBON NANOTUBES
SWNT displayed B cell epitope retained its conformational characteristics because it could be recognized by specific antibodies. Subsequently, it was reported that functionalized SWNTs could also be used for delivering oligodeoxynucleotides into target cells [154]. Further investigation in animal models is required to evaluate the immunogenicity of these SWNT-based complexes in vivo and elucidate their potential therapeutic or adjuvant properties.
20.6 THERMAL THERAPY WITH SWNTs Aside from the use of SWNTs as delivery vehicles, the physical properties of the SWNT itself can also be exploited for therapeutic applications. In a pioneering report, the intrinsic strong optical absorbance of SWNTs was used for selective cancer cell killing [155]. It was reported that continuous NIR radiation of SWNT-loaded cells can cause cell death because of excessive local heating of the SWNTs in vitro. Taking advantage of this phenomenon, selective cancer cell destruction was achieved by functionalization of SWNTs with folate, which can enable internalization of SWNTs into cells overexpressing the folate receptor on the surface (Fig. 20.6). Together, the versatile delivery capabilities of SWNTs combined with their intrinsic optical properties can lead to a new class of novel materials for cancer therapy. Subsequently, SWNTs have been used for photothermal antimicrobial therapy in Escherichia coli [156]. Upon exposure to multipulse lasers at 532 and 1064 nm with nanosecond pulse duration, notable changes in bacteria viability were observed. Thermal ablation of tumor cells with antibody-functionalized SWNTs was also reported recently by another group [157].
FIGURE 20.6 Photothermal cancer cell killing with SWNTs. (Left) Selective internalization of functionalized SWNTs into folate receptor overexpressing (FR+) cells, followed by NIR laser radiation, caused cell death. (Right) Cells that do not express the folate receptor remain viable after the same treatment. Inset: high magnification image of the cells. (Adapted from Kam et al. [155].)
P1: OTA/XYZ P2: ABC c20 JWBS037-Chen January 25, 2011
13:21
Printer Name: Yet to Come
NONRADIONUCLIDE-BASED MOLECULAR IMAGING WITH SWNTs
495
Heat release from SWNTs in a radiofrequency (rf) field was found to be cytotoxic in cell culture [158]. When treated in the rf field, human cancer cell lines incubated with functionalized SWNTs showed selective, SWNT concentration-dependent thermal destruction. In a rabbit hepatic tumor model, SWNT injection followed by rf field treatment resulted in complete tumor necrosis at 48 hours while control tumors treated with rf without SWNT injection remained completely viable. Conjugation of certain targeting ligands to enhance the SWNT uptake by cancer cells may further improve the therapeutic effect in future studies. Various other nanoparticles have been used for tumor ablation based on their intrinsic physical properties. For example, gold nanoparticles have been tested for photothermal destruction of tumor cells/tissues in vitro and in vivo [8, 159]. Inductively heating a radiolabeled, antibody-linked iron oxide nanoparticle by externally applied alternating magnetic field was shown to cause tumor necrosis in mice bearing human breast cancer tumors [160, 161]. These nanoparticles with intrinsic therapeutic potential, including SWNTs, may have far reaching impact in the treatment of cancer and other diseases, after they are functionalized for targeted drug/gene delivery and/or noninvasive imaging.
20.7 NONRADIONUCLIDE-BASED MOLECULAR IMAGING WITH SWNTs The field of molecular imaging, “the visualization, characterization and measurement of biological processes at the molecular and cellular levels in humans and other living systems” [162], has expanded tremendously over the last decade. In general, molecular imaging modalities include molecular magnetic resonance imaging (MRI), magnetic resonance spectroscopy (MRS), optical bioluminescence, optical fluorescence, targeted ultrasound, single photon emission computed tomography (SPECT), and positron emission tomography (PET) [163]. Many hybrid systems that combine two or more modalities are also commercially available and certain others are under active development [164–166]. Molecular imaging takes advantage of traditional diagnostic imaging techniques and introduces molecular imaging agents (probes) to determine the expression of indicative molecular markers at different stages of diseases. Noninvasive detection of the molecular markers can allow for much earlier diagnosis, earlier treatment, and better prognosis that will eventually lead to personalized medicine [167, 168]. Continued development and wider availability of scanners dedicated to small animal imaging studies, which can provide a similar in vivo imaging capability in mice, primates, and humans, can enable smooth transfer of knowledge and molecular measurements between species, thereby facilitating clinical translation. The physical properties and relative ease for chemical modification have made SWNTs the most versatile nanoparticle for molecule imaging applications. 20.7.1 MRI with SWNTs MRI detects the interaction of protons (or certain other nuclei) with each other and with the surrounding molecules in a tissue of interest [169]. The endogenous tissue contrast can be enhanced by administration of exogenous contrast agents [170, 171]. Traditionally, Gd3+ chelates have been used to enhance the T1 contrast [172] and iron oxide nanoparticles have been used to increase the T2 contrast [5]. Gd3+ -functionalized SWNTs have been studied as MRI contrast agents [173]. The nanoscale loading and confinement of aquated Gd3+ ion clusters within ultrashort (US;
P1: OTA/XYZ P2: ABC c20 JWBS037-Chen January 25, 2011
496
13:21
Printer Name: Yet to Come
BIOMEDICAL APPLICATIONS OF SINGLE-WALLED CARBON NANOTUBES
FIGURE 20.7 Molecular imaging with SWNTs. (a) MRI of Gd3+ filled SWNTs at different pH value. (b) NIRF images of antigen-positive and control cells treated with an SWNT–antibody conjugate. (c) (Left) Photograph of an integrin ␣v 3 -positive tumor-bearing mouse depicting the area scanned with Raman spectroscopy (black box). (Right) Raman tumor maps of mice receiving SWNT–RGD conjugate or unconjugated SWNT at 72 h postinjection. (d) Photographs of the tumors in mice and the corresponding photoacoustic subtraction images (green). (e) Two-dimensional projection of the PET images of integrin ␣v 3 -positive tumor-bearing mice at 8 h postinjection of SWNT–RGD without and with (denoted as “Block”) coinjection of an excess amount of the RGD peptide. Arrowheads indicate the tumors. (Adapted from Hartman et al. [174], Welsher et al. [186], Zavaleta et al. [204], De la Zerda et al. [215], and Liu et al. [224].)
20–80 nm) SWNTs was found to be linear superparamagnetic molecular magnets with a MRI efficacy of 40 times greater than that of the Gd3+ -based contrast agents in current clinical use [173]. In a follow-up study, it was found that these Gd3+ -containing US SWNTs are also sensitive to the pH value of the environment (Fig. 20.7a) [174]. These findings suggested that Gd3+ -containing SWNTs might be useful in the clinic for early detection of cancer, where the extracellular pH of tumors can drop to pH 7.0 or below.
P1: OTA/XYZ P2: ABC c20 JWBS037-Chen January 25, 2011
13:21
Printer Name: Yet to Come
NONRADIONUCLIDE-BASED MOLECULAR IMAGING WITH SWNTs
497
20.7.2 Fluorescence Imaging with SWNTs Due to their strong light quenching properties, direct fluorescent labeling of SWNTs has not been very successful for optical imaging applications. However, studies have shown that optical properties of SWNTs can also be manipulated without covalent modification [175, 176]. The major drawback of optical imaging in living subjects is the poor tissue penetration and intense scattering of light [177]. In the NIR region, the absorbance of all biomolecules reaches a minimum, which provides a relatively clear window for optical studies/applications [178, 179]. Therefore NIR fluorescence (NIRF) imaging can provide opportunities for rapid and cost-effective preclinical evaluation in small animal models. Many reports have investigated the NIRF imaging of SWNT and/or SWNT conjugates ex vivo [180–183], in cells (Fig. 20.7b) [184–187], and in living subjects [188]. Among all molecular imaging modalities, no single modality is perfect and sufficient to obtain all the necessary information for a particular question [163]. A combination of multiple imaging modalities can offer synergistic advantages over any single modality alone. In one report, heterostructured complexes formed from iron oxide nanoparticles and SWNTs were investigated as multimodal imaging agents [189]. Macrophage cells that engulf the DNA-wrapped complexes were imaged with both MRI and NIRF mapping. Compared with fluorescent dyes or quantum dots, the more popular fluorescent agents, the quantum efficiency of SWNTs is much lower. Therefore NIRF imaging of SWNTs is quite challenging in live animals (e.g., mice and rats). Another limitation of NIRF microscopy of SWNTs is the diffraction limit to resolution, which is more relevant for high-resolution imaging of cells. Recently, a nonperturbing, far-field optical measurement was reported, which can allow for sub-wavelength mapping of single molecule chemical reaction sites on semiconducting SWNTs [190]. X-ray fluorescence has also been applied for the first time to study macrophages exposed to crude and purified SWNTs [191].
20.7.3 Raman Spectroscopy with SWNTs Raman spectroscopy, which can differentiate the spectral fingerprint of various molecules, can have very high multiplexing capabilities [192, 193]. However, the inherently weak magnitude of the Raman effect limits its sensitivity, which severely hampered the biomedical applications of Raman spectroscopy. SWNT has an intense Raman peak produced by the strong electron–phonon coupling, which causes efficient excitation of tangential vibration in the SWNT upon light exposure [194]. A number of literature reports have focused on Raman imaging and examination of the surface-enhanced Raman scattering (SERS) of SWNTs [195–202]. In a pioneering study, SWNTs and other SERS nanoparticles were used to demonstrate whole-body Raman imaging, multiplexing, as well as in vivo tumor targeting with a small-animal Raman imaging system [203]. Subsequently, an optimized Raman microscope was used to further evaluate tumor targeting and localization of SWNTs in mice (Fig. 20.7c) [204]. The molecular target and the targeting ligand used in this study was integrin ␣v 3 and an arginine–glycine–aspartic acid (RGD, potent integrin ␣v 3 antagonist [205]) containing peptide, respectively, one of the most extensively studied and validated receptor–ligand pair over the last decade [206, 207]. One major advantage of Raman spectroscopy is its superb potential for multiplexed imaging. In a recent study, multiplexed Raman imaging of live cells with isotopically modified SWNTs was reported [208]. SWNTs with different isotope (12 C and 13 C)
P1: OTA/XYZ P2: ABC c20 JWBS037-Chen January 25, 2011
498
13:21
Printer Name: Yet to Come
BIOMEDICAL APPLICATIONS OF SINGLE-WALLED CARBON NANOTUBES
compositions, which in turn exhibit distinct Raman G-band peaks, were used for multiplexed Raman imaging in different cancer cells overexpressing specific cell surface receptors. Functionalized, macromolecular SWNTs have recently been used as multicolor Raman labels for highly sensitive, multiplexed protein detection in an arrayed format [209]. In future studies, SWNTs with controlled diameters (which also exhibit distinctive Raman peaks in their radial-breathing modes depending on the diameters of the SWNTs [210]) may also be investigated for multiplexed Raman imaging. 20.7.4 Photoacoustic Tomography (PAT) with SWNTs PAT is a cross-sectional imaging technique based on the photoacoustic effect [211]. In PAT, the tissue is usually irradiated by a short-pulsed laser beam to produce thermal and acoustic impulse responses. Locally absorbed light is converted into heat, which is further converted to a pressure rise via thermoelastic expansion of the tissue. The initial pressure rise propagates in the tissue as an ultrasonic wave, referred to as a photoacoustic wave, which is detected by ultrasonic transducers placed outside the tissue to produce electric signals. The electric signals are then amplified, digitized, and transferred to a computer to form a PAT image. A number of contrast agents for photoacoustic imaging have been suggested [212–214], yet most were not used for targeted imaging. In a recent study, SWNTs conjugated with cyclic RGD peptides were used to enhance the contrast of photoacoustic imaging of tumors in living mice [215]. Intravenous administration of the SWNT–RGD conjugate to tumorbearing mice showed eight times greater photoacoustic signal in the tumor than mice injected with SWNT alone (Fig. 20.7d). Taking advantage of the intrinsic Raman signal of the SWNTs, the in vivo PAT imaging results were further validated ex vivo with Raman microscopy. 20.7.5 A Brief Summary The major disadvantage of MRI is its inherent low sensitivity, which can only be partially compensated by working at higher magnetic fields (4.7–14 T), acquiring data for much longer time periods, and using exogenous contrast agents. Although proof-of-principle studies have been reported for molecular MRI of several targets [216], whether molecular MRI can significantly improve patient management remains to be elucidated. Whether Gd3+ -containing SWNTs can be molecularly targeted and perform well in animal studies needs to be investigated before any potential clinical applications can be in place. For clinical applications, optical imaging techniques will only be possible in limited sites such as the tissues and lesions close to the surface of the skin, tissues accessible by endoscopy, and intraoperative visualization [3]. Due to the poor fluorescence properties of SWNTs, they are not suitable for NIRF imaging in potential human studies. However, Raman spectroscopy and PAT may have potential clinical relevance/applications in the future, although the primary limitation is still light penetration beyond a few centimeters of tissue. The key advantages of Raman imaging over fluorescence imaging is the very high multiplexing capability and lack of confounding background signal from autofluorescence. In addition, the use of SWNTs as Raman tags is more advantageous over other Raman nanoparticles because of the inherent Raman signature of SWNTs, which requires no further labeling or encapsulation to produce a Raman signal. PAT compares very favorably
P1: OTA/XYZ P2: ABC c20 JWBS037-Chen January 25, 2011
13:21
Printer Name: Yet to Come
RADIONUCLIDE-BASED MOLECULAR IMAGING WITH SWNTs
499
to other imaging modalities with its precise depth information, submillimeter resolution, and nanomolar sensitivity [215]. With further improvement in background reduction, hardware and image reconstruction software, and the use of lasers with high repetition rates, it is very likely that PAT will find wide use in the future in both basic research and clinical care. The fact that SWNTs can be used as contrast agents in a number of different imaging techniques, which can significantly facilitate the cross-validation of the imaging result, is also another highly desirable trait over other contrast agents.
20.8 RADIONUCLIDE-BASED MOLECULAR IMAGING WITH SWNTs Radionuclide-based imaging techniques (PET and SPECT) have much more clinical relevance and hence are much more widely used in the clinic than molecular MRI and optical imaging. Not only is there no tissue penetration limit for these techniques, but PET and SPECT are also highly sensitive (down to the picomolar level) and quantitative [217]. In one early study, water-soluble hydroxylated SWNTs were labeled with 125 I to study the distribution in mice [218]. It was concluded that these SWNTs moved easily among the compartments and tissues of the body and behaved like small molecules, despite the fact that their apparent molecular weight was tremendously large. In a later report, 111 In-labled water-soluble SWNTs were tested for SPECT imaging [219]. Strikingly, intravenous administration of these functionalized SWNTs indicated that they were not retained in any of the reticuloendothelial system (RES) organs (e.g., liver or spleen) and were rapidly cleared from the circulation system through the renal excretion route. Because of the use of lead collimators to define the angle of incidence, SPECT imaging has a very low detection efficiency (< 10−4 times the emitted number of gamma rays) [220]. PET, on the other hand, has much higher sensitivity (greater than 10%) [221]. It was first developed in the mid-1970s [222] and dedicated PET scanners for small-animal studies were first reported in the late 1990s [223]. The biodistribution of 64 Cu-labeled SWNTs in mice has been investigated by PET, ex vivo biodistribution, and Raman spectroscopy [224]. It was found that these SWNTs are highly stable in vivo and the surface PEG chain length can significantly affect its blood concentration and biodistribution. Most importantly, efficient targeting of integrin ␣v 3 -positive tumor in mice was achieved with RGD peptide conjugated SWNTs (Fig. 20.7e). The intrinsic Raman signatures of SWNTs were used to directly probe the presence of SWNTs in mouse tissues and confirm the PET imaging results. All in vivo and ex vivo measurements confirmed that there was minimal renal uptake of the SWNT–RGD conjugate, and the majority of the conjugate was taken up by the tumor and the RES [224]. Subsequently, a few other reports also appeared on radiolabeled SWNTs (both targeted and nontargeted) [225–227]. Radiolabeled nanoparticles represent a new class of probes that have enormous potential for clinical applications. Different from other molecular imaging modalities, where typically the nanoparticle itself is detected, such as quantum dots and iron oxide nanoparticles for optical imaging and MRI, respectively, radionuclide-based imaging detects the radiolabel rather than the nanoparticle itself. The nanoparticle distribution is measured indirectly by assessing the localization of the radionuclide, which can provide quantitative measurement of the tumor targeting efficacy and pharmacokinetics only if the radiolabel on the nanoparticle is stable enough under physiological conditions. However, dissociation of the radionuclide (typically metal) from the chelator, and/or the radionuclide-containing
P1: OTA/XYZ P2: ABC c20 JWBS037-Chen January 25, 2011
500
13:21
Printer Name: Yet to Come
BIOMEDICAL APPLICATIONS OF SINGLE-WALLED CARBON NANOTUBES
TABLE 20.1 The Unique Physical and Chemical Properties of SWNTs Have Made Them Versatile Contrast Agents for a Number of Imaging Modalities Modality MRI NIRF Raman PAT PET/SPECT
Contrast Agent 3+
Gd SWNT SWNT SWNT Radioisotopes
Resolution
Sensitivity
Penetration
References
+++ ++ + +++ ++
+ ++ ++ ++ +++
+++ + + + +++
173, 174, 189 180–182, 185–191 195–204, 208, 209 215 219, 224–227
polymer coating from the nanoparticle, may occur, which can cause significant difference between the nanoparticle distribution and the radionuclide distribution. In several studies, SWNTs were reported to undergo either complete or partial renal clearance in mice, with little uptake by the liver or other organs of the RES [219, 225]. These findings defy the general trend of high RES uptake for nanoparticles [228, 229] and deserve further investigation/validation. Studies have shown that typically only molecules less than 70 kDa (a few nanometers in diameter) undergo renal clearance [230–232]. The SWNTs used in these reports are typically more than 200 nm in length, even up to a few micrometers [219, 224, 225]. It is very unlikely that SWNTs can be cleared from the kidney (dissociated radiolabel and polymer coating can, however) unless severe kidney damage has occurred. Nanoparticle-based imaging has touched upon every single modality of the molecular imaging arena [11]. The unique physical and chemical properties of SWNTs have made them versatile contrast agents for a wide variety of imaging modalities, including MRI, optical, SPECT, and PET imaging. In addition, SWNTs have also been investigated with several other imaging techniques such as Raman spectroscopy and PAT. A summary and comparison of these imaging techniques, with SWNTs or SWNT conjugates as the contrast agents, is shown in Table 20.1. It is expected that SWNT-based molecular imaging will continue to flourish for years to come.
20.9 TISSUE ENGINEERING WITH SWNTs The goal of tissue engineering is to replace diseased or damaged tissue with biologic substitutes that can restore and maintain normal function. Major advances in the areas of cell and organ transplantation [233], as well as advances in materials science and engineering, have aided in the continuing development of tissue engineering and regenerative medicine [234]. As the field of tissue engineering advances, there is a growing need for better monitoring and evaluation tools of engineered tissue along with new biomaterials to facilitate tissue growth. One class of nanomaterials that has the potential for multiple applications in tissue engineering is the SWNTs [235]. 20.9.1 Cell-Based Studies A number of reports have investigated, in cell culture, the potential of SWNTs in tissue engineering applications. Chemically functionalized water-soluble SWNT graft copolymers, when added to the culturing medium, were shown to increase the length of various neuronal
P1: OTA/XYZ P2: ABC c20 JWBS037-Chen January 25, 2011
13:21
Printer Name: Yet to Come
TISSUE ENGINEERING WITH SWNTs
501
processes [236]. These SWNT-based conjugates not only exhibited excellent biocompatibility with osteoblast cells but also appeared to modulate the cell phenotype. SWNTs grafted with polyethyleneimine, which could promote neurite outgrowth and branching, have been used as the substrate for neuron culture [237]. Subsequently, similar SWNTs were also shown to be suitable scaffold materials for osteoblast proliferation and bone formation, which represents a potential technological advance in the field of bone engineering [238]. SWNTs have been shown to support neuronal attachment and growth [239]. Simple chemical modification of the SWNTs was employed to control cell growth without interfering with the ongoing neuronal function. More importantly, electrical coupling of SWNTs and neurons was demonstrated, suggesting that SWNTs are flexible materials for tissue engineering applications that involve electrically excitable tissues such as muscles and nerves. Other groups have also demonstrated that SWNTs can be used as scaffolds for cell culture [240–242]. The capability of SWNTs in enhancing long-term cell proliferation can be of significance to in vitro cell amplification on a large scale, tissue regeneration or guided repair, as well as biomedical device applications. Composite polymeric scaffolds from alginate and SWNTs have been shown to improve the adhesion and proliferation of rat heart endothelial cells [243]. Recently, biodegradable SWNT–polymer composites were investigated for their effects on the viability of fibroblast cells [244]. Nearly 100% cell viability was observed on all crosslinked nanocomposites and cell attachment on their surfaces was comparable with that on tissue culture polystyrene, clearly indicating favorable cytocompatibility of these nanocomposites for potential use as scaffolds in tissue engineering applications. Neural stem cells (NSCs) are highly plastic neural precursors capable of adapting to environmental conditions and recreating signal transduction pathways [245]. Successful differentiation of mouse NSCs on layer-by-layer assembled SWNT composites was reported [246]. Mouse embryonic NSCs from the cortex successfully differentiated into neurons, astrocytes, and oligodendrocytes with clear formation of neurites (Fig. 20.8a). A follow-up study demonstrated that the fabrication of layer-by-layer assembled composites from SWNTs and laminin, an essential part of the human extracellular matrix, can be conducive to NSC differentiation and suitable for their successful excitation [247]. Extensive formation of functional neural networks was observed as demonstrated by the presence of synaptic connections, indicating that the protein–SWNT composite can serve as the material foundation of neural electrodes with chemical structure better adapted for long-term integration with the neural tissue. Additional studies to further understand the effect of electrical stimulus on the NSCs during the differentiation cycle and over a longer time frame are needed in the future.
20.9.2 Matrix Enhancement with SWNTs In tissue engineering, scaffolds play a pivotal role in providing temporary structural support, guiding cell growth, assisting nutrient/waste transport, and facilitating functional tissue/organ formation [248]. It is responsible for defining the space the engineered tissue occupies and aiding the process of tissue development. While popular synthetic polymers such as poly lactic acid (PLA) and co-poly lactic acid/glycolic acid (PLGA) have been widely used for tissue engineering, they lack the necessary mechanical strength [249, 250]. In addition, these polymers cannot be functionalized easily while SWNTs can.
P1: OTA/XYZ P2: ABC c20 JWBS037-Chen January 25, 2011
502
13:21
Printer Name: Yet to Come
BIOMEDICAL APPLICATIONS OF SINGLE-WALLED CARBON NANOTUBES
FIGURE 20.8 Applications of SWNTs in tissue engineering. (a) Confocal microscopy images of differentiated neurospheres grown on an SWNT composite. Neurospheres were stained for markers of NSCs (nestin), neurons (MAP2), astrocytes (GFAP), and oligodendrocytes (O4). The markers are shown in red and the cell nuclei are shown in blue. Scale bar: 20 m. (b) Representative microCT images of the rabbit femoral condyle 3 months after implantation of either a poly(propylene fumarate) (PPF) or US–SWNT scaffold. (Adapted from Jan and Kotov [246] and Sitharaman et al. [253].)
Live smooth muscle cell seeded collagen–SWNT composite matrices were investigated [251]. Cell viability in all constructs was consistently above 85% after 1 week, suggesting that such collagen–SWNT composite matrices can be used as scaffolds in tissue engineering or other medical devices. Highly porous scaffolds made of three different materials including SWNTs were fabricated to evaluate the effects of material composition and porosity on the scaffold pore structure, mechanical properties, and marrow stromal cell culture [252]. Based on the in vitro results, it was suggested that functionalized US SWNT nanocomposite scaffolds with tunable porosity and mechanical properties hold great promise for bone tissue engineering applications. Subsequently, the in vivo biocompatibility of US SWNT reinforced porous biodegradable scaffolds was tested in a rabbit model [253]. The US SWNT scaffolds, when implanted in rabbit femoral condyles or subcutaneous pockets, exhibited favorable hard and soft tissue responses at 1 and 3 months after implantation based on micro-computed tomography (microCT), histology, as well as histomorphometry (Fig. 20.8b). At 3 months, a threefold greater bone tissue ingrowth was seen in defects containing US SWNT nanocomposite scaffolds than the control scaffold, suggesting that these US SWNT scaffolds are very useful for assisting osteogenesis. The investigation of SWNTs for tissue engineering applications is still at the early development stage. As discussed in Section 20.2.3, studies have shown that enzymes can be attached to SWNTs and still retain enzymatic activity, which may allow a mixture of differently functionalized SWNTs to be added to a scaffold to create increasingly sophisticated structures. Potentially, enzymatic or protein functionalization, coupled with the
P1: OTA/XYZ P2: ABC c20 JWBS037-Chen January 25, 2011
13:21
Printer Name: Yet to Come
APPLICATIONS OF SWNTs IN BIOMEDICAL INSTRUMENTATION
503
electrochemical properties mentioned previously, may give rise to novel SWNT-based matrices that can not only provide the mechanical integrity for cell growth but can also monitor simultaneously the growth progress. 20.10 APPLICATIONS OF SWNTs IN BIOMEDICAL INSTRUMENTATION AFM is a promising technique for imaging nanoscale structures (e.g., proteins, nucleic acids, nanoparticles) with high resolution [254]. However, conventional silicon and silicon nitride AFM tips are not ideal for such applications. SWNTs, possessing small diameter, high aspect ratio, mechanical robustness, and unique chemical properties, have been investigated as AFM tips using a variety of assembly techniques such as CVD on the commercial silicon cantilever tip [255]. In an early study, SWNT-based AFM tips were used for multiplexed detection of polymorphic sites and direct determination of haplotypes in 10-kilobase sized DNA fragments [256], based on visual inspection of the AFM images of individual DNA molecules. This technology represents an important first step toward the use of nanotechnology in scanning probe microscopy in the biomedical arena. Attaching a smaller SWNT with subnanometer diameter may enable direct reading of the DNA sequence of these fragments. The high-resolution imaging capabilities of SWNT AFM probes have been demonstrated on a number of nanoparticles and biomolecules (Fig. 20.9) [254].
FIGURE 20.9 AFM with SWNTs as the tip. (a) A field emission/SEM image of SWNT bundles grown from an Si cantilever/tip assembly. (b) Detection of labeled DNA sites with SWNT tips. The arrow points to the streptavidin tag on the DNA. Scale bar: 100 nm. (c) The AFM image of a large scan area, a higher resolution image, and the crystal structure of the protein being imaged. (Adapted from Cheung et al. [255] and Woolley et al. [256].)
P1: OTA/XYZ P2: ABC c20 JWBS037-Chen January 25, 2011
504
13:21
Printer Name: Yet to Come
BIOMEDICAL APPLICATIONS OF SINGLE-WALLED CARBON NANOTUBES
Recently, a simple method to fabricate SWNT probes for AFM using the Langmuir–Blodgett (LB) technique was reported [257]. A two-step transfer process, dipping conventional AFM probes into the Langmuir monolayer followed by lifting the probes from the water surface, was able to assemble well-oriented and robust SWNTs that maintained their shape and direction even after successive scans. More importantly, it was also demonstrated that the LB method is a scalable process capable of simultaneously fabricating a large number of SWNT-modified nanoprobes. Carboxyl-terminated SWNTs have been immobilized onto the apexes of gold tips for STM [258], which gave high-resolution STM images of a diether monolayer formed on the graphite surface. In another study, it was shown that SWNTs at the apex of AFM probes can act as the template for the formation of nanoelectrodes through sputter coating with metal [259]. The resulting disk-shaped nanoelectrode at the AFM tip apex was capable of high-resolution electrochemical and topographical imaging. In addition to measuring topography, chemically functionalized SWNT-based probes can also measure the spatial arrangement of functional groups in a sample. Therefore functionalization of SWNT AFM tips offers great potential for high-resolution, chemically selective imaging of biological structures. Image-guided radiation therapy, with the goal of eradicating tumor cells while sparing healthy tissue, has been widely adopted in the clinic [260]. Image guidance at the time of treatment has been implemented using a variety of systems such as portal images, ultrasound devices, and CT scanners [261, 262]. With the exploding interest in small-animal-based research, there is an increasing need for such a system dedicated to small animals. Recently, CNT field emission technology was used to develop a micro-radiotherapy system for image-guided high precision irradiation of mice [263]. Through the field emission control of its individually addressable X-ray pixel beams, the micro-radiotherapy system can electronically shape the radiation field and form an intensity modulation pattern with temporal and spatial resolutions of approximately milliseconds and <2 mm, respectively.
20.11 OTHER APPLICATIONS OF SWNTs SWNTs have remained at the forefront of nanomaterial research since their discovery in the early 1990s, largely due to their exceptional and unusual mechanical, electrical, optical, and chemical properties. No other nanoparticle has been explored for such a diverse array of biomedical applications (Fig. 20.10). Besides the areas discussed above, SWNTs have also been investigated for certain other uses. One study showed that SWNTs can inhibit DNA duplex association and selectively induce human telomeric i-motif DNA formation by binding to the 5 -end major groove under physiological conditions [264]. This report suggested that SWNTs might have the intriguing potential to modulate human telomeric DNA structures in vivo, which may be useful for future drug design and anticancer therapy. The investigation of surface structure and characteristics of anthrax (Bacillus anthracis) spores as related to their binding by molecular species is very important to the development of countermeasure technologies for the detection and decontamination of anthrax spores [265]. SWNTs have been used as a scaffold for displaying multivalent monosaccharide ligands that bind effectively to anthrax spores with divalent cation mediation to cause significant spore aggregation [266]. Although optimization of such binding and the elucidation
P1: OTA/XYZ P2: ABC c20 JWBS037-Chen January 25, 2011
13:21
Printer Name: Yet to Come
TOXICITY OF SWNTs IN VITRO
505
FIGURE 20.10 The broad applications of SWNTs in the biomedical arena.
of related mechanistic details remain to be investigated, the aggregation of anthrax spores may be exploited for antibioterrorism applications.
20.12 TOXICITY OF SWNTs IN VITRO Like most new technologies, there are concerns about the possible side effects caused by the use of SWNTs. A large number of publications have appeared on this topic over the last several years. The potential toxicity of nanoparticles mainly comes from two aspects. First, nanoparticles can enter the body through the skin, lungs, or intestinal tract, depositing in several organs, and may cause adverse biological reactions [267]. In addition, the toxicity of nanoparticles also depends on whether they get cleared from the body and whether the host can raise an effective response to sequester or dispose of the particles [268]. Second, the toxicity can come from the material itself [269]. Although a few studies indicated that SWNTs exhibit little to no toxicity to cells [270, 271], many more reports have suggested that SWNTs do have adverse and/or toxic effects. However, the proposed mechanism for such effects varies dramatically, including increased oxidative stress and inhibition of cell proliferation due to NF-B activation [272], induction of cell apoptosis and decrease in cellular adhesion ability [273], inhibition of cell proliferation [274, 275], indirect cytotoxicity by alteration of the cell culture medium [276], induced oxidative stress and loss of antioxidants [277], expression of inducible stress responsive genes [278], as well as loss of cell viability [279]. A number of parameters were found to have considerable impact on the reactivity/toxicity of SWNTs. One study showed a dose- and time-dependent increase of intracellular reactive oxygen species and a decrease of the mitochondrial membrane potential when cells were treated with commercial SWNTs; however, incubation with purified SWNTs had no effect on the cells [280]. This report, along with a few other publications [281–283], suggested that the metal traces associated with commercial SWNTs are likely responsible for the adverse biological effects. Various studies have also suggested that the degree of SWNT functionalization [284], the diameter of the SWNTs [285, 286], the amount of adsorbed
P1: OTA/XYZ P2: ABC c20 JWBS037-Chen January 25, 2011
506
13:21
Printer Name: Yet to Come
BIOMEDICAL APPLICATIONS OF SINGLE-WALLED CARBON NANOTUBES
proteins on the SWNTs [287], as well as a combination of multiple factors [288] can affect the SWNT–cell interactions. To decipher the apparent contradictory findings of various literature reports, one study demonstrated that SWNTs interact with some tetrazolium salts (commonly used in cellbased assays) but not with the others [289]. This study brought the experimental data of many of the abovementioned reports to question. It was suggested that the cytotoxicity data of SWNTs should be verified with at least two or more independent test systems. Since SWNTs are known to also interact with many colorimetric indicator dyes frequently used in cytotoxicity assays, it was suggested that a clonogenic assay would be more accurate since no dye is used [290].
20.13 TOXICITY OF SWNTs IN RODENTS The toxicological effects of SWNTs have been studied in rodents. In an early study, pharyngeal aspiration of SWNTs elicited pulmonary effects in mice that combined acute inflammation with early onset yet progressive fibrosis and granulomas [291]. Functional respiratory deficiencies and decreased bacterial clearance were also observed in mice treated with SWNTs. A follow-up study showed that such enhanced acute inflammation and pulmonary injury with delayed bacterial clearance after SWNT exposure may lead to increased susceptibility to lung infection in mice [292]. A role for NADPH oxidase-derived reactive oxygen species in determining the course of the pulmonary response to SWNTs was then suggested [293]. Subsequently, the critical role of vitamin E (the major lipidsoluble antioxidant of the antioxidant protective system [294]) in SWNT-induced reactions was also proposed, because lowered levels of antioxidants in vitamin E-deficient mice were associated with higher sensitivity to SWNT-induced acute inflammation and enhanced profibrotic responses [295]. Since the pulmonary levels of vitamin E can be manipulated through diet, its effects on SWNT-induced inflammation may be of practical importance to the optimization of protective strategies. SWNT inhalation was found to be more effective than aspiration in causing inflammatory response, oxidative stress, collagen deposition, and fibrosis, as well as mutations of the K-ras gene locus in the mouse lungs [296]. Intratracheal instillation of SWNTs into mice caused induced alveolar macrophage activation, various chronic inflammatory responses, and severe pulmonary granuloma formation [297]. A series of signal transduction and biological events were suggested to be responsible for these effects. Interestingly, a subsequent report showed that exposure to more dispersed SWNT structures could alter the pulmonary distribution and response [298]. Pharyngeal aspiration of a dispersed preparation of SWNTs in mice caused no granulomatous lesions or epithelioid macrophages, confirming that granulomatous lesions encased by epithelioid macrophages were produced by large agglomerates in the previous intratracheal instillation or aspiration studies of SWNTs. A few other studies also reported that oral exposure or oropharyngeal aspiration of SWNTs did not exhibit significant toxic effect in mice [299, 300]. Most of the abovementioned studies used pristine SWNTs, which are irrelevant to the biomedical applications of SWNTs, in which the SWNTs would have to be functionalized for any potential use. Furthermore, SWNT-based agents will typically be intravenously injected for biomedical applications, rather than pharyngeal aspiration or inhalation. Certainly, SWNT itself is toxic to some extent just like any other nanoparticle. The most important question is not how toxic these “naked” nanoparticles are, but how to modify
P1: OTA/XYZ P2: ABC c20 JWBS037-Chen January 25, 2011
13:21
Printer Name: Yet to Come
TOXICITY OF SWNTs IN RODENTS
507
FIGURE 20.11 The distribution and clearance of SWNTs in mice. (a) Biodistribution of SWNTs functionalized with different sized PEG chains (SWNT-l-2kPEG, SWNT-l-5kPEG, and SWNT-br7kPEG) at 1 day postinjection as measured by Raman spectroscopy. (b) The concentrations of SWNTs retained in the liver and spleen of mice over a period of 3 months. Longer PEG chains on the SWNTs gave lower RES uptake. (c) No obvious toxic effect was seen in the liver and spleen of mice injected with SWNTs at 3 months postinjection. Scale bar: 50 m. (Adapted from Liu et al. [301].)
and functionalize them so that they do not exhibit any toxicity and thereby can be useful in biological systems. In a recent study, the long-term fate of functionalized SWNTs intravenously injected into animals was investigated [301]. Using the intrinsic Raman signatures of SWNTs, the blood circulation of intravenously injected SWNTs and their accumulation in various organs and tissues of mice over a period of 3 months was measured (Fig. 20.11). It was found that SWNTs functionalized by branched PEG chains enabled very long blood circulation up to 1 day, relatively low uptake in the RES, and virtually complete clearance from the main
P1: OTA/XYZ P2: ABC c20 JWBS037-Chen January 25, 2011
508
13:21
Printer Name: Yet to Come
BIOMEDICAL APPLICATIONS OF SINGLE-WALLED CARBON NANOTUBES
organs in about 2 months. Raman spectroscopy also detected SWNTs in the feces and other clearance organs, confirming the excretion and clearance of SWNTs from mice via the biliary and renal pathways. Most importantly, no toxic side effect of SWNTs in mice was observed in necropsy, histology, and blood chemistry measurements. Such a favorable in vivo toxicity profile of systematically administered, functionalized SWNTs was also confirmed by a number of other studies [302, 303].
20.14 TOXICITY OF SWNTs IN OTHER SPECIES Interestingly, the potential toxicity of SWNTs has been studied in a variety of other species since these SWNTs will eventually reach the ecosystem. In one study, acute and chronic toxicities of SWNTs were evaluated with the estuarine copepod Amphiascus tenuiremis [304]. It was found that copepods ingesting purified SWNTs exhibited no significant effect on mortality, development, and reproduction across the range of different exposure levels. However, exposure to the more complex commercial SWNTs (not purified) caused a significant increase in mortality, as well as a decrease in the fertilization rates and molting success. This study suggested that the small synthetic by-product fractions caused the increased mortality and delayed development rather than the SWNTs themselves, which corroborated with many of the previously discussed cell-based investigations [280–283]. Another study using zebrafish embryos also suggested that materials associated with raw SWNTs (likely the metal contaminants) have the potential to affect aquatic life when released into the aquatic environment while the SWNTs themselves had minimal toxicity [305]. SWNTs were found to be a respiratory toxicant in rainbow trout [306]. Although the fish were able to manage oxidative stress and osmoregulatory disturbances, pathological examination raised concerns about the cell cycle defects, neurotoxicity, and other unidentified blood-borne factors. A recent study using ciliated protozoan, widely used by ecotoxicologists because of its role in the regulation of microbial populations through the ingestion and digestion of bacteria, showed that SWNTs can be internalized, which may allow them to move up the food chain [307]. Ecotoxicity of SWNTs, along with many other nanoparticles, to aquatic organisms was studied, and the toxicity data generated on the various nanoparticles reflected a wide spectrum of sensitivity that was biological level, test, and endpoint specific [308]. Most importantly, SWNTs were found to be much less toxic than many other nanoparticles. A study on the effects of functionalized and nonfunctionalized SWNTs on the root elongation of six crop species was also carried out [309]. Not surprisingly, it was found that nonfunctionalized SWNTs affected the root length more than functionalized SWNTs.
20.15 FUTURE DIRECTIONS AND CHALLENGES Functionalized SWNTs have made far-reaching impact in the biomedical arena and it is expected that SWNTs will remain a major player in biomedical research for years to come (Fig. 20.12). With the capacity to provide enormous sensitivity, throughput, and flexibility, SWNTs have the potential to profoundly impact disease diagnosis and patient management in the near future. Big strides have been made and many proof-of-principle studies have been successfully performed, the future of SWNTs looks brighter than ever yet many
P1: OTA/XYZ P2: ABC c20 JWBS037-Chen January 25, 2011
13:21
Printer Name: Yet to Come
FUTURE DIRECTIONS AND CHALLENGES
509
FIGURE 20.12 A wide variety of functional moieties have been attached to SWNTs for a broad array of biomedical applications.
hurdles remain to be conquered. Much research effort in the following areas will be needed before clinical applications of SWNT-based agents are possible. 20.15.1 Multimodality Imaging and Early Detection of Diseases A combination of certain imaging modalities can offer synergistic advantages over any single modality alone [11, 207]. Dual-modality agents that combine PET, which is very sensitive and highly quantitative [310], and optical imaging, which can significantly facilitate ex vivo validation of the in vivo data, should be of particular interest for future SWNT-based biomedical research. The relatively large size of SWNTs may potentially allow for simultaneous multiple receptor binding of the targeting ligands on the same particle. Therefore targeting multiple, closely related receptors simultaneously may be more efficacious for molecular imaging and therapy, with both improved sensitivity and better specificity. We envision that peptides or small molecules are better targeting ligands than antibodies for SWNT conjugation, not only to keep the overall size small but also to take fully advantage of the polyvalency effect [311] since significantly more small molecules (MW < 1000) can be attached to an SWNT than the macromolecular antibodies (MW ∼150,000). Nanotechnology can have an early, paradigm-changing impact on how clinicians will detect diseases (e.g., cancer and cardiovascular diseases) in the earliest stage. One of the most pressing needs in clinical oncology is for imaging agents to detect tumors that are far smaller than what is possible with today’s technology, at a scale of 100,000 cells rather than 1,000,000,000 cells. Achieving this level of sensitivity requires better targeting efficacy of the imaging agents and signal amplification, both of which may be achieved with SWNTs. In terms of targeted imaging with SWNTs, vasculature targeting is more likely to succeed than targeting the tumor cells since SWNT-based agents may not extravasate well due to the relatively large size. Another area with near-term potential is the detection of mutations and genome instability in situ. Further work could result in an SWNT-based system capable of differentiating different types of tumors accurately and quickly, information that would be invaluable to clinicians and researchers alike. Implantable SWNT-based sensors may enable clinicians
P1: OTA/XYZ P2: ABC c20 JWBS037-Chen January 25, 2011
510
13:21
Printer Name: Yet to Come
BIOMEDICAL APPLICATIONS OF SINGLE-WALLED CARBON NANOTUBES
to more closely monitor the disease-free status of patients who have undergone treatment or individuals susceptible to the diseases because of various risk factors. SWNT-based nanosensors [17, 28] and in vivo imaging are both critical for future optimization of patient management. Ex vivo diagnostics in combination with in vivo diagnostics can provide a synergistic approach that neither strategy alone can offer. While imaging can give a whole-body perspective of the disease state in patients, detection of blood and urine markers of the disease can provide invaluable and complementary information of the biological responses to therapeutic intervention. Before treatment, after treatment, and more importantly during treatment, patient response can be evaluated by blood analysis and molecular imaging to ensure the accurate differentiation of responders from nonresponders and monitor their response to personalized therapy.
20.15.2 SWNT-Based Multifunctional Nanoplatform Because of their multifunctional capabilities, SWNTs can contain both targeting agents and therapeutic payloads. More importantly, they can be engineered to bypass a variety of biological barriers to enhance the targeting efficacy. The intrinsic physical properties of SWNTs can also offer the opportunity to utilize new approaches to therapy, such as localized heating. The future of nanomedicine lies in multifunctional nanoplatforms, which combine both therapeutic components and multimodality imaging. The ultimate goal is that nanoplatform-based agents can allow for efficient, specific in vivo delivery of therapeutic agents (drugs, genes, etc.) without systemic toxicity, and the dose delivered as well as the therapeutic efficacy can be accurately measured noninvasively over time. The versatile chemistry, as well as the unique physical properties, makes SWNT the most promising candidate for such a multifunctional nanoplatform for both multimodality molecular imaging and combination therapy (“mix-and-match” with suitably selected components for each individual application; Fig. 20.13). The most feasible applications of such nanoplatform-based agents will be in cardiovascular medicine, where there is much less biological barrier for the efficient delivery of nanoparticles, and in oncology, where the leaky tumor vasculature can allow for better tissue penetration than in normal organs/tissues.
20.15.3 Challenges Facing SWNT-Based Agents For in vitro and ex vivo applications, the advantages of state-of-the-art nanodevices (e.g., nanochips and nanosensors, including those based on SWNTs) over traditional assay methods are obvious [312, 313]. However, several barriers exist for in vivo applications of SWNT-based agents, among which are the biocompatibility, acute and chronic toxicity, pharmacokinetics and targeting efficacy, ability to escape the RES, and cost effectiveness. Recently, nanotoxicology has emerged as a new branch of toxicology for studying the undesirable effects of nanoparticles [314, 315]. Development of SWNT-based nanoplatforms for biomedical applications must proceed in tandem with the assessment of any toxicological side effects. The literature reports on SWNT-based toxicity studies clearly indicate that crude, nonfunctionalized SWNTs are toxic while pure, stably functionalized SWNTs are not. Thus, quality control of the SWNTs and robust chemistry on the SWNT surface are the two most important prerequisites for future biomedical and clinical applications of SWNTs. Further studies on well-characterized SWNTs are necessary to determine their safety profile as well as the environmental impact.
P1: OTA/XYZ P2: ABC c20 JWBS037-Chen January 25, 2011
13:21
Printer Name: Yet to Come
ACKNOWLEDGMENTS
511
FIGURE 20.13 A multifunctional SWNT-based nanoplatform incorporating multiple receptor targeting, multimodality imaging, and multiple therapeutic entities. Not all functional moieties will be necessary and only suitably selected components are needed for each individual application.
There are many commercial and regulatory challenges to be tackled with the emerging generation of more complex nanoparticles, in part owing to their multicomponent nature. SWNT-based multifunctional nanoplatforms will be difficult and expensive to manufacture on a large scale with appropriate quality, once they reach the stage of clinical investigation. However, some highly complex nanoparticles have reached the clinic for Phase I trials, indicating that complex nanoparticles can be manufactured with current good manufacturing practice (cGMP) and satisfy regulatory requirements, which is very encouraging [151]. On a nonscientific note, intellectual property-related issues may also make the eventually approved SWNT-based products quite expensive. The NCI Alliance for Nanotechnology in Cancer (http://nano.cancer.gov), a comprehensive, systematized initiative encompassing the public and private sectors, was formed in 2005 to accelerate the application of the best capabilities of nanotechnology to cancer. After establishing an interdisciplinary nanotechnology workforce, it is expected that nanotechnology will mature into a clinically useful field in the near future. Although this alliance is focused on cancer, the forefront of the battlefield against human diseases, the same techniques and agents can also be applied to a variety of other pathological disorders. After overcoming the abovementioned challenges, SWNTs have the potential to profoundly impact disease diagnosis and patient management in the future.
ACKNOWLEDGMENTS Funding was provided, in part, by the UW School of Medicine and Public Health’s Medical Education and Research Committee through the Wisconsin Partnership Program, the UW Carbone Cancer Center, and Susan G. Komen for the Cure. Ting Gao acknowledges support from Tyco Electronics.
P1: OTA/XYZ P2: ABC c20 JWBS037-Chen January 25, 2011
512
13:21
Printer Name: Yet to Come
BIOMEDICAL APPLICATIONS OF SINGLE-WALLED CARBON NANOTUBES
REFERENCES 1. Thayer, A. M. Building up nanotech research. Chem. Eng. News 2007, 85, 15–21. 2. Kawasaki, E. S.; Player, A. Nanotechnology, nanomedicine, and the development of new, effective therapies for cancer. Nanomedicine 2005, 1, 101–109. 3. Cai, W.; Hsu, A. R.; Li, Z. B.; Chen, X. Are quantum dots ready for in vivo imaging in human subjects? Nanoscale Res. Lett. 2007, 2, 265–281. 4. Michalet, X.; Pinaud, F. F.; Bentolila, L. A.; Tsay, J. M.; Doose, S.; Li, J. J.; Sundaresan, G.; Wu, A. M.; Gambhir, S. S.; Weiss, S. Quantum dots for live cells, in vivo imaging, and diagnostics. Science 2005, 307, 538–544. 5. Thorek, D. L.; Chen, A. K.; Czupryna, J.; Tsourkas, A. Superparamagnetic iron oxide nanoparticle probes for molecular imaging. Ann. Biomed. Eng. 2006, 34, 23–38. 6. Patolsky, F.; Zheng, G.; Lieber, C. M. Nanowire sensors for medicine and the life sciences. Nanomedicine 2006, 1, 51–65. 7. Lacerda, L.; Bianco, A.; Prato, M.; Kostarelos, K. Carbon nanotubes as nanomedicines: from toxicology to pharmacology. Adv. Drug Deliv. Rev. 2006, 58, 1460–1470. 8. Cai, W.; Gao, T.; Hong, H.; Sun, J. Applications of gold nanoparticles in cancer nanotechnology. Nanotechnol. Sci. and Applications 2008, 1, 17–32. 9. Hirsch, L. R.; Gobin, A. M.; Lowery, A. R.; Tam, F.; Drezek, R. A.; Halas, N. J.; West, J. L. Metal nanoshells. Ann. Biomed. Eng. 2006, 34, 15–22. 10. Ferrari, M. Cancer nanotechnology: opportunities and challenges. Nat. Rev. Cancer 2005, 5, 161–171. 11. Cai, W.; Chen, X. Nanoplatforms for targeted molecular imaging in living subjects. Small 2007, 3, 1840–1854. 12. Iijima, S. Helical microtubules of graphitic carbon. Nature 1991, 354, 56–58. 13. Iijima, S.; Ichihashi, T. Single-shell carbon nanotubes of 1-nm diameter. Nature 1993, 363, 603–605. 14. Bethune, D. S.; Kiang, C. H.; de Vries, M. S.; Gorman, G.; Savoy, R.; Vazquez, J.; Beyers, R. Cobalt-catalyzed growth of carbon nanotubes with single-atomic-layer walls. Nature 1993, 363, 605–607. 15. Yan, Y.; Chan-Park, M. B.; Zhang, Q. Advances in carbon-nanotube assembly. Small 2007, 3, 24–42. 16. Pillai, S. K.; Ray, S. S.; Moodley, M. Purification of single-walled carbon nanotubes. J. Nanosci. Nanotechnol. 2007, 7, 3011–3047. 17. Yun, Y. H.; Dong, Z.; Shanov, V.; Heineman, W. R.; Halsall, H. B.; Bhattacharya, A.; Laura Conforti, L.; Narayan, R. K.; Ball, W. S.; Schulz, M. J. Nanotube electrodes and biosensors. Nanotoday 2007, 2, 30–37. 18. Zhu, H.; Suenaga, K.; Hashimoto, A.; Urita, K.; Hata, K.; Iijima, S. Atomic-resolution imaging of the nucleation points of single-walled carbon nanotubes. Small 2005, 1, 1180–1183. 19. Porter, A. E.; Gass, M.; Muller, K.; Skepper, J. N.; Midgley, P. A.; Welland, M. Direct imaging of single-walled carbon nanotubes in cells. Nat. Nanotechnol. 2007, 2, 713–717. 20. Homma, Y.; Takagi, D.; Suzuki, S.; Kanzaki, K. I.; Kobayashi, Y. Electron-microscopic imaging of single-walled carbon nanotubes grown on silicon and silicon oxide substrates. J. Electron. Microsc. 2005, 54 (Suppl 1), i3–7. 21. Morgan, D. A.; Sloan, J.; Green, M. L. Direct imaging of o-carborane molecules within single walled carbon nanotubes. Chem. Commun. 2002, 2442–2443. 22. Liu, Z.; Yanagi, K.; Suenaga, K.; Kataura, H.; Iijima, S. Imaging the dynamic behaviour of individual retinal chromophores confined inside carbon nanotubes. Nat. Nanotechnol. 2007, 2, 422–425.
P1: OTA/XYZ P2: ABC c20 JWBS037-Chen January 25, 2011
13:21
Printer Name: Yet to Come
REFERENCES
513
23. Balasubramanian, K.; Burghard, M. Chemically functionalized carbon nanotubes. Small 2005, 1, 180–192. 24. Cinke, M.; Li, J.; Chen, B.; Cassell, A.; Delzeit, L.; Han, J.; Meyyappan, M. Pore structure of raw and purified HiPco single-walled carbon nanotubes. Chem. Phys. Lett. 2002, 365, 69–74. 25. Gruner, G. Carbon nanotube transistors for biosensing applications. Anal. Bioanal. Chem. 2006, 384, 322–335. 26. Kim, S. N.; Rusling, J. F.; Papadimitrakopoulos, F. Carbon nanotubes for electronic and electrochemical detection of biomolecules. Adv. Mater. 2007, 19, 3214–3228. 27. Wang, J. Nanomaterial-based electrochemical biosensors. Analyst 2005, 130, 421–426. 28. Sinha, N.; Ma, J.; Yeow, J. T. Carbon nanotube-based sensors. J. Nanosci. Nanotechnol. 2006, 6, 573–590. 29. Lee, C. Y.; Strano, M. S. Understanding the dynamics of signal transduction for adsorption of gases and vapors on carbon nanotube sensors. Langmuir 2005, 21, 5192–5196. 30. Snow, E. S.; Perkins, F. K.; Robinson, J. A. Chemical vapor detection using single-walled carbon nanotubes. Chem. Soc. Rev. 2006, 35, 790–798. 31. Saito, R.; Fujita, M.; Dresselhaus, G.; Dresselhaus, M. S. Electronic structure of chiral graphene tubules. Appl. Phys. Lett. 1992, 60, 2204–2206. 32. Baughman, R. H.; Zakhidov, A. A.; de Heer, W. A. Carbon nanotubes—the route toward applications. Science 2002, 297, 787–792. 33. Tans, S. J.; Devoret, M. H.; Dal, H.; Thess, A.; Smalley, R. E.; Geerligs, L. J.; Dekker, C. Individual single-wall carbon nanotubes as quantum wires. Nature 1997, 386, 474–477. 34. Bockrath, M.; Cobden, D. H.; McEuen, P. L.; Chopra, N. G.; Zettl, A.; Thess, A.; Smalley, R. E. Single-electron transport in ropes of carbon nanotubes. Science 1997, 275, 1922–1925. 35. Tans, S. J.; Verschueren, A. R. M.; Dekker, C. Room-temperature transistor based on a single carbon nanotube. Nature 1998, 393, 49–52. 36. Martel, R.; Schmidt, T.; Shea, H. R.; Hertel, T.; Avouris, P. Single- and multi-wall carbon nanotube field-effect transistors. Appl. Phys. Lett. 1998, 73, 2447–2449. 37. Kong, J.; Franklin, N. R.; Zhou, C.; Chapline, M. G.; Peng, S.; Cho, K.; Dai, H. Nanotube molecular wires as chemical sensors. Science 2000, 287, 622–625. 38. Collins, P. G.; Bradley, K.; Ishigami, M.; Zettl, A. Extreme oxygen sensitivity of electronic properties of carbon nanotubes. Science 2000, 287, 1801–1804. 39. Star, A.; Gabriel, J.-CP.; Bradley, K.; Gruener, G. Electronic detection of specific protein binding using nanotube FET devices. Nano Lett. 2003, 3, 459–463. 40. Cid, C. C.; Riu, J.; Maroto, A.; Rius, F. X. Carbon nanotube field effect transistors for the fast and selective detection of human immunoglobulin G. Analyst 2008, 133, 1005–1008. 41. Staii, C.; Johnson, A. T., Jr.; Chen, M.; Gelperin, A. DNA-decorated carbon nanotubes for chemical sensing. Nano Lett. 2005, 5, 1774–1778. 42. Kauffman, D. R.; Star, A. Electronically monitoring biological interactions with carbon nanotube field-effect transistors. Chem. Soc. Rev. 2008, 37, 1197–1206. 43. Heinze, S.; Tersoff, J.; Martel, R.; Derycke, V.; Appenzeller, J.; Avouris, P. Carbon nanotubes as schottky barrier transistors. Phys. Rev. Lett. 2002, 89, 106801. 44. Vitale, V.; Curioni, A.; Andreoni, W. Metal–carbon nanotube contacts: the link between Schottky barrier and chemical bonding. J. Am. Chem. Soc. 2008, 130, 5848–5849. 45. Chen, R. J.; Choi, H. C.; Bangsaruntip, S.; Yenilmez, E.; Tang, X.; Wang, Q.; Chang, Y. L.; Dai, H. An investigation of the mechanisms of electronic sensing of protein adsorption on carbon nanotube devices. J. Am. Chem. Soc. 2004, 126, 1563–1568. 46. Tang, X.; Bansaruntip, S.; Nakayama, N.; Yenilmez, E.; Chang, Y. L.; Wang, Q. Carbon nanotube DNA sensor and sensing mechanism. Nano Lett. 2006, 6, 1632–1636.
P1: OTA/XYZ P2: ABC c20 JWBS037-Chen January 25, 2011
514
13:21
Printer Name: Yet to Come
BIOMEDICAL APPLICATIONS OF SINGLE-WALLED CARBON NANOTUBES
47. Bradley, K.; Briman, M.; Star, A.; Gruener, G. Charge transfer from adsorbed proteins. Nano Lett. 2004, 4, 253–256. 48. Star, A.; Han, T.-R.; Gabriel, J.-C.P.; Bradley, K.; Gruener, G. Interaction of aromatic compounds with carbon nanotubes: correlation to the Hammett parameter of the substituent and measured carbon nanotube FET response. Nano Lett. 2003, 3, 1421–1423. 49. Zhou, X.; Moran-Mirabal, J. M.; Craighead, H. G.; McEuen, P. L. Supported lipid bilayer/carbon nanotube hybrids. Nat. Nanotechnol. 2007, 2, 185–190. 50. Park, D. W.; Kim, Y. H.; Kim, B. S.; So, H. M.; Won, K.; Lee, J. O.; Kong, K. J.; Chang, H. Detection of tumor markers using single-walled carbon nanotube field effect transistors. J. Nanosci. Nanotechnol. 2006, 6, 3499–3502. 51. Takeda, S.; Ozaki, H.; Hattori, S.; Ishii, A.; Kida, H.; Mukasa, K. Detection of influenza virus hemagglutinin with randomly immobilized anti-hemagglutinin antibody on a carbon nanotube sensor. J. Nanosci. Nanotechnol. 2007, 7, 752–756. 52. Wang, C. W.; Pan, C. Y.; Wu, H. C.; Shih, P. Y.; Tsai, C. C.; Liao, K. T.; Lu, L. L.; Hsieh, W. H.; Chen, C. D.; Chen, Y. T. In situ detection of chromogranin a released from living neurons with a single-walled carbon-nanotube field-effect transistor. Small 2007, 3, 1350–1355. 53. Tsai, C. C.; Yang, C. C.; Shih, P. Y.; Wu, C. S.; Chen, C. D.; Pan, C. Y.; Chen, Y. T. Exocytosis of a single bovine adrenal chromaffin cell: the electrical and morphological studies. J. Phys. Chem. B. 2008, 112, 9165–9173. 54. Someya, T.; Small, J.; Kim, P.; Nuckolls, C.; Yardley, J. T. Alcohol vapor sensors based on single-walled carbon nanotube field effect transistors. Nano Lett. 2003, 3, 877–881. 55. Star, A.; Joshi, V.; Han, T. R.; Altoe, M. V.; Gruner, G.; Stoddart, J. F. Electronic detection of the enzymatic degradation of starch. Org. Lett. 2004, 6, 2089–2092. 56. Lee, C. Y.; Strano, M. S. Amine basicity (pKb) controls the analyte binding energy on single walled carbon nanotube electronic sensor arrays. J. Am. Chem. Soc. 2008, 130, 1766– 1773. 57. Sayago, I.; Terrado, E.; Aleixandre, M.; Horrillo, M. C.; Fernandez, M. J.; Lozano, J.; Lafuente, E.; Maser, W. K.; Benito, A. M.; Martinez, M. T.; Gutierrez, J.; Munoz, E. Novel selective sensors based on carbon nanotube films for hydrogen detection. Sens Actuators B 2007, B122, 75–80. 58. Li, J.; Lu, Y.; Ng, H.; Han, J.; Meyyappan, M. Methane detection using palladium doped single-walled carbon nanotube sensors. Chem. Sens. 2004, 20, 710–711. 59. Zhang, T.; Nix, M. B.; Yoo, B.-Y.; Deshusses, M. A.; Myung, N. V. Electrochemically functionalized single-walled carbon nanotube gas sensor. Electroanalysis 2006, 18, 1153–1158. 60. Cattanach, K.; Kulkarni, R. D.; Kozlov, M.; Manohar, S. K. Flexible carbon nanotube sensors for nerve agent simulants. Nanotechnology 2006, 17, 4123–4128. 61. Bakker, E.; Qin, Y. Electrochemical sensors. Anal. Chem. 2006, 78, 3965–3984. 62. Privett, B. J.; Shin, J. H.; Schoenfisch, M. H. Electrochemical sensors. Anal. Chem. 2008, 80, 4499–4517. 63. Katz, E.; Willner, I. Integrated nanoparticle–biomolecule hybrid systems: synthesis, properties, and applications. Angew. Chem. Int. Ed. Engl. 2004, 43, 6042–6108. 64. Gooding, J. J. Nanostructuring electrodes with carbon nanotubes: a review on electrochemistry and applications for sensing. Electrochim. Acta. 2005, 50, 3049–3060. 65. Davis, J. J.; Coles, R. J.; Hill, H. A. O. Protein electrochemistry at carbon nanotube electrodes. J. Electroanal. Chem. 1997, 440, 279–282. 66. Wei, H.; Kim, S. N.; Marcus, H. L.; Papadimitrakopoulos, F. Preferential forest assembly of single-wall carbon nanotubes on low-energy electron-beam patterned Nafion films. Chem. Mater. 2006, 18, 1100–1106.
P1: OTA/XYZ P2: ABC c20 JWBS037-Chen January 25, 2011
13:21
Printer Name: Yet to Come
REFERENCES
515
67. Chattopadhyay, D.; Galeska, I.; Papadimitrakopoulos, F. Metal-assisted organization of shortened carbon nanotubes in monolayer and multilayer forest assemblies. J. Am. Chem. Soc. 2001, 123, 9451–9452. 68. Diao, P.; Liu, Z.; Wu, B.; Nan, X.; Zhang, J.; Wei, Z. Chemically assembled single-wall carbon nanotubes and their electrochemistry. Chemphyschem 2002, 3, 898–901. 69. Liu, Z.; Shen, Z.; Zhu, T.; Hou, S.; Ying, L.; Shi, Z.; Gu, Z. Organizing single-walled carbon nanotubes on gold using a wet chemical self-assembling technique. Langmuir 2000, 16, 3569–3573. 70. Cai, L.; Bahr, J. L.; Yao, Y.; Tour, J. M. Ozonation of single-walled carbon nanotubes and their assemblies on rigid self-assembled monolayers. Chem. Mater. 2002, 14, 4235–4241. 71. Du, F.; Huang, W.; Shi, Y.; Wang, Z.; Cheng, J. Real-time monitoring of NO release from single cells using carbon fiber microdisk electrodes modified with single-walled carbon nanotubes. Biosens. Bioelectron. 2008, 24, 415–421. 72. Swamy, B. E.; Venton, B. J. Carbon nanotube-modified microelectrodes for simultaneous detection of dopamine and serotonin in vivo. Analyst 2007, 132, 876–884. 73. Chen, R. S.; Huang, W. H.; Tong, H.; Wang, Z. L.; Cheng, J. K. Carbon fiber nanoelectrodes modified by single-walled carbon nanotubes. Anal. Chem. 2003, 75, 6341–6345. 74. Zhang, M.; Liu, K.; Gong, K.; Su, L.; Chen, Y.; Mao, L. Continuous on-line monitoring of extracellular ascorbate depletion in the rat striatum induced by global ischemia with carbon nanotube-modified glassy carbon electrode integrated into a thin-layer radial flow cell. Anal. Chem. 2005, 77, 6234–6242. 75. O’Connor, M.; Kim, S. N.; Killard, A. J.; Forster, R. J.; Smyth, M. R.; Papadimitrakopoulos, F.; Rusling, J. F. Mediated amperometric immunosensing using single walled carbon nanotube forests. Analyst 2004, 129, 1176–1180. 76. Wang, J.; Sun, X.; Cai, X.; Lei, Y.; Song, L.; Xie, S. Nonenzymatic glucose sensor using freestanding single-wall carbon nanotube films. Electrochem. Solid-State Lett. 2007, 10, J58– J60. 77. Yang, T.; Zhang, W.; Du, M.; Jiao, K. A PDDA/poly(2,6-pyridinedicarboxylic acid)-CNTs composite film DNA electrochemical sensor and its application for the detection of specific sequences related to PAT gene and NOS gene. Talanta 2008, 75, 987–994. 78. Yang, J.; Jiao, K.; Yang, T. A DNA electrochemical sensor prepared by electrodepositing zirconia on composite films of single-walled carbon nanotubes and poly(2,6-pyridinedicarboxylic acid), and its application to detection of the PAT gene fragment. Anal. Bioanal. Chem. 2007, 389, 913–921. 79. Wang, Y.; Wang, X.; Wu, B.; Zhao, Z.; Yin, F.; Li, S.; Qin, X.; Chen, Q. Dispersion of single-walled carbon nanotubes in poly(diallyldimethylammonium chloride) for preparation of a glucose biosensor. Sens. Actuators B 2008, B130, 809–815. 80. Wang, Y.; Joshi, P. P.; Hobbs, K. L.; Johnson, M. B.; Schmidtke, D. W. Nanostructured biosensors built by layer-by-layer electrostatic assembly of enzyme-coated single-walled carbon nanotubes and redox polymers. Langmuir 2006, 22, 9776–9783. 81. Li, Y.; Wang, P.; Wang, L.; Lin, X. Overoxidized polypyrrole film directed single-walled carbon nanotubes immobilization on glassy carbon electrode and its sensing applications. Biosens. Bioelectron 2007, 22, 3120–3125. 82. Granot, E.; Basnar, B.; Cheglakov, Z.; Katz, E.; Willner, I. Enhanced bioelectrocatalysis using single-walled carbon nanotubes (SWCNTs)/polyaniline hybrid systems in thin-film and microrod structures associated with electrodes. Electroanalysis 2006, 18, 26–34. 83. Yan, Y. M.; Yehezkeli, O.; Willner, I. Integrated, electrically contacted NAD(P)+ -dependent enzyme-carbon nanotube electrodes for biosensors and biofuel cell applications. Chemistry 2007, 13, 10168–10175.
P1: OTA/XYZ P2: ABC c20 JWBS037-Chen January 25, 2011
516
13:21
Printer Name: Yet to Come
BIOMEDICAL APPLICATIONS OF SINGLE-WALLED CARBON NANOTUBES
84. Ivnitski, D.; Artyushkova, K.; Rincon, R. A.; Atanassov, P.; Luckarift, H. R.; Johnson, G. R. Entrapment of enzymes and carbon nanotubes in biologically synthesized silica: glucose oxidase-catalyzed direct electron transfer. Small 2008, 4, 357–364. 85. Zhou, Y.; Yang, H.; Chen, H.-Y. Direct electrochemistry and reagentless biosensing of glucose oxidase immobilized on chitosan wrapped single-walled carbon nanotubes. Talanta 2008, 76, 419–423. 86. Sato, N.; Okuma, H. Development of single-wall carbon nanotubes modified screen-printed electrode using a ferrocene-modified cationic surfactant for amperometric glucose biosensor applications. Sens. Actuators B 2008, B129, 188–194. 87. Hrapovic, S.; Liu, Y.; Male, K. B.; Luong, J. H. Electrochemical biosensing platforms using platinum nanoparticles and carbon nanotubes. Anal. Chem. 2004, 76, 1083–1088. 88. Cheng, W.; Ding, L.; Lei, J.; Ding, S.; Ju, H. Effective cell capture with tetrapeptidefunctionalized carbon nanotubes and dual signal amplification for cytosensing and evaluation of cell surface carbohydrate. Anal. Chem. 2008, 80, 3867–3872. 89. Kerman, K.; Morita, Y.; Takamura, Y.; Tamiya, E. Escherichia coli single-strand binding proteinDNA interactions on carbon nanotube-modified electrodes from a label-free electrochemical hybridization sensor. Anal. Bioanal. Chem. 2005, 381, 1114–1121. 90. Hu, C.; Zhang, Y.; Bao, G.; Liu, M.; Wang, Z. L. DNA functionalized single-walled carbon nanotubes for electrochemical detection. J. Phys. Chem. B 2005, 109, 20072–20076. 91. Yang, X.; Lu, Y.; Ma, Y.; Liu, Z.; Du, F.; Chen, Y. DNA electrochemical sensor based on an adduct of single-walled carbon nanotubes and ferrocene. Biotechnol. Lett. 2007, 29, 1775– 1779. 92. Kum, M. C.; Joshi, K. A.; Chen, W.; Myung, N. V.; Mulchandani, A. Biomolecules–carbon nanotubes doped conducting polymer nanocomposites and their sensor application. Talanta 2007, 74, 370–375. 93. Sun, N.; Guan, L.; Shi, Z.; Li, N.; Gu, Z.; Zhu, Z.; Li, M.; Shao, Y. Ferrocene peapod modified electrodes: preparation, characterization, and mediation of H2 O2 . Anal. Chem. 2006, 78, 6050–6057. 94. Patolsky, F.; Weizmann, Y.; Willner, I. Long-range electrical contacting of redox enzymes by SWCNT connectors. Angew. Chem. Int. Ed. Engl. 2004, 43, 2113–2117. 95. Cai, Q. Y.; Zellers, E. T. Dual-chemiresistor GC detector employing monolayer-protected metal nanocluster interfaces. Anal. Chem. 2002, 74, 3533–3539. 96. Snow, E. S.; Perkins, F. K.; Houser, E. J.; Badescu, S. C.; Reinecke, T. L. Chemical detection with a single-walled carbon nanotube capacitor. Science 2005, 307, 1942–1945. 97. Robinson, J. A.; Snow, E. S.; Badescu, S. C.; Reinecke, T. L.; Perkins, F. K. Role of defects in single-walled carbon nanotube chemical sensors. Nano. Lett. 2006, 6, 1747–1751. 98. Kurosawa, S.; Park, J. W.; Aizawa, H.; Wakida, S.; Tao, H.; Ishihara, K. Quartz crystal microbalance immunosensors for environmental monitoring. Biosens. Bioelectron. 2006, 22, 473–481. 99. Drouvalakis, K. A.; Bangsaruntip, S.; Hueber, W.; Kozar, L. G.; Utz, P. J.; Dai, H. Peptidecoated nanotube-based biosensor for the detection of disease-specific autoantibodies in human serum. Biosens. Bioelectron. 2008, 23, 1413–1421. 100. Atashbar, M. Z.; Bejcek, B. E.; Singamaneni, S. Carbon nanotube network-based biomolecule detection. IEEE. Sens. J. 2006, 6, 524–528. 101. Penza, M.; Antolini, F.; Antisari, M. V. Carbon nanotubes as SAW chemical sensors materials. Sens. Actuators B 2004, B100, 47–59. 102. Penza, M.; Cassano, G.; Aversa, P.; Antolini, F.; Cusano, A.; Cutolo, A.; Giordano, M.; Nicolais, L. Alcohol detection using carbon nanotubes acoustic and optical sensors. Appl. Phys. Lett. 2004, 85, 2379–2381.
P1: OTA/XYZ P2: ABC c20 JWBS037-Chen January 25, 2011
13:21
Printer Name: Yet to Come
REFERENCES
517
103. Penza, M.; Cassano, G.; Aversa, P.; Antolini, F.; Cusano, A.; Consales, M.; Giordano, M.; Nicolais, L. Carbon nanotubes-coated multi-transducing sensors for VOCs detection. Sens. Actuators B 2005, B111-B112, 171–180. 104. Penza, M.; Cassano, G.; Aversa, P.; Cusano, A.; Cutolo, A.; Giordano, M.; Nicolais, L. Carbon nanotube acoustic and optical sensors for volatile organic compound detection. Nanotechnology 2005, 16, 2536–2547. 105. Penza, M.; Cassano, G.; Aversa, P.; Cusano, A.; Consales, M.; Giordano, M.; Nicolais, L. Acoustic and optical VOCs sensors incorporating carbon nanotubes. IEEE Sens. J. 2006, 6, 867–875. 106. Dharap, P.; Li, Z.; Nagarajaiah, S.; Barrera, E. V. Nanotube film based on single-wall carbon nanotubes for strain sensing. Nanotechnology 2004, 15, 379–382. 107. Wu, J.; Ji, Z.; Larade, B.; Guo, H.; Gong, X. G.; Liu, F. Computational design of carbon nanotube electromechanical pressure sensors. Phys. Rev. B: Condens. Matter Mater. Phys. 2004, 69, 153401–153406. 108. Stampfer, C.; Helbling, T.; Obergfell, D.; Schoberle, B.; Tripp, M. K.; Jungen, A.; Roth, S.; Bright, V. M.; Hierold, C. Fabrication of single-walled carbon-nanotube-based pressure sensors. Nano. Lett. 2006, 6, 233–237. 109. Ghosh, S.; Sood, A. K.; Kumar, N. Carbon nanotube flow sensors. Science 2003, 299, 1042–1044. 110. Xu, Y.; Pehrsson, P. E.; Chen, L.; Zhang, R.; Zhao, W. Double-stranded DNA single-walled carbon nanotube hybrids for optical hydrogen peroxide and glucose sensing. J. Phys. Chem. C. 2007, 111, 8638–8643. 111. Barone, P. W.; Parker, R. S.; Strano, M. S. In vivo fluorescence detection of glucose using a single-walled carbon nanotube optical sensor: design, fluorophore properties, advantages, and disadvantages. Anal. Chem. 2005, 77, 7556–7562. 112. Cusano, A.; Pisco, M.; Consales, M.; Cutolo, A.; Giordano, M.; Penza, M.; Aversa, P.; Capodieci, L.; Campopiano, S. Novel optochemical sensors based on hollow fibers and single walled carbon nanotubes. IEEE Photonics Technol. Lett. 2006, 18, 2431–2433. 113. Karachevtsev, V. A.; Glamazda, A. Y.; Leontiev, V. S.; Lytvyn, O. S.; Dettlaff-Weglikowska, U. Glucose sensing based on NIR fluorescence of DNA-wrapped single-walled carbon nanotubes. Chem. Phys. Lett. 2007, 435, 104–108. 114. Byon, H. R.; Hong, B. J.; Gho, Y. S.; Park, J. W.; Choi, H. C. Pseudo 3D single-walled carbon nanotube film for BSA-free protein chips. Chembiochem 2005, 6, 1331–1334. 115. McCormick, F. Cancer gene therapy: fringe or cutting edge? Nat. Rev. Cancer 2001, 1, 130– 141. 116. Yi, Y.; Hahm, S. H.; Lee, K. H. Retroviral gene therapy: safety issues and possible solutions. Curr. Gene Ther. 2005, 5, 25–35. 117. Pantarotto, D.; Singh, R.; McCarthy, D.; Erhardt, M.; Briand, J. P.; Prato, M.; Kostarelos, K.; Bianco, A. Functionalized carbon nanotubes for plasmid DNA gene delivery. Angew. Chem. Int. Ed. Engl. 2004, 43, 5242–5246. 118. Eliyahu, H.; Barenholz, Y.; Domb, A. J. Polymers for DNA delivery. Molecules 2005, 10, 34–64. 119. Patil, S. D.; Rhodes, D. G.; Burgess, D. J. DNA-based therapeutics and DNA delivery systems: a comprehensive review. AAPS J. 2005, 7, E61–77. 120. Singh, R.; Pantarotto, D.; McCarthy, D.; Chaloin, O.; Hoebeke, J.; Partidos, C. D.; Briand, J. P.; Prato, M.; Bianco, A.; Kostarelos, K. Binding and condensation of plasmid DNA onto functionalized carbon nanotubes: toward the construction of nanotube-based gene delivery vectors. J. Am. Chem. Soc. 2005, 127, 4388–4396.
P1: OTA/XYZ P2: ABC c20 JWBS037-Chen January 25, 2011
518
13:21
Printer Name: Yet to Come
BIOMEDICAL APPLICATIONS OF SINGLE-WALLED CARBON NANOTUBES
121. Kam, N. W.; Liu, Z.; Dai, H. Carbon nanotubes as intracellular transporters for proteins and DNA: an investigation of the uptake mechanism and pathway. Angew. Chem. Int. Ed. Engl. 2006, 45, 577–581. 122. Shi Kam, N. W.; Jessop, T. C.; Wender, P. A.; Dai, H. Nanotube molecular transporters: internalization of carbon nanotube-protein conjugates into Mammalian cells. J. Am. Chem. Soc. 2004, 126, 6850–6851. 123. Gleave, M. E.; Monia, B. P. Antisense therapy for cancer. Nat. Rev. Cancer 2005, 5, 468–479. 124. Soucek, L.; Evan, G. Myc—Is this the oncogene from Hell? Cancer Cell 2002, 1, 406–408. 125. Cui, D.; Tian, F.; Coyer, S. R.; Wang, J.; Pan, B.; Gao, F.; He, R.; Zhang, Y. Effects of antisensemyc-conjugated single-walled carbon nanotubes on HL-60 cells. J. Nanosci. Nanotechnol. 2007, 7, 1639–1646. 126. Zhao, C.; Peng, Y.; Song, Y.; Ren, J.; Qu, X. Self-assembly of single-stranded RNA on carbon nanotube: polyadenylic acid to form a duplex structure. Small 2008, 4, 656–661. 127. Fire, A.; Xu, S.; Montgomery, M. K.; Kostas, S. A.; Driver, S. E.; Mello, C. C. Potent and specific genetic interference by double-stranded RNA in Caenorhabditis elegans. Nature 1998, 391, 806–811. 128. Stevenson, M. Therapeutic potential of RNA interference. N. Engl. J. Med. 2004, 351, 1772–1777. 129. Mello, C. C.; Conte, D, Jr. Revealing the world of RNA interference. Nature 2004, 431, 338– 342. 130. Kam, N. W.; Liu, Z.; Dai, H. Functionalization of carbon nanotubes via cleavable disulfide bonds for efficient intracellular delivery of siRNA and potent gene silencing. J. Am. Chem. Soc. 2005, 127, 12492–12493. 131. Liu, Z.; Winters, M.; Holodniy, M.; Dai, H. siRNA delivery into human T cells and primary cells with carbon-nanotube transporters. Angew. Chem. Int. Ed. Engl. 2007, 46, 2023–2027. 132. Zhang, Z.; Yang, X.; Zhang, Y.; Zeng, B.; Wang, S.; Zhu, T.; Roden, R. B.; Chen, Y.; Yang, R. Delivery of telomerase reverse transcriptase small interfering RNA in complex with positively charged single-walled carbon nanotubes suppresses tumor growth. Clin. Cancer Res. 2006, 12, 4933–4939. 133. Krajcik, R.; Jung, A.; Hirsch, A.; Neuhuber, W.; Zolk, O. Functionalization of carbon nanotubes enables non-covalent binding and intracellular delivery of small interfering RNA for efficient knock-down of genes. Biochem. Biophys. Res. Commun. 2008, 369, 595–602. 134. Wang, X.; Ren, J.; Qu, X. Targeted RNA interference of cyclin A2 mediated by functionalized single-walled carbon nanotubes induces proliferation arrest and apoptosis in chronic myelogenous leukemia K562 cells. ChemMedChem 2008, 3, 940–945. 135. Lee, B. J.; Cansizoglu, A. E.; Suel, K. E.; Louis, T. H.; Zhang, Z.; Chook, Y. M. Rules for nuclear localization sequence recognition by karyopherin beta 2. Cell 2006, 126, 543–558. 136. Sturzu, A.; Heckl, S. The fluorinated and chlorinated nuclear localization sequence of the SV 40 T antigen. Chem. Biol. Drug Des. 2009, 73, 127–131. 137. Kam, N. W.; Dai, H. Carbon nanotubes as intracellular protein transporters: generality and biological functionality. J. Am. Chem. Soc. 2005, 127, 6021–6026. 138. Zeineldin, R.; Al-Haik, M.; Hudson, L. G. Role of polyethylene glycol integrity in specific receptor targeting of carbon nanotubes to cancer cells. Nano Lett. 2009, 9, 751–757. 139. Strano, M. S.; Moore, V. C.; Miller, M. K.; Allen, M. J.; Haroz, E. H.; Kittrell, C.; Hauge, R. H.; Smalley, R. E. The role of surfactant adsorption during ultrasonication in the dispersion of single-walled carbon nanotubes. J. Nanosci. Nanotechnol. 2003, 3, 81–86. 140. Liu, Z.; Sun, X.; Nakayama-Ratchford, N.; Dai, H. Supramolecular chemistry on water-soluble carbon nanotubes for drug loading and delivery. ACS Nano. 2007, 1, 50–56.
P1: OTA/XYZ P2: ABC c20 JWBS037-Chen January 25, 2011
13:21
Printer Name: Yet to Come
REFERENCES
519
141. Harris, J. M.; Chess, R. B. Effect of pegylation on pharmaceuticals. Nat. Rev. Drug Discov. 2003, 2, 214–221. 142. Feazell, R. P.; Nakayama-Ratchford, N.; Dai, H.; Lippard, S. J. Soluble single-walled carbon nanotubes as longboat delivery systems for platinum(IV) anticancer drug design. J. Am. Chem. Soc. 2007, 129, 8438–8439. 143. Dhar, S.; Liu, Z.; Thomale, J.; Dai, H.; Lippard, S. J. Targeted single-wall carbon nanotubemediated Pt(IV) prodrug delivery using folate as a homing device. J. Am. Chem. Soc. 2008, 130, 11467–11476. 144. Chen, J.; Chen, S.; Zhao, X.; Kuznetsova, L. V.; Wong, S. S.; Ojima, I. Functionalized singlewalled carbon nanotubes as rationally designed vehicles for tumor-targeted drug delivery. J. Am. Chem. Soc. 2008, 130, 16778–16785. 145. Liu, Z.; Chen, K.; Davis, C.; Sherlock, S.; Cao, Q.; Chen, X.; Dai, H. Drug delivery with carbon nanotubes for in vivo cancer treatment. Cancer Res. 2008, 68, 6652–6660. 146. Casazza, A. M.; Fairchild, C. R. Paclitaxel (Taxol) mechanisms of resistance. Cancer Treat. Res. 1996, 87, 149–171. 147. Maeda, H.; Wu, J.; Sawa, T.; Matsumura, Y.; Hori, K. Tumor vascular permeability and the EPR effect in macromolecular therapeutics. A review. J. Control. Release 2000, 65, 271–284. 148. Fadel, T. R.; Steenblock, E. R.; Stern, E.; Li, N.; Wang, X.; Haller, G. L.; Pfefferle, L. D.; Fahmy, T. M. Enhanced cellular activation with single walled carbon nanotube bundles presenting antibody stimuli. Nano Lett. 2008, 8, 2070–2076. 149. Cato, M. H.; D’Annibale, F.; Mills, D. M.; Cerignoli, F.; Dawson, M. I.; Bergamaschi, E.; Bottini, N.; Magrini, A.; Bergamaschi, A.; Rosato, N.; Rickert, R. C.; Mustelin, T.; Bottini, M. Celltype specific and cytoplasmic targeting of PEGylated carbon nanotube-based nanoassemblies. J. Nanosci. Nanotechnol. 2008, 8, 2259–2269. 150. Peer, D.; Karp, J. M.; Hong, S.; Farokhzad, O. C.; Margalit, R.; Langer, R. Nanocarriers as an emerging platform for cancer therapy. Nat. Nanotechnol. 2007, 2, 751–760. 151. Davis, M. E.; Chen, Z. G.; Shin, D. M. Nanoparticle therapeutics: an emerging treatment modality for cancer. Nat. Rev. Drug Discov. 2008, 7, 771–782. 152. Pantarotto, D.; Partidos, C. D.; Graff, R.; Hoebeke, J.; Briand, J. P.; Prato, M.; Bianco, A. Synthesis, structural characterization, and immunological properties of carbon nanotubes functionalized with peptides. J. Am. Chem. Soc. 2003, 125, 6160–6164. 153. Pantarotto, D.; Partidos, C. D.; Hoebeke, J.; Brown, F.; Kramer, E.; Briand, J. P.; Muller, S.; Prato, M.; Bianco, A. Immunization with peptide-functionalized carbon nanotubes enhances virus-specific neutralizing antibody responses. Chem. Biol. 2003, 10, 961–966. 154. Bianco, A.; Hoebeke, J.; Godefroy, S.; Chaloin, O.; Pantarotto, D.; Briand, J. P.; Muller, S.; Prato, M.; Partidos, C. D. Cationic carbon nanotubes bind to CpG oligodeoxynucleotides and enhance their immunostimulatory properties. J. Am. Chem. Soc. 2005, 127, 58–59. 155. Kam, N. W.; O’Connell, M.; Wisdom, J. A.; Dai, H. Carbon nanotubes as multifunctional biological transporters and near-infrared agents for selective cancer cell destruction. Proc. Natl. Acad. Sci. U.S.A. 2005, 102, 11600–11605. 156. Kim, J. W.; Shashkov, E. V.; Galanzha, E. I.; Kotagiri, N.; Zharov, V. P. Photothermal antimicrobial nanotherapy and nanodiagnostics with self-assembling carbon nanotube clusters. Lasers Surg. Med. 2007, 39, 622–634. 157. Chakravarty, P.; Marches, R.; Zimmerman, N. S.; Swafford, A. D.; Bajaj, P.; Musselman, I. H.; Pantano, P.; Draper, R. K.; Vitetta, E. S. Thermal ablation of tumor cells with antibodyfunctionalized single-walled carbon nanotubes. Proc. Natl. Acad. Sci. U.S.A. 2008, 105, 8697–8702. 158. Gannon, C. J.; Cherukuri, P.; Yakobson, B. I.; Cognet, L.; Kanzius, J. S.; Kittrell, C.; Weisman, R. B.; Pasquali, M.; Schmidt, H. K.; Smalley, R. E.; Curley, S. A. Carbon nanotube-enhanced
P1: OTA/XYZ P2: ABC c20 JWBS037-Chen January 25, 2011
520
159.
160.
161.
162. 163. 164.
165.
166.
167. 168. 169.
170. 171. 172. 173.
174.
175. 176.
13:21
Printer Name: Yet to Come
BIOMEDICAL APPLICATIONS OF SINGLE-WALLED CARBON NANOTUBES
thermal destruction of cancer cells in a noninvasive radiofrequency field. Cancer 2007, 110, 2654–2665. Hirsch, L. R.; Stafford, R. J.; Bankson, J. A.; Sershen, S. R.; Rivera, B.; Price, R. E.; Hazle, J. D.; Halas, N. J.; West, J. L. Nanoshell-mediated near-infrared thermal therapy of tumors under magnetic resonance guidance. Proc. Natl. Acad. Sci. U.S.A. 2003, 100, 13549–13554. DeNardo, S. J.; DeNardo, G. L.; Miers, L. A.; Natarajan, A.; Foreman, A. R.; Gruettner, C.; Adamson, G. N.; Ivkov, R. Development of tumor targeting bioprobes 111 In-chimeric L6 monoclonal antibody nanoparticles for alternating magnetic field cancer therapy. Clin. Cancer Res. 2005, 11, 7087s–7092s. DeNardo, S. J.; DeNardo, G. L.; Natarajan, A.; Miers, L. A.; Foreman, A. R.; Gruettner, C.; Adamson, G. N.; Ivkov, R. Thermal dosimetry predictive of efficacy of 111 In-ChL6 nanoparticle AMF–induced thermoablative therapy for human breast cancer in mice. J. Nucl. Med. 2007, 48, 437–444. Mankoff, D. A. A definition of molecular imaging. J. Nucl. Med. 2007, 48, 18N, 21N. Massoud, T. F.; Gambhir, S. S. Molecular imaging in living subjects: seeing fundamental biological processes in a new light. Genes Dev. 2003, 17, 545–580. Beyer, T.; Townsend, D. W.; Brun, T.; Kinahan, P. E.; Charron, M.; Roddy, R.; Jerin, J.; Young, J.; Byars, L.; Nutt, R. A combined PET/CT scanner for clinical oncology. J. Nucl. Med. 2000, 41, 1369–1379. Even-Sapir, E.; Lerman, H.; Lievshitz, G.; Khafif, A.; Fliss, D. M.; Schwartz, A.; Gur, E.; Skornick, Y.; Schneebaum, S. Lymphoscintigraphy for sentinel node mapping using a hybrid SPECT/CT system. J. Nucl. Med. 2003, 44, 1413–1420. Catana, C.; Wu, Y.; Judenhofer, M. S.; Qi, J.; Pichler, B. J.; Cherry, S. R. Simultaneous acquisition of multislice PET and MR images: initial results with a MR-compatible PET scanner. J. Nucl. Med. 2006, 47, 1968–1976. Rudin, M.; Weissleder, R. Molecular imaging in drug discovery and development. Nat. Rev. Drug Discov. 2003, 2, 123–131. Cai, W.; Rao, J.; Gambhir, S. S.; Chen, X. How molecular imaging is speeding up anti-angiogenic drug development. Mol. Cancer Ther. 2006, 5, 2624–2633. Pathak, A. P.; Gimi, B.; Glunde, K.; Ackerstaff, E.; Artemov, D.; Bhujwalla, Z. M. Molecular and functional imaging of cancer: advances in MRI and MRS. Methods Enzymol. 2004, 386, 3–60. Zhang, Z.; Nair, S. A.; McMurry, T. J. Gadolinium meets medicinal chemistry: MRI contrast agent development. Curr. Med. Chem. 2005, 12, 751–778. Pautler, R. G.; Fraser, S. E. The year(s) of the contrast agent—micro-MRI in the new millennium. Curr. Opin. Immunol. 2003, 15, 385–392. de Roos, A.; Doornbos, J.; Baleriaux, D.; Bloem, H. L.; Falke, T. H. Clinical applications of gadolinium-DTPA in MRI. Magn. Reson. Annu. 1988, 113–145. Sitharaman, B.; Kissell, K. R.; Hartman, K. B.; Tran, L. A.; Baikalov, A.; Rusakova, I.; Sun, Y.; Khant, H. A.; Ludtke, S. J.; Chiu, W.; Laus, S.; Toth, E.; Helm, L.; Merbach, A. E.; Wilson, L. J. Superparamagnetic gadonanotubes are high-performance MRI contrast agents. Chem. Commun. 2005, 3915–3917. Hartman, K. B.; Laus, S.; Bolskar, R. D.; Muthupillai, R.; Helm, L.; Toth, E.; Merbach, A. E.; Wilson, L. J. Gadonanotubes as ultrasensitive pH-smart probes for magnetic resonance imaging. Nano Lett. 2008, 8, 415–419. Didenko, V. V.; Moore, V. C.; Baskin, D. S.; Smalley, R. E. Visualization of individual singlewalled carbon nanotubes by fluorescent polymer wrapping. Nano Lett. 2005, 5, 1563–1567. Bottini, M.; Cerignoli, F.; Dawson, M. I.; Magrini, A.; Rosato, N.; Mustelin, T. Fulllength single-walled carbon nanotubes decorated with streptavidin-conjugated quantum dots
P1: OTA/XYZ P2: ABC c20 JWBS037-Chen January 25, 2011
13:21
Printer Name: Yet to Come
REFERENCES
177. 178. 179.
180. 181.
182.
183. 184. 185.
186. 187. 188.
189.
190. 191.
192. 193. 194. 195.
521
as multivalent intracellular fluorescent nanoprobes. Biomacromolecules 2006, 7, 2259– 2263. Cheong, W. F.; Prahl, S. A.; Welch, A. J. A review of the optical properties of biological tissues. IEEE J. 1990, 26, 2166–2185. Frangioni, J. V. In vivo near-infrared fluorescence imaging. Curr. Opin. Chem. Biol. 2003, 7, 626–634. Cai, W.; Shin, D. W.; Chen, K.; Gheysens, O.; Cao, Q.; Wang, S. X.; Gambhir, S. S.; Chen, X. Peptide-labeled near-infrared quantum dots for imaging tumor vasculature in living subjects. Nano Lett. 2006, 6, 669–676. Lefebvre, J.; Austing, D. G.; Bond, J.; Finnie, P. Photoluminescence imaging of suspended single-walled carbon nanotubes. Nano Lett. 2006, 6, 1603–1608. Duque, J. G.; Cognet, L.; Parra-Vasquez, A. N.; Nicholas, N.; Schmidt, H. K.; Pasquali, M. Stable luminescence from individual carbon nanotubes in acidic, basic, and biological environments. J. Am. Chem. Soc. 2008, 130, 2626–2633. Cherukuri, P.; Gannon, C. J.; Leeuw, T. K.; Schmidt, H. K.; Smalley, R. E.; Curley, S. A.; Weisman, R. B. Mammalian pharmacokinetics of carbon nanotubes using intrinsic near-infrared fluorescence. Proc. Natl. Acad. Sci. U.S.A. 2006, 103, 18882–18886. Carlson, L. J.; Krauss, T. D. Photophysics of individual single-walled carbon nanotubes. Acc. Chem. Res. 2008, 41, 235–243. Heller, D. A.; Baik, S.; Eurell, T. E.; Strano, M. S. Single-walled carbon nanotube spectroscopy in live cells: towards long-term labels and optical sensors. Adv. Mater. 2005, 17, 2793–2799. Cherukuri, P.; Bachilo, S. M.; Litovsky, S. H.; Weisman, R. B. Near-infrared fluorescence microscopy of single-walled carbon nanotubes in phagocytic cells. J. Am. Chem. Soc. 2004, 126, 15638–15639. Welsher, K.; Liu, Z.; Daranciang, D.; Dai, H. Selective probing and imaging of cells with single walled carbon nanotubes as near-infrared fluorescent molecules. Nano Lett. 2008, 8, 586–590. Jin, H.; Heller, D. A.; Strano, M. S. Single-particle tracking of endocytosis and exocytosis of single-walled carbon nanotubes in NIH-3T3 cells. Nano Lett. 2008, 8, 1577–1585. Leeuw, T. K.; Reith, R. M.; Simonette, R. A.; Harden, M. E.; Cherukuri, P.; Tsyboulski, D. A.; Beckingham, K. M.; Weisman, R. B. Single-walled carbon nanotubes in the intact organism: near-IR imaging and biocompatibility studies in Drosophila. Nano Lett. 2007, 7, 2650–2654. Choi, J. H.; Nguyen, F. T.; Barone, P. W.; Heller, D. A.; Moll, A. E.; Patel, D.; Boppart, S. A.; Strano, M. S. Multimodal biomedical imaging with asymmetric single-walled carbon nanotube/iron oxide nanoparticle complexes. Nano Lett. 2007, 7, 861–867. Cognet, L.; Tsyboulski, D. A.; Weisman, R. B. Subdiffraction far-field imaging of luminescent single-walled carbon nanotubes. Nano Lett. 2008, 8, 749–753. Bussy, C.; Cambedouzou, J.; Lanone, S.; Leccia, E.; Heresanu, V.; Pinault, M.; MayneL’hermite, M.; Brun, N.; Mory, C.; Cotte, M.; Doucet, J.; Boczkowski, J.; Launois, P. Carbon nanotubes in macrophages: imaging and chemical analysis by X-ray fluorescence microscopy. Nano Lett. 2008, 8, 2659–2663. Cao, Y. C.; Jin, R.; Mirkin, C. A. Nanoparticles with Raman spectroscopic fingerprints for DNA and RNA detection. Science 2002, 297, 1536–1540. Sun, L.; Yu, C.; Irudayaraj, J. Surface-enhanced Raman scattering based nonfluorescent probe for multiplex DNA detection. Anal. Chem. 2007, 79, 3981–3988. Jorio, A.; Saito, R.; Dresselhaus, G.; Dresselhaus, M. S. Determination of nanotubes properties by Raman spectroscopy. Philos. Trans. A Math. Phys. Eng. Sci. 2004, 362, 2311–2336. Zhang, L.; Aite, S.; Yu, Z. Unique laser-scanning optical microscope for low-temperature imaging and spectroscopy. Rev. Sci. Instrum. 2007, 78, 083701.
P1: OTA/XYZ P2: ABC c20 JWBS037-Chen January 25, 2011
522
13:21
Printer Name: Yet to Come
BIOMEDICAL APPLICATIONS OF SINGLE-WALLED CARBON NANOTUBES
196. Zhang, C.; Abdijalilov, K.; Grebel, H. Surface enhanced Raman with anodized aluminum oxide films. J. Chem. Phys. 2007, 127, 044701. 197. Anderson, N.; Hartschuh, A.; Cronin, S.; Novotny, L. Nanoscale vibrational analysis of singlewalled carbon nanotubes. J. Am. Chem. Soc. 2005, 127, 2533–2537. 198. Hartschuh, A.; Qian, H.; Meixner, A. J.; Anderson, N.; Novotny, L. Nanoscale optical imaging of excitons in single-walled carbon nanotubes. Nano Lett. 2005, 5, 2310–2313. 199. Hartschuh, A.; Sanchez, E. J.; Xie, X. S.; Novotny, L. High-resolution near-field Raman microscopy of single-walled carbon nanotubes. Phys. Rev. Lett. 2003, 90, 095503. 200. Atalay, H.; Lefrant, S. Analysing one isolated single walled carbon nanotube in the near-field domain with selective nanovolume Raman spectroscopy. J. Nanosci. Nanotechnol. 2004, 4, 749–761. 201. Doorn, S. K.; Zheng, L.; O’Connell, M. J.; Zhu, Y.; Huang, S.; Liu, J. Raman spectroscopy and imaging of ultralong carbon nanotubes. J. Phys. Chem. B 2005, 109, 3751–3758. 202. Itkis, M. E.; Perea, D. E.; Jung, R.; Niyogi, S.; Haddon, R. C. Comparison of analytical techniques for purity evaluation of single-walled carbon nanotubes. J. Am. Chem. Soc. 2005, 127, 3439–3448. 203. Keren, S.; Zavaleta, C.; Cheng, Z.; de la Zerda, A.; Gheysens, O.; Gambhir, S. S. Noninvasive molecular imaging of small living subjects using Raman spectroscopy. Proc. Natl. Acad. Sci. U.S.A. 2008, 105, 5844–5849. 204. Zavaleta, C.; de la Zerda, A.; Liu, Z.; Keren, S.; Cheng, Z.; Schipper, M.; Chen, X.; Dai, H.; Gambhir, S. S. Noninvasive Raman spectroscopy in living mice for evaluation of tumor targeting with carbon nanotubes. Nano Lett. 2008, 8, 2800–2805. 205. Cai, W.; Chen, X. Anti-angiogenic cancer therapy based on integrin avb3 antagonism. AntiCancer Agents Med. Chem. 2006, 6, 407–428. 206. Cai, W.; Chen, X. Multimodality molecular imaging of tumor angiogenesis. J. Nucl. Med. 2008, 49(Suppl 2), 113S–128S. 207. Cai, W.; Niu, G.; Chen, X. Imaging of integrins as biomarkers for tumor angiogenesis. Curr. Pharm. Des. 2008, 14, 2943–2973. 208. Liu, Z.; Li, X.; Tabakman, S. M.; Jiang, K.; Fan, S.; Dai, H. Multiplexed multicolor Raman imaging of live cells with isotopically modified single walled carbon nanotubes. J. Am. Chem. Soc. 2008, 130, 13540–13541. 209. Chen, Z.; Tabakman, S. M.; Goodwin, A. P.; Kattah, M. G.; Daranciang, D.; Wang, X.; Zhang, G.; Li, X.; Liu, Z.; Utz, P. J.; Jiang, K.; Fan, S.; Dai, H. Protein microarrays with carbon nanotubes as multicolor Raman labels. Nat. Biotechnol. 2008, 26, 1285–1292. 210. Rao, A. M.; Richter, E.; Bandow, S.; Chase, B.; Eklund, P. C.; Williams, K. A.; Fang, S.; Subbaswamy, K. R.; Menon, M.; Thess, A.; Smalley, R. E.; Dresselhaus, G.; Dresselhaus, M. S. Diameter-selective Raman scattering from vibrational modes in carbon nanotubes. Science 1997, 275, 187–191. 211. Wang, L. V. Prospects of photoacoustic tomography. Med. Phys. 2008, 35, 5758–5767. 212. Kim, G.; Huang, S. W.; Day, K. C.; O’Donnell, M.; Agayan, R. R.; Day, M. A.; Kopelman, R.; Ashkenazi, S. Indocyanine-green-embedded PEBBLEs as a contrast agent for photoacoustic imaging. J. Biomed. Opt. 2007, 12, 044020. 213. Ku, G.; Wang, L. V. Deeply penetrating photoacoustic tomography in biological tissues enhanced with an optical contrast agent. Opt. Lett. 2005, 30, 507–509. 214. Liao, C. K.; Huang, S. W.; Wei, C. W.; Li, P. C. Nanorod-based flow estimation using a high-frame-rate photoacoustic imaging system. J. Biomed. Opt. 2007, 12, 064006. 215. De la Zerda, A.; Zavaleta, C.; Keren, S.; Vaithilingam, S.; Bodapati, S.; Liu, Z.; Levi, J.; Smith, B. R.; Ma, T. J.; Oralkan, O.; Cheng, Z.; Chen, X.; Dai, H.; Khuri-Yakub, B. T.; Gambhir, S. S.
P1: OTA/XYZ P2: ABC c20 JWBS037-Chen January 25, 2011
13:21
Printer Name: Yet to Come
REFERENCES
216. 217. 218. 219.
220. 221.
222. 223.
224.
225.
226.
227.
228. 229. 230. 231. 232. 233. 234.
523
Carbon nanotubes as photoacoustic molecular imaging agents in living mice. Nat. Nanotechnol. 2008, 3, 557–562. Sosnovik, D. E.; Weissleder, R. Emerging concepts in molecular MRI. Curr. Opin. Biotechnol. 2007, 18, 4–10. Cai, W.; Chen, X. Multimodality imaging of vascular endothelial growth factor and vascular endothelial growth factor receptor expression. Front Biosci. 2007, 12, 4267–4279. Wang, H.; Wang, J.; Deng, X.; Sun, H.; Shi, Z.; Gu, Z.; Liu, Y.; Zhao, Y. Biodistribution of carbon single-wall carbon nanotubes in mice. J. Nanosci. Nanotechnol. 2004, 4, 1019–1024. Singh, R.; Pantarotto, D.; Lacerda, L.; Pastorin, G.; Klumpp, C.; Prato, M.; Bianco, A.; Kostarelos, K. Tissue biodistribution and blood clearance rates of intravenously administered carbon nanotube radiotracers. Proc. Natl. Acad. Sci. U.S.A. 2006, 103, 3357–3362. Chatziioannou, A. F. Instrumentation for molecular imaging in preclinical research: Micro-PET and Micro-SPECT. Proc. Am. Thorac. Soc. 2005, 2, 510–511, 533–536. Visser, E. P.; Disselhorst, J. A.; Brom, M.; Laverman, P.; Gotthardt, M.; Oyen, W. J.; Boerman, O. C. Spatial resolution and sensitivity of the Inveon small-animal PET scanner. J. Nucl. Med. 2009, 50, 139–147. Phelps, M. E.; Hoffman, E. J.; Mullani, N. A.; Ter-Pogossian, M. M. Application of annihilation coincidence detection to transaxial reconstruction tomography. J. Nucl. Med. 1975, 16, 210–224. Cherry, S. R.; Shao, Y.; Silverman, R. W.; Meadors, K.; Siegel, S.; Chatziioannou, A.; Young, J. W.; Jones, W. F.; Moyers, J. C.; Newport, D.; Boutefnouchet, A.; Farquhar, T. H.; Andreaco, M.; Paulus, M. J.; Binkley, D. M.; Nutt, R.; Phelps, M. E. MicroPET: a high resolution PET scanner for imaging small animals. IEEE Trans. Nucl. Sci. 1997, 44, 1161–1166. Liu, Z.; Cai, W.; He, L.; Nakayama, N.; Chen, K.; Sun, X.; Chen, X.; Dai, H. In vivo biodistribution and highly efficient tumour targeting of carbon nanotubes in mice. Nat. Nanotechnol. 2007, 2, 47–52. McDevitt, M. R.; Chattopadhyay, D.; Kappel, B. J.; Jaggi, J. S.; Schiffman, S. R.; Antczak, C.; Njardarson, J. T.; Brentjens, R.; Scheinberg, D. A. Tumor targeting with antibodyfunctionalized, radiolabeled carbon nanotubes. J. Nucl. Med. 2007, 48, 1180–1189. McDevitt, M. R.; Chattopadhyay, D.; Jaggi, J. S.; Finn, R. D.; Zanzonico, P. B.; Villa, C.; Rey, D.; Mendenhall, J.; Batt, C. A.; Njardarson, J. T.; Scheinberg, D. A. PET imaging of soluble yttrium-86-labeled carbon nanotubes in mice. PLoS ONE 2007, 2, e907. Villa, C. H.; McDevitt, M. R.; Escorcia, F. E.; Rey, D. A.; Bergkvist, M.; Batt, C. A.; Scheinberg, D. A. Synthesis and biodistribution of oligonucleotide-functionalized, tumor-targetable carbon nanotubes. Nano Lett. 2008, 8, 4221–4228. Brannon-Peppas, L.; Blanchette, J. O. Nanoparticle and targeted systems for cancer therapy. Adv. Drug Deliv. Rev. 2004, 56, 1649–1659. Chavanpatil, M. D.; Khdair, A.; Panyam, J. Nanoparticles for cellular drug delivery: mechanisms and factors influencing delivery. J. Nanosci. Nanotechnol. 2006, 6, 2651–2663. Caliceti, P.; Veronese, F. M. Pharmacokinetic and biodistribution properties of poly(ethylene glycol)-protein conjugates. Adv. Drug Deliv. Rev. 2003, 55, 1261–1277. Wu, A. M.; Senter, P. D. Arming antibodies: prospects and challenges for immunoconjugates. Nat. Biotechnol. 2005, 23, 1137–1146. Soo Choi, H.; Liu, W.; Misra, P.; Tanaka, E.; Zimmer, J. P.; Itty Ipe, B.; Bawendi, M. G.; Frangioni, J. V. Renal clearance of quantum dots. Nat. Biotechnol. 2007, 25, 1165–1170. Cortesini, R. Progress in tissue engineering and organogenesis in transplantation medicine. Exp. Clin. Transplant. 2003, 1, 102–111. Nolan, K.; Millet, Y.; Ricordi, C.; Stabler, C. L. Tissue engineering and biomaterials in regenerative medicine. Cell Transplant. 2008, 17, 241–243.
P1: OTA/XYZ P2: ABC c20 JWBS037-Chen January 25, 2011
524
13:21
Printer Name: Yet to Come
BIOMEDICAL APPLICATIONS OF SINGLE-WALLED CARBON NANOTUBES
235. Harrison, B. S.; Atala, A. Carbon nanotube applications for tissue engineering. Biomaterials 2007, 28, 344–353. 236. Ni, Y.; Hu, H.; Malarkey, E. B.; Zhao, B.; Montana, V.; Haddon, R. C.; Parpura, V. Chemically functionalized water soluble single-walled carbon nanotubes modulate neurite outgrowth. J. Nanosci. Nanotechnol. 2005, 5, 1707–1712. 237. Hu, H.; Ni, Y.; Mandal, S. K.; Montana, V.; Zhao, B.; Haddon, R. C.; Parpura, V. Polyethyleneimine functionalized single-walled carbon nanotubes as a substrate for neuronal growth. J. Phys. Chem. B 2005, 109, 4285–4289. 238. Zanello, L. P.; Zhao, B.; Hu, H.; Haddon, R. C. Bone cell proliferation on carbon nanotubes. Nano Lett. 2006, 6, 562–567. 239. Liopo, A. V.; Stewart, M. P.; Hudson, J.; Tour, J. M.; Pappas, T. C. Biocompatibility of native and functionalized single-walled carbon nanotubes for neuronal interface. J. Nanosci. Nanotechnol. 2006, 6, 1365–1374. 240. Meng, J.; Song, L.; Kong, H.; Zhu, G.; Wang, C.; Xu, L.; Xie, S.; Xu, H. Using single-walled carbon nanotubes nonwoven films as scaffolds to enhance long-term cell proliferation in vitro. J. Biomed. Mater. Res. A 2006, 79, 298–306. 241. Aoki, N.; Akasaka, T.; Watari, F.; Yokoyama, A. Carbon nanotubes as scaffolds for cell culture and effect on cellular functions. Dent. Mater. J. 2007, 26, 178–185. 242. Malarkey, E. B.; Fisher, K. A.; Bekyarova, E.; Liu, W.; Haddon, R. C.; Parpura, V. Conductive single-walled carbon nanotube substrates modulate neuronal growth. Nano Lett. 2009, 9, 264–268. 243. Yildirim, E. D.; Yin, X.; Nair, K.; Sun, W. Fabrication, characterization, and biocompatibility of single-walled carbon nanotube-reinforced alginate composite scaffolds manufactured using freeform fabrication technique. J. Biomed. Mater. Res. B Appl. Biomater. 2008, 87, 406–414. 244. Shi, X.; Sitharaman, B.; Pham, Q. P.; Spicer, P. P.; Hudson, J. L.; Wilson, L. J.; Tour, J. M.; Raphael, R. M.; Mikos, A. G. In vitro cytotoxicity of single-walled carbon nanotube/biodegradable polymer nanocomposites. J. Biomed. Mater. Res. A. 2008, 86, 813–823. 245. Temple, S. The development of neural stem cells. Nature 2001, 414, 112–117. 246. Jan, E.; Kotov, N. A. Successful differentiation of mouse neural stem cells on layer-by-layer assembled single-walled carbon nanotube composite. Nano Lett. 2007, 7, 1123–1128. 247. Kam, N. W.; Jan, E.; Kotov, N. A. Electrical stimulation of neural stem cells mediated by humanized carbon nanotube composite made with extracellular matrix protein. Nano Lett. 2009, 9, 273–278. 248. Hollister, S. J. Porous scaffold design for tissue engineering. Nat. Mater. 2005, 4, 518–524. 249. Langer, R. 1994 Whitaker Lecture: polymers for drug delivery and tissue engineering. Ann. Biomed. Eng. 1995, 23, 101–111. 250. Gunatillake, P. A.; Adhikari, R. Biodegradable synthetic polymers for tissue engineering. Eur. Cell Mater. 2003, 5, 1–16; discussion 16. 251. MacDonald, R. A.; Laurenzi, B. F.; Viswanathan, G.; Ajayan, P. M.; Stegemann, J. P. Collagencarbon nanotube composite materials as scaffolds in tissue engineering. J. Biomed. Mater. Res. A. 2005, 74, 489–496. 252. Shi, X.; Sitharaman, B.; Pham, Q. P.; Liang, F.; Wu, K.; Edward, Billups W.; Wilson, L. J.; Mikos, A. G. Fabrication of porous ultra-short single-walled carbon nanotube nanocomposite scaffolds for bone tissue engineering. Biomaterials 2007, 28, 4078–4090. 253. Sitharaman, B.; Shi, X.; Walboomers, X. F.; Liao, H.; Cuijpers, V.; Wilson, L. J.; Mikos, A. G.; Jansen, J. A. In vivo biocompatibility of ultra-short single-walled carbon nanotube/biodegradable polymer nanocomposites for bone tissue engineering. Bone 2008, 43, 362–370.
P1: OTA/XYZ P2: ABC c20 JWBS037-Chen January 25, 2011
13:21
Printer Name: Yet to Come
REFERENCES
525
254. Hafner, J. H.; Cheung, C. L.; Woolley, A. T.; Lieber, C. M. Structural and functional imaging with carbon nanotube AFM probes. Prog. Biophys. Mol. Biol. 2001, 77, 73–110. 255. Cheung, C. L.; Hafner, J. H.; Lieber, C. M. Carbon nanotube atomic force microscopy tips: direct growth by chemical vapor deposition and application to high-resolution imaging. Proc. Natl. Acad. Sci. U.S.A. 2000, 97, 3809–3813. 256. Woolley, A. T.; Guillemette, C.; Li Cheung, C.; Housman, D. E.; Lieber, C. M. Direct haplotyping of kilobase-size DNA using carbon nanotube probes. Nat. Biotechnol. 2000, 18, 760– 763. 257. Lee, J. H.; Kang, W. S.; Choi, B. S.; Choi, S. W.; Kim, J. H. Fabrication of carbon nanotube AFM probes using the Langmuir–Blodgett technique. Ultramicroscopy 2008, 108, 1163– 1167. 258. Nishino, T.; Ito, T.; Umezawa, Y. Carbon nanotube scanning tunneling microscopy tips for chemically selective imaging. Anal. Chem. 2002, 74, 4275–4278. 259. Burt, D. P.; Wilson, N. R.; Weaver, J. M.; Dobson, P. S.; Macpherson, J. V. Nanowire probes for high resolution combined scanning electrochemical microscopy–atomic force microscopy. Nano. Lett. 2005, 5, 639–643. 260. Mackie, T. R.; Tome, W. Advanced image-guided external beam radiotherapy. Cancer Treat. Res. 2008, 139, 7–39. 261. Dawson, L. A.; Jaffray, D. A. Advances in image-guided radiation therapy. J. Clin. Oncol. 2007, 25, 938–946. 262. Xing, L.; Thorndyke, B.; Schreibmann, E.; Yang, Y.; Li, T. F.; Kim, G. Y.; Luxton, G.; Koong, A. Overview of image-guided radiation therapy. Med. Dosim. 2006, 31, 91–112. 263. Wang, S.; Liu, Z.; Sultana, S.; Schreiber, E.; Zhou, O.; Chang, S. A novel high resolution micro-radiotherapy system for small animal irradiation for cancer research. Biofactors 2007, 30, 265–270. 264. Li, X.; Peng, Y.; Ren, J.; Qu, X. Carboxyl-modified single-walled carbon nanotubes selectively induce human telomeric i-motif formation. Proc. Natl. Acad. Sci. U.S.A. 2006, 103, 19658–19663. 265. Young, J. A.; Collier, R. J. Anthrax toxin: receptor binding, internalization, pore formation, and translocation. Annu. Rev. Biochem. 2007, 76, 243–265. 266. Wang, H.; Gu, L.; Lin, Y.; Lu, F.; Meziani, M. J.; Luo, P. G.; Wang, W.; Cao, L.; Sun, Y. P. Unique aggregation of anthrax (Bacillus anthracis) spores by sugar-coated single-walled carbon nanotubes. J. Am. Chem. Soc. 2006, 128, 13364–13365. 267. Singh, S.; Nalwa, H. S. Nanotechnology and health safety–toxicity and risk assessments of nanostructured materials on human health. J. Nanosci. Nanotechnol. 2007, 7, 3048–3070. 268. Dobrovolskaia, M. A.; McNeil, S. E. Immunological properties of engineered nanomaterials. Nat. Nanotechnol. 2007, 2, 469–478. 269. Derfus, A. M.; Chan, W. C. W.; Bhatia, S. N. Probing the cytotoxicity of semiconductor quantum dots. Nano Lett. 2004, 4, 11–18. 270. Davoren, M.; Herzog, E.; Casey, A.; Cottineau, B.; Chambers, G.; Byrne, H. J.; Lyng, F. M. In vitro toxicity evaluation of single walled carbon nanotubes on human A549 lung cells. Toxicol. In Vitro 2007, 21, 438–448. 271. Yehia, H. N.; Draper, R. K.; Mikoryak, C.; Walker, E. K.; Bajaj, P.; Musselman, I. H.; Daigrepont, M. C.; Dieckmann, G. R.; Pantano, P. Single-walled carbon nanotube interactions with HeLa cells. J. Nanobiotechnology 2007, 5, 8. 272. Manna, S. K.; Sarkar, S.; Barr, J.; Wise, K.; Barrera, E. V.; Jejelowo, O.; Rice-Ficht, A. C.; Ramesh, G. T. Single-walled carbon nanotube induces oxidative stress and activates nuclear transcription factor-kappaB in human keratinocytes. Nano Lett. 2005, 5, 1676–1684.
P1: OTA/XYZ P2: ABC c20 JWBS037-Chen January 25, 2011
526
13:21
Printer Name: Yet to Come
BIOMEDICAL APPLICATIONS OF SINGLE-WALLED CARBON NANOTUBES
273. Cui, D.; Tian, F.; Ozkan, C. S.; Wang, M.; Gao, H. Effect of single wall carbon nanotubes on human HEK293 cells. Toxicol. Lett. 2005, 155, 73–85. 274. Jacobsen, N. R.; Pojana, G.; White, P.; Moller, P.; Cohn, C. A.; Korsholm, K. S.; Vogel, U.; Marcomini, A.; Loft, S.; Wallin, H. Genotoxicity, cytotoxicity, and reactive oxygen species induced by single-walled carbon nanotubes and C(60) fullerenes in the FE1-Mutatrade markMouse lung epithelial cells. Environ. Mol. Mutagen. 2008, 49, 476–487. 275. Raja, P. M.; Connolley, J.; Ganesan, G. P.; Ci, L.; Ajayan, P. M.; Nalamasu, O.; Thompson, D. M. Impact of carbon nanotube exposure, dosage and aggregation on smooth muscle cells. Toxicol. Lett. 2007, 169, 51–63. 276. Casey, A.; Herzog, E.; Lyng, F. M.; Byrne, H. J.; Chambers, G.; Davoren, M. Single walled carbon nanotubes induce indirect cytotoxicity by medium depletion in A549 lung cells. Toxicol. Lett. 2008, 179, 78–84. 277. Sharma, C. S.; Sarkar, S.; Periyakaruppan, A.; Barr, J.; Wise, K.; Thomas, R.; Wilson, B. L.; Ramesh, G. T. Single-walled carbon nanotubes induces oxidative stress in rat lung epithelial cells. J. Nanosci. Nanotechnol. 2007, 7, 2466–2472. 278. Sarkar, S.; Sharma, C.; Yog, R.; Periakaruppan, A.; Jejelowo, O.; Thomas, R.; Barrera, E. V.; Rice-Ficht, A. C.; Wilson, B. L.; Ramesh, G. T. Analysis of stress responsive genes induced by single-walled carbon nanotubes in BJ Foreskin cells. J. Nanosci. Nanotechnol. 2007, 7, 584–592. 279. Kisin, E. R.; Murray, A. R.; Keane, M. J.; Shi, X. C.; Schwegler-Berry, D.; Gorelik, O.; Arepalli, S.; Castranova, V.; Wallace, W. E.; Kagan, V. E.; Shvedova, A. A. Single-walled carbon nanotubes: geno- and cytotoxic effects in lung fibroblast V79 cells. J. Toxicol. Environ. Health A. 2007, 70, 2071–2079. 280. Pulskamp, K.; Diabate, S.; Krug, H. F. Carbon nanotubes show no sign of acute toxicity but induce intracellular reactive oxygen species in dependence on contaminants. Toxicol. Lett. 2007, 168, 58–74. 281. Rotoli, B. M.; Bussolati, O.; Bianchi, M. G.; Barilli, A.; Balasubramanian, C.; Bellucci, S.; Bergamaschi, E. Non-functionalized multi-walled carbon nanotubes alter the paracellular permeability of human airway epithelial cells. Toxicol. Lett. 2008, 178, 95–102. 282. Kagan, V. E.; Tyurina, Y. Y.; Tyurin, V. A.; Konduru, N. V.; Potapovich, A. I.; Osipov, A. N.; Kisin, E. R.; Schwegler-Berry, D.; Mercer, R.; Castranova, V.; Shvedova, A. A. Direct and indirect effects of single walled carbon nanotubes on RAW 264.7 macrophages: role of iron. Toxicol. Lett. 2006, 165, 88–100. 283. Kaiser, J. P.; Wick, P.; Manser, P.; Spohn, P.; Bruinink, A. Single walled carbon nanotubes (SWCNT) affect cell physiology and cell architecture. J. Mater. Sci. Mater. Med. 2008, 19, 1523–1527. 284. Sayes, C. M.; Liang, F.; Hudson, J. L.; Mendez, J.; Guo, W.; Beach, J. M.; Moore, V. C.; Doyle, C. D.; West, J. L.; Billups, W. E.; Ausman, K. D.; Colvin, V. L. Functionalization density dependence of single-walled carbon nanotubes cytotoxicity in vitro. Toxicol. Lett. 2006, 161, 135–142. 285. Kang, S.; Herzberg, M.; Rodrigues, D. F.; Elimelech, M. Antibacterial effects of carbon nanotubes: size does matter! Langmuir. 2008, 24, 6409–6413. 286. Kang, S.; Pinault, M.; Pfefferle, L. D.; Elimelech, M. Single-walled carbon nanotubes exhibit strong antimicrobial activity. Langmuir 2007, 23, 8670–8673. 287. Dutta, D.; Sundaram, S. K.; Teeguarden, J. G.; Riley, B. J.; Fifield, L. S.; Jacobs, J. M.; Addleman, S. R.; Kaysen, G. A.; Moudgil, B. M.; Weber, T. J. Adsorbed proteins influence the biological activity and molecular targeting of nanomaterials. Toxicol. Sci. 2007, 100, 303–315. 288. Nimmagadda, A.; Thurston, K.; Nollert, M. U.; McFetridge, P. S. Chemical modification of SWNT alters in vitro cell-SWNT interactions. J. Biomed. Mater. Res. A 2006, 76, 614–625.
P1: OTA/XYZ P2: ABC c20 JWBS037-Chen January 25, 2011
13:21
Printer Name: Yet to Come
REFERENCES
527
289. Worle-Knirsch, J. M.; Pulskamp, K.; Krug, H. F. Oops they did it again! Carbon nanotubes hoax scientists in viability assays. Nano Lett. 2006, 6, 1261–1268. 290. Herzog, E.; Casey, A.; Lyng, F. M.; Chambers, G.; Byrne, H. J.; Davoren, M. A new approach to the toxicity testing of carbon-based nanomaterials—the clonogenic assay. Toxicol. Lett. 2007, 174, 49–60. 291. Shvedova, A. A.; Kisin, E. R.; Mercer, R.; Murray, A. R.; Johnson, V. J.; Potapovich, A. I.; Tyurina, Y. Y.; Gorelik, O.; Arepalli, S.; Schwegler-Berry, D.; Hubbs, A. F.; Antonini, J.; Evans, D. E.; Ku, B. K.; Ramsey, D.; Maynard, A.; Kagan, V. E.; Castranova, V.; Baron, P. Unusual inflammatory and fibrogenic pulmonary responses to single-walled carbon nanotubes in mice. Am. J. Physiol. Lung. Cell. Mol. Physiol. 2005, 289, L698–708. 292. Shvedova, A. A.; Fabisiak, J. P.; Kisin, E. R.; Murray, A. R.; Roberts, J. R.; Tyurina, Y. Y.; Antonini, J. M.; Feng, W. H.; Kommineni, C.; Reynolds, J.; Barchowsky, A.; Castranova, V.; Kagan, V. E. Sequential exposure to carbon nanotubes and bacteria enhances pulmonary inflammation and infectivity. Am. J. Respir. Cell. Mol. Biol. 2008, 38, 579–590. 293. Shvedova, A. A.; Kisin, E. R.; Murray, A. R.; Kommineni, C.; Castranova, V.; Fadeel, B.; Kagan, V. E. Increased accumulation of neutrophils and decreased fibrosis in the lung of NADPH oxidase-deficient C57BL/6 mice exposed to carbon nanotubes. Toxicol. Appl. Pharmacol. 2008, 231, 235–240. 294. Eggermont, E. Recent advances in vitamin E metabolism and deficiency. Eur. J. Pediatr. 2006, 165, 429–434. 295. Shvedova, A. A.; Kisin, E. R.; Murray, A. R.; Gorelik, O.; Arepalli, S.; Castranova, V.; Young, S. H.; Gao, F.; Tyurina, Y. Y.; Oury, T. D.; Kagan, V. E. Vitamin E deficiency enhances pulmonary inflammatory response and oxidative stress induced by single-walled carbon nanotubes in C57BL/6 mice. Toxicol. Appl. Pharmacol. 2007, 221, 339–348. 296. Shvedova, A. A.; Kisin, E. R.; Murray, A. R.; Johnson, V. J.; Gorelik, O.; Arepalli, S.; Hubbs, A. F.; Mercer, R. R.; Keohavong, P.; Sussman, N.; Jin, J.; Yin, J.; Stone, S.; Chen, B. T.; Deye, G.; Maynard, A.; Castranova, V.; Baron, P. A.; Kagan, V. E. Inhalation vs. aspiration of single-walled carbon nanotubes in C57BL/6 mice: inflammation, fibrosis, oxidative stress, and mutagenesis. Am. J. Physiol. Lung Cell. Mol. Physiol. 2008, 295, L552–565. 297. Chou, C. C.; Hsiao, H. Y.; Hong, Q. S.; Chen, C. H.; Peng, Y. W.; Chen, H. W.; Yang, P. C. Single-walled carbon nanotubes can induce pulmonary injury in mouse model. Nano Lett. 2008, 8, 437–445. 298. Mercer, R. R.; Scabilloni, J.; Wang, L.; Kisin, E.; Murray, A. R.; Schwegler-Berry, D.; Shvedova, A. A.; Castranova, V. Alteration of deposition pattern and pulmonary response as a result of improved dispersion of aspirated single-walled carbon nanotubes in a mouse model. Am. J. Physiol. Lung Cell. Mol. Physiol. 2008, 294, L87–97. 299. Szendi, K.; Varga, C. Lack of genotoxicity of carbon nanotubes in a pilot study. Anticancer Res. 2008, 28, 349–352. 300. Mangum, J. B.; Turpin, E. A.; Antao-Menezes, A.; Cesta, M. F.; Bermudez, E.; Bonner, J. C. Single-walled carbon nanotube (SWCNT)-induced interstitial fibrosis in the lungs of rats is associated with increased levels of PDGF mRNA and the formation of unique intercellular carbon structures that bridge alveolar macrophages in situ. Part. Fibre Toxicol. 2006, 3, 15. 301. Liu, Z.; Davis, C.; Cai, W.; He, L.; Chen, X.; Dai, H. Circulation and long-term fate of functionalized, biocompatible single-walled carbon nanotubes in mice probed by Raman spectroscopy. Proc. Natl. Acad. Sci. U.S.A. 2008, 105, 1410–1415. 302. Schipper, M. L.; Nakayama-Ratchford, N.; Davis, C. R.; Kam, N. W.; Chu, P.; Liu, Z.; Sun, X.; Dai, H.; Gambhir, S. S. A pilot toxicology study of single-walled carbon nanotubes in a small sample of mice. Nat. Nanotechnol. 2008, 3, 216–221.
P1: OTA/XYZ P2: ABC c20 JWBS037-Chen January 25, 2011
528
13:21
Printer Name: Yet to Come
BIOMEDICAL APPLICATIONS OF SINGLE-WALLED CARBON NANOTUBES
303. Hamad, I.; Christy Hunter, A.; Rutt, K. J.; Liu, Z.; Dai, H.; Moein, Moghimi S. Complement activation by PEGylated single-walled carbon nanotubes is independent of C1q and alternative pathway turnover. Mol. Immunol. 2008, 45, 3797–3803. 304. Templeton, R. C.; Ferguson, P. L.; Washburn, K. M.; Scrivens, W. A.; Chandler, G. T. Lifecycle effects of single-walled carbon nanotubes (SWNTs) on an estuarine meiobenthic copepod. Environ. Sci. Technol. 2006, 40, 7387–7393. 305. Cheng, J.; Flahaut, E.; Cheng, S. H. Effect of carbon nanotubes on developing zebrafish (Danio rerio) embryos. Environ. Toxicol. Chem. 2007, 26, 708–716. 306. Smith, C. J.; Shaw, B. J.; Handy, R. D. Toxicity of single walled carbon nanotubes to rainbow trout, (Oncorhynchus mykiss): respiratory toxicity, organ pathologies, and other physiological effects. Aquat. Toxicol. 2007, 82, 94–109. 307. Ghafari, P.; St.-Denis, C. H.; Power, M. E.; Jin, X.; Tsou, V.; Mandal, H. S.; Bols, N. C.; Tang, X. S. Impact of carbon nanotubes on the ingestion and digestion of bacteria by ciliated protozoa. Nat. Nanotechnol. 2008, 3, 347–351. 308. Blaise, C.; Gagne, F.; Ferard, J. F.; Eullaffroy, P. Ecotoxicity of selected nano-materials to aquatic organisms. Environ. Toxicol. 2008, 23, 591–598. 309. Canas, J. E.; Long, M.; Nations, S.; Vadan, R.; Dai, L.; Luo, M.; Ambikapathi, R.; Lee, E. H.; Olszyk, D. Effects of functionalized and non-functionalized single-walled carbon nanotubes on root elongation of select crop species. Environ. Toxicol. Chem. 2008, 27, 1922–1931. 310. Phelps, M. E. PET: the merging of biology and imaging into molecular imaging. J. Nucl. Med. 2000, 41, 661–681. 311. Mammen, M.; Chio, S.; Whitesides, G. M. Polyvalent interactions in biological systems: implications for design and use of multivalent ligands and inhibitors. Angew. Chem. Int. Ed. Engl. 1998, 37, 2755–2794. 312. Grodzinski, P.; Silver, M.; Molnar, L. K. Nanotechnology for cancer diagnostics: promises and challenges. Expert. Rev. Mol. Diagn. 2006, 6, 307–318. 313. Sahoo, S. K.; Parveen, S.; Panda, J. J. The present and future of nanotechnology in human health care. Nanomedicine 2007, 3, 20–31. 314. Oberdorster, G.; Oberdorster, E.; Oberdorster, J. Nanotoxicology: an emerging discipline evolving from studies of ultrafine particles. Environ. Health Perspect. 2005, 113, 823–839. 315. Kagan, V. E.; Bayir, H.; Shvedova, A. A. Nanomedicine and nanotoxicology: two sides of the same coin. Nanomedicine 2005, 1, 313–316.
P1: OTA/XYZ P2: ABC c21 JWBS037-Chen February 1, 2011
0:17
Printer Name: Yet to Come
CHAPTER 21
Multifunctional Nanoparticles for Multimodal Molecular Imaging YANGLONG HOU and RUI HAO Department of Advanced Materials and Nanotechnology, College of Engineering, Peking University, Beijing, China
21.1 INTRODUCTION Inorganic nanoparticles (NPs) are coming forth as potential probes in biomedical imaging and therapy [1, 2]; their comparable size to biofunctional molecules has allowed for ultrasensitive detection of biomolecular targets [3]. Of the various inorganic NPs, magnetic NPs show a unique magnetic resonance (MR) contrast enhancement effect, enabling noninvasive MR imaging of cancer, cardiovascular disease, and cell trafficking [4]. On the other hand, quantum dots (QDs), noble metal, and fluorescent dye hybrid NPs are representative examples of optical nanoprobes [5]. However, to retrieve accurate biological information on a subcellular level is difficult by individually using any of the imaging modalities described above due to their low sensitivity and/or low resolution. Therefore the technology to obtain the detailed information of biological targets becomes an increasingly important need for understanding basic biological phenomena and for early stage diagnosis of various diseases. Current single-model imaging tends to be inadequate. Note that multifunctional NPs can be engineered as platforms for effective multimodal imaging, overcoming the shortcomings that are present for single-model imaging methods. Multifunctional nanoparticle-based probes have great potential for early detection, accurate diagnosis, and personalized treatment of diseases. For example, it has been demonstrated that the simultaneous use of positron emission tomography (PET) and computed tomography (CT) allowed for early detection of cancer in dual modalities [6]. Except for multimodal imaging, the function of therapy can also be integrated into one particle [7]. The multifunctional NPs are emerging as versatile probes for molecular imaging and therapy. So far, as shown in Figure 21.1, there are several types of multifunctional NPs for multimodal probes: (a) functionalized nanoparticle, (b) core–shell nanoparticle, (c) dimer, (d) complex core–shell nanostructure, (e) core–satellite nanostructure, and (f) hybrid nanostructure. In this chapter, we review the current state-of-the-art multifunctional nanoplatforms for multimodal molecular imaging. Single-model molecular imaging is not emphasized here. Nanoplatform-Based Molecular Imaging Edited by Xiaoyuan Chen C 2011 John Wiley & Sons, Inc. Copyright
529
P1: OTA/XYZ P2: ABC c21 JWBS037-Chen February 1, 2011
530
0:17
Printer Name: Yet to Come
MULTIFUNCTIONAL NANOPARTICLES FOR MULTIMODAL MOLECULAR IMAGING
(a)
(b)
(c)
(d)
(e)
(f)
FIGURE 21.1 Several types of nanostructures for multimodal imaging: (a) functionalized nanoparticle, (b) core–shell nanoparticle, (c) dimer, (d) complex core–shell nanostructure, (e) core–satellite nanostructure, and (f) hybrid nanostructure.
21.2 NANOPARTICLE-BASED PROBES 21.2.1 Magnetic NPs Magnetic NPs exhibit a variety of unique magnetic behavior that is drastically different from that of their bulk counterparts due to the nanoscale effect. Magnetic NPs exhibit a significant potential in biomedical applications such as molecular imaging, drug delivery, thermal therapy, biosensing, and bioseparation. In magnetic resonance imaging (MRI), the larger magnetic moment of NPs could create a larger magnetic field heterogeneity, which affects the surrounding water in living subjects. It is used to produce predominantly T2 relaxation effects, resulting in a negative signal on a T2-weighted image. Superparamagnetic iron oxide (SPIO) NPs were employed as an MRI contrast agent after surface modification of the NPs. Coating materials here are usually biocompatible polymers, such as dextran and polyethylene glycol (PEG). At the present time, there are several commercial SPIO NPs, for example, AMI-25 (Feridex, Endorem) [8], SHU 555A (Resovist) [9], and AMI-227 (Combidex, Sinerem) [10]. Some other magnetic NPs (e.g., FeCo [11], MnFe2 O4 [12]), which offering high magnetic moments and r1/r2 relaxivities, have demonstrated longlasting positive-contrast enhancement for vascular MRI. After conjugation with targeting molecules, functionalized NPs can also be used as probes for targeting specific species on the surface of cells not only as an MRI contrast agent. Montet et al. [13] prepared cross linked iron oxide (CLIO) NPs, which were conjugated with Cy5.5 (an infrared optical fluorescence dye) and bound to cyclic arginine–glycine–aspartic acid (RGD); the resulting NPs were used to target ␣v 3 integrins on BT-20 tumor. In this study, the BT-20 tumor cell was successfully observed by fluorescence-based imaging and MRI. A similar route was explored by employing Cy5.5 for fluorescence imaging and CLIO as an MRI contrast
P1: OTA/XYZ P2: ABC c21 JWBS037-Chen February 1, 2011
0:17
Printer Name: Yet to Come
NANOPARTICLE-BASED PROBES
531
agent [14]. The difference between the two examples is that the latter chose peptides (EPPT) which specifically recognize uMUC-1 as a targeting group. This multimodal probe could not only detect orthotopically implanted preclinical models of adenocarcinomas but could also track in situ tumor response to chemotherapy in vivo. By using Cy5.5 as the fluorescence probe, thermally crosslinked superparamagnetic iron oxide (TCL-SPIO) NPs were also developed for dual-modality in vivo cancer imaging of Lewis lung carcinoma tumor allograft in mice. Although the NPs here weren’t conjugated with a targeting group, the tumor was unambiguously detected after intravenous injection in T2-weighted MR images as a 68% signal drop as well as in optical fluorescence images. Cancer immunotherapy approach has been developed to assist the body’s immune system to selectively recognize and kill malignant tumor cells. A dual-modality nanoparticle system is capable of selectively labeling and imaging CTL cells expressing T-cell receptors (TCRs) that recognize cognate MHC–peptide complexes on the surface of antigen-presenting tumor cells [15], as shown in Figure 21.2. In this work, it has been demonstrated that the ex vivo T-cell selectivity and reporter functionality allowed development of NPs for both MRI and fluorescence imaging. PET technology is an imaging modality that uses the signal emitted by positron-emitting radiotracers to construct images of tracer distribution in vivo [16]. (By combining PET with MRI, more information on biological structures could be obtained. Jarrett et al. [17] reported 64 Cu radiolabeling of dextran sulfate coated SPIO NPs, which were prepared by coordinating the 64 Cu to the chelating bifunctional ligand, p-SCNBz-DOTA, followed by integrating the NPs. The multifunctional NPs are targeted toward inflammatory events such as atherosclerotic plaques. Polyaspartic acid (PASP) coated iron oxide NPs [18] were also synthesized for integrin ␣v 3 recognition and positron-emitting radionuclide 64 Cu labeling by conjugating with cyclic arginine–glycine–aspartic (RGD) peptides and the macrocyclic chelating agent DOTA. In this study, multimodal probes were successfully obtained for dual PET/MRI of tumor integrin ␣v 3 expression in vivo using a small-animal model. As is well known, the lymphatic system is an important first line in defense against infection, and it is also used for the metastasis of malignant cancer; hence accurate localization of sentinel lymph nodes (SLNs) is critical for cancer staging [19]. A PET/MRI probe was explored as an in vivo dual-modality imaging agent in an SLN model biological subject [20]. Herein, serum albumin (SA) coated MnFe2 O4 was applied as the MRI probe. In this procedure, the carboxylic and amine groups of SA were crosslinked by using N-(3-dimethylaminopropyl)-N-ethylcarbodiimide hydrochloride and N-hydroxysulfosuccinimide, and then 124 I, which is a PET radionuclide, was conjugated to the tyrosine residue in SA. In the in vivo experiment, two different types of LNs could clearly be identified and accurately localized in a PET/MR fusion image. 21.2.2 Quantum Dots QDs are usually II–VI or III–V semiconductor NPs with a dimensional size of about 10 nm. Compared with small molecule fluorophores, QDs show a narrow, tunable, emission spectrum, broader excitation energy, and better environmental stability. QDs are widely used as biomarkers in molecular imaging and biosensing. Fluorescence imaging/MRI probes were prepared by simple modification of quantum dots, followed by the conjugation of molecular ions such as Gd3+ . Programmed cell death (PCD) is a process of organized cell suicide and occurs in various forms; PCD plays an important role in the physiology as well as the pathology of multicellular organisms. One of the most explored molecular imaging
P1: OTA/XYZ P2: ABC c21 JWBS037-Chen February 1, 2011
532
0:17
Printer Name: Yet to Come
MULTIFUNCTIONAL NANOPARTICLES FOR MULTIMODAL MOLECULAR IMAGING
(a) NH2 NH
2 NH
2
NH2 NH2
SIA
NH2
MHC
NH2 NH2 NH2
Alexa Fluor 647
Polyethylene Glycol
(1) Alexa Fluor 647
O
(2) SATA
N H
Neutravidin
Biotin MHC Complex
Iron Oxide Nanoparticle
SH Neutravidin
Targeting Molecule
in
(b)
vid
(c)
C IA H -S G-M G E E -P -P NP NP
r Co
-P
-P
HC
M
EG
-P
NP
Zeta Potential
–27.1
–28
–22.2
Diameter (nm)
64.8
67.8
71.0
PDI
0.136
0.163
0.105
NP
50 14
80
Amide II (1480 cm–1)
Number (%)
16
25
16 4 15 2 46 11 05
29 28 16 26
Ne
EG
NP
PEG (1105, 1546, 1642 2860, 2916 cm–1) Amide I (1650 cm–1)
Absorbance (AU)
A
SI
EG
e
NP
ra ut
20 15 10 5 0
4000
3000 2000 Wavelength (cm –1)
1000
10
100 Diameter (nm)
1000
FIGURE 21.2 (a) Schematic illustration of synthesis of NP-PEG-MHC-AF647. Iron oxide nanoparticles were coated with functionalized PEG to which neutravidin was covalently bound via a thioether linkage. Biotinylated peptide–MHC was attached to the PEG termini, lending the particle targeting specificity for CTLs. Neutravidin was prelabeled with the fluorophore Alexa Fluor 647. (b) Surface modification of nanoparticles with PEG and MHC–peptide verified by FTIR. (c) Hydrodynamic size and zeta-potential study of nanoparticle constructs at physiologic pH. (Reproduced with permission from Gunn et al. [15].)
targets of PCD is phosphatidylserine. Prinzen et al. [21] prepared AnxA5-functionalized and Gd3+ conjugated QDs for fluorescence imaging and MRI of phosphatidylserine expressed cell. Jin et al. [22] reported the preparation of Gd3+ -functionalized near-infrared (NIR)emitting QDs as in vivo NIR fluorescence imaging and MRI dual functional probe. The as-prepared probe was prepared based on hydrophobic glutathione coated CdSeTe/CdS QDs bound with Gd3+ –DOTA complex. In an in vivo experiment, a phantom containing 5 mL of Gd3+ –DOTA QDs (10 mM) was embedded into a mouse abdomen to test the probe.
P1: OTA/XYZ P2: ABC c21 JWBS037-Chen February 1, 2011
0:17
Printer Name: Yet to Come
NANOPARTICLE-BASED PROBES
533
21.2.3 Noble Metal NPs Noble metal NPs such as gold and silver particles have attracted significant attention owing to their potential applications in molecular imaging, in biosensing, and in plasmons. The strong optical absorption and scattering of noble metal nanoparticles are attributed to the localized surface plasmon resonance [23, 24]. For example, the resonant extinction, which can be tuned from the visible to the near-infrared region by controlling the shape and structure, allows the nanoparticles to serve as molecular contrast agents. The localized heating of resonant absorption drives new thermal therapies and drug delivery mechanisms. The sensitivity of these resonances to their environment leads to simple affinity sensors for the detection of low-level molecular analytes. Coupled with their general lack of toxicity, noble metal nanoparticles are a highly promising class of nanomaterials for new biomedical applications. For example, DTDTPA (dithiolated diethylenetriaminepentaacetic acid) on the surface of gold NPs facilitates conjugation of gadolinium complex; the resulting NPs generate a pronounced enhancement in contrast [25]. The multimodality of Au-Gd-DTDTPA suggests the possible use of these systems in living organisms for diagnostic (MRI) and/or therapeutic applications. In the follow-up work [26], gold@gadolinium chelate complex can freely circulate in the blood pool without undesirable accumulation in the lung, liver, and spleen, suggesting that it is a promising multimodal probe for in vivo MRI and X-ray imaging synchronously. Silver nanoparticles were also applied to the areas of microbial resistance, wound healing, surface enhanced Raman scattering, and metal enhanced fluorescence [27]; however, the multimodal probe based on silver NPs has rarely been reported.
21.2.4 Silica and Other NPs Owing to their good chemical stability and biocompatibility, silica NPs have been widely used in biomedicine. The chemical preparation of silica NPs is easy, low-cost, and convenient; moreover, silica NPs can easily be modified by a silylation agent and encapsulated functional molecular materials. Recently, multifunctional magnetic SiO2 @Ag sandwichlike nanostructures were reported; they are is composed of a magnetic silica core and a continuous Ag shell. The magnetic Ag shells integrated the broad NIR absorption property and superparamagnetic property into one particle, providing promising multimodal probes and magnetic-field targeted photothermal therapy agents [28], as shown in Figure 21.3. MCM-41 type materials, which possess a hexagonal array of one-dimensional channels with diameters of 2–10 nm [29], have been introduced into biomedicine in recent years. For example, Gd-DTPA (DTPA=diethylenetriamine pentaacetic acid) and rhodamine B grafted MCM-4 nanospheres were prepared [30]. In this study, the resulting NPs are a highly efficient T1 contrast agent for intravascular MR imaging and an excellent T2 contrast agent for MR imaging of soft tissues when the NPs were applied at a higher dosage. Polymer or other kinds of organic nanomaterials were also explored as multimodal probes. The complex NPs constructed from Gd and rhodamine B [31], which were linked lipid monomers with diacetylene bonds, were labeled to tumor cells. In this study, labeled cells were inoculated subcutaneouosly into the flanks of C3H mice, and used for imaging the labeled tumor cells via MRI and optical imaging. Cartier et al. [32] reported the synthesis of polyglycidylmethacrylate (EPMA) NPs for multimodality in vivo imaging and detection of the cationic tracers. 111 In, 68 Ga, and Gd3+ were directly bound without the necessity of a chelating conjugate; 68 Ga can be used for PET, while 111 In can be used for quantitative
P1: OTA/XYZ P2: ABC c21 JWBS037-Chen February 1, 2011
534
0:17
Printer Name: Yet to Come
MULTIFUNCTIONAL NANOPARTICLES FOR MULTIMODAL MOLECULAR IMAGING
FIGURE 21.3 Schematic Illustration of the fabrication procedure of magnetic Ag nanoshells. (Reproduced with permission from Chen et al. [28].)
gamma scintigraphy and Gd3+ for MR imaging. Rhodamine B was also incorporated to examine specific interactions with blood cells. Metal organic framework (MOF) is a new class of hybrid material built from metal ions with well-defined coordination geometry and organic bridging ligands. MOF nanorods containing Gd3+ , Eu3+ , or Td3+ were applied to improve the water signals in T1-weighted images efficiently [33, 34]; the results showed high luminescence upon UV excitation with characteristic red and green luminescence from Eu3+ and Tb3+ .
21.3 COMPLEX NANOSTRUCTURE-BASED PROBES 21.3.1 Core–Shell NPs Core–shell NPs, which enrich different components and chemical structures, showed layered spherical structures. Different layers of core–shell NPs can be designed for various needs, possessing multifunctional properties in one unit. Hence it is easy to manufacture multimodal probes by using core–shell nanostructures. Of the core–shell nanostructures, silica-based core–shell NPs have been widely used because of the flexible synthetic route and surface modification. For time-resolved detection of fluorescence cell imaging, Wu et al. [35] prepared multifunctional iron oxide@silica NPs through the copolymerization of Eu3+ complex, 4,4 -bis(1 ,1 ,1 -trifluoro-2 ,4 -butanedion-4 -yl) chlrorosulfoO-terphenyl-Eu3+ (APS-BTBCT-Eu3+ ), free (3-aminopropyl) triethoxysilane (APS), and tetraethyl orthosilicate (TEOS) in the presence of poly(vinylpyrrolidone) (PVP) stabilized magnetic Fe3 O4 NPs. By using as analogous route, fluorescein-isothiocyanate (FITC) functionalization of iron oxide@silica NPs was achieved to improve the specificity of SPIO for targeting glioma cells [36]. Hsiao et al. [37] synthesized small spherical bifunctional Gd-Dye@mesoporous silica NPs (MSN), which were conjugated with FITC and diethylenetriaminepentaacetic acid (DTPA). The functionalized NPs are paramagnetic and emit green fluorescence. MSN was evaluated for its capability as an effective bifunctional fluorescent and T1-enhanced tracker in human mesenchymal stem cells. QDs are good fluorescent agents. The major drawbacks for clinical translation of QDs are inefficient delivery, potential toxicity, and lack of quantification [1]. However, with further improvement of the conjugation strategy, it is expected that QDs may achieve optimal tumor targeting efficacy with acceptable toxicity profile for further application.
P1: OTA/XYZ P2: ABC c21 JWBS037-Chen February 1, 2011
0:17
Printer Name: Yet to Come
COMPLEX NANOSTRUCTURE-BASED PROBES
535
FIGURE 21.4 Iron oxide@Au nanoparicle and its blue and red shift achieved by adjusting the coating layer. (Reproduced with permission from Xu et al. [43].)
Yang et al. [38] developed QDs with silica coating for multimodal imaging. Paramagnetic Gd(III) functionalization via a metal chelating silane coupling agent (TSPETE) to the yellow fluorescent, silica-coated CdS:Mn@ZnS core–shell quantum dots resulted in multimodal NPs that can be imaged optically and by MRI. After silica-coated NPs were covalently linked to Gd3+ ion chelator [39], tetra-azacyclo-dodecane-tetra-acetic acid (DOTA), QDs@silica and Au@silica nanoparticles for dual-modality imaging were obtained. Similar core–shell NPs exhibited high photoluminescence quantum yield, good positive MRI contrast, low cytotoxicity, and intracellular delivery in viable cells [40]. Mn2+ can be used as a T1-weighted MRI probe, while CdSe quantum dots are extremely high luminescence material; however, it is difficult to directly dope Mn2+ into CdSe. Wang et al. [41] maintained optical/MRI dual-modality core–shell structural probes with luminescent CdSe nanoparticles as the core and a paramagnetic Mn-doped ZnS surface. If Gd2 O3 nanoparticles are used for T1-weighted imaging, a dual functional probe can be prepared by conjugating with a fluorescent group. Gadolinium oxide@polysiloxane NPs were functionalized by the organic dye PEG for in vivo dual-modality magnetic resonance and fluorescence imaging [42]. Gold nanostructures, which possess good biocompatibility and surface plasmon resonance, can be used as optical probes. By combining iron oxide with gold, magic iron oxide@Au nanostructures [43] exhibit great potential for biomedical applications due to their integration of magnetic and optical properties (Fig. 21.4). Iron oxide@Au NPs were applied to molecular specific bimodal contrast agents for both magnetic and optical imaging [44]. Kim et al. [45] designed a silica@Au core–shell nanostructure that embeded iron oxide nanoparticle in the gold layer. Anti-HER2/neu was linked to the core–shell structure for targeted MRI and NIR photothermal therapy of cancer cells. A novel multifunctional magnetic particle@silica@Au core–shell structure was reported for targeted cancer detection by MRI as well as for a synchronous cancer therapy via therapeutic antibody and hyperthermia effects [46]. In this study, MnFe2 O4 NP clusters were embedded in silica and further coated with a gold nanolayer. 21.3.2 Heterostructures Similar to core–shell structures, heterostructural NPs are integral domains of different functional components and show multiple properties within one particle. Moreover, heterostructures introduce anisotropy into NPs. For example, they offer more functional surfaces than core–shell NPs, so it is easy to attach different kinds of molecules for specific purposes. Primary research is focused on dimer structures in biomedicine. Choi et al. [47] prepared FePt-Au dimer NPs. In this procedure, FePt NPs were first prepared and Au NPs were added on to the FePt NPs by heteroepitaxial growth. By Au-S linkage modification, the
P1: OTA/XYZ P2: ABC c21 JWBS037-Chen February 1, 2011
536
0:17
Printer Name: Yet to Come
MULTIFUNCTIONAL NANOPARTICLES FOR MULTIMODAL MOLECULAR IMAGING
dimer structures showed high water solubility and biocompatibility. While Au was used for chip-based biosensing, the FePt part was used for MRI because of its superparamagnetism. Sun and co-workers demonstrated that the biocompatible dimer Au–Fe3 O4 NPs can be suitable for linking different functional molecules to each end of the structure through proper surface functionalization [48]. Similar dimer NPs composed of various combinations of metal (Au, Ag, Pt, or Ni) and oxide (Fe3 O4 or MnO) can readily be synthesized from thermal decomposition of mixtures of metal–oleate and metal–oleylamine complexes [49]. Au-MnO and Au-Fe3 O4 dimer NPs have been tested as multimodal imaging probes. Another kind of dimer structure, Fe3 O4 @SiO2 NPs, were prepared first [50]; then Fe3 O4 @SiO2 NPs attached with FITC were produced by using the as-prepared core–shell structures and FITC as precursor. These NPs are fluorescent, magnetic, and porous, exhibiting potential applications in bimodal imaging probes and drug reservoirs.
21.4 NANOHYBRID-BASED PROBES Nanohybrids are a kind of composite material prepared by simple linkage of chemical bonds and/or molecular interactions. Recently, Cheon and co-workers fabricated new “core–satellite” structured dual functional NPs by employing dye doped silica as core and multiple magnetic NPs as satellites [51]. The integrated structures can be used for simultaneous optical and MR imaging of neuroblastoma cells expressing polysialic acids (PSAs) (Fig. 21.5). Yang et al. [52] synthesized antibody conjugated anticancer drug doxorubicin (DOX)– magnetic PLGA NPs (HER-DMPNP) for the detection and treatment of cancer. The welltailored DMPNP was prepared using a surfactant through a nanoemulsion method, in which magnetic NPs were embedded in polymeric NPs through hydrophobic interaction. DOX acts as both a drug and fluorescent group here; the NPs showed excellent sensitivity as MR probes for detection of cancer cells. Cetuximab conjugated fluorescent magnetic nanohybrids were also developed to detect cancer using MR and optical imaging, as shown in Figure 21.6 [53]. Another fluorescent magnetic nanohybrid (FMNH), which was stabilized by pyrenelabeled PCL-b-PMAA, was explored for simultaneous cancer-targeted magnetic resonance (MR) or optical imaging and magnetically-guided drug delivery [54]. This platform was composed of inorganic nanocrystals of Fe3 O4 or CdSe/ZnS, in addition to DOX embedded in biodegradable PLGA polymer NPs. Pegylated folate was coated on the surface of the polymer NPs for active targeting of cancer cells. Luminescent [Ru(bpy)3 ]Cl2 (bpy = 2,2 -bypyridine) was used to build the multifunctional probes. Recently, Ru(bpy):GdIII @SiO2 complex NPs have been demonstrated to be photostable, radio-opaque, and paramagnetic, emerging as multimodal imaging probes, visualized using CT, MRI, and diffuse optical tomography. The surface reactive groups of these probes can be modified to contain both ligands and antibodies, allowing for the detection of cellular events in vivo [55]. Robust luminescent and paramagnetic hybrid silica NPs showed high payloads of magnetic centers for in vitro optical and T1- and T2-weighted MR imaging; in similar work [56], hybrid silica NPs containing [Ru(bpy)3 ]Cl2 were employed for the fluorescent part, and a monolayer coating of silylated Gd complexes were used for MRI. By using layer-by-layer (LbL) self-assembly [57], the developed NPs were further deposited with cationic Gd(III)-DOTA or FITC oligomer and polystyrenesulfonate as an anion layer via electrostatic interactions. By using in vitro MR imaging, it was demonstrated that LbL particles efficiently targeted cancer cells with attached K7 RGD peptide
P1: OTA/XYZ P2: ABC c21 JWBS037-Chen February 1, 2011
0:17
Printer Name: Yet to Come
NANOHYBRID-BASED PROBES
537
(a)
(b)
(c)
(d)
FIGURE 21.5 Core–satellite heterostructures: (a) schematic diagram for the synthesis of core–satellite DySiO2 –(Fe3 O4 )n NPs.(b–d) TEM images of (b) rhodamine-doped silica (DySiO2 ), (c) Fe3 O4 , and (d) core–satellite DySiO2 –(Fe3 O4 )n NPs. (Reproduced with permission from Lee et al. [51].)
FIGURE 21.6 Schematic illustration for preparation of fluorescence magnetic nanohybrids as multimodal imaging agents for cancer detection. (Reproduced with permission from Yang et al. [53].)
P1: OTA/XYZ P2: ABC c21 JWBS037-Chen February 1, 2011
538
0:17
Printer Name: Yet to Come
MULTIFUNCTIONAL NANOPARTICLES FOR MULTIMODAL MOLECULAR IMAGING
noncovalently. This strategy not only provides NPs with transnormally high MR relaxivities but also provides an efficient route for noncovalent functionalization with affinity molecules. 21.5 CONCLUSION AND PERSPECTIVES Molecular imaging is an interdisciplinary research involving chemistry, physics, biology, engineering, medicine, nanoscience, and nanotechnology. Progress on the next generation of molecular probes will result from fusing multiple fluorescent dyes, drug carrier, and multiple magnetic NPs into a single nanoprobe that provides superior fluorescence, MR imaging, and drug delivery capabilities through the synergistic enhancement of its respective components. The development of technology in multimodal imaging is providing an alternative route to retrieve biological information in exquisite detail. Each of the most frequently used techniques, such as MRI, optical imaging, PET, and CT, have inherent advantages and disadvantages. Multifunctional NPs with simultaneous optical and MR imaging and carrier capability are being developed for high sensitivity and resolution of optical microscopy with the ability for MRI and therapy. These nanoplatforms enable the same biological structures of a specimen to be studied at dramatically different resolutions. The synthesis and surface chemistry of multifunctional nanoparticles are a rapidly developing area of research that focuses more attention on the in vivo requirements of biocompatibility and pharmacokinetics. The multimodal probes based on multifunctional NPs are expected to give accurate imaging and provide a research model for biological and medical application.
REFERENCES 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. 14.
Cai, W. B.; Chen, X. Y. Small 2007, 3, 1840. McCarthy, J. R.; Weissleder, R. Adv. Drug Deliv. Rev. 2008, 60, 1241. Bulte, J. W. M.; Kraitchman, D. L. NMR Biomed. 2004, 17, 484. Jun, Y. W.; Seo, J. W.; Cheon, J. Acc. Chem. Res. 2008, 41, 179. Sharrna, P.; Brown, S.; Walter, G.; Santra, S.; Moudgil, B. Adv. Colloid Interface Sci. 2006, 123, 471. Jorn, A. D.; Wouter, V. V.; Corstens Frans, H. M.; OyenWim, J. G. Cancer Imaging 2007, 7, 77. Liong, M.; Lu, J.; Kovochich, M.; Xia, T.; Ruehm, S. G.; Nel, A. E.; Tamanoi, F.; Zink, J. I. ACS Nano 2008, 2, 889. Weissleder, R.; Stark, D. D.; Engelastad, B. L.; Bacon, B. R.; Compton, C. C.; White, D. L.; Jacobs, P.; Lewis, J. Am. J. Roentgenol. 1989, 152, 167. Reimer, P.; Rummeny, E. J.; Daldrup, H. E.; Balzer, T.; Tombach, B.; Berns, T.; Peters, P. E. Radiology 1995, 195, 489. Jung, C. W.; Jacobs, P. Magn. Reson. Imaging 1995, 13, 661. Lee, J. H.; Huh, Y. M.; Jun, Y. W.; Seo, J. W.; Jang, J. T.; Song, H. T.; Kim, S.; Cho, E. J.; Yoon, H. G.; Suh, J. S.; Cheon, J. Nat. Med. 2007, 13, 95. Seo, W. S.; Lee, J. H.; Sun, X.; Suzuki, Y.; Mann, D.; Liu, Z.; Terashima, M.; Yang, P. C.; McConnell, M. V.; Nishimura, D. G.; Dai, H. Nat. Mater. 2006, 5, 971. Montet, X.; Montet-Abou, K.; Reynolds, F.; Weissleder, R.; Josephson, L. Neoplasia 2006, 8, 214. Grimm, J.; Potthast, A.; Wunder, A.; Moore, A. Int. J. Cancer 2003, 106, 806.
P1: OTA/XYZ P2: ABC c21 JWBS037-Chen February 1, 2011
0:17
Printer Name: Yet to Come
REFERENCES
539
15. Gunn, J.; Wallen, H.; Veiseh, O.; Sun, C.; Fang, C.; Cao, J. H.; Yee, C.; Zhang, M. Q. Small 2008, 4, 712. 16. West, C. M. L.; Jones, T.; Price, P. Nat. Rev. Cancer 2004, 4, 457. 17. Jarrett, B. R.; Gustafsson, B.; Kukis, D. L.; Louie, A. Y. Bioconjug. Chem. 2008, 19, 1496. 18. Lee, H. Y.; Li, Z.; Chen, K.; Hsu, A. R.; Xu, C. J.; Xie, J.; Sun, S. H.; Chen, X. Y. J. Nucl. Med. 2008, 49, 1371. 19. Wallace, A. M.; Hoh, C. K.; Ellner, S. J.; Darrah, D. D.; Schulteis, G.; Vera, D. R. Ann. Surg. Oncol. 2007, 14, 913. 20. Choi, J. S.; Park, J. C.; Nah, H.; Woo, S.; Oh, J.; Kim, K. M.; Cheon, G. J.; Chang, Y.; Yoo, J.; Cheon, J. Angew. Chem. Int. Ed. 2008, 47, 6259. 21. Prinzen, L.; Miserus, R.; Dirksen, A.; Hackeng, T. M.; Deckers, N.; Bitsch, N. J.; Megens, R. T. A.; Douma, K.; Heemskerk, J. W.; Kooi, M. E.; Frederik, P. M.; Slaaf, D. W.; van Zandvoort, M.; Reutelingsperger, C. P. M. Nano Lett. 2007, 7, 93. 22. Jin, T.; Yoshioka, Y.; Fujii, F.; Komai, Y.; Seki, J.; Seiyama, A. Chem. Commun. 2008, 5764. 23. Liao, H. W.; Nehl, C. L.; Hafner, J. H. Nanomedicine 2006, 1, 201. 24. Hu, M.; Chen, J.; Li, Z. Y.; Au, L.; Hartland, G. V.; Li, X.; Marquez, M.; Xia, Y. Chem. Soc. Rev. 2006, 35, 1084. 25. Debouttiere, P. J.; Roux, S.; Vocanson, F.; Billotey, C.; Beuf, O.; Favre-Reguillon, A.; Yi, L.; Pellet-Rostaing, S.; Lamartine, R.; Perriat, P.; Tillement, O. Adv. Funct. Mater. 2006, 16, 2330. 26. Alric, C.; Taleb, J.; Le Duc, G.; Mandon, C.; Billotey, C.; Le Meur-Herland, A.; Brochard, T.; Vocanson, F.; Janier, M.; Perriat, P.; Roux, S.; Tillement, O. J. Am. Chem. Soc. 2008, 130, 5908. 27. Nair, L. S.; Laurencin, C. T. J. Biomed. Nanotechnol. 2007, 3, 301. 28. Chen, M.; Kim, Y. N.; Lee, H. M.; Li, C.; Cho, S. O. J. Phys. Chem. C 2008, 112, 8870. 29. Kresge, C. T.; Leonowicz, M. E.; Roth, W. J.; Vartuli, J. C.; Beck, J. S. Nature 1992, 359, 710. 30. Taylor, K. M. L.; Kim, J. S.; Rieter, W. J.; An, H.; Lin, W. L.; Lin, W. B. J. Am. Chem. Soc. 2008, 130, 2154. 31. Vuu, K.; Xie, J. W.; McDonald, M. A.; Bernardo, M.; Hunter, F.; Zhang, Y. T.; Li, K.; Bednarski, M.; Guccione, S. Bioconjug. Chem. 2005, 16, 995. 32. Cartier, R.; Kaufner, L.; Paulke, B. R.; Wustneck, R.; Pietschmann, S.; Michel, R.; Bruhn, H.; Pison, U. Nanotechnology 2007, 18, 195102. 33. Rieter, W. J.; Taylor, K. M. L.; An, H. Y.; Lin, W. L.; Lin, W. B. J. Am. Chem. Soc. 2006, 128, 9024. 34. Taylor, K. M. L.; Jin, A.; Lin, W. B. Angew. Chem. Int. Ed. 2008, 47, 7722. 35. Wu, J.; Ye, Z. Q.; Wang, G. L.; Yuan, J. L. Talanta 2007, 72, 1693. 36. Meng, X. X.; Wan, J. Q.; Jing, M.; Zhao, S. G.; Cai, W.; Liu, E. Z. Acta Pharmacol. Sinica 2007, 28, 2019. 37. Hsiao, J. K.; Tsai, C. P.; Chung, T. H.; Hung, Y.; Yao, M.; Liu, H. M.; Mou, C. Y.; Yang, C. S.; Chen, Y. C.; Huang, D. M. Small 2008, 4, 1445. 38. Yang, H. S.; Santra, S.; Walter, G. A.; Holloway, P. H. Adv. Mater. 2006, 18, 2890. 39. Gerion, D.; Herberg, J.; Bok, R.; Gjersing, E.; Ramon, E.; Maxwell, R.; Kurhanewicz, J.; Budinger, T. F.; Gray, J. W.; Shuman, M. A.; Chen, F. F. J. Phys. Chem. C 2007, 111, 12542. 40. Bakalova, R.; Zhelev, Z.; Aoki, I.; Masamoto, K.; Mileva, M.; Obata, T.; Higuchi, M.; Gadjeva, V.; Kanno, I. Bioconjug. Chem. 2008, 19, 1135. 41. Wang, S.; Jarrett, B. R.; Kauzlarich, S. M.; Louie, A. Y. J. Am. Chem. Soc. 2007, 129, 3848. 42. Bridot, J. L.; Faure, A. C.; Laurent, S.; Riviere, C.; Billotey, C.; Hiba, B.; Janier, M.; Josserand, V.; Coll, J. L.; Vander Elst, L.; Muller, R.; Roux, S.; Perriat, P.; Tillement, O. J. Am. Chem. Soc. 2007, 129, 5076.
P1: OTA/XYZ P2: ABC c21 JWBS037-Chen February 1, 2011
540
0:17
Printer Name: Yet to Come
MULTIFUNCTIONAL NANOPARTICLES FOR MULTIMODAL MOLECULAR IMAGING
43. Xu, Z.; Hou, Y.; Sun, S. J. Am. Chem. Soc. 2007, 129, 8698. 44. Larson, T. A.; Bankson, J.; Aaron, J.; Sokolov, K. Nanotechnology 2007, 18, 8. 45. Kim, J.; Park, S.; Lee, J. E.; Jin, S. M.; Lee, J. H.; Lee, I. S.; Yang, I.; Kim, J. S.; Kim, S. K.; Cho, M. H.; Hyeon, T. Angew. Chem. Int. Ed. 2006, 45, 7754. 46. Lee, J.; Yang, J.; Ko, H.; Oh, S. J.; Kang, J.; Son, J. H.; Lee, K.; Lee, S. W.; Yoon, H. G.; Suh, J. S.; Huh, Y. M.; Haam, S. Adv. Funct. Mater. 2008, 18, 258. 47. Choi, J. S.; Jun, Y. W.; Yeon, S. I.; Kim, H. C.; Shin, J. S.; Cheon, J. J. Am. Chem. Soc. 2006, 128, 15982. 48. Xu, C.; Xie, J.; Ho, D.; Wang, C.; Kohler, N.; Walsh, E. G.; Morgan, J. R.; Chin, Y. E.; Sun, S. Angew. Chem. Int. Ed. 2008, 47, 173. 49. Choi, S. H.; Na, H. B.; Park, Y. I.; An, K.; Kwon, S. G.; Jang, Y.; Park, M.; Moon, J.; Son, J. S.; Song, I. C.; Moon, W. K.; Hyeon, T. J. Am. Chem. Soc. 2008, 130, 15573. 50. Wu, S. H.; Lin, Y. S.; Hung, Y.; Chou, Y. H.; Hsu, Y. H.; Chang, C.; Mou, C. Y. Chembiochem 2008, 9, 53. 51. Lee, J. H.; Jun, Y. W.; Yeon, S. I.; Shin, J. S.; Cheon, J. Angew. Chem. Int. Ed. 2006, 45, 8160. 52. Yang, J.; Lee, C. H.; Park, J.; Seo, S.; Lim, E. K.; Song, Y. J.; Suh, J. S.; Yoon, H. G.; Huh, Y. M.; Haam, S. J. Mater. Chem. 2007, 17, 2695. 53. Yang, J.; Lim, E. K.; Lee, H. J.; Park, J.; Lee, S. C.; Lee, K.; Yoon, H. G.; Suh, J. S.; Huh, Y. M.; Haam, S. Biomaterials 2008, 29, 2548. 54. Kim, J.; Lee, J. E.; Lee, S. H.; Yu, J. H.; Lee, J. H.; Park, T. G.; Hyeon, T. Adv. Mater. 2008, 20, 478. 55. Santra, S.; Bagwe, R. P.; Dutta, D.; Stanley, J. T.; Walter, G. A.; Tan, W.; Moudgil, B. M.; Mericle, R. A. Adv. Mater. 2005, 17, 2165. 56. Rieter, W. J.; Kim, J. S.; Taylor, K. M. L.; An, H. Y.; Lin, W. L.; Tarrant, T.; Lin, W. B. Angew. Chem. Int. Ed. 2007, 46, 3680. 57. Kim, J. S.; Rieter, W. J.; Taylor, K. M. L.; An, H.; Lin, W. L.; Lin, W. B. J. Am. Chem. Soc. 2007, 129, 8962.
P1: OTA/XYZ P2: ABC c22 JWBS037-Chen January 25, 2011
13:25
Printer Name: Yet to Come
CHAPTER 22
Multifunctional Nanoparticles for Cancer Theragnosis SEULKI LEE, ICK CHAN KWON, and KWANGMEYUNG KIM Biomedical Research Center, Korea Institute of Science and Technology, Seoul, South Korea
Cancer theragnosis is a new concept in next-generation cancer treatment that combines early-stage diagnosis and efficient therapy. Recently, nanotechnology and molecular imaging have been combined to generate multifunctional nanoparticles that simultaneously facilitate cancer diagnosis and therapy (cancer theragnosis). These multifunctional nanoparticles show great promise in the emerging field of personalized medicine, because they allow detection as well as monitoring of an individual patient’s cancer at an early stage and delivery of anticancer agents over an extended period for enhanced therapeutic efficacy. Moreover, noninvasive and real-time monitoring of the multifunctional nanoparticles could enable clinical management of cancer for early diagnostic and response to drug treatment assessment. The application of cancer theragnosis is still in its infancy, but cancer theragnosis with multifunctional nanoparticles presents a promising new strategy in cancer treatment, permitting simultaneous early-stage diagnosis, drug delivery, and real-time monitoring for therapeutic efficacy. The hypothesis offered in this chapter is that cancer theragnosis, if properly integrated with established cancer research, provides extraordinary opportunities to meet these challenges.
22.1 INTRODUCTION Although many researchers have paid much attention to cancer treatment, no substantial progress has been made over the past 50 years. This is because current cancer treatments the lack selectivity of anticancer agents in cancer cells. The anticancer agents used in cancer treatment are randomly distributed in the body because they have no tumor selectivity, resulting in a relatively low therapeutic index. Therefore we need specific delivery systems designed to safely shuttle anticancer agents to the target cancer cells. For this purpose, various nanoparticles of polymers, metals, and semiconductors have been used as new drug delivery systems or imaging agents that have novel and unique properties, Nanoplatform-Based Molecular Imaging Edited by Xiaoyuan Chen C 2011 John Wiley & Sons, Inc. Copyright
541
P1: OTA/XYZ P2: ABC c22 JWBS037-Chen January 25, 2011
542
13:25
Printer Name: Yet to Come
MULTIFUNCTIONAL NANOPARTICLES FOR CANCER THERAGNOSIS
such as enhanced tumor targeting ability, prolonged in vivo biodistribution, and biocompatibility [1–3]. Various nanoparticles are known to accumulate at the tumor site by the unique vascular structural changes associated with the pathology of cancer, providing opportunities for the use of nanoscale delivery systems [4–7]. The blood vessels in tumors have several abnormalities, including a relatively high proportion of proliferating endothelial cells, increased tortuosity, deficiency in pericytes, and aberrant basement membrane formation. The defective vascular structure, created due to the rapid vascularization for serving fast-growing cancers, renders them permeable to macromolecules or nanoparticles. This passive targeting phenomenon discovered by Maeda and colleagues is called the “enhanced permeation and retention effect (EPR effect)” [8, 9]. It is now well established that macromolecules and nanosized particulates (e.g., polymer conjugates, polymeric micelles, dendrimers, and liposomes) accumulate passively in solid tumor tissues by the EPR effect, which increases the therapeutic index but decreases side effects. However, in vivo studies have shown that the tumor specificity of nanoparticles is not as high as anticipated. Nanoparticles introduced into the circulating blood are quickly removed from the body by the immune system. Indeed, a number of nanoparticles with different characteristics (e.g., different surface chemistry, size, surface charge, and molecular weight) have proved unsatisfactory in in vivo tests with regard to stability, biodistribution, and tumor targeting specificity [10, 11]. One promising approach that enhances the cancerspecific target efficacy is to develop nanoparticle constructs containing a targeting moiety, such as antibody, aptamer, and small molecule ligand, for optimized tumor homing with reduced nonspecific accumulation [12–14]. These targeting moieties showed selective delivery of nanoparticles into tumors or selective detection in tumors. Despite their improved selectivity, they still have to overcome an obstacle—a large percentages of nanoparticles continue to accumulate in the liver and spleen. Apparently, nanoparticulate delivery systems are not working well as expected. Furthermore, the final in vivo destination of nanoparticles is largely unknown, because it is difficult to acquire direct and noninvasive in vivo characteristics of nanoparticles. Recently, nanotechnology and molecular imaging have been combined to generate multifunctional nanoparticles that simultaneously facilitate cancer theragnosis [15]. Theragnostic nanoparticles containing therapeutic agents and imaging probes show great promise in the emerging field of cancer theragnosis, because they allow noninvasive imaging of in vivo characteristics of various nanoparticles as well as delivering anticancer agents over an extended period for enhanced therapeutic efficacy (Fig. 22.1) [16]. For these reasons, we believe that cancer theragnosis with multifunctional nanoparticles is a promising new strategy in cancer treatment, permitting simultaneous early-stage diagnosis, drug delivery, and real-time monitoring for therapeutic efficacy. The basic rationale is that nanometer-sized particles, such as polymeric micelles/particles, semiconductor quantum dots, and iron oxide nanocrystals, have theragnostic and multifunctional characteristics that are not available from one molecule or bulk solids. When nanosized particle structure is formed with or without tumor targeting ligands such as monoclonal antibodies, peptides, or small molecules, these nanoparticles can be used as a tumor-targeting drug delivery system as well as tumorspecific imaging agents, simultaneously. Moreover, real-time and noninvasive monitoring of the theragnostic nanoscale particles could enable researchers to rapidly decide whether the multifunctional nanoparticle is effective in cancer treatment. For these reasons, we believe that cancer theragnosis with theragnostic nanoparticles present a promising new strategy in cancer treatment, permitting simultaneous early-stage diagnosis, drug delivery, and real-time monitoring for therapeutic efficacy.
P1: OTA/XYZ P2: ABC c22 JWBS037-Chen January 25, 2011
13:25
Printer Name: Yet to Come
CONCEPT OF CANCER THERAGNOSIS
543
FIGURE 22.1 Design of multifunctional nanoparticles with specific targeting moieties, therapeutic agents, and imaging probes in cancer treatment.
22.2 CONCEPT OF CANCER THERAGNOSIS 22.2.1 Cancer as Theragnostic Nanoparticle Target One of the current challenges in cancer treatment is tumor-targeting delivery of imaging agents and anticancer agents. In particular, cancer researchers have been actively exploring nanoparticles made of lipid-based micelles, natural/synthetic polymeric particles, and inorganic particles for cancer diagnosis and therapy [2, 17, 18]. This is because nanoparticles are known to accumulate at the tumor site by the so-called enhanced permeation and retention (EPR) effect, resulting in efficient passive targeting in solid tumor tissues (Fig. 22.2). Tumors show several abnormalities of blood vessel structure in comparison with physiological normal blood vessels. These include relatively high angiogenesis with increased proliferation of endothelial cells, increased tortuosity, and aberrant basement membrane formation [19, 20]. In general, the rapid formation of blood vessels in the tumor area leads to a discontinuous endothelium, with gaps several hundred nanometers in size between cells. The transportation of macromolecules from blood vessel to tumor occurs via discontinuous endothelium, vesicular vacuolar organelles, and fenestrations. However, it still remains debatable which pathways are predominantly responsible for tumor hyperpermeability and macromolecular transvascular transport. The optimum size of nanoparticles is not yet precisely known for effective accumulation in a tumor via the EPR effect. Some studies using liposomes and nanoparticles have shown that the cutoff size of the gap in the tumor endothelium is as large as 200 nm to
P1: OTA/XYZ P2: ABC c22 JWBS037-Chen January 25, 2011
544
13:25
Printer Name: Yet to Come
MULTIFUNCTIONAL NANOPARTICLES FOR CANCER THERAGNOSIS
FIGURE 22.2 Schematic representation of tumor-specific accumulation via “active targeting,” which is achieved by functionalizing the surface of nanoparticles with ligands that promote cellspecific recognition, and “Passive targeting,” which is achieved by extravasation of nanoparticles through increased permeability of the tumor vasculature.
1.2 mm [21, 22]. In addition, the direct observation of tumor vasculature has demonstrated a tumor-dependent gap size between the endothelium cells from 200 nm to 2 mm [23, 24]. These size ranges appear to be an indication that drug-loaded nanoparticles may accumulate in the tumor area. Consistent with this, administered liposomal formulations with DOX entrapment have been described to exhibit favorable pharmacokinetics due to EPRmediated tumor targeting, as compared with free DOX [25]. Furthermore, DOX-loaded polymeric nanoparticles showed long circulation in the blood for more than 3 days, and gradual accumulation in tumors via the EPR effect [26, 27]. In theory, genes and proteins could be delivered to primary or metastasized tumors via the EPR effect. This suggests that a wide variety of polymeric nanoparticles may be used for tumor targeting of anticancer drugs. However, it should be noted that the vessel permeability that forms a cornerstone of the EPR effect varies during tumor progression. In addition, transvascular transportation of polymeric nanomedicines will depend on the tumor type and anatomical location, as well as the physicochemical properties of the utilized polymer in nanomedicines. Moreover, specific tumor accumulation of various nanoparticles via active targeting is achieved by chemical attachment of targeting moieties that strongly interact with tumor-associated or tumor-specific antigens located on the target cancer cells. Myriad nanoparticles have several advantages such as large surface area (ideal for efficient modification with a wide range of imaging moieties), prolonged plasma circulation time, enhanced stability, improved targeting ability, and reduced nonspecific binding. For specific targeting to tumors, these nanoparticles could be chemically modified with tumor-targeting proteins (mainly antibodies and their fragments), nucleic acids (aptamer), or other receptor ligands such as peptides, vitamins, and carbohydrates [28–31]. The use of tumor-targeting moieties
P1: OTA/XYZ P2: ABC c22 JWBS037-Chen January 25, 2011
13:25
Printer Name: Yet to Come
CONCEPT OF CANCER THERAGNOSIS
545
allows facilitated cellular uptake of the drug by receptor-mediated endocytosis, which is an active process requiring a significantly lower gradient of internalized substance across the plasma membrane when compared with simple endocytosis, as well as delivering the multifunctional nanoparticles containing therapeutic agents or imaging probes in cancer theragnosis [32]. 22.2.2 Molecular Imaging in Cancer Theragnosis The physicochemical properties can determine nanoparticles’ behavior in a biological system, such as biodistribution, safety, targeting efficacy, and multifunctional activity in drug delivery systems and molecular imaging. General traditional techniques (electron microscopy, dynamic light scattering, energy dispersive X-ray, zeta potential, etc.) can be used to characterize the physicochemical properties of nanoparticles. However, the in vitro physicochemical properties of nanosystems do not always match with in vivo behaviors of nanoparticles, due to the in vivo biological microenvironment. Therefore characterization of nanoparticles must be carried out under in vivo conditions, not only in vitro. Unfortunately, there is only limited information about the interaction between nanosystems and the biological microenvironment. The physicochemical properties of optimum nanoparticles for drug carriers and imaging probes have not been fully characterized. Despite the benefits that nanosystems have contributed to cancer treatment, some applications remain to be improved; for example, specific targeting to the active site, efficient drug delivery inside the target cells or tissues, and early-stage diagnosis. So resonable and well-established approaches for in vivo characterization of nanosystems are required for novel nanocarriers to be used in drug delivery and molecular imaging. Rationally designed nanosystems, based on our understanding of their interactions with the biological environment and molecular mechanisms in vivo, are necessary for achieving efficient drug delivery and molecular imaging. Consequently, it is very important to understand how multifunctional nanoparticles directly or indirectly interact with target cancer cells. It is necessary to learn how they interact with cancer cells, how the particles are targeted, how they detect early-stage cancer, and how they efficiently treat the tumor. All of theses processes can now be visualized via “molecular imaging” techniques at the molecular level of resolution. Molecular imaging has recently emerged and is a rapidly growing field; it is providing new opportunities to directly visualize in vivo characteristics of multifunctional nanoparticles. Currently, in vivo molecular imaging techniques, including magnetic resonance imaging (MRI) [33, 34], positron emission tomography (PET) [35, 36], computed tomography (CT) [37], and near-infrared (NIR) fluorescence imaging [38, 39], are extensively used to characterize the in vivo behaviors of nanosized drug carriers and imaging probes (Table 22.1). Positron emission tomography (PET) and single photon emission computed tomography (SPECT) are two important nuclear imaging techniques for clinical applications that show high sensitivity and sufficient tissue penetration. In the case of small laboratory animals, microPET and microSPECT are also designed for high-resolution imaging at the molecular level [35, 36]. For target-specific imaging, antibodies, peptides, and nucleic acids are generally labeled for nuclear imaging. Radionuclides for SPECT such as 99m Tc, 123 I, and 125 I have relatively long half-lives compared to those for PET, in which the radiotracers are labeled with positron emitting radioactive atoms such as 11 C, 13 N, 15 O, and 18 F. Also, MRI provides both functionality as well as anatomical information. Recently, a novel noninvasive molecular magnetic resonance imaging (mMRI) technology was developed for
546 Pseudoquantification
Yes Yes Yes
<10 cm
1.2 mm
SPECT
Fluorescence 1 mm imaging
No limit cm No limit
50 m 50 m 1.2 mm
CT Ultrasound PET
Yes
Yes
No limit
10–100 m
MRI
Quantification
No limit
Depth
Resolution
mTc- or 11 In-labeled compounds
Near-infrared fluorochromes, QDs
99
Paramagnetic chelates, magnetic particles Iodinated molecules, Gold Microbubbles 18 F-, 64 Cu-, or 11 Clabeled compounds
Imaging agents
Tumor, liver Angiography Tumor, pharmacokinetics, pharmacodynamics, biodistribution Tumor pharmacokinetics, pharmacodynamics, biodistribution Tumor, biodistribution, molecular events, physiological events
Tumor, liver
Imaging Targets
Photodynamic therapy
No
Tumor-targeting drug delivery system Hyperthermia therapy US responsive drug delivery No
Therapy
13:25
Imaging Method
TABLE 22.1 Characteristics of Currently Available Imaging Techniques
P1: OTA/XYZ P2: ABC c22 JWBS037-Chen January 25, 2011 Printer Name: Yet to Come
P1: OTA/XYZ P2: ABC c22 JWBS037-Chen January 25, 2011
13:25
Printer Name: Yet to Come
MULTIFUNCTIONAL NANOPARTICLES FOR CANCER THERAGNOSIS
547
visualization of the biological process at the cellular and the molecular levels based on advances in MRI physics and target-specific contrast imaging agents [33, 34]. Although it has limited tissue penetration (a few centimeters), optical imaging is a very powerful and noninvasive technique that provides excellent spatial and temporal resolution. The light in the visible wavelength range is used for conventional and intravital microscopic observation. The near-infrared (NIR) region around 650–900 nm has been used to image deep tissue, since there is lower light absorption from hemoglobin, water, and lipids [37, 40]. In general, various nanoparticles have been directly labeled with a wide range of imaging probes according to the imaging modality via conventional chemistry. For example, fluorescent probes for optical imaging, paramagnetic agents for MR imaging, and radiolabeled probes for PET imaging were labeled to nanoparticles. Novel nanosystems with imaging probes can be used in real-time monitoring and visualization in a noninvasive way, allowing for clinical uses in animals and humans. With the help of various imaging systems, the most important characteristics of nanosystems in vivo were studied, including in vivo ADME (absorption, distribution, metabolism, and excretion), targeting efficacy, and their multifunctional properties as drug carriers and imaging probes. Also, in vivo experimental uncertainties arising from interanimal variations are greatly reduced, because each animal serves as its own “control” for consecutive analyses at the same condition. The different imaging technologies revealed the information on circulating or localized nanoparticles, which has close relevance with the in vivo fate of different nanoparticles in small animals or humans, in that drug carriers and imaging probes are passively targeted or nonspecifically localized.
22.3 MULTIFUNCTIONAL NANOPARTICLES FOR CANCER THERAGNOSIS 22.3.1 Polymeric Multifunctional Nanoparticles for Cancer Theragnosis Polymeric nanoparticles have been applied as effective theragnostic nanoparticles capable of enhancing the diagnostic and therapeutic efficacy of imaging probes and therapeutic agents in cancer treatment. Modern polymer chemistry has produced an increasing number of polymeric nanoparticles for cancer theragnosis, including dendrimers, natural/synthetic amphiphilic polymers, and grafted or branched polymers (Fig. 22.3). When they form a self-assembled nanoparticle, they could exhibit unique physicochemical characteristics in the aqueous condition, possibly having unusual rheological features, a small hydrodynamic radius, or thermodynamic stability [41]. In particular, polymeric nanoparticles exhibited ideal in vivo characteristics, such as biocompatibility and biodegradability, and prolonged circulation time in the bloodstream. These unique in vivo characteristics of polymeric nanoparticles have been extensively exploited in cancer treatment. This is because biocompatible and prolonged circulating polymeric nanoparticles in vivo allow them to accumulate and extravasate into the tumor tissue, where a disorganized vasculature and defective vascular architecture is usual. Recently, a fusion technology of polymer chemistry and molecular imaging approaches has led to the generation of cancer theragnosis based on novel polymeric nanoparticles. Most imaging probes currently used in clinical applications are physically/chemically introduced into the polymeric nanoparticles. For example, PET/SPECT radionuclides (11 C, 18 F, and 64 Cu), MRI imaging probe (such as Gd3+ ), and optical imaging probes (fluorochromes,
P1: OTA/XYZ P2: ABC c22 JWBS037-Chen January 25, 2011
548
13:25
Printer Name: Yet to Come
MULTIFUNCTIONAL NANOPARTICLES FOR CANCER THERAGNOSIS
FIGURE 22.3 Various polymeric nanoparticles for cancer theragnosis. Physical loading or chemical conjugation of therapeutic agents and imaging probes on polymeric nanoparticles were used for cancer theragnosis.
quantum dot, and photoproteins) were simply introduced into the polymeric nanoparticles by physical complex or chemical conjugation. Polymeric nanoparticles containing imaging probes could be noninvasively visualized in animals and humans and the in vivo characteristics of polymeric nanoparticles—biodistribution, tumor-targeting efficacy, and therapeutic efficacy—are monitored by various imaging modalities. In addition, different imaging probe-labeled polymeric particles have presented distinguishing properties, such as prolonged plasma half-lives, enhanced stability, reduced toxicity, and improved target specificity, compared to small molecular weight imaging probes. In order to make theragnostic and multifunctional nanoparticles, various polymeric nanoparticles were chemically or physically modified with radionuclides (11 C, 18 F, and 64 Cu), Mn2+ , Gd3+ , and optical imaging probes (fluorochromes, quantum dots, and photoproteins). Simultaneously, anticancer agents of hydrophobic chemical drugs, peptides, proteins, and DNAs were encapsulated or chemically grafted to the polymeric nanoparticles. After intravenous injection, multifunctional polymeric nanoparticles exhibited tumortargeting ability with low nonspecific uptake by other tissues. Simultaneously, they could be monitored in real-time and visualized in a noninvasive way, allowing for clinical uses in animals and humans. Radionuclide imaging is commonly divided into two general modalities: single photon emission computed tomography (SPECT) and positron emission tomography (PET) [42]. For PET applications, first-choice positron-emitting radionuclides are 18 F, 11 C, 13 N, and 15 O. These elements are able to emit a positive electron from the nucleus. In particular, the principal applied radionuclide for PET is Fluor-18 (18 F), which is known for its ideal
P1: OTA/XYZ P2: ABC c22 JWBS037-Chen January 25, 2011
13:25
Printer Name: Yet to Come
MULTIFUNCTIONAL NANOPARTICLES FOR CANCER THERAGNOSIS
549
half-life (1.83 h) [43]. Recently, Weissleder and co-workers reported the synthesis and in vivo characterization of an 18 F modified trimodal nanoparticle (18 F-CLIO) [44]. First, 18 FPEG3 was prepared using Synthra RN Plus automated synthesis. Second, 18 F-PEG3 was added to azido-CLIO. The reaction mixture was incubated for 40 min via a “Click” reaction and filtered to remove the 18 F-CLIO from unreacted 18 F-PEG3 [45].Purified 18 F-CLIO was injected into the tail vein of three mice, and dynamic PET imaging was performed to determine organ biodistribution and blood clearance over time. Figure 22.4 demonstrates one imaging series in which data were collected over the course of 16 h after intravenous injection of 18 F-CLIO and shows images acquired at 2, 7, and 16 h. Initially, high sustained activity was detected in the blood pool, demonstrated by high PET signal in the cardiac chambers. Over time, this signal declined, while an increase of activity was observed in the liver and spleen. The low renal and bladder uptake emphasizes that 18 F-CLIO is stable in vivo, and the tracer is not excreted via the kidneys. Taken together, these PET data follow the typical pharmacokinetics of CLIO, showing redistribution of the blood pool signal to the liver and lymphatic organs over the course of several hours (Fig. 22.4, d–f). This nanoparticle is biodegradable and had previously been shown to be broken down into elemental components in a weeks to months time frame. A major future application of the above approach is to use nanoparticle platforms for targeted molecular imaging. Nanoparticles allow multivalent attachment of small affinity ligands, vastly increasing avidity. Kwon and co-workers recently reported hydrophobically modified glycol chitosan with hydrophobic bile acids which stably formed self-assembled nanoparticles in aqueous condition. Various anticancer drugs, such as Taxol, cisplatin, camptothecin (CPT), docetaxel, and therapeutic peptide drugs were easily encapsulated into the hydrophobic cores of glycol chitosan nanoparticles due to their stable nanoparticle structure. After encapsulation of anticancer therapeutic agents, these glycol chitosan nanoparticles were modified with near-infrared (NIR) fluorescence imaging probe, Cy5.5, for optical imaging simultaneously [26, 46–49]. In vivo NIR fluorescence images showed that anticancer-agent-encapsulated glycol chitosan nanoparticles presented a prolonged circulation time of 1 week and preferential accumulation in the target tumor area compared to normal tissues. From the optical imaging data, tumor-targeting efficacy of glycol chitosan nanoparticles was four to sevenfold higher than that of other organs, providing decisive evidence that the nanoparticles specifically targeted tumors. It can be emphasized that the glycol chitosan nanoparticles progressively accumulated in the tumor tissues only based on the EPR effect within 24 h and lasted up to 10 days with considerably low uptake in the liver and spleen as compared with that of conventional nanoparticle-based formulations. Therefore it is deduced that cancer theragnosis with multifunctional polymeric nanoparticles presents a promising new strategy in cancer treatment, permitting simultaneous early-stage diagnosis, drug delivery, and real-time monitoring for therapeutic efficacy. Photodynamic therapy (PDT) is a method for cancer treatment that is a combination of photosensitizers and light. When irradiated with light, the photosensitizers produce singlet oxygen or other reactive oxygen species via a transfer of energy to molecular oxygen, which damages adjacent tissues. However, PDT has barriers: limited tumor selectivity of PDT agents and phototoxicity to skin and eyes. Choi et al. [50] recently proposed a proteasemediated PDT (PM-PDT) to avoid unexpected toxicities. Based on the concept of enzymatic cleavable imaging probe, they used PGC polymer conjugated with chlorine e6 (Ce6), as a photosensitizer that exhibits fluorescence signal and singlet oxygen generation (SOG). Normally, they show photosensitizer–photosensitizer self-quenching (Fig. 22.5a). When the
P1: OTA/XYZ P2: ABC c22 JWBS037-Chen January 25, 2011
550
13:25
Printer Name: Yet to Come
MULTIFUNCTIONAL NANOPARTICLES FOR CANCER THERAGNOSIS
(a)
(b)
(c)
(d)
FIGURE 22.4 Dynamic PET/CT imaging of BALB/c mouse after injection with 18 F-CLIO. Fused PET/CT coronal images (a) at 2 h and (b) 16 h after injection of 18 F-CLIO. Three-dimensional fused PET-CT images at (c) 2 h and (d) 16 h postinjection. Arrow (green) indicates blood pool region of interest (ROI) and asterisk indicates liver ROI. (Reproduced with permission from Devaraj et al. [44]. Copyright © 2006 American Chemical Society.)
PGC polymer conjugated with Ce6 contacts with protease CaB, Ce6-conjugated-PGC (Ce6PGC) can be cleaved and induces stronger fluorescence activation. In its native state, SOG of Ce6-PGC was only 13% compared to free Ce6, however, its SOG dramatically increased up to 79% with CaB treatment. In addition, Ce6-PGC showed high accumulation in an HT1080 tumor at 24 h postinjection (Fig. 22.5b), and its fragments were mainly distributed in the cellular cytoplasm. Apoptosis and significant tissue loss were observed in large areas of the tumor, after light illumination. The proposed PM-PDT strategy demonstrates visualization of the target and local drug concentration before therapy, and could be used as a primary anticancer treatment or as an adjuvant to other therapeutic options. However, this PDT approach based on protease cleavage is restricted by the limited tissue penetration of light. However, it illustrates key issues of concern, including a defined biological rationale, safety, pharmacokinetics, more information in preclinical trials, and manufacture on a large scale for commercial challenges. Once optimized, we hope that the multifunctional nanoparticle systems will provide improved treatment for life-threatening and debilitating diseases.
P1: OTA/XYZ P2: ABC c22 JWBS037-Chen January 25, 2011
13:25
Printer Name: Yet to Come
MULTIFUNCTIONAL NANOPARTICLES FOR CANCER THERAGNOSIS
551
(a)
(b)
(c)
(d)
FIGURE 22.5 (a) Schematic representation of PM-PDT strategy. Without protease, Ce6 molecules are self-quenched. After proteolytic cleavage of the peptide backbone, Ce6 and SOG were released and fluorescent properties were exhibited after light excitation. (b) Biodistribution of Ce6 in bilateral flank tumors. (c) Fluorescence images of tumor tissues that were stained after treatment with Ce6conjugated L-PGC and light. (d) TUNEL staining of Ce6-conjugated L-PGC-treated tumors 24 h after light illumination. (Reproduced with permission from Choi et al. [50], Copyright © 2006 American Association for Cancer Research.)
22.3.2 Multifunctional Magnetic Nanoparticles for Cancer Imaging and Therapy Magnetic nanoparticles (MNPs) have become important tools for magnetic resonance imaging (MRI) contrast enhancement in many cancers [51, 52] and have potential in current clinical diagnostic and therapeutic methods. In order to impart stability, biocompatibility, and functionality for targeting, MNPs require surface coating or conjugation of various materials [53]. Commonly, lipids [54], proteins [55], dendrimers [56], PEG [57], and ® ® polysaccharides [58] were used for the MNP’s surface coating. Ferridex , Combidex , ® Resovist , and AMI-288/gerumoxytrol, are clinically used as MNPs, which have been based on dextran or a similar carbohydrate coating. However, these MNPs still have limitations; they are fairly nonspecific for cancer diagnosis, and their signal enhancement is still unsatisfactory. To improve specificity and signal enhancement of MR contrast agents, recently, there has been a growing interest in the development of targeted MNP-based probes for tumor diagnostics. Furthermore, MNPs have been reported as drug carriers for delivering and releasing drugs [59]. Surface modified multifunctional MNPs with targeting moieties and/or therapeutic drugs can be targeted to tumors or other lesions at molecular or cellular events [18, 56, 60–63], and further allow simultaneous diagnosis and treatment of the diseased tissues with more efficient targeting, precise diagnosis, and therapeutic applications. As previously described, MNPs can serve both as contrast enhancement agents in MRI and as drug carriers in controlled drug delivery, targeted at cancer diagnosis and therapy,
P1: OTA/XYZ P2: ABC c22 JWBS037-Chen January 25, 2011
552
13:25
Printer Name: Yet to Come
MULTIFUNCTIONAL NANOPARTICLES FOR CANCER THERAGNOSIS
simultaneously [64, 65]. Conjugation MNPs with therapeutic drugs or targeting moieties are an effective method of theragnosis, which combines diagnosis and therapeutic applications. For example, Kohler et al. [66, 67] fabricated MNPs covalently bound with methotrexate (MTX), a therapeutic drug that can target folate receptor-overexpressing (FR+) cancer cells. The conjugation of MNPs was performed with MTX or MTX-PEG through amidation between COOH of MTX or PEG-MTX and the amine groups on the MNP surface. In addition to a therapeutic effect, MTX was conjugated to PAMAM nanoparticles with folic acid analog to enhance cytotoxicity to FR+ cancer cells [68]. After a 2-h incubation, MTXconjugated MNPs showed higher internalization into FR+ HeLa (ca. 10-fold) and MCF-7 (ca. 20-fold) cells in comparison with normal cells. The chemical conjugation of MTX on MNPs provides stable linkage in the physiological condition. However, after internalization of MTX-conjugated MNPs into FR+ cancer cells, MTX could be released by low pH and intracellular enzymes; the released MTX induces apoptosis in cancer cells. When MTXconjugated MNPs were introduced into normal cells, they greatly reduced the toxicity of MTX, suggesting MTX-conjugated MNPs exhibited severe toxicity against FR+ cancer cells, but did not exhibit side effects in normal tissues. Polymer coating on the surface of MNPs has enhanced stability and biocompatibility in vivo. However, it is possible that the coated polymers can be removed under harsh conditions in vivo and MNPs could aggregate [69, 70]. Recently, thermally crosslinked superparamagnetic iron oxide nanoparticles (TCL-SPIONs) have been developed by using poly(3-(trimethoxysilyl)propyl methacrylate-r-PEG methyl ether methacrylate-r-Nacryloxysuccinimide), which is an antibiofouling PEG copolymer that prevents nonspecific adsorption of plasma protein [71]. The antibiofouling PEG copolymers have a COOH group on TCL-SPIONs, which can be further converted to an amine group as the functional group. Finally, Cy5.5 is conjugated to amine modified TCL-SPION to obtain Cy5.5 dye-labeled TCL-SPIONs for use as a multimodal (MR/fluorescence) imaging probe [72]. Without any targeting moieties, TCL-SPIONs showed high accumulation in the tumor area and demonstrated MRI and fluorescence images. More recently, researchers developed a novel multifunctional drug carrier based on TCL-SPIONs as the imaging agent, DOX@TCL-SPIONs [73]. DOX is a well-known anticancer agent that was successfully incorporated in the polymeric shell of TCL-SPIONs via electrostatic interactions between positively charged DOX and the negatively charged polymer. Loaded DOX was released quicker at mildly acidic pH of tumor areas than at neutral pH of vascular compartments. Thus when DOX@TCL-SPIONs were systemically administered, DOX@TCL-SPIONs showed much lower toxicity in normal organs than free DOX, but showed a therapeutic response in cancer cells. In vivo MRI and fluorescence images indicated that DOX@TCLSPIONs could efficiently detect a Lewis lung carcinoma (LLC) and sufficiently deliver DOX to the tumor tissue—thereby showing excellent anticancer activity. Coupled with drug delivery to tumor areas, antisense oligonucleotides have become a powerful tool for effective antisense therapy. The problems, however, are associated with rapid degradation by exonuclease and poor diffusion across the cell membrane, which must be solved for clinical antisense therapy [74, 75]. Dendrimers are excellent candidates for antisense oligonucleotide delivery agents as nonviral vectors since they are safer, simpler to use, and more easily mass-produced than other viral vectors [76, 77]. To solve the problems in clinical antisense therapy, Pan et al. [78] fabricated polyamidoamine (PAMAM) dendrimer-conjugated MNPs as a gene delivery carrier. The MNP’s surface was coated with 3-aminopropyl-trimethoxysilane (APTS) to make G0 dMNP, and then further added to excessive methacrylate. After rinsing, the methacrylate group of G0 dMNP was further
P1: OTA/XYZ P2: ABC c22 JWBS037-Chen January 25, 2011
13:25
Printer Name: Yet to Come
MULTIFUNCTIONAL NANOPARTICLES FOR CANCER THERAGNOSIS
553
converted to amine by using excessive ethylenediamine. Stepwise growth using methacrylate and ethylenediamine was repeated until the desired number of generations from 1.0 to 5.0 (G1.0 -G5.0 ) was achieved. As the generation of dendrimer increased, adsorption of antisense survivin oligonuelcotide (asOND) also increased on dMNP via electrostatic interactions. The asOND-G5.0 dMNP with more positive charges was highly internalized and crossed tumor cell membranes, thereby inhibiting the growth of tumor cells and showing a strong inhibition effect on expression of survivin gene and protein. This study demonstrated that dMNPs are helpful to protect asOND from degradation by enzymes inside cells as a gene delivery system. The siRNA molecules are short double-stranded nucleic acid molecules that can act as mediators of RNA interference (RNAi) within the cytoplasm of cells. For clinical therapeutic application of siRNAs, efficient delivery methods of siRNAs into the target cells are required due to the limited entrance of naked siRNAs into cells. Various delivery methods have been developed including the use of lipid-based agents [79], antibody–protein fusion proteins [80], liposomes [81], and nanoparticles [82]. For clinical trials, a noninvasive approach of siRNA delivery to target tissues is needed to optimize experimental treatment strategies. In vivo imaging of siRNA is currently confined to bioluminescence imaging of siRNA-mediated silencing [83]. Recently, a multifunctional MNP was reported as a noninvasive dual-modality imaging probe for the simultaneous in vivo transfer of siRNA and imaging of siRNA accumulation in tumors by both MRI and NIRF optical imaging [84]. To achieve this, siRNA molecules as the therapeutic target gene are covalently linked to Cy5.5-labeled MNPs, and further modified with myristoylated polyarginine peptide (MPAP, membrane translocation peptide) for successful intracellular delivery [85]. The results demonstrated in vivo tracking of these multifunctional MNPs in the tumor by MR and NIRF imaging; thus they could simultaneously deliver and detect siRNA-based therapeutic agents in vivo. However, it is also suggested that detailed investigation is further needed to clarify the precise mechanisms mediating the efficient gene silencing in vivo. MNPs have been used as clinical tools for numerous diseases in the MRI. In the near future, a new generation of multifunctional MNPs combined therapeutic drugs, targeting moieties, and MR contrast agents will allow for the investigation of diseases across a number of platforms and accumulation of vast amounts of information in clinics. 22.3.3 Theragnostic Inorganic Nanoparticles for Cancer Theragnosis Quantum dots (QDs) as semiconductor nanocrystals have been increasingly applied as biological imaging and labeling probes [86–88]. QDs have several advantages over conventional organic fluorophores: narrow photoluminescence spectra, high quantum yield, low photobleaching, and resistance to chemical degradation [89]. In addition, QDs have intrinsic fluorescence emission spectra wavelengths (400–2000 nm) that depend on their size and composition [90]. Appropriate modification and optimization with targeting moieties (e.g., antibodies, aptamers, peptides) has resulted in more sensitive and specific targeted imaging and diagnostic modalities [91, 92]. More recently, QDs have been designed to carry therapeutic agents for simultaneous imaging and therapeutic applications [93]. These combined imaging and therapy nanoparticles represent an exciting advance in the field of nanomedicine. In this section, we describe several multifunctional QD nanoparticles for simultaneous cancer imaging and therapy. For simultaneous imaging and therapeutic uses, new smart multifunctional QD nanoparticles have recently been developed that are capable of simply sensing the release of
P1: OTA/XYZ P2: ABC c22 JWBS037-Chen January 25, 2011
554
13:25
Printer Name: Yet to Come
MULTIFUNCTIONAL NANOPARTICLES FOR CANCER THERAGNOSIS
therapeutics by a change in the fluorescence of the imaging modality. Bagalkot et al. [94] designed a QD–aptamer (QD-Apt) conjugate that could concurrently deliver anticancer drugs to prostate cancer (PCa) cells and sense drug delivery based on a fluorescence resonance energy transfer (FRET) mechanism. The QD-Apt (DOX) system was prepared by functionalizing the surface of fluorescent QDs (donor) with an aptamer—targeting unit— that recognized the prostate specific membrane antigen (PSMA) expressed in LNCaP cells. Also, DOX was intercalated into the double-stranded RNA aptamers. In this system, the QD fluorescence was quenched by DOX and the DOX fluorescence was quenched by the double-stranded RNA aptamers. Therefore reporting from this construct was dependent on two donor–quencher pairs, where the fluorescence was quenched through FRET. Upon administration, the targeted cancer cells internalized the particles and DOX was gradually released into the cells, which switched the fluorescence of the QD and DOX to the “ON” state. These multifunctional nanoparticles simultaneously delivered DOX to PSMAexpressing LNCaP cells and the efficacy of DOX delivery was monitored by activating QD fluorescence. Although this mechanism significantly enhanced therapeutic efficacy against the targeted LNCaP cells compared to nontargeted PC3 cells, it is clear that these QD-Apt (DOX) multifunctional nanoparticles should be further evaluated in vivo prior to additional medical applications. It has been reported that QDs in PDT can act as either photosensitizers themselves or activators of another photosensitizer by serving as the energy donor [95–97].The energy transfer between QDs and cell molecules (as triple oxygen, reducing equivalents, pigments) potentially could generate reactive oxygen species to provoke apoptosis in cells. As a good photosensitizer, QDs have several characteristics, such as a constant composition, simple and inexpensive synthesis, and no toxicity in the absence of light but potential cytotoxicity under UV irradiation. Unlike the visible emission of most conventional photosensitizers, QDs can be tuned to emit in the NIR region, which can be useful in PDT for deep-seated tumors since NIR light is not scattered and absorbed by tissue [97]. Even though QDs can generate singlet oxygen, the quantum yield of QD-generated singlet oxygen is very low (less than 5%) compared to classic photosensitizer (40–60%) [95, 98]. To reach a comparatively high steady-state level of singlet oxygen and induce apoptosis in target tissues, prolonged and repetitive exposure to light irradiation is required. This problem could be overcome by conjugating QDs with a phthalocyanine (PC4) photosensitizer [95, 98]. PC4 can be directly activated for excitation wavelengths between 550 and 630 nm. However, conjugation of QDs allowed indirect excitation of PC4 at 488 nm through a fluorescence resonance energy transfer from the QD to PC and increased the photosensitizing power; thus PC4 emission could be observed at 680 nm. This study indicated that QDs can also be used to activate other established photosensitizers and are no longer restricted to photosensitizers themselves. Metallic gold nanomaterials have been extensively studied for potential applications in the emerging and highly interdisciplinary field of nanotechnology [99–104]. Gold nanomaterials have several attractive characteristics for diagnostic applications, including: (1) biocompatibility and stability, (2) unique tunable optical properties, and (3) easy conjugation of biomolecules (i.e., antibody) to the surface for tumor-specific targeting. Thus many biodiagnostic applications of gold nanomaterials have been developed for bioassays or colorimetric assays to quantify protein [101, 103], polynucleotides [102], and bacteria [105]. However, the application of gold nanomaterials for therapeutic purposes is largely an undeveloped field. Two characteristics of gold nanomaterials make them particularly suitable for therapeutic applications: antibodies and other bioactive molecules can easily be
P1: OTA/XYZ P2: ABC c22 JWBS037-Chen January 25, 2011
13:25
Printer Name: Yet to Come
MULTIFUNCTIONAL NANOPARTICLES FOR CANCER THERAGNOSIS
555
conjugated to the surface of gold nanomaterials, and gold nanomaterials have absorption efficiencies that are four to five orders of magnitude greater than conventional photothermal dyes and are not affected by photobleaching [106]. More complex shapes with excitation wavelengths of 800–1200 nm could absorb the NIR light and then convert it to heat. Once gold nanomaterials are densely accumulated at the target site, they are activated via the absorption of irradiation of an appropriate wavelength, thereby causing irreversible thermal cellular destruction [106–108]. This property can provide an opportunity for therapeutic treatment in deep tissues. In this section, we discuss the NIR-absorbing gold nanomaterials for photothermal therapy. The first trial of gold nanoshells in hyperthermal therapy was introduced by Hirsch et al. [107]. The average temperature of tumor cells treated with gold nanoshells and NIR light increased up to 38 ◦ C at a depth of 2.5 mm beneath the dermal surface, and irreversible photothermal destruction was observed and confined to the tumor area. In the follow-up work by Halas and co-workers, gold nanoshells were conjugated with HER2 antibody to actively target breast carcinoma cells [109]. In this study, NIR irradiation caused a rise in the temperature of the target regions between 40 and 50 ◦ C, which selectively destroyed the carcinomas. In addition, the survival rate of mice treated with HER–gold nanoshells and NIR irradiation was excellent compared with controls (nonspecific antibody or NIR light alone). Optical coherence tomography (OCT) enables high-resolution and crosssectional subsurface imaging of biological tissue with micron-scale resolution [110]. Gold nanoshells are suitable as OCT imaging contrast agents because of their biocompatibility, tunable optical properties, and small size, yielding easily detected backscattering of NIR light [111]. Based on these results, Gobin and co-workers combined OCT imaging with NIR treatment using NIR absorbing and scattering nanoshells (ca. 120 nm in diameter, 12 nm shell thickness) in an in vivo model [112]. After intravenous injection into colon carcinoma (CT-26)-bearing mice, PEG–gold nanoshells passively accumulated in colon tumor tissues. Nearly 12.5 ppm of nanoshells per gram of tumor tissues was determined by neutron activation analysis, compared to 0 ppm for tumors treated with PBS alone. Due to the accumulation of nanoshells, there was an enhanced brightness in the image of the tumor tissue. However, no enhancement was observed in the normal tissue of mice treated with nanoshells, meaning that nanoshells do not extravasate into normal tissue. After treatment with nanoshells, OCT imaging, and irradiation with an NIR laser, tumors were greatly reduced in size, and the median survival of mice was significantly greater than either the control or PBS sham groups. Radiation therapy is an approach to the treatment of many tumors. However, radiation therapy alone cannot eradicate all regional cancers because of the intrinsic resistance of some cancers to ionizing irradiation [113]. In particular, intratumoral hypoxia changes tumor cellular and microenvironmental conditions, resulting in tumor aggressiveness and resistance to radiation therapy [114]. An alternative strategy to overcome radioresistance and enhance radiation therapy is mild temperature hyperthermia, which has direct antitumor effects and tumor microenvironment effects, mediated through mitigation of hypoxia, that contribute to an increased therapeutic index of radiation therapy [115, 116]. However, this approach is underutilized in routine clinical practice because of the invasive means of achieving and maintaining hyperthermia, the time commitment involved in a treatment that lasts about an hour, the lack of good thermal dosimetry, and the inability to achieve localized hyperthermic temperature [117]. More recently, Diagaradjane et al. [118] suggested the use of optically activated gold nanoshells as a novel and minimally invasive method to induce mild temperature hyperthermia. To modulate in vivo tumor radiation response
P1: OTA/XYZ P2: ABC c22 JWBS037-Chen January 25, 2011
556
13:25
Printer Name: Yet to Come
MULTIFUNCTIONAL NANOPARTICLES FOR CANCER THERAGNOSIS
using gold nanoshell-mediated hyperthermia, they designed gold nanoshells that have a 120-nm core diameter and a 12–15-nm thick shell, and showed an optical absorption peak between 780 and 800 nm. Thiolated polyethylene glycol (SH-PEG) was further immobilized onto the surface of the gold nanoshells to enhance circulation time in the blood. After the injection of gold nanoshells in human colorectal cancer (HCT 116)bearing mice via the tail vein, localized hyperthermia was carried out 20–24 h postinjection. Irradiation with a laser power setting of 0.6 W achieved a T s of ∼10 ± 1.5 ◦ C and ∼8 ± 0.5 ◦ C in the tumor core and base, respectively, whereas irradiation without gold nanoshells yielded a T s of ∼2.5 ± 3.5 ◦ C in the tumor core, the base of the tumor, and in irradiated muscle of the contralateral thigh. The temperature difference (T) of ∼11 ◦ C demonstrated by real-time magnetic resonance temperature imaging (MRTI) is consistent with the thermocouple measurements. Indeed, the T1-weighted images showed a temperature distribution in tumor tissue that corresponded to the exposure time after laser illumination. To confirm whether gold nanoshell-mediated hyperthermia enhances the efficacy of radiation therapy, they performed a tumor regrowth delay assay. Tumor regrowth delays of approximately 4, 9, 17, and 29 days for the control, hyperthermia, radiation, and thermoradiotherapy groups were observed, respectively. Histological data revealed that large necrotic regions were observed in the tumor core after thermoradiotherapy. These results indicate that hyperthermia mediated by gold nanoshells markedly improved the therapeutic index of radiation therapy. By changing the shape of gold nanoparticles to nanorods, the absorption and scattering wavelength changes from the visible to the NIR region and their absorption and scattering cross sections also increase. Indeed, the surface plasmon field absorption of gold nanorods is the strongest of all the different shapes of gold and silver nanoparticles [119, 120]. Unlike gold nanospheres with one visible absorption band around 520 nm, gold nanorods showed two surface plasmon absorption bands: a strong long-wavelength band due to the longitudinal oscillation of electrons and a weak short-wavelength band around 520 nm due to the transverse electronic oscillation. Based on the increase in the rod’s aspect ratio, the longitudinal absorption band shifts from the visible to the NIR region. For this reason, gold nanorods can be used as a novel contrast agent for dual molecular imaging using simple dark-field microscopy and selective photothermal therapy using an NIR laser. For selective photothermal therapy using an NIR laser, Huang et al. [121] have prepared antiepidermal growth factor receptor (anti-EGFR) antibody-conjugated gold nanorods (aspect ratio, 3.9; 800-nm absorption band). When nanorods were incubated into cancer cells, their cellular uptake was twice as high in the two malignant cells (HOC 313 clone 8 and HSC 3) compared to the nonmalignant cells (HaCaT) due to the targeting of EGFR on the malignant cell surface. Under irradiation by the NIR laser (800 nm), photothermal destruction was observed in both malignant and nonmalignant cells. However, increased uptake of the nanorods by the two EGFR-overexpressing malignant cells reduced the laser energy required to cause death in normal cells.
22.4 PROSPECTS AND CHALLENGES OF CANCER THERAGNOSIS Great steps are being taken every day in molecular imaging and nanotechnology toward changing the scale and methods of theragnosis for a better future of cancer therapy and imaging. Theragnosis is a multidisciplinary field undergoing rapid expansion with the promise of new developments in medicine, chemistry, molecular imaging, and nanotechnology.
P1: OTA/XYZ P2: ABC c22 JWBS037-Chen January 25, 2011
13:25
Printer Name: Yet to Come
REFERENCES
557
Cancer theragnosis is the next-generation in cancer therapy and combines early-stage diagnosis and efficient therapy. As demonstrated above, theragnosis is based on multifunctional nanoparticles, due to the lack of tumor selectivity and relatively low therapeutic index of general imaging probes and therapeutic agents. Different nanoparticle-based carriers have been elaborated for passive or active targeting of cancer with various therapeutic agents and imaging probes in spite of the fact that not all of these targeting mechanisms are actually clearly understood. Thus the approach with nanoparticles for specific diseases needs more research on the molecular mechanism of the cancer area, to allow specific targeting of a disease area and early-stage diagnosis. Another concern is the lack of precise diagnosis relating to the imaging agents. In particular, various multimodal imaging probes have been tested to overcome the limitations of either imaging method when used alone. Novel multifunctional nanoparticle design with specific disease targeting in addition to multimodal imaging ability will significantly contribute to the improvement of both disease imaging and therapy.
REFERENCES 1. Duncan, R. The dawning era of polymer therapeutics. Nat. Rev. Drug Discov. 2003, 2, 347–360. 2. Ferrari, M. Cancer nanotechnology: opportunities and challenges. Nat. Rev. Cancer 2005, 5, 161–171. 3. Peer, D.; Karp, J. M.; Hong, S.; Farokhzad, O. C.; Margalit, R.; Langer, R. Nanocarriers as an emerging platform for cancer therapy. Nat. Nanotechnol. 2007, 2, 751–760. 4. Torchilin, V. P. Micellar nanocarriers: pharmaceutical perspectives. Pharm. Res. 2007, 24, 1–16. 5. Gros, L.; Ringsdorf, H.; Schupp, H. Polymeric anti-tumor agents on a molecular and on a cellular-level. Angew. Chem. Int. Ed. Engl. 1981, 20, 305–325. 6. Jones, M.; Leroux, J. Polymeric micelles—a new generation of colloidal drug carriers. Eur. J. Pharm. Biopharm. 1999, 48, 101–111. 7. Torchilin, V. P. Structure and design of polymeric surfactant-based drug delivery systems. J. Control. Release 2001, 73, 137–172. 8. Matsumura, Y.; Maeda, H. A new concept for macromolecular therapeutics in cancer chemotherapy: mechanism of tumoritropic accumulation of proteins and the antitumor agent smancs. Cancer Res. 1986, 46, 6387–6392. 9. Maeda, H. The enhanced permeability and retention (EPR) effect in tumor vasculature: the key role of tumor-selective macromolecular drug targeting. Adv. Enzyme Regul. 2001, 41, 189–207. 10. Papahadjopoulos, D.; Allen, T. M.; Gabizon, A.; Mayhew, E.; Matthay, K.; Huang, S. K.; et al. Sterically stabilized liposomes: improvements in pharmacokinetics and antitumor therapeutic efficacy. Proc. Natl. Acad. Sci. U.S.A. 1991, 88, 11460–11464. 11. Matsumura, Y.; Hamaguchi, T.; Ura, T.; Muro, K.; Yamada, Y.; Shimada, Y.; et al. Phase I clinical trial and pharmacokinetic evaluation of NK911, a micelle-encapsulated doxorubicin. Br. J. Cancer 2004, 91, 1775–1781. 12. Elliott, R. L.; Elliott, M. C.; Wang, F.; Head, J. F. Breast carcinoma and the role of iron metabolism. A cytochemical, tissue culture, and ultrastructural study. Ann. N. Y. Acad. Sci. 1993, 698, 159–166. 13. Goren, D.; Horowitz, A. T.; Tzemach, D.; Tarshish, M.; Zalipsky, S.; Gabizon, A. Nuclear delivery of doxorubicin via folate-targeted liposomes with bypass of multidrug-resistance efflux pump. Clin. Cancer Res. 2000, 6, 1949–1957.
P1: OTA/XYZ P2: ABC c22 JWBS037-Chen January 25, 2011
558
13:25
Printer Name: Yet to Come
MULTIFUNCTIONAL NANOPARTICLES FOR CANCER THERAGNOSIS
14. Dhar, S.; Gu, F. X.; Langer, R.; Farokhzad, O. C.; Lippard, S. J. Targeted delivery of cisplatin to prostate cancer cells by aptamer functionalized Pt(IV) prodrug-PLGA-PEG nanoparticles. Proc. Natl. Acad. Sci. U.S.A. 2008, 105, 17356–17361. 15. Ozdemir, V.; Williams-Jones, B.; Glatt, S. J.; Tsuang, M. T.; Lohr, J. B.; Reist, C. Shifting emphasis from pharmacogenomics to theragnostics. Nat. Biotechnol. 2006, 24, 942–946. 16. Blanco, E.; Kessinger, C. W.; Sumer, B. D.; Gao, J. Multifunctional micellar nanomedicine for cancer therapy. Exp. Biol. Med. (Maywood) 2009, 234, 123–131. 17. Brigger, I.; Dubernet, C.; Couvreur, P. Nanoparticles in cancer therapy and diagnosis. Adv. Drug Deliv. Rev. 2002, 54, 631–651. 18. Lee, J. H.; Huh, Y. M.; Jun, Y. W.; Seo, J. W.; Jang, J. T.; Song, H. T.; et al. Artificially engineered magnetic nanoparticles for ultra-sensitive molecular imaging. Nat. Med. 2007, 13, 95–99. 19. Jain, R. K. Transport of molecules in the tumor interstitium: a review. Cancer Res. 1987, 47, 3039–3051. 20. Baban, D. F.; Seymour, L. W. Control of tumour vascular permeability. Adv. Drug Deliv. Rev. 1998, 34, 109–119. 21. Yokoyama, M.; Okano, T.; Sakurai, Y.; Ekimoto, H.; Shibazaki, C.; Kataoka, K. Toxicity and antitumor activity against solid tumors of micelle-forming polymeric anticancer drug and its extremely long circulation in blood. Cancer Res. 1991, 51, 3229–3236. 22. Yuan, F.; Dellian, M.; Fukumura, D.; Leunig, M.; Berk, D. A.; Torchilin, V. P.; et al. Vascular permeability in a human tumor xenograft: molecular size dependence and cutoff size. Cancer Res. 1995, 55, 3752–3756. 23. Hobbs, S. K.; Monsky, W. L.; Yuan, F.; Roberts, W. G.; Griffith, L.; Torchilin, V. P.; et al. Regulation of transport pathways in tumor vessels: role of tumor type and microenvironment. Proc. Natl. Acad. Sci. U.S.A. 1998, 95, 4607–4612. 24. Hashizume, H.; Baluk, P.; Morikawa, S.; McLean, J. W.; Thurston, G.; Roberge, S.; et al. Openings between defective endothelial cells explain tumor vessel leakiness. Am. J. Pathol. 2000, 156, 1363–1380. 25. Moghimi, S. M.; Hunter, A. C.; Murray, J. C. Nanomedicine: current status and future prospects. FASEB. J. 2005, 19, 311–330. 26. Kim, J. H.; Kim, Y. S.; Kim, S.; Park, J. H.; Kim, K.; Choi, K.; et al. Hydrophobically modified glycol chitosan nanoparticles as carriers for paclitaxel. J. Control. Release 2006, 111, 228– 234. 27. Park, J. H.; Kwon, S.; Lee, M.; Chung, H.; Kim, J. H.; Kim, Y. S.; et al. Self-assembled nanoparticles based on glycol chitosan bearing hydrophobic moieties as carriers for doxorubicin: in vivo biodistribution and anti-tumor activity. Biomaterials 2006, 27, 119–126. 28. Silletti, S.; Kessler, T.; Goldberg, J.; Boger, D. L.; Cheresh, D. A. Disruption of matrix metalloproteinase 2 binding to integrin alpha vbeta 3 by an organic molecule inhibits angiogenesis and tumor growth in vivo. Proc. Natl. Acad. Sci. U.S.A. 2001, 98, 119–124. 29. Torchilin, V. P.; Lukyanov, A. N.; Gao, Z.; Papahadjopoulos-Sternberg, B. Immunomicelles: targeted pharmaceutical carriers for poorly soluble drugs. Proc. Natl. Acad. Sci. U.S.A. 2003, 100, 6039–6044. 30. Jeong, Y. I.; Seo, S. J.; Park, I. K.; Lee, H. C.; Kang, I. C.; Akaike, T.; et al. Cellular recognition of paclitaxel-loaded polymeric nanoparticles composed of poly(gamma-benzyl l-glutamate) and poly(ethylene glycol) diblock copolymer endcapped with galactose moiety. Int. J. Pharm. 2005, 296, 151–161. 31. Ross, J. F.; Chaudhuri, P. K.; Ratnam, M. Differential regulation of folate receptor isoforms in normal and malignant tissues in vivo and in established cell lines. Physiologic and clinical implications. Cancer 1994, 73, 2432–2443.
P1: OTA/XYZ P2: ABC c22 JWBS037-Chen January 25, 2011
13:25
Printer Name: Yet to Come
REFERENCES
559
32. Sethuraman, V. A.; Bae, Y. H. TAT peptide-based micelle system for potential active targeting of anti-cancer agents to acidic solid tumors. J. Control. Release 2007, 118, 216–224. 33. Sun, C.; Lee, J. S.; Zhang, M. Magnetic nanoparticles in MR imaging and drug delivery. Adv. Drug Deliv. Rev. 2008, 60, 1252–1265. 34. Perez, J. M.; Josephson, L.; O’Loughlin, T.; Hogemann, D.; Weissleder, R. Magnetic relaxation switches capable of sensing molecular interactions. Nat. Biotechnol. 2002, 20, 816–820. 35. Nahrendorf, M.; Zhang, H.; Hembrador, S.; Panizzi, P.; Sosnovik, D. E.; Aikawa, E.; et al. Nanoparticle PET-CT imaging of macrophages in inflammatory atherosclerosis. Circulation. 2008, 117, 379–387. 36. Burton, J. B.; Johnson, M.; Sato, M.; Koh, S. B.; Mulholland, D. J.; Stout, D.; et al. Adenovirusmediated gene expression imaging to directly detect sentinel lymph node metastasis of prostate cancer. Nat. Med. 2008, 14, 882–888. 37. Segal, E.; Sirlin, C. B.; Ooi, C.; Adler, A. S.; Gollub, J.; Chen, X.; et al. Decoding global gene expression programs in liver cancer by noninvasive imaging. Nat. Biotechnol. 2007, 25, 675–680. 38. Tallury, P.; Payton, K.; Santra, S. Silica-based multimodal/multifunctional nanoparticles for bioimaging and biosensing applications. Nanomedicine 2008, 3, 579–592. 39. Klohs, J.; Wunder, A.; Licha, K. Near-infrared fluorescent probes for imaging vascular pathophysiology. Basic. Res. Cardiol. 2008, 103, 144–151. 40. Licha, K.; Hessenius, C.; Becker, A.; Henklein, P.; Bauer, M.; Wisniewski, S.; et al. Synthesis, characterization, and biological properties of cyanine-labeled somatostatin analogues as receptor-targeted fluorescent probes. Bioconjug. Chem. 2001, 12, 44–50. 41. Gref, R.; Minamitake, Y.; Peracchia, M. T.; Trubetskoy, V.; Torchilin, V.; Langer, R. Biodegradable long-circulating polymeric nanospheres. Science 1994, 263, 1600–1603. 42. Hamoudeh, M.; Kamleh, M. A.; Diab, R.; Fessi, H. Radionuclides delivery systems for nuclear imaging and radiotherapy of cancer. Adv. Drug Deliv. Rev. 2008, 60, 1329–1346. 43. Kirby, A. M.; George, Mikhaeel N. The role of FDG PET in the management of lymphoma: practical guidelines. Nucl. Med. Commun. 2007, 28, 355–357. 44. Devaraj, N. K.; Keliher, E. J.; Thurber, G. M.; Nahrendorf, M.; Weissleder, R. 18 F labeled nanoparticles for in vivo PET-CT imaging. Bioconjug. Chem. 2009, 20, 397–401. 45. Lewis, W. G.; Magallon, F. G.; Fokin, V. V.; Finn, M. G. Discovery and characterization of catalysts for azide-alkyne cycloaddition by fluorescence quenching. J. Am. Chem. Soc. 2004, 126, 9152–9153. 46. Park, K.; Kim, J. H.; Nam, Y. S.; Lee, S.; Nam, H. Y.; Kim, K.; et al. Effect of polymer molecular weight on the tumor targeting characteristics of self-assembled glycol chitosan nanoparticles. J. Control. Release 2007, 122, 305–314. 47. Min, K. H.; Park, K.; Kim, Y. S.; Bae, S. M.; Lee, S.; Jo, H. G.; et al. Hydrophobically modified glycol chitosan nanoparticles-encapsulated camptothecin enhance the drug stability and tumor targeting in cancer therapy. J. Control. Release 2008, 127, 208–218. 48. Kim, J. H.; Kim, Y. S.; Park, K.; Lee, S.; Nam, H. Y.; Min, K. H.; et al. Antitumor efficacy of cisplatin-loaded glycol chitosan nanoparticles in tumor-bearing mice. J. Control. Release 2008, 127, 41–49. 49. Kim, J. H.; Kim, Y. S.; Park, K.; Kang, E.; Lee, S.; Nam, H. Y.; et al. Self-assembled glycol chitosan nanoparticles for the sustained and prolonged delivery of antiangiogenic small peptide drugs in cancer therapy. Biomaterials 2008, 29, 1920–1930. 50. Choi, Y.; Weissleder, R.; Tung, C. H. Selective antitumor effect of novel protease-mediated photodynamic agent. Cancer Res. 2006, 66, 7225–7229. 51. Schellenberger, E. A.; Reynolds, F.; Weissleder, R.; Josephson, L. Surface-functionalized nanoparticle library yields probes for apoptotic cells. Chembiochem 2004, 5, 275–279.
P1: OTA/XYZ P2: ABC c22 JWBS037-Chen January 25, 2011
560
13:25
Printer Name: Yet to Come
MULTIFUNCTIONAL NANOPARTICLES FOR CANCER THERAGNOSIS
52. Peer, A.; Strauss, S.; Witz, E.; Manor, H.; Eidelman, A. Use of percutaneous nephrostomy in hydronephrosis of pregnancy. Eur. J. Radiol. 1992, 15, 220–223. 53. Qin, J.; Laurent, S.; Jo, Y. S.; Roch, A.; Mikhaylova, M.; Bhujwalla, Z. M.; et al. A high-performance magnetic resonance imaging T2 contrast agent. Adv. Mater. 2007, 19, 1874–1878. 54. Nitin, N.; LaConte, L. E.; Zurkiya, O.; Hu, X.; Bao, G. Functionalization and peptide-based delivery of magnetic nanoparticles as an intracellular MRI contrast agent. J. Biol. Inorg. Chem. 2004, 9, 706–712. 55. Wilhelm, C.; Billotey, C.; Roger, J.; Pons, J. N.; Bacri, J. C.; Gazeau, F. Intracellular uptake of anionic superparamagnetic nanoparticles as a function of their surface coating. Biomaterials. 2003, 24, 1001–1011. 56. Bulte, J. W.; Douglas, T.; Witwer, B.; Zhang, S. C.; Strable, E.; Lewis, B. K.; et al. Magnetodendrimers allow endosomal magnetic labeling and in vivo tracking of stem cells. Nat. Biotechnol. 2001, 19, 1141–1147. 57. Illum, L.; Church, A. E.; Butterworth, M. D.; Arien, A.; Whetstone, J.; Davis, S. S. Development of systems for targeting the regional lymph nodes for diagnostic imaging: in vivo behaviour of colloidal PEG-coated magnetite nanospheres in the rat following interstitial administration. Pharm. Res. 2001, 18, 640–645. 58. Kellar, K. E.; Fujii, D. K.; Gunther, W. H.; Briley-Saebo, K.; Bjornerud, A.; Spiller, M.; et al. NC100150 Injection, a preparation of optimized iron oxide nanoparticles for positive-contrast MR angiography. J. Magn. Reson. Imaging 2000, 11, 488–494. 59. Kircher, M. F.; Mahmood, U.; King, R. S.; Weissleder, R.; Josephson, L. A multimodal nanoparticle for preoperative magnetic resonance imaging and intraoperative optical brain tumor delineation. Cancer Res. 2003, 63, 8122–8125. 60. Weissleder, R.; Moore, A.; Mahmood, U.; Bhorade, R.; Benveniste, H.; Chiocca, E. A.; et al. In vivo magnetic resonance imaging of transgene expression. Nat. Med. 2000, 6, 351–355. 61. Kang, H. W.; Josephson, L.; Petrovsky, A.; Weissleder, R.; Bogdanov, A., Jr. Magnetic resonance imaging of inducible E-selectin expression in human endothelial cell culture. Bioconjug. Chem. 2002, 13, 122–127. 62. Zhao, M.; Beauregard, D. A.; Loizou, L.; Davletov, B.; Brindle, K. M. Non-invasive detection of apoptosis using magnetic resonance imaging and a targeted contrast agent. Nat. Med. 2001, 7, 1241–1244. 63. Huh, Y. M.; Jun, Y. W.; Song, H. T.; Kim, S.; Choi, J. S.; Lee, J. H.; et al. In vivo magnetic resonance detection of cancer by using multifunctional magnetic nanocrystals. J. Am. Chem. Soc. 2005, 127, 12387–12391. 64. Alexiou, C.; Arnold, W.; Klein, R. J.; Parak, F. G.; Hulin, P.; Bergemann, C.; et al. Locoregional cancer treatment with magnetic drug targeting. Cancer Res. 2000, 60, 6641–6648. 65. Bulte, J. W.; Hoekstra, Y.; Kamman, R. L.; Magin, R. L.; Webb, A. G.; Briggs, R. W.; et al. Specific MR imaging of human lymphocytes by monoclonal antibody-guided dextran-magnetite particles. Magn. Reson. Med. 1992, 25, 148–157. 66. Kohler, N.; Sun, C.; Wang, J.; Zhang, M. Methotrexate-modified superparamagnetic nanoparticles and their intracellular uptake into human cancer cells. Langmuir 2005, 21, 8858–8864. 67. Kohler, N.; Sun, C.; Fichtenholtz, A.; Gunn, J.; Fang, C.; Zhang, M. Methotrexate-immobilized poly(ethylene glycol) magnetic nanoparticles for MR imaging and drug delivery. Small 2006, 2, 785–792. 68. Duthie, S. J. Folic-acid-mediated inhibition of human colon-cancer cell growth. Nutrition 2001, 17, 736–737. 69. Zhang, Y.; Kohler, N.; Zhang, M. Surface modification of superparamagnetic magnetite nanoparticles and their intracellular uptake. Biomaterials 2002, 23, 1553–1561.
P1: OTA/XYZ P2: ABC c22 JWBS037-Chen January 25, 2011
13:25
Printer Name: Yet to Come
REFERENCES
561
70. Moghimi, S. M.; Hunter, A. C.; Murray, J. C. Long-circulating and target-specific nanoparticles: theory to practice. Pharmacol. Rev. 2001, 53, 283–318. 71. Lee, H.; Lee, E.; Kim, do K.; Jang, N. K.; Jeong, Y. Y.; Jon, S. Antibiofouling polymer-coated superparamagnetic iron oxide nanoparticles as potential magnetic resonance contrast agents for in vivo cancer imaging. J. Am. Chem. Soc. 2006, 128, 7383–7389. 72. Lee, H.; Yu, M. K.; Park, S.; Moon, S.; Min, J. J.; Jeong, Y. Y.; et al. Thermally cross-linked superparamagnetic iron oxide nanoparticles: synthesis and application as a dual imaging probe for cancer in vivo. J. Am. Chem. Soc. 2007, 129, 12739–12745. 73. Yu, M. K.; Jeong, Y. Y.; Park, J.; Park, S.; Kim, J. W.; Min, J. J.; et al. Drug-Loaded Superparamagnetic Iron Oxide Nanoparticles for Combined Cancer Imaging and Therapy In Vivo. Angew. Chem. Int. Ed. Engl. 2008, 47, 5362–5365. 74. Denhardt, D. T. Mechanism of action of antisense RNA. Sometime inhibition of transcription, processing, transport, or translation. Ann. N. Y. Acad. Sci. 1992, 660, 70–76. 75. Tarkanyi, I.; Horvath, A.; Szatmari, I.; Eizert, H.; Vamosi, G.; Damjanovich, S.; et al. Inhibition of human telomerase by oligonucleotide chimeras, composed of an antisense moiety and a chemically modified homo-oligonucleotide. FEBS. Lett. 2005, 579, 1411–1416. 76. Majoros, I. J.; Myc, A.; Thomas, T.; Mehta, C. B.; Baker, J. R, Jr. PAMAM dendrimer-based multifunctional conjugate for cancer therapy: synthesis, characterization, and functionality. Biomacromolecules 2006, 7, 572–579. 77. Radu, D. R.; Lai, C. Y.; Jeftinija, K.; Rowe, E. W.; Jeftinija, S.; Lin, V. S. A polyamidoamine dendrimer-capped mesoporous silica nanosphere-based gene transfection reagent. J. Am. Chem. Soc. 2004, 126, 13216–13217. 78. Pan, B.; Cui, D.; Sheng, Y.; Ozkan, C.; Gao, F.; He, R.; et al. Dendrimer-modified magnetic nanoparticles enhance efficiency of gene delivery system. Cancer Res. 2007, 67, 8156–8163. 79. Landen, C. N., Jr., Chavez-Reyes, A.; Bucana, C.; Schmandt, R.; Deavers, M. T.; LopezBerestein, G.; et al. Therapeutic EphA2 gene targeting in vivo using neutral liposomal small interfering RNA delivery. Cancer Res. 2005, 65, 6910–6918. 80. Song, E.; Zhu, P.; Lee, S. K.; Chowdhury, D.; Kussman, S.; Dykxhoorn, D. M.; et al. Antibody mediated in vivo delivery of small interfering RNAs via cell-surface receptors. Nat. Biotechnol. 2005, 23, 709–717. 81. Zimmermann, T. S.; Lee, A. C.; Akinc, A.; Bramlage, B.; Bumcrot, D.; Fedoruk, M. N.; et al. RNAi-mediated gene silencing in non-human primates. Nature 2006, 441, 111–114. 82. Bartlett, D. W.; Davis, M. E. Physicochemical and biological characterization of targeted, nucleic acid-containing nanoparticles. Bioconjug. Chem. 2007, 18, 456–468. 83. Takeshita, F.; Minakuchi, Y.; Nagahara, S.; Honma, K.; Sasaki, H.; Hirai, K.; et al. Efficient delivery of small interfering RNA to bone-metastatic tumors by using atelocollagen in vivo. Proc. Natl. Acad. Sci. U.S.A. 2005, 102, 12177–12182. 84. Medarova, Z.; Pham, W.; Farrar, C.; Petkova, V.; Moore, A. In vivo imaging of siRNA delivery and silencing in tumors. Nat. Med. 2007, 13, 372–377. 85. Mae, M.; Langel, U. Cell-penetrating peptides as vectors for peptide, protein and oligonucleotide delivery. Curr. Opin. Pharmacol. 2006, 6, 509–514. 86. Bruchez, M., Jr., Moronne, M.; Gin, P.; Weiss, S.; Alivisatos, A. P. Semiconductor nanocrystals as fluorescent biological labels. Science 1998, 281, 2013–2016. 87. Seydel, C. Quantum dots get wet. Science 2003, 300, 80–81. 88. Gao, X.; Cui, Y.; Levenson, R. M.; Chung, L. W.; Nie, S. In vivo cancer targeting and imaging with semiconductor quantum dots. Nat. Biotechnol. 2004, 22, 969–976. 89. Michalet, X.; Pinaud, F. F.; Bentolila, L. A.; Tsay, J. M.; Doose, S.; Li, J. J.; et al. Quantum dots for live cells, in vivo imaging, and diagnostics. Science. 2005, 307, 538–544.
P1: OTA/XYZ P2: ABC c22 JWBS037-Chen January 25, 2011
562
13:25
Printer Name: Yet to Come
MULTIFUNCTIONAL NANOPARTICLES FOR CANCER THERAGNOSIS
90. Akerman, M. E.; Chan, W. C.; Laakkonen, P.; Bhatia, S. N.; Ruoslahti, E. Nanocrystal targeting in vivo. Proc. Natl. Acad. Sci. U.S.A. 2002, 99, 12617–12621. 91. Medintz, I. L.; Clapp, A. R.; Mattoussi, H.; Goldman, E. R.; Fisher, B.; Mauro, J. M. Selfassembled nanoscale biosensors based on quantum dot FRET donors. Nat. Mater. 2003, 2, 630–638. 92. Young, S. H.; Rozengurt, E. Qdot nanocrystal conjugates conjugated to bombesin or ANG II label the cognate G protein-coupled receptor in living cells. Am. J. Physiol. Cell Physiol. 2006, 290, C728–C732. 93. Yezhelyev, M. V.; Gao, X.; Xing, Y.; Al-Hajj, A.; Nie, S.; O’Regan, R. M. Emerging use of nanoparticles in diagnosis and treatment of breast cancer. Lancet Oncol. 2006, 7, 657–667. 94. Bagalkot, V.; Zhang, L.; Levy-Nissenbaum, E.; Jon, S.; Kantoff, P. W.; Langer, R.; et al. Quantum dot-aptamer conjugates for synchronous cancer imaging, therapy, and sensing of drug delivery based on bi-fluorescence resonance energy transfer. Nano Lett. 2007, 7, 3065–3070. 95. Samia, A. C.; Chen, X.; Burda, C. Semiconductor quantum dots for photodynamic therapy. J. Am. Chem. Soc. 2003, 125, 15736–15737. 96. Willard, D. M.; Van Orden, A. Quantum dots: resonant energy-transfer sensor. Nat. Mater. 2003, 2, 575–576. 97. Bakalova, R.; Ohba, H.; Zhelev, Z.; Ishikawa, M.; Baba, Y. Quantum dots as photosensitizers? Nat. Biotechnol. 2004, 22, 1360–1361. 98. Samia, A. C.; Dayal, S.; Burda, C. Quantum dot-based energy transfer: perspectives and potential for applications in photodynamic therapy. Photochem. Photobiol. 2006, 82, 617–625. 99. Daniel, M. C.; Astruc, D. Gold nanoparticles: assembly, supramolecular chemistry, quantumsize-related properties, and applications toward biology, catalysis, and nanotechnology. Chem. Rev. 2004, 104, 293–346. 100. Love, J. C.; Estroff, L. A.; Kriebel, J. K.; Nuzzo, R. G.; Whitesides, G. M. Self-assembled monolayers of thiolates on metals as a form of nanotechnology. Chem. Rev. 2005, 105, 1103– 1169. 101. Thanh, N. T.; Rosenzweig, Z. Development of an aggregation-based immunoassay for antiprotein A using gold nanoparticles. Anal. Chem. 2002, 74, 1624–1628. 102. Elghanian, R.; Storhoff, J. J.; Mucic, R. C.; Letsinger, R. L.; Mirkin, C. A. Selective colorimetric detection of polynucleotides based on the distance-dependent optical properties of gold nanoparticles. Science 1997, 277, 1078–1081. 103. Nam, J. M.; Thaxton, C. S.; Mirkin, C. A. Nanoparticle-based bio-bar codes for the ultrasensitive detection of proteins. Science 2003, 301, 1884–1886. 104. Hirsch, L. R.; Jackson, J. B.; Lee, A.; Halas, N. J.; West, J. L. A whole blood immunoassay using gold nanoshells. Anal. Chem. 2003, 75, 2377–2381. 105. Zhao, X.; Hilliard, L. R.; Mechery, S. J.; Wang, Y.; Bagwe, R. P.; Jin, S.; et al. A rapid bioassay for single bacterial cell quantitation using bioconjugated nanoparticles. Proc. Natl. Acad. Sci. U.S.A. 2004, 101, 15027–15032. 106. Loo, C.; Lin, A.; Hirsch, L.; Lee, M. H.; Barton, J.; Halas, N.; et al. Nanoshell-enabled photonicsbased imaging and therapy of cancer. Technol. Cancer Res. Treat. 2004, 3, 33–40. 107. Hirsch, L. R.; Stafford, R. J.; Bankson, J. A.; Sershen, S. R.; Rivera, B.; Price, R. E.; et al. Nanoshell-mediated near-infrared thermal therapy of tumors under magnetic resonance guidance. Proc. Natl. Acad. Sci. U.S.A. 2003, 100, 13549–13554. 108. O’Neal, D. P.; Hirsch, L. R.; Halas, N. J.; Payne, J. D.; West, J. L. Photo-thermal tumor ablation in mice using near infrared-absorbing nanoparticles. Cancer Lett. 2004, 209, 171–176. 109. Loo, C.; Lowery, A.; Halas, N.; West, J.; Drezek, R. Immunotargeted nanoshells for integrated cancer imaging and therapy. Nano. Lett. 2005, 5, 709–711.
P1: OTA/XYZ P2: ABC c22 JWBS037-Chen January 25, 2011
13:25
Printer Name: Yet to Come
REFERENCES
563
110. Fujimoto, J. G.; Brezinski, M. E.; Tearney, G. J.; Boppart, S. A.; Bouma, B.; Hee, M. R.; et al. Optical biopsy and imaging using optical coherence tomography. Nat. Med. 1995, 1, 970–972. 111. Oldenburg, S. J.; Averitt, R. D.; Westcott, S. L.; Halas, N. J. Chem. Phys. Lett. 1998, 288, 243–247. 112. Gobin, A. M.; Lee, M. H.; Halas, N. J.; James, W. D.; Drezek, R. A.; West, J. L. Near-infrared resonant nanoshells for combined optical imaging and photothermal cancer therapy. Nano Lett. 2007, 7, 1929–1934. 113. Moeller, B. J.; Richardson, R. A.; Dewhirst, M. W. Hypoxia and radiotherapy: opportunities for improved outcomes in cancer treatment. Cancer Metastasis Rev. 2007, 26, 241–248. 114. Harris, A. L. Hypoxia—a key regulatory factor in tumour growth. Nat. Rev. Cancer 2002, 2, 38–47. 115. Kampinga, H. H.; Dikomey, E. Hyperthermic radiosensitization: mode of action and clinical relevance. Int. J. Radiat. Biol. 2001, 77, 399–408. 116. Roti, Roti JL. Introduction: radiosensitization by hyperthermia. Int. J. Hyperthermia 2004, 20, 109–114. 117. Moros, E. G.; Corry, P. M.; Orton, C. G. Thermoradiotherapy is underutilized for the treatment of cancer. Med. Phys. 2007, 34, 1–4. 118. Diagaradjane, P.; Shetty, A.; Wang, J. C.; Elliott, A. M.; Schwartz, J.; Shentu, S.; et al. Modulation of in vivo tumor radiation response via gold nanoshell-mediated vascular-focused hyperthermia: characterizing an integrated antihypoxic and localized vascular disrupting targeting strategy. Nano Lett. 2008, 8, 1492–1500. 119. Hao, E.; Schatz, G. C. Electromagnetic fields around silver nanoparticles and dimers. J. Chem. Phys. 2004, 120, 357–366. 120. Hao, E.; Schatz, G. C.; Hupp, J. T. Synthesis and optical properties of anisotropic metal nanoparticles. J. Fluoresc. 2004, 14, 331–341. 121. Huang, X.; El-Sayed, I. H.; Qian, W.; El-Sayed, M. A. Cancer cell imaging and photothermal therapy in the near-infrared region by using gold nanorods. J. Am. Chem. Soc. 2006, 128, 2115–2120.
P1: OTA/XYZ P2: ABC c23 JWBS037-Chen January 25, 2011
13:28
Printer Name: Yet to Come
CHAPTER 23
Nanoparticles for Combined Cancer Imaging and Therapy VAISHALI BAGALKOT, MI KYUNG YU, and SANGYONG JON School of Life Sciences, Gwangju Institute of Science and Technology, Gwangju, South Korea
This chapter describes the use of various nanoparticles for combined cancer imaging and therapy. Nanoparticles (<100 nm) synthesized from materials such as polymers, metals, ceramics, and lipids can bridge diagnostics with therapy in a programmable and specific manner. The chapter begins with background on nanoparticle-based imaging agents, a discussion on in vivo fate of nanoparticles, and tumor-specific targeting. The multifunctional nanoparticles can deliver not only imaging agents but anticancer chemical or biological drugs, leading to early cancer diagnosis and treatments that are listed as various examples in the chapter. The chapter concludes with a discussion on the prospects and consideration of such multifunctional nanoparticles in clinical applications.
23.1 INTRODUCTION Despite advancements in imaging modalities and therapeutics, cancer mortality still remains the top ranked challenge worldwide. On the one hand, various imaging modalities used in the clinic, including MRI, CT, and PET, rely on conventional contrast agents for better resolution but even with these early cancer detection is hard to achieve. On the other hand, chemical anticancer drugs suffer from severe adverse effects in cancer treatments due to the lack of specificity over normal tissue. Thus there is a keen necessity for early cancer detection and efficient therapy based on a new paradigm. To this end, nanoparticlebased nanomedicine has recently emerged with the anticipation that nanoparticles may act as novel imaging or drug delivering agents in a cancer-specific manner, resulting in early cancer detection and better therapeutic outcome. Because inorganic nanoparticles (i.e., magnetic, semiconductor, and metal nanoparticles) possess unique physical properties such as superparamagnetic behavior and unprecedented optical properties, they have shown promise as imaging agents in animal models. Furthermore, various cancer-targeting ligands
Nanoplatform-Based Molecular Imaging Edited by Xiaoyuan Chen C 2011 John Wiley & Sons, Inc. Copyright
565
P1: OTA/XYZ P2: ABC c23 JWBS037-Chen January 25, 2011
566
13:28
Printer Name: Yet to Come
NANOPARTICLES FOR COMBINED CANCER IMAGING AND THERAPY
including antibodies, peptides, and aptamers can be deposited onto nanoparticles so as to impart cancer specificity. Cancer chemotherapy benefits from nanosized drug carriers that can target tumor sites, enhance drug efficacy, reduce nonspecific toxicities as compared to free drugs, and thus raise the therapeutic index. Thus a combination of inorganic nanoparticles, targeting ligands, and therapeutic agents into a single entity has been attempted to construct multifunctional nanoparticles enabling combined cancer imaging and therapy [1]. The main goals that can be achieved by using “whole in one approach” are to be able to detect, treat, and track the progress of cancer treatment. At the forefront of imaging techniques used widely for cancer detection are fluorescence optical imaging and magnetic resonance imaging (MRI), where quantum dots (QDs) and magnetic nanoparticles such as superparamagnetic iron oxide nanoparticles (SPIONs) have revolutionized imaging with enhanced photostability, enhanced contrast, precise molecular targeting capabilities, and the ability to carry drugs [2]. Here, the main focus is on SPIONs, QDs, and gold nanoparticles with various examples for combined imaging and therapy, in vivo fate, and future directions. Table 23.1 summarizes key properties and therapeutic uses of major nanoparticles in two categories: organic and inorganic nanoparticles.
23.2 DESIGN PRINCIPLES OF NANOPARTICLES FOR IN VIVO IMAGING AND THERAPY 23.2.1 Effect of Surface Characteristics and Size of Particles on Biodistribution Usually for cancer therapy, systemic administration of drug carriers is preferred as this delivery route can effectively target both primary and metastatic tumors. As soon as nanoparticles enter the bloodstream, various serum proteins and complement factors tend to be adsorbed onto the particles’ surfaces nonspecifically. The resulting opsonized nanoparticles during circulation are encountered by macrophages of the reticuloendothelial system (RES organs comprised of liver, spleen, lungs, and bone marrow) especially the Kupffer cells of the liver [3], which recognize and then internalize the nanoparticles through the scavenger receptors on the surfaces of macrophages; thus most nanoparticles are lost from circulation [4, 5]. Both size and surface charge of particles influence the opsonization and are key properties to target the systemically administered nanoparticles effectively to the tumor (Fig. 23.1). Therefore surface engineering is necessary for evading macrophages so nanoparticles can be distributed to the target sites in the body. The surface charge of nanoparticles also influences the pharmacokinetics. Negatively charged particles show an increased uptake by phagocytes of the liver and increased clearance, whereas positively charged particles can form aggregates with serum proteins that can lead to blocking of capillaries (lung) [6, 7]. Positively charged polymers like poly-l-lysine coated magnetic nanoparticles that can nonspecifically adsorb to cells due to charge interactions with cell membrane were found to have reduced circulation time [8] nuetral or zwitterionic charged particles are preferred for longer blood circulation times. For example, the body’s own red blood cells evade uptake by macrophages and survive for a few months; their surfaces are covered by oligosaccharides and their curvature helps them easily deform to escape splenic filtration [9, 10]. Such oligosaccharide (dextran) coated iron oxide nanoparticles are well known and clinically approved as MR contrast agents [11, 12]. The surface charge on the nanoparticle can also
P1: OTA/XYZ P2: ABC c23 JWBS037-Chen January 25, 2011
13:28
Printer Name: Yet to Come
DESIGN PRINCIPLES OF NANOPARTICLES FOR IN VIVO IMAGING AND THERAPY
567
TABLE 23.1 Various Nanoparticles Used in Cancer Imaging and Therapy Nanoparticles
Property
Therapeutic Use
Organic Liposomes
r Self-assembling colloid structures
r r r r r
composed of lipid bilayers surrounding an inner aqueous compartment. r Can encapsulate both hydrophilic and hydrophobic drugs. Polymeric micelles
Polymer drug conjugates
Dendrimers
r Self-assembly of block copolymers of two or more chains with different hydrophobicity that spontaneously assemble in an aqueous environment, forming core shell structures. r Hydrophobic core can load poorly soluble drugs with high loading capacity (5–25% weight).
nanoparticles
r Poly(d,l-lactic-co-glycolic acid)-block-poly(ethylene glycol)
r PEG-antivascular endothelial
are chemically conjugated to polymers such as N-(2hydroxypropyl)methacrylamide (HPMA) and poly(ethylene glycol) (PEG). r Prolong the blood circulation time of drugs (∼ several hours) to allow passive tumor targeting of drugs.
growth factor aptamer, PEG-interferon-alpha-2a r PEG-l-asparaginase r Polyglutamate paclitaxel, etc.
r Globular highly branched synthetic
r Polyamidoamine (PAA)
r Size tunable optical properties, high photo- and chemical stability, broad excitation and narrow emission, multispectral imaging.
Iron oxide nanoparticles
r Pluronic block copolymers r Polymer–lipid hybrid
r Small molecule therapuetic agents
polymers consisting of an initiator core and multiple layers ending in a variety of functional groups. r Can be loaded with drugs in the core or covalently conjugated to the end functional groups Inorganic Quantum dots
Liposomal daunorubicin Liposomal vincristine Liposomal doxorubicin Liposomal annamycin Liposomal cisplatin
r Superparamagnetic nature, used as contrast agents for magnetic resonance imaging (MRI). r Thermotherapy.
dendrimers, for example, folate targeted methotrexate loaded G5 PAA dendrimer.
r Quantum dot conjugated HER2 antibody,TAT peptide, cell adhesion molecules, PSMA antibody, etc.
r Feridex IV, Combidex—clinical. r Amino-CLIO—magneto/ optical probes for in vivo animal imaging.
P1: OTA/XYZ P2: ABC c23 JWBS037-Chen January 25, 2011
568
13:28
Printer Name: Yet to Come
NANOPARTICLES FOR COMBINED CANCER IMAGING AND THERAPY
TABLE 23.1 (Continued) Nanoparticles Gold nanoparticles and nanoshells
Carbon Nanotubes
Property
Therapeutic Use
r Tunable surface plasmon
r Colloidal gold nanoparticles. r Gold nanoshells.
resonance in the near-infrared (NIR) region. r Used for NIR imaging and photothermal therapy.
r Thermal ablation therapy. r Transportation of DNA cargoes into cells.
r Folic acid adsorbed nanotubes for targeted cancer therapy in animal mouse model.
be reduced by linking it with a polymer such as poly ethylene glycol (PEG), creating stealth surfaces that reduce interactions with opsonins or serum proteins [13–18]. PEG is known as an antibiofouling polymer that shields surface charge and increases hydrophilicity with its excellent aqueous solubility, flexibility of its polymer chains, low toxicity, and immunogenicity. It is nonbiodegradable. For nonbiodegradable polymers the molecular weights should be below 45–50 kDa, that is, the threshold for renal clearance [19]. Thiolated PEG, block copolymers consisting of polyethylene imine (PEI)-g-(PEG), and PEG derivatized phospholipids have been used to modify surfaces of QDs and render them water soluble, dispersible, and stable in vivo [20–27]. PEG has also been attached to SPIONs via silanol grafting; for example, our own research group has reported on the development of protein
FIGURE 23.1 Particle size and surface charge influence their biodistribution.
P1: OTA/XYZ P2: ABC c23 JWBS037-Chen January 25, 2011
13:28
Printer Name: Yet to Come
DESIGN PRINCIPLES OF NANOPARTICLES FOR IN VIVO IMAGING AND THERAPY
569
resistant poly(TMSMA-r-PEGMA) copolymer comprised of silanol anchoring groups and PEG side chains [28]. On the other hand, the particle size also greatly influences the biodistribution of particles. It was shown that particles in the range of 100–200 nm and greater than 200 nm have significant protein adsorption compared to particles below 100 nm [29]. Particles greater than 400 nm are easily captured in the spleen, whereas particles less than 5 nm are rapidly cleared by renal filtration [30, 31]. It is reported that particles less than 50 nm in size can penetrate the fenestrae of the liver [30]. Particles either too small or too big cannot circulate for long in the blood. The property of long circulation time is extremely important for nanoparticles to effectively target tumor sites with an increased chance to distribute to tumor sites [32]. Tumor blood vessels are distinct from normal vessels, in that the endothelial cells in tumors possess wide fenestrations, ranging from 200 nm to 1.2 mm [31, 33–35]. For particles to effectively extravasate from the tumor blood vessels into the interstitium of the tumor is governed by the size of the open interendothelial gap junctions and small sized (<200 nm) particles are most effective for extravasating the tumor microvessels [36, 37]. For example, it was reported by Liu and colleagues that tumoral uptake can be correlated to long circulation time in blood; liposomes with a size range between 100 and 200 nm showed fourfold higher rate of uptake in tumor compared to liposomes greater than 300 nm or less than 50 nm in size [30]. The optimum size that can maintain the pharmacokinetics of the nanoparticles for efficient targeted delivery would probably fall within the 50 nm to less than 100 nm range. To summarize, small size less than 100 nm and neutral charged PEGylated nanoparticles have long residence time and are effective for tumor targeting. 23.2.2 Tumor-Targeting Principles Using Nanoparticles Targeting nanoparticles to a tumor can be either passive or active. Passive targeting forms the basis for the majority of nanocarriers (liposomes, PEGylated liposomes, albumin bound paclitaxel, polymer drug conjugates) that are either approved clinically or are in further clinical development [38–45]. The surface modified nanoparticles endowed with additional targeting ligands enable the nanocarriers to interact with tumor cells actively by delivering drugs intracellularly and are retained in the tumors as compared to nontargeted counterparts [46]. Passive targeting exploits the inherent characteristics of the tumors that grow fast and develop leaky vasculature, coupled with poor lymphatic drainage allows for permeation and retention of nanoparticles within the tumor region. For nanoparticles of proper size, it is possible to effectively target them to tumor sites through the enhanced permeation and retention (EPR) effect. Thus by tuning the size of nanoparticles, their distribution to organs such as the liver, spleen, or lymph nodes or their retention in blood circulation (vascular compartment) can be controlled. Specifically, MR imaging of various RES organs and tumors has been possible through iron oxide nanoparticles of various sizes. For example, dextran coated iron oxide nanoparticles (Feridex® ) 80–150 nm in size are used for liver imaging and the same nanoparticles in the size range of 20–40 nm (Combidex® ) are used for metastatic lymph node imaging [47]. In the case of QDs, Ballou and co-workers reported that QDs when coated with polyethylene glycol were found to stay in the blood circulation for long time periods (half-life more than 3 hours) as PEG is a well-known antibiofouling polymer that can prevent opsonization with serum proteins; the size of these PEG coated probes were greater than the renal excretion limit (reported to be around 5nm cutoff) but small enough to circulate for a long time and avoid uptake by the RES [48]. Another study by Ashwath and colleagues reported that efficient PEGylation of the
P1: OTA/XYZ P2: ABC c23 JWBS037-Chen January 25, 2011
570
13:28
Printer Name: Yet to Come
NANOPARTICLES FOR COMBINED CANCER IMAGING AND THERAPY
QD surface with additional immune shielding of monoclonal antibody (mAb) through blocking the Fc region of the antibody by Fc-specific F(ab)2 fragments can reduce uptake of these conjugates by the liver and spleen, and indeed such QD–mAb conjugates targeted to cell adhesion molecules (CAMs) could specifically label the retinal endothelium in an in vivo diabetic rat model [49]. QDs have also been used to differentiate between tumor vasculature and perivascular cells and matrix with greater clarity as they are comparatively more photostable than organic dyes [50]. For example, in an elegant example that was reported by Kim and co-workers, type-II QDs around 19 nm (that emit in the near-infrared (NIR) region when injected intradermally) can move rapidly into nearby sentinel lymph nodes and such QD probes can be helpful during an intraoperative surgery to easily locate the lymph nodes [51]. Active tumor targeting can be achieved by specifically designing drug carriers with specific ligands that recognize biophysical characteristics unique to cancer cells (tumor-specific or tumor-associated antigens). Several classes of targeting molecules have been utilized such as antibodies, aptamers, peptides (e.g., cRGD), (e.g., folic acids), and carbohydrates [52]. Various actively targeted QD bioconjugates with antibodies specific for PSMA (prostate-specific membrane antigen), HER-2 (human epidermal receptor-2), cell adhesion molecules (CAMs), and specific peptides such as integrin (␣v 3 ) binding RGD peptide and alpha fetoprotein have been reported [49, 53–57]. Magnetic iron oxide nanoparticles have been conjugated to monoclonal antibodies (mAbs), HerceptinTM (mAb to HER-2), RGD peptides, folic acids, and Tat peptides [58–63]. 23.2.3 Softness and Hardness of Nanoparticles In Sections 23.2.1 and 23.2.2, the effects of surface characteristics and sizes of nanoparticles on biodistribution and tumor targeting were discussed. An additional factor to consider is the particle deformability, an ability to change its shape (i.e., squeezed or elongated) without destruction or loss of its entity. Red blood cells (RBCs) are one of the best example to explain the deformability issue. Despite their much larger size (6–8 m in diameter) than the cutoff size in the spleen (∼400 nm), RBCs pass through the spleen without difficulty and circulate for a long time in the blood (weeks). This long circulation time is believed to be possible because RBCs are soft enough to elongate or deform their shape while passing through the fenestrae of various organs, suggesting that deformability is an another key property to achieve long blood circulation time. Most nanoparticles used for imaging or therapeutic purposes can be divided into two types: soft and hard nanoparticles in which the former have high and the latter have low deformability, respectively. Liposomes, micelles or polymer micelles, and gels with loosely entangled matrix can be categorized as soft nanoparticles in that they are able to deform their original shapes under specific conditions. On the other hand, most inorganic nanoparticles made from metals, metal oxides, and semiconductors can be considered hard nanoparticles because of the difficulty in shape deformation. Based on our own and other’s experiments (either published or unpublished data), the hard nanoparticles are likely to accumulate in the spleen and liver more than their soft counterparts for a given size. Recently, we reported that PEGylated gold nanoparticles could be used as CT contrast agents in vivo, in which we observed unexpected high accumulation of the nanoparticles in spleen even though their hydrodynamic size was ∼30 nm, far smaller than the cutoff of the organ [64]. Similar results were obtained with SPIONs ∼20–30 nm in size for MRI using mice (unpublished results). Although polymeraggregate-based nanoparticles such as PLGA, PLA, and PCL are made of polymers, they
P1: OTA/XYZ P2: ABC c23 JWBS037-Chen January 25, 2011
13:28
Printer Name: Yet to Come
MULTIFUNCTIONAL NANOPARTICLES FOR COMBINED CANCER IMAGING AND THERAPY
571
can be considered hard nanoparticles, not soft ones according to their biodistribution profiles [65, 66]. In summary, unlike soft nanoparticles, it seems to be desirable to make inorganic nanoparticles as small as possible to achieve long circulation times by evading the splenic and liver filtration systems, resulting in better tumor targeting or higher tumor accumulation.
23.3 MULTIFUNCTIONAL NANOPARTICLES FOR COMBINED CANCER IMAGING AND THERAPY: VARIOUS EXAMPLES The development of multifunctional nanoparticle platforms with both diagnostic and therapeutic capabilities may allow in vivo monitoring of biodistribution of the nanocarriers and tumor response to therapy. Thus there has been a tremendous amount of interest in developing the next generation of multifunctional molecular agents for the diagnosis and treatment of cancer. The multifunctional nanoparticles in general should have antibiofouling surfaces that allow for long blood circulation times and ensure the EPR (enhanced permeation and retention) effect. The surface coating must be stable under in vivo conditions and prevent aggregation of nanoparticles and embolization, and they should be within a preferred size (5–50 nm) range to avoid uptake by the macrophages of the RES. Nanoparticles used in cancer imaging are mainly iron oxide nanoparticles, such as MRI probes, QDs (flourescence based), and gold nanoparticles (surface plasmon resonance based) as probes for optical imaging. Below, we discuss relevant examples cited in the literature and from our own research group for the development of these combined targeted nanoparticles for cancer imaging and therapy. 23.3.1 Superparamagnetic Iron Oxide Nanoparticles (SPIONs) Among inorganic nanoparticles, SPIONs are widely used in biomedical applications such as MRI and drug delivery and therapy (hyperthermia) [60, 67–76]. Bulk magnetic materials consist of multiple domains, where the magnetic dipoles are arranged and their orientation in the presence of a magnetic field determines the overall magnetic property of the material. As the size of these materials gets smaller, in the range below 20 nm, they would consist of only a single domain resulting in superparamagnetic behavior with no remnant magnetization after removal of the magnetic field [77]. SPIONs consist of a core of magnetite iron oxide (Fe3 O4 ) covered with biocompatible polymeric layers, since the bare iron oxide surface can form aggregates in the blood, leading to plasma protein binding to the surface of particles and clearance by the macrophages of the RES and also aggregation, which reduces the innate superparamagnetic properties [78, 79]. Various synthetic polymers have been used to engineer the surface of SPIONs to prevent biofouling and aggregation under physiological conditions (i.e., high salt and protein concentrations) and to maintain good dispersibility such as dextran, polyethylene glycol (PEG), and polyvinylpyrrolidone (PVP), all of which are known to be biocompatible and result in long blood-circulating SPIONs [80–84]. Monocrystalline iron oxide nanoparticles (MIONs) and crosslinked iron oxide (CLIO) nanoparticles are typical examples of dextran-coated SPIONs having a magnetite core and are widely used in in vivo as well as in vitro MRI [85–89]. The advantages of using magnetic nanoparticles are (1) the superparamagnetic nature allows them to be visualized by magnetic resonance (MR) imaging; (2) they can be guided to target sites by means of an external field; (3) they can be heated to provide hyperthermia for cancer therapy,
P1: OTA/XYZ P2: ABC c23 JWBS037-Chen January 25, 2011
572
13:28
Printer Name: Yet to Come
NANOPARTICLES FOR COMBINED CANCER IMAGING AND THERAPY
and, most importantly, (4) they can be degraded into nontoxic iron ions in vivo. Inorganic nanoparticles have unique size-dependent physical properties that can be highly useful for imaging and therapy applications, but a major concern is their long-term residence in the body, leading to toxicity. In the case of SPIONs, they can be degraded into iron ions under the acidic conditions in the body and enter into the normal iron metabolic pathways, thus causing no apparent long-term toxicities. Some key properties that must be considered for their use as drug delivery carriers are drug-loading efficiency, efficient drug release from the nanoparticles, dispersibility in aqueous conditions, and their biocompatibility and degradability. Iron oxide nanoparticles can be used as drug delivery carriers, both as magnetic-fielddependent (magnetic drug targeting) and magnetic-field-independent, for accumulation of the nanoparticles at target sites. The former—drug targeting by release of drugs from magnetic carriers in the presence of an external magnetic field—is well known and has been used for local cancer treatment in Phase I clinical trials in patients. There are several considerations that need to be addressed, such as the external magnetic field strength needed to retain the particles at the target site, depth of the target tissue, blood flow, vascular supply, aggregation of the particles (leading to embolization), and whether the particles can only be used to target tissues close to the surface of the body. Such remotely triggered release of drugs would be beneficial when they can be implanted at a tumor site and can release drugs or be individually targeted to the tumor site. With the development of instrumentation to apply high-power radiofrequency field to patients, it should be possible in the future to use remote actuation platforms that can guide drug delivery with minimum collateral tissue damage. The SPIONs absorb the radiofrequency electromagnetic field, which is minimally absorbed or scattered by surrounding tissue compared to near-infrared radiation, and due to the alignment of the magnetic dipoles with the external field, it can produce heat sufficient to cause DNA melting and thus release of DNA tethered cargo, as reported by Derfus and colleagues who designed an electromagnetic field triggered release of nanoparticle-tethered dye in pulsatile and multistage profiles. Multifunctional nanoparticles were embedded in a matrigel plug and the plug was implanted subcutaneously in living mice acting as a tumor model, and the release of flourescein-labeled 18 bp oligonucleotide by electromagnetic field (EMF) exposure was studied (Fig. 23.2a) [90]. The EMF application to the model tumor implants resulted in release of model cargo, and penetration into surrounding tissue was approximately sixfold higher compared to unexposed controls (Fig. 23.2b,c). In addition, the magnetic particles could be visualized noninvasively by MRI (Fig. 23.3d). The above study establishes a modular approach that can be tuned to release drug molecules at target sites. To use such remotely triggered nanoparticles as potential vehicles for diagnosis and therapy, it would be essential for a large number of these particles to accumulate at the tumor site preferentially, and to develop high power radiofrequency instruments and particles with greater magnetization cores. Drug delivery carriers that are independent of a magnetic field—iron oxide nanoparticles such as SPIONs—are designed to accumulate in the tumor tissue based on size and long circulation time; this is called passive targeting. Specifically, targeting ligands such as peptides, antibodies, and aptamers can be attached to the polymer surface of these nanoparticles for efficient drug delivery, known as active targeting. Recently, our research group reported on the development of novel antibiofouling polymer coated poly(TMSMAr-PEGMA) SPIONs containing a silanol anchoring part that could be thermally crosslinked
P1: OTA/XYZ P2: ABC c23 JWBS037-Chen January 25, 2011
13:28
Printer Name: Yet to Come
MULTIFUNCTIONAL NANOPARTICLES FOR COMBINED CANCER IMAGING AND THERAPY
573
FIGURE 23.2 Remotely triggered release from nanoparticles in vivo. Nanoparticles were mixed with matrigel and injected subcutaneously near the posterior mammary fat pad of mice, forming model tumors (a). Application of EMF to implants containing 18 bp tethers resulted in release of model drugs and penetration far into surrounding tissue (b) when compared to unexposed controls (c, scale bar = 100 micrometers). These mice were imaged with a 7T MRI scanner, and transverse section shown in (d) depicts image contrast due to presence of nanoparticles (arrow). (Reproduced with permission from Ref. [90]).
to form stable polymer layers, had antibiofouling properties because of the PEG part, were stable and aqueous dispersible, and could passively target tumors. TCL-SPIONs showed efficiency as MR contrast agent for in vivo cancer imaging and Cy 5.5 was conjugated to TCL-SPION as a dual (MR/optical) cancer imaging probe. Further more, TCL-SPION was reported as a novel drug delivering MR contrast agent in a combined imaging and therapy platform (Fig. 23.3a) [91]. The anticancer drug doxorubicin was incorporated in the polymeric shell of TCL-SPION through electrostatic interaction between positively charged Dox and the negatively charged polymer coating layers; the size of Dox-loaded TCL-SPION (Dox@TCL-SPION) was 21 ± 6 nm with optimal loading amount of Dox to TCL-SPION about 2 wt%. When MRI was performed in Lewis lung carcinoma (LLC) bearing mice 4.5 h after intravenous injection of Dox@TCL-SPION, there was a significant darkening of the tumor area with relative signal enhancement of about 58%, indicating large accumulation of nanoparticles within the tumor (Fig. 23.3b). Dox@TCL-SPION was found to be effective therapeutically; when evaluated in LLC tumor bearing mice, it resulted in 63% tumor inhibition compared to controls whereas free Dox at 8 fold higher dose showed only 38% tumor inhibition (Fig. 23.3c). Independent of the magnetic field or any targeting ligands, Dox@TCL-SPION could passively target tumors and aid in tumor detection by MR imaging because of its high stability and antibiofouling characteristics of the polymer shell. At the same time, it could deliver sufficient amounts of anticancer drugs that were released from the nanoparticles and exhibited anticancer activity. Thus the location of the tumor, proper delivery of drug to the tumor, and therapeutic response of the
P1: OTA/XYZ P2: ABC c23 JWBS037-Chen January 25, 2011
574
13:28
Printer Name: Yet to Come
NANOPARTICLES FOR COMBINED CANCER IMAGING AND THERAPY
1
2
3
4
5
FIGURE 23.3 Smart drug loaded SPIONs. (a) Polymer (poly(TMSMA-r-PEGMA)) coated and thermally crosslinked SPION that can be loaded with Dox into the polymeric shell through electrostatic interactions between positively charged Dox and negatively charged polymer layers, amine terminated, and conjugated to carboxyl TCL-SPION through EDC/NHS chemistry. (b) T2-weighted fast-spin echo images (time of repetition/time of echo: 4200 ms/102 ms) taken at 0 h and 4.5 h after injection of Dox@TCL-SPION at the level of LLC tumor on the right back of the mouse. The dashed circle with white arrow indicates the allograft tumor region. (c) Antitumor efficacy of Dox@TCLSPION in LLC allograft animal model. Excised tumors from mice euthanized after the 19th day of treatment with: 1, control; 2, TCL-SPION (12.5 mg Fe kg−1 ); 3, Dox (0.64 mg kg−1 ); 4, Dox (5 mg kg−1 ); 5, Dox@TCL-SPION (12.5 mg Fe kg−1 , 0.64 mg Dox kg−1 ). (Reproduced with permission from Yu et al. [91].) (d) T2 transverse relaxation times of LNCaP and PC3 cells and cells incubated with TCL-SPION and TCL-SPION-Apt bioconjugates. (e) Iron (Prussian blue) stain of LNCaP and PC3 cells after incubation with TCL-SPION-Apt bioconjugates for 12 h. The blue color within the LNCaP cells incubated with TCL-SPION-Apt bioconjugates represents intracellular uptake of the nanoparticles. Data below represents MTT cell proliferation assay: LNCaP cells and PC3 cells were incubated with TCL-SPION-Apt bioconjugates (0.1 mg/mL), TCL-SPION-Apt(Dox) (0.1 mg/mL, 5 M of Dox), and free dox (5 M ) for 3 h. The cells were washed three times with phosphate buffer and further incubated in media for an additional 48 h before the MTT assay. (Reproduced with permission from Wang et al. [96].)
tumor could be answered by using simple but smart, drug-loaded SPIONs. For example, it was shown by Tapan and colleagues that water-dispersible oleic acid (OA)-pluroniccoated iron oxide nanoparticles anchor at the OA–water interface and confer aqueous dispersibility and antibiofouling properties; additionally, particles could be loaded with waterinsoluble anticancer agents such as taxol where the drug can partition into the OA shell [92].
P1: OTA/XYZ P2: ABC c23 JWBS037-Chen January 25, 2011
13:28
Printer Name: Yet to Come
MULTIFUNCTIONAL NANOPARTICLES FOR COMBINED CANCER IMAGING AND THERAPY
575
Although the OA–pluronic stabilized iron oxide magnetic nanoparticles (mean size ∼193 nm with 8% drug loading) showed sustained release of drugs, good water dispersibility, and antiproliferative effect for breast and prostate cancer cells in vitro, it is also important to consider the size of the nanoparticles for their use in vivo, as large size particles greater than 100 nm are susceptible to uptake by RES organs. In previous examples, drugs were either ionically bound or entrapped in the polymer layers of magnetic nanoparticles, whereas drugs can also be covalently linked to the polymer layers. For example, it was reported by Kohler and co-workers that the anticancer drug methotrexate (MTX) covalently attached to amine functionalized iron oxide nanoparticles through amide bonds is stable under in vivo conditions [93]. The conjugate showed specific targeting to 9L glioma cells and induced cell apoptosis and could be detected by in vitro magnetic resonance (MR) imaging; thus such conjugates could be used for both diagnostic and therapeutic purposes. Nanoparticles have high surface area that can be used to load multiple ligands and drugs, and the multifunctional construct can deliver large numbers of drugs per biorecognition event. Further more, if such nanoparticles are made hollow, the interior void of these nanoparticles can also be loaded with large amounts of therapeutic drugs. Jongmin and colleagues have reported the development of hollow manganese oxide nanoparticles (HMONs) that were 20 nm in diameter for MR imaging and drug delivery by loading the anticancer drug Dox into the hollow core of these nanoparticles [94]. It was hypothesized that particles could mediate a sustained release profile of Dox, as their antiproliferative effects in breast cancer cell lines were only slightly lower than equivalent doses of Dox. SPIONs have been used as diagnostic agents for cancer imaging in organs such as the liver and lymph nodes, mainly based on size [95]. To detect metastatic cancers, it is necessary to specifically identify cancer cells based on their unique characteristics, such as overexpression of receptors and antigens compared to normal cells through the use of targeting ligands. Targeting allows the preferential delivery of therapeutic, diagnostic, or imaging agents to diseased tissues. Over the last two decades, molecular targeted diagnostic and therapeutic agents have dramatically improved cancer diagnosis and treatment [98–105]. Preclinical data have shown that targeted nanoparticle systems accumulate preferentially in the target tissue versus their nontargeted counterparts, demonstrating the vast potential of targeted nanoparticles [65, 106, 107]. Recently, we reported on the development of novel multifunctional TCL-SPION that can detect prostate cancer cells (PCa) as well as deliver targeted chemotherapeutic agent directly to the cancer cells; A10 RNA aptamer (Apt), which binds the extracellular domain of the prostate-specific membrane antigen (PSMA), was used to engineer targeted nanoparticles for PCa therapy and imaging [65, 106, 108] (Fig. 23.3a). PSMA is a well-established marker for PCa cells, and its expression is primarily PCa specific, with relatively lower levels seen in the normal prostate, brain, salivary glands, and small intestine [109]. A10 aptamer can be used as a targeting ligand and to deliver doxorubicin through intercalation of Dox into the CG sequence in the aptamer [108, 110, 111]. TCL-SPION-Apt bioconjugates were differentially taken up in PCa cells (LNCaP cell line) line expressing the PSMA protein as visualized by iron prussian blue stain compared to nonexpressing PC3 cell line (Fig. 23.3e). TCL-SPION-Apt bioconjugates also showed potential as targeted MR contrast agents in vitro: when cells (LNCaP and PC3) were incubated with bioconjugates, washed, and suspended in matrigel and measured for transverse relaxation time (T2) using a singlesided NMR probe, it was found that T2 relaxation time was altered significantly in PSMA expressing LNCaP cells, enabling their high sensitivity detection (Figure 23.3d). When analyzed for in vitro cytotoxicity after Dox loading onto TCL-SPION-Apt bioconjugates,
P1: OTA/XYZ P2: ABC c23 JWBS037-Chen January 25, 2011
576
13:28
Printer Name: Yet to Come
NANOPARTICLES FOR COMBINED CANCER IMAGING AND THERAPY
it was significantly enhanced against the targeted LNCaP cells as compared to the nontargeted PC3 cells (Fig. 23.3e). Thus novel multifunctional TCL-SPION consisting of TCL-SPION-Apt bioconjugates that can both detect and treat PCa cells in vitro is reported. Since there is a lack of sensitive and specific imaging agents as well as a paucity of effective therapeutic approaches for disseminated PCa, it is possible that multifunctional nanoparticle technologies, such as the TCL-SPION-Apt bioconjugates, may be a suitable approach for both the detection and treatment of disseminated PCa [112–115]. More broadly, the unique advantages of such multifunctional nanoparticles with diagnostic and therapeutic capabilities include the following: targeted delivery of therapeutics to disease cells only, observation of therapeutic delivery, and detection of therapeutic response. Magnetic nanoparticles have also been used to deliver biological drugs such as siRNA, antisense oligo, and also therapeutic peptides such as chlorotoxin (CTX) that can inhibit tumor invasiveness, especially for gliomas that are highly invasive. For example, Medarova and colleagues have reported the development of dual-purpose probes for simultaneous noninvasive imaging and delivery of siRNA to tumors [97]. The probe consisted of dextran-coated SPIONs (for MRI) labeled with Cy5.5 dye (for near-infrared (NIR) flourescence imaging) and conjugated to a synthetic siRNA duplex targeting a gene of interest. The probe (MN-NIRF-siGFP) was also modified with myristoylated polyarginine peptides (MPAPs) that aid in membrane translocation (Fig. 23.4a). The probe was injected intravenously and the delivery to tumors (24 h postinjection) was analyzed by in vivo MRI and NIRF imaging; images acquired after administration of the contrast agent showed significant drop in T2 relaxivity and high intensity NIRF signal in the tumors as a result of accumulation of the probe (Fig. 23.4a,b). To assess its therapeutic capability for gene silencing, the probe was conjugated to siRNA targeting an antiapoptotic gene encoding “survivin,” a member of the inhibitor of apoptosis protein family showing tumor restricted expression. When analyzed for gene silencing of survivin by RT-PCR, the survivin transcript levels were found to be 97% lower than control magnetic nanoparticle or mismatch control treated mice, substantial silencing was achieved in tumors (Fig. 23.4d). Thus through long circulation (favorable biodistribution to tumors) and the MR imaging sensitivity of magnetic nanoparticles, it was possible to simultaneously deliver and detect siRNA based therapeutic agents in vivo. Similarly, polyamidoamine (PAMAM) dendrimer-modified magnetic nanoparticles (8 nm) have been employed to deliver antisense survivin oligonucleotides; the composites were found to enter into human breast cancer and liver cancer cell lines (MCF, MDA-MB-435, and HepG2) [116]. Such dendrimer nanoparticle composites could solve the problems of poor diffusion of antisense oligo nucleotides across cell membranes and rapid exonuclease degradation. 23.3.2 QDs QDs also known as artificial atoms, have discrete energy levels and the spacing between these energy levels can be controlled precisely by variations in size [117]. They are semiconductor particles varying in size from 2 to 40 nm and containing roughly 200–10,000 atoms. They are composed of groups II–VI (e.g., CdSe) or III–V (e.g., InP) materials. Semiconductor nanocrystals known as QDs have been increasingly utilized as biological imaging and labeling probes because of their unique optical properties, including broad absorption with narrow photoluminescence spectra (ability to use multicolor QDs to image multiple targets), high quantum yield (makes them bright under photon-limited conditions in vivo), low photobleaching (well suited for continuous live cell imaging or tracking), and resistance to chemical degradation [118–121]. For biological applications, QDs have to be
P1: OTA/XYZ P2: ABC c23 JWBS037-Chen January 25, 2011
13:28
Printer Name: Yet to Come
MULTIFUNCTIONAL NANOPARTICLES FOR COMBINED CANCER IMAGING AND THERAPY
577
FIGURE 23.4 Novel dual-purpose probes for simultaneous noninvasive imaging and delivery of siRNAs to tumors. The probe consisted of magnetic nanoparticles labeled with near-infrared Cy5.5 dye (NIRF) and linked through two different linkers to membrane translocation peptides (MPAP) and siRNA molecules targeting survivin (sisurvivin). (a) Application of MN-NIRF-sisurvivin in LS174T human colorectal adenocarcinoma bearing mice tumor models; a high-intensity NIRF signal was observed in vivo by optical imaging after injection of MN-NIRF-sisurvivin probe confirming its delivery to the target tumor tissue. (b) When mice were imaged under MR imaging after administration of the probe, there was a significant drop in T2 relaxivity in images indicating the delivery of the probe. (c) Representative TUNEL assay images of tumor sections counterstained with DAPI and consecutively stained with H&E stain after administration of probe; distinct high density of apoptotic nuclei (green) were seen in tumors treated with MN-NIRF-sisurvivin probe as compared to tumor tissue treated with control magnetic nanoparticles that do not show apoptotic nuclei. (d) Quantitative RT-PCR analysis of survivin expression in LS174T tumors after injection with either MN-NIRF-sisurvivin, a mismatch control, or the parental magnetic nanoparticle (MN). (Reproduced with permission from Medarova et al. [97].)
water dispersible over a broad pH range and ionic strengths. Such water-soluble QDs have been generated by two broad procedures. Cap exchange is a method where hydrophobic ligands (TOPO) on the surface of organic synthesized QDs are exchanged with hydrophilic moieties that coordinate with surface atoms on the outer shell of the QDs. Hydrophilic ligands such as alkyl amines, phosphines, and mono or bidentate thiols have been used. Surface capping is another modification method, where the hydrophobic ligands are retained on the QD surface but additional amphiphilic phospholipids, triblock copolymers, and saccharides can bind to the QD ligands through the hydrophobic part and expose the hydrophilic part to render them water dispersible; the ligands are stable in aqueous media for months to years, and this is a common method used in the production of commercially available QDs [122]. The first significant uses of QDs in vivo were to image and map
P1: OTA/XYZ P2: ABC c23 JWBS037-Chen January 25, 2011
578
13:28
Printer Name: Yet to Come
NANOPARTICLES FOR COMBINED CANCER IMAGING AND THERAPY
draining lymph nodes and angiogenic vessels and blood pool imaging [18, 48, 51, 123]. In vivo flourescence imaging with QDs that emit in the near-infrared region is advantageous as there is minimal tissue absorption and scattering in this region of the spectrum and an optically bright signal can be obtained. Major limitations for in vivo use of QDs are their uptake by the RES due to opsonization and the use of PEG to increase circulation times of QDs by preventing opsonization; long chain (5000-Da) methoxy-PEG-coated QDs showed long circulation times up to 70 min compared to 750- or 3400-Da PEG-coated QDs that showed circulation time of less than 12 min [48]. In an interesting study, Jackson and co-workers showed that, by saturating RES uptake by using several nanomolar concentrations of QDs, the QDs can be specifically targeted to tumors through uptake by tumor infiltrating macrophages [124]. The study shows a dose–response relationship between QD uptake by the RES organs and their presence in the tumors; when QDs were intravenously injected at low doses (3–8 nmol), the majority of QDs (Qdot ITK Amino/PEG, Quantum Dot Corp.) are sequestered in the liver, spleen, and lymph nodes, whereas at higher doses (17 nmol), increasing quantities of QDs are present within experimental brain tumors due to the macrophages and microglia that carry the QDs and colocalize with glioma cells and outline small metastatic lesions. The surface modification of QDs with antibodies, aptamers, peptides, or small molecules that bind to antigens present on the target cells or tissues has resulted in the development of sensitive and specific targeted imaging and diagnostic modalities for in vitro and in vivo applications [125, 126]. More recently, QDs have been engineered to carry distinct classes of therapeutic agents for simultaneous imaging and therapeutic applications [17, 44, 127]. These combined imaging–therapy nanoparticles represent an exciting advance in the field of nanomedicine. It would be ideal to engineer “smart” multifunctional nanoparticles that are capable of performing these tasks while sensing the delivery of drugs in a simple and easily detectable manner. It would be important to maintain simplicity in design and engineering to assure the possible development and scale-up of these systems for research and medical applications. To explore this, we have reported a simple proof-of-concept of QD–aptamer (QD–Apt) conjugate that can image and deliver anticancer drugs to PCa cells and sense the delivery of drugs to the targeted tumor cells based on the mechanism of fluorescence resonance energy transfer (FRET) phenomena occurring between two fluorophores, in this case Dox (flourescent anticancer drug) and QDs (Fig. 23.5a) [118–121]. A10 RNA aptamers that served as targeting (specific for PSMA antigen overexpressed on prostate cancer cells) and drug (Dox) carrying molecules (Dox can intercalate into double-stranded CG sequences of RNA and DNA) were conjugated to the surface of QDs. The assembly of this conjugate results in Bi-FRET complex, a donor–acceptor model FRET between QD and Dox, where the fluorescence of QD is quenched as a result of Dox absorbance, and a donor–quencher model FRET between Dox and aptamer, where Dox is quenched by double–stranded RNA aptamer. Therefore both QD and Dox of the conjugate are in the fluorescence “OFF” state when the QD-Apt is loaded with Dox (QD-Apt(Dox)). After the particle is taken up by targeted cancer cells, Dox is gradually released from the conjugate, which induces the activation of QD and Dox fluorescence to the “ON” state. Time course studies with QD-Apt(Dox) revealed that, at 1.5 h of further incubation, more Dox was released from the QD-Apt conjugates, resulting in substantial increase in flourescence of both QD and Dox that signifies the “ON” state of the Bi-FRET system (Fig. 23.5b). This simple multifunctional nanoparticle system can potentially deliver Dox to the targeted cells; the cytotoxicity of the QD-Apt(Dox) conjugate was significantly enhanced against the targeted LNCaP cell line as compared to the nontargeted PC3 cell line, and simultaneously
P1: OTA/XYZ P2: ABC c23 JWBS037-Chen January 25, 2011
13:28
Printer Name: Yet to Come
MULTIFUNCTIONAL NANOPARTICLES FOR COMBINED CANCER IMAGING AND THERAPY
579
(a)
QDot QDot
(b) QDot
FIGURE 23.5 FRET-based nanoprobe for cancer imaging, therapy, and sensing of drug delivery. (a) Schematic of QDs aptamer Dox bioconjugate based on Bi-FRET mechanism, donor–acceptor FRET between QDs (FRET 1) and Dox and donor–quencher FRET between Dox and aptamer (FRET 2). (b) In vitro imaging of QD-aptamer-Dox bioconjugates in PSMA overexpressing LNCaP cell lines at 0 h incubation and 1.5 h postincubation. (Reproduced with permission from Bagalkot et al. [108].)
we could sense the delivery of Dox by activating the fluorescence of QD, which concurrently images the cancer cells. QDs have also been used to deliver biological drugs such as siRNA. For example, it has been shown by Gao and colleagues that semiconductor QDs with proton absorbing polymeric coatings with a balanced composition of tertiary amine and carboxylic acid groups provide steric and electrostatic interactions that are highly responsive to the acidic organelles and are also well suited for siRNA binding and cellular entry [128]. The QD covered with polymer having tertiary amines provides positive charge that can electrostatically interact with negatively charged siRNA and such positively charged QDs can enter into cells by well-known fluid phase endocytosis or macropinocytosis. The key features for nanoparticle-based delivery agents for siRNA is their cellular entry and endosomal escape to release the siRNA into cytoplasm and coupling with RNAinduced silencing complex (RISC). The positive charges on the polymer can induce the proton sponge effect. The weak conjugate base of these groups (tertiary amines, carboxyls) can absorb protons in acidic organelles such as endosomes and lead to an osmotic pressure build up across the organelle membrane, causing it to swell and finally rupture, releasing the trapped materials into the cytoplasm.
P1: OTA/XYZ P2: ABC c23 JWBS037-Chen January 25, 2011
580
13:28
Printer Name: Yet to Come
NANOPARTICLES FOR COMBINED CANCER IMAGING AND THERAPY
Homing peptide
Internalization
PEG QD core
Nucleus
Endosome Escape
siRNA
Interaction with RISC (mRNA degradation)
FIGURE 23.6 Design of a multifunctional nanoparticle for siRNA delivery. Because of their photostable fluorescence and multivalency, QDs are suitable vehicles for ferrying siRNA into live cells in vitro and in vivo. Conjugation of homing peptides (along with the siRNA cargo) to the QD surface allows targeted internalization in tumor cells. Once internalized, these particles must escape the endolysomal pathway and reach the cytoplasm to interact with the RNA-induced silencing complex (RISC), which leads to degradation of mRNA homologous to the siRNA sequence. (Reproduced with permission from Derfus et al. [129].)
For drug (chemotherapeutics) or gene (antisense, siRNA, and plasmid) delivery, it is important to improve the pharmacokinetics, optimize the target-to-nontarget ratio, prevent degradation, tune optimal release, and maintain unaltered bioactivities. Nanoparticles are thus favorable for the above reasons as they have large surfaces that can be used to attach ligands for targeting, drug molecules, and imaging agents into a single particle, and one such example using targeted QDs conjugates to deliver siRNA has been reported by Derfus et al. [129]. A PEGlyated QD core was used as a scaffold to conjugate siRNA and tumor-homing peptides (F3) to the functional groups on the surface of the QD [129]. Most importantly, siRNA attached by a reducible disulfide crosslinker (sulfo SPDP) showed greater silencing efficiency than when attached by a nonreducible crosslinker. When evaluated in an EGFP model system (EGFP transfected HeLa cells), F3/siRNA-QDs showed significant knockdown of EGFP signal (Fig. 23.6).
23.3.3 Gold Nanoparticles/Nanoshells Gold nanoparticles, also called noble metals, have huge potential in cancer diagnosis and therapy because of their property of “surface plasmon resonance” (SPR) enhanced light scattering and absorption [130]. The surface plasmon resonance phenomenon occurs for spherical nanoparticles much smaller than the wavelength of light, when an electromagnetic field at a certain frequency induces a resonant coherent oscillation of the metal free electrons across the nanoparticle surface. This occurs at visible wavelengths for noble metals such as gold, silver, and copper and the SPR frequency depends on many factors such as size, shape, solvent, surface functionalization, and refractive index of the surrounding medium [131]. Because of their unique optical properties that can be tuned by changing the size and shape,
P1: OTA/XYZ P2: ABC c23 JWBS037-Chen January 25, 2011
13:28
Printer Name: Yet to Come
MULTIFUNCTIONAL NANOPARTICLES FOR COMBINED CANCER IMAGING AND THERAPY
581
they are widely used as contrast agents in imaging; they are nontoxic and biocompatible compared to heavy metal bearing QDs. SPR can be tuned for optical imaging from the visible to near-infrared region; the gold surface is amenable to functionalization (strong affinity toward thiols, disulfides, and amines), which provides attachment sites for bioconjugation of various targeting ligands, drugs, and genes. They are resistant to photodecomposition, and, most importantly, they can convert absorbed light into heat and are being used for photothermal cancer therapy. Gold nanoparticles have tunable plasmon resonance in the visible region and are widely used for in vitro labeling and thermal therapy [132]. For in vivo imaging and photothermal therapy applications that require penetration of light into deeper tissues, the nanoparticles must absorb and scatter light in the NIR region (650–900 nm); in this region the body tissue has minimum absorption and scattering, and light penetration only occurs up to a few centimeters. The near-infrared (NIR) absorbing gold nanoparticles, called “nanoshells,” are composed of a silica core with a surrounding layer of gold shell; they were first reported by the Jennifer L. West research group. Gold nanoshells, nanorods, and nanocages can be designed to absorb and scatter light in the NIR [133]. Such gold nanoshells have been used in vivo for optical contrast enhancement and simultaneous photothermal ablation of tumors [134]. 23.3.4 Imaging Nanoparticles Embedded in Polymers or Liposomes Nanomedicines such as polymer-coated liposomes (Doxil® /Caelyx® ), polymeric drugs (Copaxone® ), antibodies (Herceptin® , AvastinTM ), polymer–protein conjugates (Oncospar® ), and the nanoparticle drug AbraxaneTM are already bringing clinical benefits to cancer patients [17]. Research is currently being directed at development of multicomponent nanomedicines for simultaneous diagnostic imaging and therapy; imaging agents such as iron oxide nanoparticles, gold nanoparticles, and QDs combined with therapeutic drugs into a single vehicle can track the outcome of drug therapy and precisely inform the location of tumor by imaging, demonstrating efficacy by proper delivery of drugs to the tumor site through proper drug release. Below, we discuss a few related examples of such nanocomposite structures built from the previous cancer technologies. For example, Ji-Ho Park and colleagues reported the development of hybrid tricomponent PEGylated phospholipid micellar nanoparticles consisting of magnetic nanoparticles, QDs, and the anticancer drug doxorubicin [135]. The micelles had a hydrodynamic size of 60–70 nm with closely packed iron oxide nanoparticles and QDs. The near-infrared (NIR) flourescence of QDs (subnanomolar) in the presence of iron oxide nanoparticles (submicromolar) was not altered (though quenched) and was still detectable; the superparamagnetic nature of the embedded iron oxide nanoparticles remained unchanged. As polymeric sterically stabilized micelles composed of amphiphilic block copolymers are in general found to be stable in vivo, attributed to the PEG part, such hybrid micellar nanoparticles exhibit long blood circulation time in vivo and were detected by both NIR fluorescence and MR imaging at 20 h postinjection in mice bearing human breast cancer (MDA-MB-435) tumors. Its applicability toward drug delivery was studied in vitro in cell lines, and the Dox-loaded F3 peptide targeted hybrid micelles were indeed toxic to breast cancer cells compared to naked hybrid micelles without drug, demonstrating that the toxicity comes directly from the drug and not the heavy metal (QD) bearing micelles. The study shows that hybrid nanostructures can be designed to incorporate two or more imaging agents and drugs in a biocompatible long circulating tumor-targeted micelle-based delivery platform. In a similar study, Nasongkla and colleagues report the development of
P1: OTA/XYZ P2: ABC c23 JWBS037-Chen January 25, 2011
582
13:28
Printer Name: Yet to Come
NANOPARTICLES FOR COMBINED CANCER IMAGING AND THERAPY
(a)
(b)
(c)
FIGURE 23.7 Multifunctional polymeric micelles as cancer-targeted, MRI-ultrasensitive drug delivery systems. (a) Amphiphilic block copolymer core–shell micelles with hydrophilic shell that ensures stability of particles in aqueous solution and the hydrophobic core that was loaded with both chemotherapeutic drug and SPIO nanoparticles; the cRGD targeting peptide was attached to the maleimide end functional groups of the polymer. (b) Transmission electron microscopy (TEM) images of 16% cRGD-DOXO-SPIO-loaded polymeric micelles; inset shows cryo-TEM of the same sample (scale bar is 20 nm). (c) Inhibition of SLK cell growth in the presence of different formulations of SPIO-loaded micelles with or without 1 M DOXO concentration after 4 h incubation time. (Reproduced with permission from Nasongkla et al. [136].)
multifunctional bicomponent polymeric micelles made from amphiphilic block copolymer of poly(ethylene glycol)-block-poly(d,l-lactide) (Fig. 23.7a,b) [136]. The cores of the micelles were loaded with SPIONs and the anticancer drug doxorubicin; cyclic RGD peptides were conjugated to the surface of the micelle that specifically targets the tumor through recognition of the unique cancer-specific biomarker, “integrin,” that is highly expressed on the endothelial cells of the tumor vasculature and is a well-known biomarker for development of targeted drug delivery platforms. The bicomponent integrin targeted micelles at nanomolar concentration showed good MRI visibility and specific targeting to integrin expressing SLK endothelial cells. Furthermore, when tested for its cytotoxicity with increasing cRGD percentage on micelles, the cytotoxic nature also increased (Fig. 23.7c). The study elegantly demonstrates that with increasing targeting efficacy the cytotoxicity of the hybrid drug encapsulated micelles also increased. Similarly, Yang and colleagues reported multifunctional amphiphilic block copolymers that were encapsulated with magnetic nanoparticles and the anticancer drug doxorubicin,
P1: OTA/XYZ P2: ABC c23 JWBS037-Chen January 25, 2011
13:28
Printer Name: Yet to Come
MULTIFUNCTIONAL NANOPARTICLES FOR COMBINED CANCER IMAGING AND THERAPY
583
targeted with anti-HER antibody that was attached to the carboxyl groups on the surface of the polymer [137]. The HER targeted polymeric magnetic nanohybrids were detected by MR imaging both in vitro in NIH3T6.7 cell line and in vivo in nude mice bearing NIH3T6.7 breast cancer tumor models. The magnetic polymeric nanohybrids contain the chemotherapeutic Dox and anti-HER antibody showed synergistic effect on tumor inhibition and, indeed, when intravenously injected, it resulted in effective tumor growth inhibition in vivo in nude mice bearing human xenograft breast cancer (NIH3T6.7) tumor models; the efficient targeting of the polymeric hybrids to overexpressed HER2/neu receptors on NIH3T6.7 cells in the tumor model and release of Dox caused the tumors to shrink and showed exceptional therapeutic efficacy. In the above examples, polymeric micelles made from amphiphilic block copolymers were loaded with diagnostic imaging agents and chemotherapy drugs and were specifically targeted to the tumors through the attachment of specific targeting ligands; thus such targeted multifunctional platforms that can deliver drugs specifically and image tumors noninvasively offer exciting oppurtunities for further development of these systems to monitor the outcome of therapy. Liposomes are vesicles made up of phospholipid bilayers that can encapsulate drugs both in an inner aqueous compartment and a lipid bilayer, can improve biodistribution and drug clearance, and can aid in superior drug delivery to tumor sites. FDA approved pegylatedSTEALTH liposomal formulation containing doxorubicin (Doxil) is already used clinically, and novel antibody targeted liposomal nanoparticles (immunoliposomes against HER2 and EGFR) are in preclinical development [138]. Recently, these targeted drug-loaded liposomal structures have also been combined with imaging agents such as QDs as reported in a study by Wang and colleagues who conjugated QDs to liposomes targeted with anti-HER2 antibody (immunoliposomes, ILs) and loaded them with the anticancer drug doxorubicin; the multifunctional QD-IL hybrid nanoparticles (∼212 nm) were assembled with imaging, targeting, and therapeutic modalities [139]. The hybrid QD-IL nanoparticles containing Dox showed anticancer activity against HER2-overexpressing SK-BR-3 cells in vitro. The in vivo pharmacokinetic study revealed that QD conjugated to immunoliposomes showed long circulation time of ∼3 h without further signs of gross toxicity: there was no weight loss in mice during the 3-month study period. When the QD-IL hybrid nanoparticles were evaluated for targeted cancer imaging (flourescence imaging) in vivo in nude mice bearing HER2-overexpressing MCF7/HER2 xenografts, flourescent signals were detected in both mononuclear phagocytic system (MPS) organs, such as liver and spleen that mediate liposome clearance, and tumors. The study also showed that tumor localization for QD-tagged liposomes was mostly due to passive targeting by the EPR effect, as both targeted and nontargeted hybrid nanoparticles showed a similar percentage of flourescence within tumors at 24 h. The study points out that targeted QD-tagged liposome hybrid nanoparticles can be used to study nanoparticle biodistribution and to monitor drug delivery in real-time in vivo conditions and can also label tumor cells; the nonspecific uptake by the MPS organs could be due to the large size of these nanoparticles and this system may be improved through use of much smaller QDs or through designing a smaller sized construct that can evade size-based clearance in the MPS organs and can be optimized to accumulate with higher percentage in the targeted tumor site. Biodegradable, biocompatible polymers such as chitosan have also been used to encapsulate QDs and monitor the delivery of siRNA targeting HER2/neu gene as reported by Tan and colleagues. The QDs (anionic) were encapsulated within chitosan (200 kDa)
P1: OTA/XYZ P2: ABC c23 JWBS037-Chen January 25, 2011
584
13:28
Printer Name: Yet to Come
NANOPARTICLES FOR COMBINED CANCER IMAGING AND THERAPY
(cationic) through charge–charge interactions and the QD encapsulated chitosan was mixed with HER2 siRNA, with 85% conjugation efficiency of siRNA observed when the N/P ratio (chitosan amino group/RNA phosphate group) was around 8; the resulting HER2 siRNA complexed QD encapsulated chitosan nanoparticles were also conjugated to HER2 antibody for specific targeting to HER2 overexpressing SKBR3 cells [127]. The chitosan QD hybrid nanoparticles showed targeted delivery of HER2 siRNA to SKBR3 breast cancer cells and gene silencing effect of conjugated siRNA in vitro, and the particles could be tracked due to the presence of flourescent QDs. The study shows that such QD encapsulated biocompatible chitosan-based hybrid nanoparticles could be useful for future monitoring of targeted gene silencing studies in vivo.
23.4 PROSPECTS AND CHALLENGES FOR TRANSLATION INTO CLINICS A variety of examples of multifunctional nanoparticles have demonstrated their potential for use in combined cancer imaging and therapy. With those nanoparticles we may be able to know where a tumor is located, whether drugs are properly delivered to the tumor, and what the therapeutic response of the tumor is [91]. To translate such emerging, promising nanomedicines into the clinic for human treatment, several limitations must be overcome. A major issue is that the biodistribution of the nanoparticles on tumor-to-normal tissues is still unsatisfactory. Despite largely increased tumor accumulation of anticancer drugs by nanoparticle-based systems as compared to conventional anticancer therapy, major sites of uptake are still the liver and spleen, suggesting that more burden may be imposed on those organs. The other issue for current inorganic nanoparticle-based nanomedicine is the high risk of associated toxicity with their long-term residence in the body without degradation [91]. In the case of conventional therapy, most drugs can rapidly be cleared from the body after nonspecific biodistribution throughout the body, which may result in lower overall toxicity than in the case of nanoparticle systems. To resolve the latter issue, biodegradable inorganic nanoparticles such as SPIONs could be used, which seem to have promise for clinical applications because of their biodegradability, good biocompatibility, and abilities to deliver drugs and to produce strong MR signal. For other inorganic nanoparticles such as QDs and gold, more extensive studies with regard to long-term toxicity are necessary to meet requirements for clinical applications. To resolve the unsatisfactory biodistribution profiles of most inorganic nanoparticles, we need to reduce particle sizes to fall into the 5–10-nm range, including coating layers; furthermore, we need to develop more efficient coating technologies and materials that can evade macrophages as well as splenic and liver filters so as to achieve optimized circulation time for higher tumor accumulation but lower residence in the body. As accumulation of knowledge on nanoparticles’ toxicities eases people’s concerns about nanomedicine, it is anticipated that a few of the current multifunctional nanoparticles will be able to advance into clinical phases in the near future.
REFERENCES 1. Ferrari, M. Cancer nanotechnology: opportunities and challenges. Nat. Rev. Cancer 2005, 5, 161–171. 2. Azzazy, H. M.; Mansoor, M. M.; Kazmierczak, S. C. From diagnostics to therapy: prospects of quantum dots. Clin. Biochem. 2007, 40, 917–927.
P1: OTA/XYZ P2: ABC c23 JWBS037-Chen January 25, 2011
13:28
Printer Name: Yet to Come
REFERENCES
585
3. Owens, D. E.; Peppas, N. A. Opsonization, biodistribution, and pharmacokinetics of polymeric nanoparticles. Int. J. Pharm. 2006, 307, 93–102. 4. Opanasopit, P.; Nishikawa, M.; Hashida, M. Factors affecting drug and gene delivery: effects of interaction with blood components. Crit. Rev. Ther. Drug Carrier Syst. 2002, 19, 191–233. 5. Nagayama, S.; Ogawara, K.; Fukuoka, Y.; Higaki, K.; Kimura, T. Time-dependent changes in opsonin amount associated on nanoparticles alter their hepatic uptake characteristics. Int. J. Pharm. 2007, 342, 215–221. 6. Levchenko, T. S.; Rammohan, R.; Lukyanov, A. N.; Whiteman, K. R.; Torchilin, V. P. Liposome clearance in mice: the effect of a separate and combined presence of surface charge and polymer coating. Int. J. Pharm. 2002, 240, 95–102. 7. Zhang, J. S.; Liu, F.; Huang, L. Implications of pharmacokinetic behavior of lipoplex for its inflammatory toxicity. Adv. Drug Deliv. Rev. 2005, 57, 689–698. 8. Papisov, M. I.; Bogdanov, A.; Schaffer, B.; Nossiff, N.; Shen, T.; Weissleder, R.; Brady, T. J. Colloidal magnetic resonance contrast agents: effect of particle surface on biodistribution. J. Magn. Magn. Mater. 1993, 122, 383–386. 9. Moghimi, S. M. Exploiting bone marrow microvascular structure for drug delivery and therapies. Adv. Drug Deliv. Rev. 1995, 17, 61–73. 10. Moghimi, S. M. Mechanisms of splenic clearance of blood cells and particles: towards development of new splenotropic agents. Adv. Drug Deliv. Rev. 1995, 17, 103–115. 11. Shen, T.; Weissleder, R.; Papisov, M.; Bogdanov, A., Jr.; Brady, T. J. Monocrystalline iron oxide nanocompounds (MION): physicochemical properties. Magn. Reson. Med. 1993, 29, 599–604. 12. Josephson, L.; Tung, C. H.; Moore, A.; Weissleder, R. High-efficiency intracellular magnetic labeling with novel superparamagnetic-Tat peptide conjugates. Bioconjug. Chem. 1999, 10, 186–191. 13. Bhadra, D.; Bhadra, S.; Jain, S.; Jain, N. K. A PEGylated dendritic nanoparticulate carrier of fluorouracil. Int. J. Pharm. 2003, 257, 111–124. 14. Fawaz, F.; Bonin, F.; Guyot, M.; Lagueny, A.-M.; Hatem, F.; Devissaguet, J.-P. Influence of poly(dl-lactide) nanocapsules on the biliary clearance and enterohepatic circulation of indomethacin in the rabbit. Pharm. Res. 1993, 10, 750–756. 15. Gref, R.; Luck, M.; Quellec, P.; Marchand, M.; Dellacherie, E.; Harnisch, S.; Blunk, T.; Muller, R. H. Stealth corona-core nanoparticles surface modified by polyethylene glycol (PEG): influences of the corona (PEG chain length and surface density) and of the core composition on phagocytic uptake and plasma protein adsorption. Colloids Surf. B Biointerfaces 2000, 18, 301–313. 16. Otsuka, H.; Nagasaki, Y.; Kataoka, K. PEGylated nanoparticles for biological and pharmaceutical applications. Adv. Drug Deliv. Rev. 2003, 55, 403–419. 17. Zhang, L.; Gu, F. X.; Chan, J. M.; Wang, A. Z.; Langer, R. S.; Farokzad, O. C. Nanoparticles in medicine: therapeutic applications and developments. Clin. Pharmacol. Ther. 2008, 83, 761–769. 18. Akerman, M. E.; Chan, W. C. W.; Laakkonen, P.; Bhatia, S. N.; Ruoslahti, E. Nanocrystal targeting in vivo. Proc. Natl. Acad. Sci. U.S.A. 2002, 99, 12617–12621. 19. Jorgensen, K. E.; Moller, J. V. Use of flexible polymers as probes of glomerular pore size. Am. J. Physiol. Renal Physiol. 1979, 236, 103–111. 20. Uyeda, H. T.; Medintz, I. L.; Jaiswal, J. K.; Simon, M. S.; Mattoussi, H. Synthesis of compact multidentate ligands to prepare stable hydrophilic quantum dot fluorophores. J. Am. Chem. Soc. 2005, 127, 3870–3878. 21. Kang, E.-C.; Ogura, A.; Kataoka, K.; Nagasaki, Y. Preparation of water-soluble PEGylated semiconductor nanocrystals. Chem. Lett. 2004, 33, 840–841.
P1: OTA/XYZ P2: ABC c23 JWBS037-Chen January 25, 2011
586
13:28
Printer Name: Yet to Come
NANOPARTICLES FOR COMBINED CANCER IMAGING AND THERAPY
22. Dixit, S. K.; Goicochea, N. L.; Daniel, M, C.; Murali, A.; Bronstein, L.; De, M.; Stein, B.; Rotello, V. M.; Kao, C. C.; Dragnea, B. Quantum dot encapsulation in viral capsids. Nano Lett. 2006, 6, 1993–1999. 23. Derfus, A. M.; Chan, W. C. W.; Bhatia, S. N. Intracellular delivery of quantum dots for live cell labelling and organelle tracking. Adv. Mater. 2004, 16, 961–966. 24. Hu, F. Q.; Ran, Y. L.; Zhou, Z. A. Preparation of bioconjugates of CdTe nanocrystals for cancer marker detection. Nanotechnology 2006, 17, 2972–2977. 25. Duan, H.; Nie, S. Cell-penetrating quantum dots based on multivalent and endosome-disrupting surface coatings. J. Am. Chem. Soc. 2007, 129, 3333–3338. 26. Dubertret, B.; Skourides, P.; Norris, D. J.; Noireaux, V.; Brivanlou, A. H.; Libchaber, A. In vivo imaging of quantum dots encapsulated in phospholipid micelles. Science 2002, 298, 1759–1762. 27. Srinivasan, C.; Lee, J.; Papadimitrakopoulos, F.; Silbart, L. K.; Zhao, M.; Burgess, D. J. Labelling and intracellular tracking of functionally active plasmid DNA with semiconductor quantum dots. Mol. Ther. 2006, 14, 192–200. 28. Lee, H.; Lee, E.; Kim, D. K.; Jang, N. K.; Jeong, Y. Y.; Jon, S. Antibiofouling polymer-coated superparamagnetic iron oxide nanoparticles as potential magnetic resonance contrast agents for in vivo cancer imaging. J. Am. Chem. Soc. 2006, 128, 7383–7389. 29. Fang, C.; Shi, B.; Pei, Y.-Y.; Hong, M.-H.; Wu, J.; Chen, H.-Z. In vivo tumor targeting of tumor necrosis factor-R-loaded stealth nanoparticles: effect of MePEG molecular weight and particle size. Eur. J. Pharm. Sci. 2006, 27, 27–36. 30. Liu, D.; Mori, A.; Huang, L. Role of liposome size and RES blockade in controlling biodistribution and tumor uptake of GM1-containing liposomes. Biochim. Biophys. Acta. 1992, 1104, 95–101. 31. Choi, H. S.; Liu, W.; Misra, P.; Tanaka, E.; Zimmer, J. P.; Itty Ipe B.; Bawendi, M. G.; Frangioni, J. V. Renal clearance of quantum dots. Nat. Biotechnol. 2007, 25, 1165–1170. 32. Moghimi, S. M.; Hunter, A. C.; Murray, J. C. Long-circulating and target specific nanoparticles: theory to practice. Pharm. Rev. 2001, 53, 283–318. 33. Hobbs, S. K.; Monsky, W. L.; Yuan, F.; Roberts, W. G.; Griffith, L.; Torchilin, V. P.; Jain, R. K. Regulation of transport pathways in tumor vessels: role of tumor type and microenvironment. Proc. Natl. Acad. Sci. U.S.A. 1998, 95, 4607–4612. 34. Maeda, H.; Fang, J.; Inutsuka, T.; Kitamoto, Y. Vascular permeability enhancement in solid tumor: various factors, mechanisms involved and its implications. Int. Immunopharmacol. 2003, 3, 319–328. 35. Dvorak, H. F. Leaky tumor vessels: consequences for tumor stroma generation and for solid tumor therapy. Prog. Clin. Biol. Res. 1990, 354A, 317–330. 36. Yuan, F.; Leunig, M.; Huang, S. K.; Berk, D. A.; Papahadjopoulos, D.; Jain, R. K. Microvascular permeability and interstitial penetration of sterically stabilized (stealth) liposomes in a human tumor xenograft. Cancer Res. 1994, 54, 3352–3356. 37. Yuan, F.; Dellian, M.; Fukumura, D.; Leunig, M.; Berk, D. A.; Torchilin, V. P.; Jain, R. K. Vascular permeability in a human tumor xenograft: molecular size dependence and cutoff size. Cancer Res. 1995, 55, 3752–3756. 38. Wang, X.; Yang, L.; Chen, Z. G.; Shin, D. M. Application of nanotechnology in cancer therapy and imaging. CA Cancer J. Clin. 2008, 58, 97–110. 39. Barry, A. J. Clinical trials of targeted alpha therapy for cancer. Rev Recent Clin. Trials 2008, 3, 185–191. 40. Duncan, R. The dawning era of polymer therapeutics. Nat. Rev. Drug Discov. 2003, 2, 347– 360.
P1: OTA/XYZ P2: ABC c23 JWBS037-Chen January 25, 2011
13:28
Printer Name: Yet to Come
REFERENCES
587
41. Harris, J. M.; Chess, R. B. Effect of pegylation on pharmaceuticals. Nat. Rev. Drug Discov. 2003, 2, 214–221. 42. Torchilin, V. P. Recent advances with liposomes as pharmaceutical carriers. Nat. Rev. Drug Discov. 2005, 4, 145–160. 43. Moghimi, S. M.; Hunter, A. C.; Murray, J. C. Nanomedicine: current status and future prospects. FASEB J. 2005, 19, 311–330. 44. Brigger, I.; Dubernet, C.; Coureur, P. Nanoparticles in cancer therapy and diagnosis. Adv. Drug Deliv. Rev. 2002, 54, 631–651. 45. Vauthier, C.; Dubernet, C.; Fattal, E.; Pinto-Alphandary, H.; Couvreur, P. Poly(alkylcyanoacrylates) as biodegradable materials for biomedical applications. Adv. Drug Deliv. Rev. 2003, 55, 519–548. 46. Marcucci, F.; Lefoulon, F. Active targeting with particulate drug carriers in tumor therapy: fundamentals and recent progress. Drug Discov. Today 2004, 9, 219–228. 47. Wang, Y. X.; Hussain, S. M.; Krestin, G. P. Superparamagnetic iron oxide contrast agents: physicochemical characteristics and applications in MR imaging. Eur. Radiol. 2001, 11, 2319–2331. 48. Ballou, B.; Lagerholm, B. C.; Ernst, L. A.; Bruchez, M. P.; Waggoner, A. S. Noninvasive imaging of quantum dots in mice. Bioconjug. Chem. 2004, 15, 79–86. 49. Jayagopal, A.; Russ, P. K.; Frederick, H. R. Surface engineering of quantum dots for in vivo vascular imaging. Bioconjug. Chem. 2007, 18, 1424–1433. 50. Stroh, M.; Zimmer, J. P.; Duda, D. G.; Levchenko, T. S.; Cohen, K. S.; Brown, E. B.; Scadden, D. T.; Torchilin, V. P.; Bawendi, M. G.; Fukumura, D.; Jain, R. K. Quantum dots spectrally distinguish multiple species within the tumor milieu in vivo. Nat. Med. 2005, 11, 678–682. 51. Kim, S.; Lim, Y. T.; Soltesz, E. G.; De Grand, A. M.; Lee, J.; Nakayama, A.; Parker, J. A.; Mihaljevic, T.; Laurence, R. G.; Dor, D. M.; Cohn, L. H.; Bawendi, M. G.; Frangioni, J. V. Nearinfrared fluorescent type II quantum dots for sentinel lymph node mapping. Nat. Biotechnol. 2004, 22, 93–97. 52. Cho, K.; Wang, X.; Nie, S.; Chen, Z. G.; Shin, D, M. Therapeutic nanoparticles for drug delivery in cancer. Clin. Cancer Res. 2008, 14, 1310–1316. 53. Gao, X.; Cui, Y.; Levenson, R. M.; Chung, L. W.; Nie, S. In vivo cancer targeting and imaging with semiconductor quantum dots. Nat. Biotechnol. 2004, 22, 969–976. 54. Wu, X.; Liu, H.; Haley, K. N.; Treadway, J. A.; Larson, J. P.; Ge, N.; Peale, F.; Bruchez, M. P. Immunofluorescent labeling of cancer marker Her2 and other cellular targets with semiconductor quantum dots. Nat. Biotechnol. 2003, 21, 41–46. 55. Tada, H.; Higuchi, H.; Wanatabe, T. M.; Ohuchi, N. In vivo real-time tracking of single quantum dots conjugated with monoclonal anti-HER2 antibody in tumors of mice. Cancer Res. 2007, 67, 1138–1144. 56. Cai, W.; Shin, D. W.; Chen, K.; Gheysens, O.; Cao, Q.; Wang, S. X.; Gambhir, S. S.; Chen, X. Peptide-labeled near-infrared quantum dots for imaging tumor vasculature in living subjects. Nano Lett. 2006, 6, 669–676. 57. Yu, X. Immunofluorescence detection with quantum dot bioconjugates for hepatoma in vivo. J. Biomed. Optics 2007, 12, 014008–014005. 58. Cerdan, S.; Lotscher, H. R.; Kunnecke, B.; Seelig, J. Monoclonal antibody-coated magnetite particles as contrast agents in magnetic resonance imaging of tumors. Magn. Reson. Med. 1989, 12, 151–163. 59. Artemov, D.; Mori, N.; Ravi, R.; Bhujwalla, Z. M. Magnetic resonance molecular imaging of the HER-2/neu receptor. Cancer Res. 2003, 63, 2723–2727.
P1: OTA/XYZ P2: ABC c23 JWBS037-Chen January 25, 2011
588
13:28
Printer Name: Yet to Come
NANOPARTICLES FOR COMBINED CANCER IMAGING AND THERAPY
60. Artemov, D. Molecular magnetic resonance imaging with targeted contrast agents. J. Cell Biochem. 2003, 90, 518–524. 61. Zhang, C.; Jugold, M.; Woenne, E. C.; Lammers, T.; Morgenstern, B.; Mueller, M. M.; Zentgraf, H.; Bock, M.; Eisenhut, M.; Semmler, W.; Kiessling, F. Specific targeting of tumor angiogenesis by RGD-conjugated ultrasmall superparamagnetic iron oxide particles using a clinical 1.5-T magnetic resonance scanner. Cancer Res. 2007, 67, 1555–1562. 62. Sun, C.; Sze, R.; Zhang, M. Folic acid–PEG conjugated superparamagnetic nanoparticles for targeted cellular uptake and detection by MRI. J. Biomed. Mater. Res. Part A 2006, 78A, 550–557. 63. Lewin, M.; Carlesso, N.; Tung, C.-H.; Tang, X.-W.; Cory, D.; Scadden, D. T.; Weissleder, R. Tat peptide-derivatized magnetic nanoparticles allow in vivo tracking and recovery of progenitor cells. Nat. Biotechnol. 2000, 18, 410–414. 64. Kim, D.; Park, S.; Lee, J. H.; Jeong, Y. Y.; Jon, S. Antibiofouling polymer-coated gold nanoparticles as a contrast agent for in vivo X-ray computed tomography imaging. J. Am. Chem. Soc. 2007, 129, 7661–7665. 65. Farokhzad, O. C.; Cheng, J.; Teply, B. A.; Sherifi, I.; Jon, S.; Kantoff, P. W.; Richie, J. P.; Langer, R. Targeted nanoparticle–aptamer bioconjugates for cancer chemotherapy in vivo. PNAS 2006, 103, 6315–6320. 66. Cheng, J.; Teply, B. A.; Sherifi, I.; Sung, J.; Luther, G.; Gu, F. X.; Levy-Nissenbaum, E.; Radovic-Moreno, A. F.; Langer, R.; Farokhzad, O. C. Formulation of functionalized PLGAPEG nanoparticles for in vivo targeted drug delivery. Biomaterials 2007, 28, 869–876. 67. Baghi, M.; Mack, M. G.; Hambek, M.; Rieger, J.; Vogl, T.; Gstoettner, W.; Knecht, R. The efficacy of MRI with ultrasmall superparamagnetic iron oxide particles (USPIO) in head and neck cancers. Anticancer Res. 2005, 25, 3665–3670. 68. Martina, M. S.; Fortin, J. P.; Menager, C.; Clement, O.; Barratt, G.; Grabielle-Madelmont, C.; Gazeau, F.; Cabuil, V.; Lesieur, S. Generation of superparamagnetic liposomes revealed as highly efficient MRI contrast agents for in vivo imaging. J. Am. Chem. Soc. 2005, 127, 10676–10685. 69. Blasberg, R. G. Molecular imaging and cancer. Mol. Cancer Ther. 2003, 2, 335–343. 70. Kroft, L. J.; de Roos, A. Blood pool contrast agents for cardiovascular MR imaging. J. Magn. Reson. Imaging. 1999, 10, 395–403. 71. Bonnemain, B. Superparamagnetic agents in magnetic resonance imaging: physicochemical characteristics and clinical applications. A review. J. Drug Target. 1998, 6, 167–174. 72. Bellin, M. F.; Beigelman, C.; Precetti-Morel, S. Iron oxide-enhanced MR lymphography: initial experience. Eur. J. Radiol. 2000, 34, 257–264. 73. Kohler, N.; Sun, C.; Wang, J.; Zhang, M. Methotrexate-modified superparamagnetic nanoparticles and their intracellular uptake into human cancer cells. Langmuir. 2005, 21, 8858– 8864. 74. Gupta, A. K.; Curtis, A. S. Surface modified superparamagnetic nanoparticles for drug delivery: interaction studies with human fibroblasts in culture. J. Mater. Sci. Mater. Med. 2004, 15, 493–496. 75. Gupta, A. K.; Gupta, M. Synthesis and surface engineering of iron oxide nanoparticles for biomedical applications. Biomaterials 2005, 26, 3995–4021. 76. Ito, A.; Kuga, Y.; Honda, H.; Kikkawa, H.; Horiuchi, A.; Watanabe, Y.; Kobayashi, T. Magnetite nanoparticle-loaded anti-HER2 immunoliposomes for combination of antibody therapy with hyperthermia. Cancer Lett. 2004, 212, 167–175. 77. Nagesha, D.; Srinivas, H. D.; Sridhar Amiji, M. Multifunctional Magnetic Nanosystems for Tumor Imaging, Targeted Drug Delivery, and Thermal Medicine. Springer Science LLC: New York, 2008.
P1: OTA/XYZ P2: ABC c23 JWBS037-Chen January 25, 2011
13:28
Printer Name: Yet to Come
REFERENCES
589
78. Raynal, I.; Prigent, P.; Peyramaure, S.; Najid, A.; Rebuzzi, C.; Corot, C. Macrophage endocytosis of superparamagnetic iron oxide nanoparticles: mechanisms and comparison of ferumoxides and ferumoxtran-10. Invest. Radiol. 2004, 39, 56–63. 79. Rogers, W. J.; Basu, P. Factors regulating macrophage endocytosis of nanoparticles: implications for targeted magnetic resonance plaque imaging. Atherosclerosis 2005, 178, 67–73. 80. Kohler, N.; Fryxell, G. E.; Zhang, M. A bifunctional poly(ethylene glycol) silane immobilized on metallic oxide-based nanoparticles for conjugation with cell targeting agents. J. Am. Chem. Soc. 2004, 126, 7206–7211. 81. Kaufman, C. L.; Mangay, W.; Madison, R. L.; Tract, S. L.; Mike, T.; Ho, C. Superparamagnetic iron oxide particles transactivator protein-fluorescein isothiocyanate particle labeling for in vivo magnetic resonance imaging detection of cell migration: uptake and durability. Transplantation 2003, 76, 1043–1046. 82. Berry, C. C.; Wells, S.; Charles, S.; Aitchison, G.; Curtis, A. S. Cell response to dextranderivatised iron oxide nanoparticles post internalisation. Biomaterials 2004, 25, 5405–5413. 83. D’Souza, A. J.; Schowen, R. L.; Topp, E. M. Polyvinylpyrrolidone-drug conjugate: synthesis and release mechanism. J. Control. Release 2004, 94, 91–100. 84. Zhang, Y.; Kohler, N.; Zhang, M. Surface modification of superparamagnetic magnetite nanoparticles and their intracellular uptake. Biomaterials 2002, 23, 1553–1561. 85. Choi, H.; Choi, S. R.; Zhou, R.; Kung, H. F.; Chen, I. W. Iron oxide nanoparticles as magnetic resonance contrast agent for tumor imaging via folate receptor-targeted delivery. Acad. Radiol. 2004, 11, 996–1004. 86. Josephson, L.; Tung, C. H.; Moore, A.; Weissleder, R. High-efficiency intracellular magnetic labeling with novel superparamagnetic–Tat peptide conjugates. Bioconjug. Chem. 1999, 10, 186–191. 87. Hogemann, D.; Josephson, L.; Weissleder, R.; Basilion, J. P. Improvement of MRI probes to allow efficient detection of gene expression. Bioconjug. Chem. 2000, 11, 941–946. 88. Lewin, M.; Carlesso, N.; Tung, C.-H.; Tang, X.-W.; Cory, D.; Scadden, D. T.; Weissleder, R. Tat peptide-derivatized magnetic nanoparticles allow in vivo tracking and recovery of progenitor cells. Nat. Biotechnol. 2000, 18, 410–414. 89. Josephson, L.; Kircher, M. F.; Mahmood. U.; Tang, Y.; Weissleder, R. Near-infrared fluorescent nanoparticles as combined MR/optical imaging probes. Bioconjug. Chem. 2002, 13, 554–560. 90. Derfus, A. M.; Maltzahn, G. V.; Harris, T. J.; Duza, T.; Vecchio, K. S.; Ruoslahti, E.; Bhatia, S. N. Remotely triggered release from magnetic nanoparticles. Adv. Mater. 2007, 19, 3932– 3936. 91. Yu, M. K.; Jeong, Y. Y.; Park, J.; Park, S.; Kim, J. W.; Min, J. J.; Kim, K.; Jon, S. Drug-loaded superparamagnetic iron oxide nanoparticles for combined cancer imaging and therapy in vivo. Angew. Chem. Int. Ed. 2008, 47, 5362–5365. 92. Jain, T. K.; Morales, M. A.; Sahoo, S. K.; Leslie-Pelecky, D. L.; Labhasetwar, V. Iron oxide nanoparticles for sustained delivery of anticancer agents. Mol. Pharm. 2005, 2, 194–205. 93. Kohler, N.; Sun, C.; Fichtenholtz, A.; Gunn, J.; Fang, C.; Zhang, M. Methotrexate-immobilized poly(ethylene glycol) magnetic nanoparticles for MR imaging and drug delivery. Small 2006, 2, 785–792. 94. Shin, J.; Anisur, R. M.; Ko, M. K.; Im, G. H.; Lee, J. H.; Lee, I. S. Hollow manganese oxide nanoparticles as multifunctional agents for magnetic resonance imaging and drug delivery. Angew. Chem. Int. Ed. Engl. 2009, 48, 321–324. 95. Thorek, D. L. J.; Chen, A. K.; Czupryna, J.; Tsourkas, A. Superparamagnetic iron oxide nanoparticle probes for molecular imaging. Ann. Biomed. Eng. 2006, 34, 23–38. 96. Wang, A. Z.; Bagalkot, V.; Vasilliou, C. C.; Gu, F.; Alexis, F.; Zhang, L.; Shaikh, M.; Yuet, K.; Cima, M. J.; Langer, R.; Kantoff, P. W.; Bander, N. H.; Jon, S.; Farokhzad, O. C.
P1: OTA/XYZ P2: ABC c23 JWBS037-Chen January 25, 2011
590
97. 98. 99.
100.
101.
102.
103.
104. 105. 106.
107.
108.
13:28
Printer Name: Yet to Come
NANOPARTICLES FOR COMBINED CANCER IMAGING AND THERAPY
Superparamagnetic iron oxide nanoparticle–aptamer bioconjugates for combined prostate cancer imaging and therapy. Chem. Med. Chem. 2008, 3, 1311–1315. Medarova, Z.; Pham, W.; Farrar, C.; Petkova, V.; Moore, A. In vivo imaging of siRNA delivery and silencing in tumors. Nat. Med. 2007, 13, 372–377. Conti, P. S.; Lilien, D. L.; Hawley, K.; Keppler, J.; Grafton, S. T.; Bading, J. R. PET and [18 F]-FDG in oncology: a clinical update. Nucl. Med. Biol. 1996, 23, 717–735. Romond, E. H.; Perez, E. A.; Bryant, J.; Suman, V. J.; Geyer. C. E. Jr., Davidson, N. E.; TanChiu, E.; Martino, S.; Paik, S.; Kaufman, P. A.; Swain, S, M.; Pisansky, T, M.; Fehrenbacher, L.; Kutteh, L. A.; Vogel, V. G.; Visscher, D. W.; Yothers, G.; Jenkins, R. B.; Brown, A. M.; Dakhil, S. R.; Mamounas, E. P.; Lingle, W. L.; Klein, P. M.; Ingle, J. N.; Wolmark, N. Trastuzumab plus adjuvant chemotherapy for operable HER2-positive breast cancer. N. Engl. J. Med. 2005, 353, 1673–1684. Piccart-Gebhart, M. J.; Procter, M.; Leyland-Jones, B.; Goldhirsch, A.; Untch, M.; Smith, I.; Gianni, L.; Baselga, J.; Bell, R.; Jackisch, C.; Cameron, D.; Dowsett, M.; Barrios, C. H.; Steger, G.; Huang, C. S.; Andersson, M.; Inbar, M.; Lichinitser, M.; L´ang, I.; Nitz, U.; Iwata, H.; Thomssen, C.; Lohrisch, C.; Suter, T. M.; R¨uschoff, J.; Suto, T.; Greatorex, V.; Ward, C.; Straehle, C.; McFadden, E.; Dolci, M. S.; Gelber, R. D.; Herceptin Adjuvant (HERA) Trial Study Team. Trastuzumab after adjuvant chemotherapy in HER2-positive breast cancer. N. Engl. J. Med. 2005, 353, 1659–1672. Coiffier, B.; Lepage, E.; Briere, J.; Herbrecht, R.; Tilly, H.; Bouabdallah, R.; Morel, P.; Van Den, Neste, E.; Salles, G.; Gaulard, P.; Reyes, F.; Lederlin, P.; Gisselbrecht, C. CHOP chemotherapy plus rituximab compared with CHOP alone in elderly patients with diffuse large-B-cell lymphoma. N. Engl. J. Med. 2002, 346, 235–242. Czuczman, M. S.; Weaver, R.; Alkuzweny, B.; Berlfein, J.; Grillo-Lopez, A. J. Prolonged clinical and molecular remission in patients with low-grade or follicular non-Hodgkin’s lymphoma treated with rituximab plus CHOP chemotherapy: 9-year follow-up. J. Clin. Oncol. 2004, 22, 4711–4716. Hiddemann, W.; Kneba, M.; Dreyling, M.; Schmitz, N.; Lengfelder, E.; Schmits, R.; Reiser, M.; Metzner, B.; Harder, H.; Hegewisch-Becker, S.; Fischer, T.; Kropff, M.; Reis, H. E.; Freund, M.; Wormann, B.; Fuchs, R.; Planker, M.; Schimke, J.; Eimermacher, H.; Trumper, L.; Aldaoud, A.; Parwaresch, R.; Unterhalt, M. Frontline therapy with rituximab added to the combination of cyclophosphamide, doxorubicin, vincristine, and prednisone (CHOP) significantly improves the outcome for patients with advanced-stage follicular lymphoma compared with therapy with CHOP alone: results of a prospective randomized study of the German Low-Grade Lymphoma Study Group. Blood 2005, 106, 3725–3732. Zhang, H.; Berezov, A.; Wang, Q.; Zhang, G.; Drebin, J.; Murali, R.; Greene, M. I. ErbB receptors: from oncogenes to targeted cancer therapies. J. Clin. Invest. 2007, 117, 2051–2058. Hudis, C. A. Trastuzumab—mechanism of action and use in clinical practice. N. Engl. J. Med. 2007, 357, 39–51. Farokhzad, O. C.; Jon, S.; Khademhosseini, A.; Tran, T.-N.; La Van, D. A.; Langer, R. Nanoparticle-aptamer bioconjugates: a new approach for targeting prostate cancer cells. Cancer Res. 2004, 64, 7668–7672. Kim, S. H.; Joeng, J. H.; Chun, K. W.; Park, T. G. Target-specific cellular uptake of PLGA nanoparticles coated with poly(l-lysine)–poly(ethylene glycol)–folate conjugate. Langmuir 2005, 21, 8852–8857. Bagalkot, V.; Zhang, L.; Levy-Nissenbaum, E.; Jon, S.; Kantoff, P. W.; Langer, R.; Farokhzad, O. C. Quantum dot–aptamer conjugates for synchronous cancer imaging, therapy, and sensing of drug delivery based on bi-fluorescence resonance energy transfer. Nano Lett. 2007, 7, 3065–3070.
P1: OTA/XYZ P2: ABC c23 JWBS037-Chen January 25, 2011
13:28
Printer Name: Yet to Come
REFERENCES
591
109. Israeli, R. S.; Powell, C. T.; Corr, J. G.; Fair, W. R.; Heston, W. D. Expression of the prostatespecific membrane antigen. Cancer Res. 1994, 54, 1807–1811. 110. Bagalkot, V.; Farokhzad, O. C.; Langer, R.; Jon, S. An aptamer–doxorubicin physical conjugate as a novel targeted drug-delivery platform. Angew. Chem. Int. Ed. 2006, 45, 8149–8152. 111. Zhang, L.; Radovic-Moreno, A. F.; Alexis, F.; Gu, F. X.; Basto, P. A.; Bagalkot, V.; Jon, S.; Langer, R. S.; Farokhzad, O. C. Co-delivery of hydrophobic and hydrophilic drugs from nanoparticle-aptamer bioconjugates. Chem. Med. Chem. 2007, 2, 1268–1271. 112. Hricak, H.; Dooms, G. C.; Jeffrey, R. B.; Avallone, A.; Jacobs, D.; Benton, W. K.; Narayan, P.; Tanagho, E. A. Prostatic carcinoma: staging by clinical assessment, CT, and MR imaging. Radiology 1987, 162, 331–336. 113. Kurhanewicz, J.; Vigneron, D. B.; Males, R. G.; Swanson, M. G.; Yu, K. K.; Hricak, H. The prostate: MR imaging and spectroscopy. Present and future. Radiol. Clin. North Am. 2000, 38, 115–138. 114. Tempany, C.M.; Zhou, X.; Zerhouni, E. A.; Rifkin, M. D.; Quint, L. E.; Piccoli, C. W.; Ellis, J. H.; McNeil, B. J. Staging of prostate cancer: results of Radiology Diagnostic Oncology Group project comparison of three MR imaging techniques. Radiology 1994, 192, 47–54. 115. Wolf, J. S. Jr., Cher, M.; Dall’era, M.; Presti, J. C., Jr., Hricak, H.; Carroll, P. R. The use and accuracy of cross-sectional imaging and fine needle aspiration cytology for detection of pelvic lymph node metastases before radical prostatectomy. J. Urol. 1995, 153, 993–999. 116. Pan, B.; Cui, D.; Sheng, Y.; Ozkan, C.; Gao, F.; He, R.; Li, Q.; Xu, P.; Huang, T. Dendrimermodified magnetic nanoparticles enhance efficiency of gene delivery system. Cancer Res. 2007, 67, 8156–8163. 117. Alivisatos, P. The use of nanocrystals in biological detection. Nat. Biotechnol. 2004, 22, 47–52. 118. Michalet, X.; Pinaud, F. F.; Bentolila, L. A.; Tsay, J. M.; Doose, S.; Li, J. J.; Sundaresan, G.; Wu, A. M.; Gambhir, S. S.; Weiss, S. Quantum dots for live cells, in vivo imaging, and diagnostics. Science 2005, 307, 538–544. 119. Medintz, I. L.; Uyeda, H. T.; Goldman, E. R.; Mattoussi, H. Quantum dot bioconjugates for imaging, labelling and sensing. Nat. Mater. 2005, 4, 435–446. 120. Yao, J.; Larson, D. R.; Vishwasrao, H. D.; Zipfel, W. R.; Webb, W. W. Blinking and nonradiant dark fraction of water-soluble quantum dots in aqueous solution. PNAS 2005, 102, 14284–14289. 121. Gopalakrishnan, G.; Danelon, C.; Izewska, P.; Prummer, M.; Bolinger, P. Y.; Geissbuhler, I.; Demurtas, D.; Dubochet, J.; Vogel, H. Multifunctional lipid/quantum dot hybrid nanocontainers for controlled targeting of live cells. Angew. Chem. Int. Ed. 2006, 45, 5478–5483. 122. Hezinger, A. F.; Tessmar, J.; Gopferich, A. Polymer coating of quantum dots—a powerful tool towards diagnostics and sensorics. Eur. J. Pharm. Biopharm. 2008, 68, 138–152. 123. Smith, B. R.; Cheng, Z.; De, A.; Koh, A. L.; Sinclair, R.; Gambhir, S. S. Real-time intravital imaging of RGD-quantum dot binding to luminal endothelium in mouse tumor neovasculature. Nano Lett. 2008, 8, 2599–2606. 124. Jackson, H.; Muhammad, O.; Daneshvar, H.; Nelms, J.; Popescu, A.; Vogelbaum, M. A.; Bruchez, M.; Toms, S. A. Quantum dots are phagocytized by macrophages and colocalize with experimental gliomas. Neurosurgery 2007, 60, 524–530. 125. Chu, T. C.; Marks, J. W.; Lavery, L. A.; Faulkner, S.; Rosenblum, M. G.; Ellington, A. D.; Levy, M. Aptamer:toxin conjugates that specifically target prostate tumor cells. Cancer Res. 2006, 66, 5989–5992. 126. Smith, A. M.; Duan, H.; Mohs, A. M.; Nie, S. Bioconjugated quantum dots for in vivo molecular and cellular imaging. Adv. Drug Deliv. Rev. 2008, 60, 1226–1240. 127. Tan, W. B.; Jiang, S.; Zhang, Y. Quantum-dot based nanoparticles for targeted silencing of HER2/neu gene via RNA interference. Biomaterials 2007, 28, 1565–1571.
P1: OTA/XYZ P2: ABC c23 JWBS037-Chen January 25, 2011
592
13:28
Printer Name: Yet to Come
NANOPARTICLES FOR COMBINED CANCER IMAGING AND THERAPY
128. Yezhelyev, M. V.; Qi, L.; O’Regan, R. M.; Nie, S.; Gao, X. Proton-sponge coated quantum dots for siRNA delivery and intracellular imaging. J. Am. Chem. Soc. 2008, 130, 9006–9012. 129. Derfus, A. M.; Chen, A. A.; Min, D. H.; Ruoslahti, E.; Bhatia, S. N. Targeted Quantum Dot Conjugates for siRNA Delivery. Bioconjug. Chem. 2007, 18, 1391–1396. 130. Kim, K. Y. Nanotechnology platforms and physiological challenges for cancer therapeutics. Nanomedicine 2007, 3, 103–110. 131. Jain, P. K.; El-Sayed, I. H.; El-Sayed, M. A. Au nanoparticles target cancer Nano Today 2007, 2, 18–29. 132. Ghosh, P.; Han, G.; De, M.; Kim, C. K.; Rotello, V. M. Gold nanoparticles in delivery applications. Adv. Drug Deliv. Rev. 2008, 60, 1307–1315. 133. Son, S. J.; Bai, X.; Lee, S. B. Inorganic hollow nanoparticles and nanotubes in nanomedicine Part 2: imaging, diagnostic, and therapeutic applications. Drug Discov. Today 2007, 12, 657–663. 134. Gobin, A. M.; Lee, M. H.; Halas, N. J.; James, W. D.; Drezek, R. A.; West, J. L. Near-infrared resonant nanoshells for combined optical imaging and photothermal cancer therapy. Nano Lett. 2007, 7, 1929–1934. 135. Park, J. H.; Von Maltzahn, G.; Ruoslahti, E.; Bhatia, S. N.; Sailor, M. J. Micellar hybrid nanoparticles for simultaneous magnetofluorescent imaging and drug delivery. Angew. Chem. Int. Ed. Engl. 2008, 47, 7284–7288. 136. Nasongkla, N.; Bey, E.; Ren, J.; Ai, H.; Khemtong, C.; Guthi, J. S.; Chin, S. F.; Sherry, A. D.; Boothman, D. A.; Gao, J. Multifunctional polymeric micelles as cancer-targeted, MRIultrasensitive drug delivery systems. Nano Lett. 2006, 6, 2427–2430. 137. Yang, J.; Lee, C. H.; Ko, H. J.; Suh, J. S.; Yoon, H. G.; Lee, K.; Huh, Y. M.; Haam, S. Multifunctional magneto-polymeric nanohybrids for targeted detection and synergistic therapeutic effects on breast cancer. Angew. Chem. Int. Ed. Engl. 2007, 46, 8836–8839. 138. Allen, T. M. Ligand-targeted therapeutics in anticancer therapy. Nat. Rev. Cancer 2002, 2, 750–763. 139. Weng, K. C.; Noble, C. O.; Papahadjopoulos-Sternberg, B.; Chen, F. F.; Drummond, D. C.; Kirpotin, D. b.; Wang, D.; Hom, Y. K.; Hann, B.; Park, J. W. Targeted tumor cell internalization and imaging of multifunctional quantum dot-conjugated immunoliposomes in vitro and in vivo. Nano Lett. 2008, 8, 2851–2857.
P1: OTA/XYZ P2: ABC c24 JWBS037-Chen January 25, 2011
13:31
Printer Name: Yet to Come
CHAPTER 24
Multimodal Imaging and Therapy with Magnetofluorescent Nanoparticles JASON R. McCARTHY and RALPH WEISSLEDER Center for Molecular Imaging Research, Harvard Medical School and Massachusetts General Hospital, Charlestown, Massachusetts, USA
Magnetic nanoparticles and their magnetofluorescent analogs have enjoyed widespread use in the molecular imaging of numerous diseases, including cancer and atherosclerosis. These particles are highly useful for in vivo fluorescent and magnetic resonance imaging. Further functionalization of the nanoscaffold with radionuclides allows for these agents to be visualized by nuclear imaging, such as PET or SPECT, as well. A number of strategies have been developed to target the nanoagents to specific cells via the inclusion of affinity ligands, such as antibodies, aptamers, peptides, or small molecules. This chapter summarizes some of the recent advances of this class of nanomaterials.
24.1 INTRODUCTION Advances in nanotechnology have allowed for the creation of nanoagents designed for specific applications. Nanometer sized metal oxide nanoparticles, in particular, iron oxides, have proved clinically useful in the diagnosis of disease. This usefulness is derived from their biocompatibility and low toxicity, in addition to their small size and large surface area, which allows them to remain in circulation for extended periods of time versus small molecules [1], and the ability to functionalize their surface with targeting, diagnostic, and therapeutic moieties. While iron oxides (magnetite [Fe3 O4 ] and maghemite [␥ -Fe2 O3 ]) have been utilized for magnetic separations for fifty years, it was not until 1978 that Ohgushi demonstrated that they were capable of shortening the relaxation times of water in a controlled and useful manner [2]. Aside from their use in magnetic resonance imaging (MRI), iron oxide nanoparticles have been used in a number of other applications, including hyperthermia [3, 4], as spoilers in magnetic resonance spectroscopy [5], and as sensors for metabolites and other biomolecules [6–10]. In this chapter we discuss the targeting of multifunctional multimodal magnetic nanoparticles, including the relevant chemistries and Nanoplatform-Based Molecular Imaging Edited by Xiaoyuan Chen C 2011 John Wiley & Sons, Inc. Copyright
593
P1: OTA/XYZ P2: ABC c24 JWBS037-Chen January 25, 2011
594
13:31
Printer Name: Yet to Come
MULTIMODAL IMAGING AND THERAPY WITH MAGNETOFLUORESCENT NANOPARTICLES
ligands. In addition, we explore the utility of these nanoagents in the imaging and therapy of cancer and cardiovascular disease. 24.1.1 Biocompatible Magnetic Nanoparticles Numerous magnetic nanoparticle compositions have been investigated for potential biomedical use (Fig. 24.1) [11]. These particles rely on magnetic phenomena only present on the nanoscale, such as superparamagnetism and enhanced magnetic moments. While a variety of nanoparticle preparations have been investigated, those containing iron have found the widest utility. Other metallic nanoparticles, such as cobalt and nickel, or metal-alloy nanoparticles, such as FePt and FeCo, may serve as highly efficient magnetic resonance (MR) imaging agents with high saturation magnetization, but are unstable in aqueous suspension, readily oxidizing in the absence of protective coatings [12, 13]. Superparamagentic iron oxide nanoparticles are the most widely studied magnetic nanoparticles and are commonly synthesized in either the aqueous or organic phase. While high-temperature decomposition of organometallic precursors in organic solvents results in nanoparticles with well controlled size and morphology, these particles must be functionalized with coating materials that ensure biocompatibility and aqueous stability prior to use in biological applications [14–18]. Aqueous syntheses, on the other hand, involve the neutralization of acidic iron salts in the presence of the appropriate coating material and are thus more facile [19]. Recent research has also attempted to enhance the magnetic
FIGURE 24.1 Magnetic nanoparticles fall into two main classes: metallic and metal oxides. Metallic nanoparticles are those composed of elemental metals, such as Co, Fe, Mn, and Ni, or metal alloys, such as FePt and FeCo. Applicable metal oxide nanoparticles are mainly composed of iron oxides or iron oxides doped with diamagnetic metals, such as Co2+ , Fe2+ , Mn2+ , and Ni2+ .
P1: OTA/XYZ P2: ABC c24 JWBS037-Chen January 25, 2011
13:31
Printer Name: Yet to Come
SYNTHESIS OF MULTIFUNCTIONAL NANOPARTICLES
595
properties of iron oxides by doping the particle core with dicationic metal ions, such as Co2+ , Fe2+ , Mn2+ , or Ni2+ [15]. Lee et al. [15] have demonstrated that MnFe2 O4 nanoparticles exhibit higher magnetic susceptibility than similarly sized magnetite nanoparticles and display no apparent toxicity in vitro. Since surface coatings are an integral part of nanoparticle synthesis and functionalization, a number of materials have been utilized for the stabilization of iron oxide nanoparticles. These have included bisphosphonates [20], dendrimers [21], lipids [22, 23], liposomes [24, 25], polyethylene glycol [26], polyacrylamide [27], polysaccharides [28, 29], and proteins [30, 31]. Current clinically utilized preparations are commonly based on dextran or other polysaccharide coating materials due to their affinity for iron oxide, as well as their previous use as plasma expanders [32–41]. Dextran-coated iron oxide nanoparticles have proved highly useful in clinical magnetic resonance imaging. In particular, these particles have been utilized to increase the accuracy of nodal cancer staging [42–44], to better delineate primary tumors [45], detect metastases [46, 47], and image angiogenesis [48]. In an effort to extend the utility of these preparations, targeting of the nanoparticles to sites of interest has been attempted with ligands, such as antibodies [49]. Unfortunately, these preparations have one main drawback, in that the dextran is in equilibrium with the surface of the nanoparticle, so it is possible for the coating material, including the conjugated targeting ligands, to dissociate from the surface. In order to circumvent this dissociation, Josephson and co-workers have crosslinked the dextran using epichlorohydrin [19]. This results in crosslinked iron oxide (CLIO) particles that are superbly stable, even under harsh conditions. In addition, they have aminated the particle surface to allow for facile functionalization via amide bond formation. While these particle have been highly useful as a model compound for many experimental applications, the synthesis of novel preparations featuring high-affinity biodegradable coatings are currently in development.
24.2 SYNTHESIS OF MULTIFUNCTIONAL NANOPARTICLES One of the main advantages of utilizing nanoparticles in applications such as molecular imaging and drug delivery is that they can be functionalized with multiple targeting, imaging, and therapeutic moieties, or combinations thereof. This is due, in part, to the large surface area to volume ratio of nanoparticles, as well as the development of chemistries that allow for facile conjugation of these ligands. These multifunctional nanoparticles also benefit from the multivalency induced by the inclusion of multiple copies of targeting ligands on the particle surface. As numerous copies of receptors are typically displayed on the surface of targets of interest, multivalent nanoagents can readily bind to more than one receptor, thereby increasing the overall binding affinity of the nanoparticle preparation. Below, we describe the functionalization of the nanoparticle surface, including relevant chemistries, and the inclusion of targeting and imaging ligands. 24.2.1 Functionalization of Nanoparticle Coatings The methodologies utilized for the decoration of the nanoparticle surface are directly related to the functional groups present on the coating material. Most often, these are either amines or carboxylic acids, although these groups are readily exchanged (see Scheme 24.1a). Amine-coated particles can be reacted with cyclic anhydrides, such as glutaric or succinic
P1: OTA/XYZ P2: ABC c24 JWBS037-Chen January 25, 2011
596
13:31
Printer Name: Yet to Come
MULTIMODAL IMAGING AND THERAPY WITH MAGNETOFLUORESCENT NANOPARTICLES
H N
NH2
H N
OH
H N
NH2
ii
(a)
O
O
O
O
O O
Succinimidyl ester
Iodoacetate
N
I O
O O O
Succinimidyl ester
N O
HS
Pyridyldithiol
S
N
S
O O
NH2
O
O
Succinimidyl ester
N O
Maleimide
N
O O O
Succinimidyl ester
_
N O
N O
NH2
O
O
O O
Succinimidyl ester
O
O3S
O
(b)
O
Amine
O
O
N O
O
H2 N
Succinimidyl ester
N O
O O H 2N
Amine
HS
Maleimide
N
O O O
SO3
N
O
_
O
Amine
O
_
H 2N
NH2
O
O3S
N O O
O SH
Maleimide
N O
O N
Maleimide
HS
O
SCHEME 24.1 Conventional bioconjugation strategies. (a) Conversion between amine and carboxylic acid functionalized nanoparticles: i, cyclic anhydrides, such as glutaric or succinic anhydride; ii, (a) EDC/sulfo-NHS, then (b) diamine, such as ethylenediamine. (b) Generic heterobifunctional linkers for conjugation of ligands to the nanoparticle surface. Blue = amine, red = sulfo-NHS activated carboxylic acid, and orange = thiol functionalized ligand or nanoparticle. Conventional heterobifunctional or homobifunctional crosslinking reagents can be found in Refs. 75 and 76. Discrete PEG heterobifunctional or homobifunctional crosslinking reagents can be found in Ref. 77. PEG heterobifunctional or homobifunctional crosslinking reagents can be found in Refs. 78–80.
P1: OTA/XYZ P2: ABC c24 JWBS037-Chen January 25, 2011
13:31
Printer Name: Yet to Come
SYNTHESIS OF MULTIFUNCTIONAL NANOPARTICLES
597
anhydride, to yield the corresponding carboxylic acid, while acid moieties can be reacted with diamines following activation with 1-ethyl-3-[3-dimethylaminopropyl]carbodiimide (EDC) and N-hydroxysuccinimide (NHS) or N-hydroxysulfosuccinimide (sulfo-NHS).
Conventional Bioconjugation Strategies Amine-modified surface coatings allow for the most facile and diverse functionalization reactions (Scheme 24.1b). Isothiocyanate or succinimidyl ester containing ligands are readily reacted with amines, whereas carboxylic acids can be reacted with the amines in the presence of activating reagents, such as EDC and sulfo-NHS. In order to functionalize the surface with thiol-containing ligands, the amines must be modified with thiol-reactive moieties, such as N-succinimidyl-3-(2pyridyldithio)propionate (SPDP), succinimidyliodoacetate (SIA), or succinimidyl-4-(Nmaleimidomethyl)cyclohexane-1-carboxylate (SMCC). The use of SPDP is advantageous in that it facilitates the determination of the number of free amines on the particle’s surface. Reaction of the particle with a large excess of SPDP is expected to label nearly all of the free amines. If this particle is then treated with dithiothreitol (DTT), the SPDP reacts with the free thiol, liberating pyridine-2-thione, which can be monitored at 343 nm and quantified. During reaction with thiol-containing ligands, pyridine-2-thione is also released; thus the degree of conjugation can be calculated. While SPDP is highly useful, this reductive lability may also be detrimental in vitro or in vivo, as reductive environments can cleave ligands from the nanoparticle surface. Whereas amine-coated particles are readily reacted with a number of ligands, carboxylic acid coatings require activation prior to reaction with amines or alcohols. As described above, these moieties are often activated with EDC and sulfo-NHS at pH 6.0, which prolongs the lifetime activated ester in aqueous media, prior to amide bond formation with amines. Esterification of the acids is a much more involved process, requiring the precipitation of the particles from the aqueous suspension using acetone followed by acetone washes to remove most traces of water. The particles are then suspended in anhydrous dimethylsulfoxide (DMSO) and reacted with thionyl chloride to form the corresponding acid chloride. This particle can then be reacted with the hydroxyl containing ligand in the presence of a base catalyst to yield the desired ester. Orthogonal Bioconjugation Strategies While the conventional bioconjugation methodologies listed above are, in most cases, adequate for the functionalization of nanoparticles, the introduction of multiple functionalities may require the use of orthogonal conjugation strategies. One reaction that has been utilized is the Huisgen 1,3-dipolar cycloaddition, pioneered by Sharpless and co-workers [50]. This “click reaction” involves the copper(I) catalyzed reaction of an alkyne with an azide, resulting in the formation of a stable triazole. This reaction requires mild conditions and can be undertaken in aqueous solutions. Importantly, it is highly specific and high yielding and can take place in the presence of other functional groups, such as amines and carboxylic acids. It also does not produce any undesirable side products, such as dicyclohexylurea, which is a result of carboxylic acid activation with dicyclohexylcarbodiimide. The click reaction has been demonstrated with carboxylic acid functionalized CLIO [51, 52]. Activation of the acid moieties with EDC and sulfo-NHS in 2-(N-morpholino) ethanesulfonicacid (MES) buffer at pH 6.0, followed by reaction with azido propylamine or propargylamine, yields azide or alkyne functionalized particles, respectively. Concomitantly, a number of relevant ligands were modified to include the corresponding reactive groups. The succinimidyl ester of VivoTag 680, a near-infrared fluorophore, and fluorescein
P1: OTA/XYZ P2: ABC c24 JWBS037-Chen January 25, 2011
598
13:31
Printer Name: Yet to Come
MULTIMODAL IMAGING AND THERAPY WITH MAGNETOFLUORESCENT NANOPARTICLES
isothiocyanate were readily reacted with the azide or alkynyl amine, whereas functionalization of estradiol, paclitaxel, and disperse red 1 required activation of the ligand hydroxyl group with 1,1 -carbonyldiimidazole (CDI), followed by reaction with the amine. The conjugation reactions were facilitated by the addition of copper(I) iodide, with reactions lasting between 5 and 8 hours at 37 ◦ C, resulting in 90% conversion. 24.2.2 Targeting of Magnetic Nanoparticles Nanoparticles can be targeted to sites of interest using a number of different strategies. Commonly, these have involved the modification of the particle surface with antibodies, aptamers, peptides, or small molecules. Due to the finite area of the nanoparticle surface, the availability of functional groups, and the size of the targeting agents, differing numbers of each ligand can be appended. For example, on CLIO the maximum number of antibodies that can be conjugated is 1–2, whereas peptides or small molecules can be added in 5- to 100-fold greater numbers. This increase in valency subsequently increases the apparent affinity of the nanoagent for its target, versus the free ligand. Thus this section focuses on the utility of peptides and small molecules in the targeting of magnetic nanoparticles.
Peptides Peptide targeting ligands can be derived from a number of different sources. These include natural products, phage display, and in silico data mining, although all three sources are intimately related. Those that occur in nature have the advantage of evolutionary mechanisms, which optimize the affinity of a peptide for a specific substrate. For example, Jaffer and co-workers have developed a fluorescent imaging agent for thrombosis based on a peptide sequence gleaned from the N terminus of ␣2 -antiplasmin, which is crosslinked into acute clots by activated factor XIII [53, 54]. As compared to a control peptide, this fluorescently labeled peptide, GNQEQVSPLTLLK, readily bound to and was able to image thrombi in vivo in a murine model of thrombosis. Analogous to natural selection, phage display utilizes heterogeneous libraries of bacteriophage, viruses that infect bacterial cells, to identify peptide sequences that bind to targets of interest. The phages are engineered via standard recombinant DNA techniques to express short peptide sequences on the virus coat protein. Libraries of these phages can encompass over 1010 distinct sequences. These phages can then be screened against proteins, cells, or tissues of interest. In its simplest form, phage screening involves the immobilization of proteins on the surface of a plate, which is then incubated with the phages. The unbound phages are removed by washing, and the bound phages are expanded and sequenced to identify the peptide of interest. While this method is fairly facile, with the binding partner of the bound phages immediately known, there are a number of drawbacks, such as the possibility that the protein is not in its native conformation. Cell-based phage screens have thus been developed to circumvent this problem. An added advantage is that the screens can be conducted with no a priori knowledge and can be biased toward phages that have been internalized. This leads to peptide sequences that are highly useful for molecular imaging due to the signal amplification that occurs via intracellular accumulation of the agent. Cell-based assays can also be accomplished under flow conditions that mimic the sheer stress encountered in vivo. Whereas weakly bound phages will be washed away, phages that are more avidly bound lead to peptide sequences that are more efficacious. One last modification that has been made to phage display is the ability to screen in vivo in relevant model animals. In vivo, the phages are presented with a number of physiological
P1: OTA/XYZ P2: ABC c24 JWBS037-Chen January 25, 2011
13:31
Printer Name: Yet to Come
MULTIMODAL NANOPARTICLES
599
barriers; therefore those that reach the target tissues will result in imaging agents with higher efficacies. While all of the above methods rely on selection processes to identify targeting peptide sequences, one alternative is to mine electronic resources. Recently, we have created a searchable peptide database, PepBank (http://pepbank.mgh.harvard.edu/), that contains sequences culled from publicly available sources and databases, such as Medline, Pubmed, UniProt, ASPD, and full text PDF files. This database allows users to search for specific peptide sequences, or to search for a specific target to see what peptides are associated with it. The main detractor to this method is that it is limited to those peptides that have already been identified and added to the database.
Small Molecules While antibodies, aptamers, and peptides are most often used in the targeting of nanomaterials, one portion of the chemical space that has been overlooked is small molecules. This is due, in part, to the lack of a general method to conjugate and screen small molecule libraries. To this end, Weissleder et al. [55] have synthesized a library, based on CLIO, containing 146 small molecule modifications. These fluoresceinmodified particles were screened in a high-throughput manner using a modified robotic system. In the initial screen, the nanoagents were incubated with five different cell types, including endothelial cells (HUVEC, human umbilical vein endothelial cells), primary resting human macrophages, granulocyte macrophage colony-stimulated (GM-CSF) human macrophages, a human macrophage-like cell line (U937), and human pancreatic ductal carcinoma cells (PaCa-2). Cellular binding and uptake were determined by fluorescence microscopy, flow cytometry, or a fluorescein immunoassay, and the results were plotted in heat map form to obviate the effect of the small molecule modifications (Fig. 24.2). One topic of interest was whether any of the preparations could be used to identify activated versus resting macrophages, which is potentially useful for in vivo imaging and drug delivery. Based on the initial screen, the authors identified bentri-modified CLIO as having high affinity for the resting macrophages, with little uptake in activated macrophages, while glycine-modified CLIO exhibited an inverse relationship. The particle preparations were subsequently scaled up and tested in primary human macrophages activated with a number of proinflammatory agents, such as GM-CSF, which mimics macrophages in immune disease, oxidized LDL, which mimics foam cells in atherosclerosis, and lipopolysaccharide, which mimics macrophages in infection. While CLIO-bentri was readily internalized by the nonactivated macrophages (about fourfold greater than aminated CLIO starting material), CLIO-gly demonstrated significant uptake in all of the activated cell lines. The authors further propose that the results of this study could lead to novel classes of nanomaterials based on small molecule modifications that are useful in the diagnosis and treatment of disease.
24.3 MULTIMODAL NANOPARTICLES While iron oxide nanoparticle platforms are useful in magnetic resonance (MR) imaging, due to their intrinsic superparamagnetic character, these particles can be further functionalized with diagnostic or therapeutic moieties to increase their overall utility. Multimodal nanoparticles, or those bearing more than one type of complementary ligand, including those used in optical, nuclear, and MR imaging, allow for the maximum amount of data to be extracted from a single nanoparticle preparation. The following are recent examples
P1: OTA/XYZ P2: ABC c24 JWBS037-Chen January 25, 2011
600
13:31
Printer Name: Yet to Come
MULTIMODAL IMAGING AND THERAPY WITH MAGNETOFLUORESCENT NANOPARTICLES
FIGURE 24.2 Heat map representing cellular uptake of different nanoparticle preparations. Columns from right to left: 1, pancreatic cancer cells (PaCa-2); 2, macrophage cell line (U937); 3, resting primary human macrophages; 4, activated primary human macrophages; 5, human umbilical vein endothelial cells (HUVECs). Red refers to the lowest accumulation and green refers to the highest accumulation. (Reproduced with permission from Weissleder et al. [55].)
P1: OTA/XYZ P2: ABC c24 JWBS037-Chen January 25, 2011
13:31
Printer Name: Yet to Come
MULTIMODAL NANOPARTICLES
601
TABLE 24.1 Multimodal Magnetofluorescent Imaging Agents Target
Targeting Ligand
References
VCAM-1
Anti-VCAM-1 antibody VHS peptide VINP peptide Anti-E-selectrin antibody E-selectin binding peptide Small-cell lung cancer specific aptamer Anti-␣v 3 antibody RGD peptides Muc-1 peptide Annexin V Tat peptide Intrinsic avidity Small molecules Hepsin peptide CREKA peptide F3 peptide Herceptin Monoclonal antibody A7 Chimeric L6 antibody Plectin-1 targeted peptides Bombesin-like peptide Peptide–major histocompatibility complex
71 72 73, 74 81, 82 83 84 85, 86, 87 85 88 89, 90 19, 91, 92 64, 65, 70 55 93 57 94, 95 96, 97 98 99 56 100 101
E-selectin Small-cell lung cancer ␣v 3 Integrin Muc-1 Phosphatidylserine Cell label Macrophages Macrophage subtypes Hepsin Tumor vasculature Her2/neu Colorectal carcinoma Tumor associated antigen Plectin-1 Bombesin Cytotoxic T lymphocytes
of the synthesis and utility of multimodal imaging and therapy using targeted magnetic nanoparticles. (See Table 24.1.) 24.3.1 Multimodal Imaging of Cancers
Early Detection of Pancreatic Ductal Adenocarcinoma Pancreatic ductal adenocarcinoma (PDAC) is the fourth leading cause of cancer deaths in the United States. This is due to late diagnosis of the disease as well as the resistance of the cancer to existing therapies. Current methods of early detection are not sensitive or are unreliable; therefore the development of novel diagnostic strategies is needed. To this end, Kelly et al. [56] have utilized a phage display approach to identify peptide sequences that distinguish PDAC cells from normal pancreatic ductal cells. Using a genetically engineered mouse model that spontaneously develops PDAC due to the activation of Kras and deletion of p53 or Ink4a/Arf tumor suppressors, the researchers were able to isolate primary cell lines from emerging PDAC. Screening for phages internalized by PDAC, followed by counterselection against normal pancreatic cells, yielded a phage clone (KTLLPTP) that was later determined to bind to plectin-1, an intermediate filament and important crosslinking element of the cytoskeleton. The specificity of the resulting phage clone was further tested in vitro on frozen sections of human pancreas at different stages of PDAC progression using fluorescein-labeled phages. In all cases, the phages identified areas of diseased tissue, whereas a control (unrelated phage) displayed no binding. The peptide was then utilized to develop a plectin-1 targeted PDAC imaging agent. In order to conjugate the plectin-1 targeted peptide (PTP) to Cy5.5-labeled CLIO, the authors
P1: OTA/XYZ P2: ABC c24 JWBS037-Chen January 25, 2011
602 (a)
13:31
Printer Name: Yet to Come
MULTIMODAL IMAGING AND THERAPY WITH MAGNETOFLUORESCENT NANOPARTICLES
PTP-NP-Cy5.5
PTP-NP (red) Vasculature (blue)
(b)
(c) % Injected dose
4.0 PTP-NP Control-NP 3.0 2.0 1.0
us cl e Sp le en Ki dn In ey te st in e Li ve Pa No r nc rm re al as
M
or
ea rt Sk in
H
Lu
Tu m
Control-NP (red) Vasculature (blue)
ng
0
Control-NP-Cy5.5
FIGURE 24.3 (a) Intravital confocal microscopy of early pancreatic lesions imaged using targeted nanoagent (red, top) or control nanoagent (red, bottom) and AF750-labeled blood pool agent (blue). (b) Low-magnification view of pancreatic fluorescence shows distribution of targeted nanoagent in distinct areas of the pancreas. White light overlay provides anatomic correlation (left). Dotted line outlines the pancreas. (c) Biodistribution of targeted nanoagent and control nanoagent. (Reproduced with permission from Kelly et al [56].)
modified the N terminus of the peptide with the peptide sequence GGSK(FITC)GC. Reaction of amine-coated CLIO with SIA, followed by reaction with the peptide, yielded the nanoagent. Inclusion of fluorescein (FITC) allows for the number of peptides per particle to be calculated using the extinction coefficient of fluorescein. The utility of the nanoagent was tested in 9-week-old Kras/p53L/+ mice, which normally have small PDAC, yet no outward signs of illness. Twenty-four hours after intravenous (IV) injection of the agent the mice were anesthetized, and the abdominal cavity was surgically exposed. The mice were then imaged by intravital fluorescence microscopy (IVFM) and areas of focal fluorescence signal were observed within the pancreas (Fig. 24.3), which was further correlated with ex vivo fluorescence reflectance imaging (FRI). Ex vivo MRI of diseased pancreas also demonstrated focal regions of decreased signal, indicative of magnetofluorescent nanoparticle accumulation.
Targeted Imaging of the Tumor Vasculature Phage display has also been used by Ruoslahti and co-workers to identify a tumor vasculature targeting peptide [57]. In vivo phage screening in tumor-bearing transgenic mice yielded a peptide sequence, CREKA, that readily bound to and identified transgenic or xenograph tumors when fluorescently labeled and injected intravenously. Unexpectedly, the peptide formed a meshwork within the tumor stroma and highlighted the tumor vasculature. As tumors contain clotted plasma proteins that organize into a similar meshwork, the authors hypothesized that the peptide bound these clots. In order to test this they injected the peptide into fibrinogen-deficient mice, which lack the fibrin meshwork in tumors, and demonstrated that the agent did not bind. The peptide was next conjugated to amino dextran-coated superparamagnetic iron oxide (SPIO) particles via the N-terminal cysteine. Activation of the particle surface with N-[␣-maleimidoacetoxy]succinimide ester (AMAS), followed by addition of the
P1: OTA/XYZ P2: ABC c24 JWBS037-Chen January 25, 2011
13:31
Printer Name: Yet to Come
MULTIMODAL NANOPARTICLES
603
fluorescein-modified peptide, resulted in the inclusion of 8000 peptides per particle. Interestingly, when the nanoparticle preparation was injected into MDA-MB-435 tumor-bearing mice, the particles did not accumulate in the tumor. Instead, they demonstrated preferential localization to reticuloendothelial system (RES) tissues. The authors therefore utilized “decoy” Ni(II) chelate-containing liposomes in order to deplete plasma opsonins, which resulted in significant accumulation in the tumor vasculature. In fact, over 20% of the vessels within the tumor were filled with fluorescent masses, which could be stained for fibrin content, indicating thrombus formation. The authors thereby deduced that the CREKAlabeled particles were inducing clot formation, which subsequently bound more nanoagent and amplified the tumoral targeting. Tumor targeting was lastly demonstrated in xenograph bearing mice by fluorescence reflectance imaging. Cy7-labeled CREKA-nanoparticles were injected intravenously following injection of the decoy liposomes. In one cohort the mice also received heparin, which prevents clot formation. The tumors were clearly delineated in the nonheparin group, while those receiving the anticoagulant demonstrated a significant decrease in fluorescence signal.
Imaging of Tumoral ␣ v 3 Integrin While positron emission tomography is a highly useful imaging modality for the determination of in vivo biodistribution of an agent due to its high sensitivity, it suffers from the inability to accurately determine agent localization in the absence of identifiable anatomical structures. To overcome this, methods have been developed in which PET is utilized in tandem with other imaging modalities, such as X-ray computed tomography and MRI, which allows for the coregistration of images, and the more accurate determination of agent localization. One of the main disadvantages of these multimodal techniques is the need to administer more than one type of imaging agent (e.g., 18 F-fluorodeoxyglucose as a radiotracer and an iodinated agent for contrast in CT). Chen and co-workers have circumvented this problem by developing bifunctional nanoagents that are simultaneously useful for both MR and PET imaging [58]. Polyaspartic acid (PASP)-coated nanoparticles were synthesized in the aqueous phase by neutralization of iron salts by a solution of PASP in ammonia. After workup, the 45-nm hydrodynamic diameter amine-coated particles were functionalized with 1,4,7,10-tetraazacyclododecaneN, N , N ,N -tetraacetic acid (DOTA) after its activation with EDC and sulfo-NHS, and a maleimide-terminated polyethylene glycol. Lastly, the integrin targeting peptide, RGD, was conjugated to the particle surface via reaction with the maleimide, and the nanoagent was radiolabeled with 64 Cu. When the targeted agent was injected into U87MG tumor-bearing mice, the tumor was clearly visible from 1 to 21 h by PET (Fig. 24.4), with maximum accumulation after 4 h (10.1 ± 2.1 %ID/g). This uptake could be abrogated by coinjection of a blocking dose of the RGD peptide (10 mg/kg). As well, a control particle lacking the RGD peptide showed significantly lower tumoral accumulation at all time points. A similar experiment was performed to investigate the MRI capabilities of the targeted nanoagent. As compared to the control particle or the coinjection of a blocking dose of the RGD peptide, the targeted peptide showed significant accumulation within the tumor, as indicated by the MR signal decrease, further correlating the PET findings. 24.3.2 Multimodal Imaging in Cardiovascular Disease (CVD)
Macrophage Infiltration in CVD Macrophages partake in all facets of CVD, including the initiation and progression of atherosclerosis, as well as the infiltration of the myocardium after infarction. In atherosclerosis, macrophages can constitute up to 20% of the cells within
P1: OTA/XYZ P2: ABC c24 JWBS037-Chen January 25, 2011
604
13:31
Printer Name: Yet to Come
MULTIMODAL IMAGING AND THERAPY WITH MAGNETOFLUORESCENT NANOPARTICLES
(a)
(b)
FIGURE 24.4 (a) Decay-corrected whole-body coronal PET images of nude mouse bearing human U87MG tumor at 1, 4, and 21 h after injection of 3.7 MBq of 64 Cu- DOTA-IO, 64 Cu-DOTA-IO-RGD, or 64 Cu-DOTA-IO-RGD with 10 mg of c(RGDyK) peptide per kilogram (300 mg of iron-equivalent IO particles per mouse). (b) Time–activity curves of U87MG tumors after injection of 3.7 MBq of 64 Cu-DOTA-IO, 64 Cu-DOTA-IO-RGD, or 64 Cu-DOTA-IO-RGD with blocking dose of c(RGDyK). (Reproduced with permission from Lee et al. [58].)
lesions and are responsible for the production of various proteases, which degrade the extracellular matrix and promote plaque rupture [59–61]. Due to the high morbidity and mortality associated with CVD, novel techniques to diagnose and treat patients prior to the presentation of clinical symptoms is needed. Dextran-coated iron oxide nanoparticles possess an innate avidity for inflammatory macrophages. In fact, this property has been clinically utilized for the detection of suspect lesions in symptomatic patients with severe internal carotid artery stenosis [62, 63]. Recently, these nanoparticles have been further functionalized for use in multimodal imaging. Jaffer and co-workers have demonstrated the localization of Cy5.5-labeled CLIO nanoparticles in apolipoprotein E deficient mice (apoE−/− ) ex vivo by FRI, and in vivo by MRI and IVFM [64, 65]. Nahrendorf et al. [66] have built on this initial work by further decorating the particle surface with 64 Cu chelates for PET imaging, via reaction of CLIO with diethylenetriaminepentaacetic (DTPA) dianhydride, followed by radiolabeling. Once injected into atherosclerotic lesion-laden mice, the authors were readily able to identify probe uptake in the aortic arch and root, as determined by PET/CT and MRI (Fig. 24.5). The aortas of the mice were then excised and digested in order to determine the cellular distribution of the agent within the lesions by flow cytometry. Macrophages and monocytes contributed to over 74% of the observed activity within the lesions, while 17% was contained within neutrophils, and the remainder within endothelial cells, smooth muscle cells, and lymphocytes. The avidity of CLIO for macrophages was further utilized by us for the development of a novel near-infrared light activated therapeutic (NILAT) nanoagent [67–69]. As macrophages contribute to all stages of atherogenesis, focal macrophage ablation can potentially result in the stabilization of inflamed lesions. Conjugation of a potent mesotetrahydroxyphenylchlorin analog and Alexa Fluor 750 (AF750) to the CLIO backbone
P1: OTA/XYZ P2: ABC c24 JWBS037-Chen January 25, 2011
13:31
Printer Name: Yet to Come
MULTIMODAL NANOPARTICLES
605
FIGURE 24.5 64 Cu-TNP distributes to atherosclerotic lesions. (a, b) PET-CT shows enhancement of the posterior aortic root (arrow). (c–f), En face Oil Red O staining of the excised aorta depicts plaque-loaded vessel segments, which colocalize with areas of high 64 Cu-TNP uptake on autoradiography. (e, f) Zoomed image of the root and arch. Arrows depict a plaque-laden segment of the root with high activity, which corresponds to the in vivo signal seen in (b). (g–i) Preinjection and postinjection MRIs of the aortic root (inset). The dotted line in the long-axis views demonstrates slice orientation for short-axis root imaging. (i) Signal intensity (pseudocolored with identical scaling for preinjection and postinjection image) decreased significantly after injection of 64 Cu-TNP, which was quantified by calculation of the contrast-to-noise ratio (CNR). (k) Near-infrared fluorescence reflectance imaging (NIRF) of excised aortas shows accumulation of the probe in plaques residing in the root (arrow), thoracic aorta, and carotid bifurcation (arrowheads), further corroborating the PET signal observed in these vascular territories. (Reproduced with permission from Nahrendorf et al. [66].)
P1: OTA/XYZ P2: ABC c24 JWBS037-Chen January 25, 2011
606
13:31
Printer Name: Yet to Come
MULTIMODAL IMAGING AND THERAPY WITH MAGNETOFLUORESCENT NANOPARTICLES
yielded a therapeutic agent that could be optically imaged in the near-infrared region of the electromagnetic spectrum. The absorption maxima of the diagnostic and therapeutic portions of the agent are spectrally distinct, allowing for the determination of the localization of the nanoparticle without extraneous cell death. When incubated with murine macrophages, the nanoagent exhibited time-dependent uptake, as determined by flow cytometry, with an LC50 of 14 nM when irradiated at 650 nm. The particles were next tested in apoE−/− mice. Twenty-four hours after IV injection of the agent, the carotid artery was surgically exposed and particle localization was determined by IVFM. Following the imaging session, the lesions were irradiated with a 650-nm laser, the incision was closed, and the mouse was allowed to recover. One week after therapy, the mice were injected with AF750-labeled CLIO to allow for the determination of changes in lesional macrophage burden by IVFM. As compared to the pretherapy image, the fluorescence signal from the treated lesions decreases significantly, while the control group increases (target to background ratio of 4.1 pretreatment to 0.6 post-treatment in the treated group, versus 5.2 pretreatment to 13.0 post-treatment for the control group). Therapeutic efficacy was examined histologically 3 weeks post-therapy. The treated lesions demonstrated a thickened fibrous cap and decreased inflammation. While these results are certainly promising, further experimentation is being undertaken to assess therapeutic efficacy. Macrophage infiltration into infarcted myocardium has also been investigated by fluorescence molecular tomography (FMT) and MRI using Cy5.5-labeled CLIO. Postinfarction macrophage content is significant, in that it is implicated in the healing of the myocardium, as well as other conditions, such as myocarditis, heart failure, and transplant rejection. Sosnovik et al. [70] have induced myocardial infarctions in mice via left coronary artery ligation, followed 48 h later by injection of the agent. The particles were allowed 48 h to localize, which was then analyzed by MRI, revealing left ventricular dilation with thinning and akinesis of the ventricle walls. Negative contrast enhancement was obvious in the hypocontractile areas of the myocardium, indicative of nanoparticle localization. The mice were then imaged by FMT, where two areas of nanoagent uptake were observed in the infarcted mice, one of which was clearly accumulation in the liver (Fig. 24.6). The second was located over the heart and thorax. Nanoparticle accumulation was further correlated ex vivo using MRI, FRI, and histology. This report is particularly significant in that it is the first demonstration of the ability to noninvasively image cellular and subcellular events in the heart, including the potential to affect the care of the postinfarction patient.
Vascular Cell Adhesion Molecule-1 Vascular cell adhesion molecule-1 (VCAM-1) is a glycoprotein expressed on activated endothelial cells and smooth muscle cells that partakes in the initiation and progression of atherosclerosis by facilitating the adhesion of leukocytes and their transmigration to the atheroma. Thus VCAM-1 may serve as an ideal target for the detection and imaging of atherosclerosis. While a number of methodologies have been developed to image VCAM-1 expression using radiolabeled antibodies, these reports have suffered from low signal-to-background ratios. In order to circumvent this, Tsourkas et al. [71] developed a VCAM-1 antibody targeted magnetofluorescent nanoparticle bearing Cy5.5 for fluorescence imaging and approximately three anti-VCAM-1 antibodies per particle. This nanoagent was assayed in vitro in cells expressing high (murine heart endothelial cells, MHECs) or low levels of VCAM-1 (murine dermal endothelial cells, MDECs), and displayed significant binding to the MHEC, as determined by flow cytometry. This agent was further tested in vivo in a murine acute inflammation model, in which tumor necrosis factor-␣ (TNF-␣) is injected into one ear, while the contralateral
P1: OTA/XYZ P2: ABC c24 JWBS037-Chen January 25, 2011
13:31
Printer Name: Yet to Come
MULTIMODAL NANOPARTICLES
(a)
(b)
(c)
(d)
(e)
(f)
607
FIGURE 24.6 FMT of myocardial macrophage inflltration in vivo. Reconstructed coronal slices from the three-dimensional FMT data set have been superimposed on white light images of the mice. Slices 2 to 4 in the FMT data set intersected the heart, slice 1 passed anterior to it, and slices 5 to 8 passed posterior to it. (a) Long-axis MRI slice in an infarcted mouse corresponding to slice 2 from the fluorescence data set of that mouse (b). (c) Slice 5 from the fluorescence data set of the infarcted mouse. The corresponding slices (d = slice 2, E = slice 5) of a sham-operated mouse are shown. Fluorescence intensity over the heart was significantly greater in the infarcted mice than the sham-operated mice (f). (Reproduced with permission from Sosnovik et al. [70].)
ear is able to serve as a control. After injection of the nanoparticles, the vasculature of the treated ear demonstrated significant binding of the agent, with maximum accumulation after 6 h. While the results obtained above were promising, the system suffered from one serious flaw. The antibody-targeted nanoparticles were not able to enter the cell, thus amplification of the signal by cell internalization was not possible. In order to overcome this shortcoming, Kelly and co-workers have used phage display to identify peptide sequences that bound to and were internalized by MHECs [72–74]. In their earliest work, they utilized phage libraries bearing cyclic peptides and identified a highly efficacious sequence (CVHSPNKKC) [72]. This sequence was further modified with an C-terminal linker (GGSKGC) for inclusion of fluorophores and conjugation to CLIO. The fluorescently labeled peptide was shown to be rapidly internalized by MHECs, versus a scrambled control, which could be abrogated by preincubation of the cells with a VCAM-1 antibody. Conjugation of the peptide to CLIO (4 peptides per particle) resulted in further efficacy, with 10-fold greater uptake in MHECs than CLIO modified with the control peptide. In vivo testing of the nanoagent in the TNF-␣ stimulated ear model resulted in accumulation within the inflamed ear after 4 h, with negligible binding in the control ear, and significant fluorescence signal still present after 24 h. These findings were further corroborated histologically, with colocalization of the nanoparticles and an anti-VCAM-1 antibody. This utility of this nanoagent was further demonstrated in atherosclerotic apoE−/− mice. Injection of the targeted particle followed by MR imaging revealed decreases in signal from the aortic wall, as expected for areas of uptake. This observation was not duplicated in mice that received the control nanoagent or in wild-type mice devoid of lesions.
P1: OTA/XYZ P2: ABC c24 JWBS037-Chen January 25, 2011
608
13:31
Printer Name: Yet to Come
MULTIMODAL IMAGING AND THERAPY WITH MAGNETOFLUORESCENT NANOPARTICLES
The authors followed up this work with the identification of a linear peptide targeting ligand (VHPKQHR) for VCAM-1 [73, 74]. The authors surmised that a linear peptide sequence would allow for the inclusion of more targeting ligands per nanoparticle than the previously reported cyclic, disulfide constrained derivative. Upon conjugation of the peptide to CLIO (>10 peptides per particle) the agent showed 20-fold increased binding to MHECs than the VHS peptide modified agent, which was correlated in the TNF-␣ stimulated ear model. The authors were also able to demonstrate that the synthesized nanoagent could bind to VCAM-1 in human specimens. The nanoparticles were incubated with freshly resected human endarterectomy specimens for 24 h followed by MRI and FRI. In the T2-weighted images, areas of particle binding were readily observed as decreases in signal intensity, whereas intense fluorescence signal was observed in the FRI. The authors conclude that these nanoagents are capable not only of imaging atherosclerotic lesions but of detecting the degree of inflammatory activation.
24.4 CONCLUSION AND FUTURE PERSPECTIVES Magnetic nanoparticles, in particular, iron oxides, have found clinical applicability in a number of (pre)clinical MR imaging applications. This utility has been based on the superparamagnetic character of the nanoagents, in combination with the particles natural avidity for inflammatory cell types. With the emergence of multimodal imaging, the ability to develop novel nanoagents bearing multiple functionalities will allow for an increased amount of information to be learned about specific diseases, thereby increasing the quality of patient care. In addition, the inclusion of therapeutic moieties will enable the targeted diagnosis and treatment of a number of diseases with one nanoparticulate platform.
ACKNOWLEDGMENTS This work was supported by NIH grants U01-HL080731 (RW), U54-CA119349 (RW), and U54-CA126515 (RW). We thank all CMIR lab members for many helpful discussions.
REFERENCES 1. Maeda, H.; Wu, J.; Sawa, T.; Matsumura, Y.; Hori, K. Tumor vascular permeability and the EPR effect in macromolecular therapeutics: a review. J. Control. Release 2000, 65, 271–284. 2. Ohgushi, M.; Nagayama, K.; Wada, A. Dextran-magnetite: a new relaxation reagent and its application to T2 measurements in gel systems. J. Magn. Reson. 1978, 29, 599–601. 3. Halbreich, A.; Roger, J.; Pons, J. N.; et al. Biomedical applications of maghemite ferrofluid. Biochimie 1998, 80, 379–390. 4. Pankhurst, Q. A.; Connolly, J.; Jones, S. K.; Dobson, J. Applications of magnetic nanopartiocles in biomedicine. J. Phys. D: Appl. Phys. 2003, 36, R167–R181. 5. Le Duc, G.; Vander, Elst L; Colet, J. M.; et al. Ultrasmall particulate iron oxides as contrast agents for magnetic resonance spectroscopy: a dose–effect study. J. Magn. Reson. Imaging. 2001, 13, 619–626. 6. Grimm, J.; Perez, J. M.; Josephson, L.; Weissleder, R. Novel nanosensors for rapid analysis of telomerase activity. Cancer Res. 2004, 64, 639–643.
P1: OTA/XYZ P2: ABC c24 JWBS037-Chen January 25, 2011
13:31
Printer Name: Yet to Come
REFERENCES
609
7. Kim, G. Y.; Josephson, L.; Langer, R.; Cima, M. J. Magnetic relaxation switch detection of human chorionic gonadotrophin. Bioconjug. Chem. 2007, 18, 2024–2028. 8. Lee, H.; Sun, E.; Ham, D.; Weissleder, R. Chip-NMR biosensor for detection and molecular analysis of cells. Nat. Med. 2008, 14, 869–874. 9. Perez, J. M.; Josephson, L.; O’Loughlin, T.; Hogemann, D.; Weissleder, R. Magnetic relaxation switches capable of sensing molecular interactions. Nat. Biotechnol. 2002, 20, 816–820. 10. Taktak, S.; Sosnovik, D.; Cima, M. J.; Weissleder, R.; Josephson, L. Multiparameter magnetic relaxation switch assays. Anal. Chem. 2007, 79, 8863–8869. 11. Sun, C.; Lee, J. S.; Zhang, M. Magnetic nanoparticles in MR imaging and drug delivery. Adv. Drug Deliv. Rev. 2008, 60, 1252–1265. 12. Gao, J.; Liang, G.; Zhang, B.; Kuang, Y.; Zhang, X.; Xu, B. FePt@CoS(2) yolk-shell nanocrystals as a potent agent to kill HeLa cells. J. Am. Chem. Soc. 2007, 129, 1428–1433. 13. Hong, R.; Fischer, N. O.; Emrick, T.; Rotello, V. M. Surface PEGylation and ligand exchange chemistry of FePt nanoparticles for biological applications. Chem. Mater. 2005, 17, 4617–4621. 14. Hou, Y.; Xu, Z.; Sun, S. Controlled synthesis and chemical conversions of FeO nanoparticles. Angew. Chem. Int. Ed. Engl. 2007, 46, 6329–6332. 15. Lee, J. H.; Huh, Y. M.; Jun, Y. W.; et al. Artificially engineered magnetic nanoparticles for ultra-sensitive molecular imaging. Nat. Med. 2007, 13, 95–99. 16. Qin, J.; Laurent, S.; Jo, Y. S.; et al. A high-performance magnetic resonance imaging T2 contrast agent. Adv. Mater. 2007, 19, 1874–1878. 17. Rockenberger, J.; Scher, E. C.; Alivisatos, A. P. A new nonhydrolytic single-precursor approach to surfactant-capped nanocrystals of transition metal oxides. J. Am. Chem. Soc. 1999, 121, 11595–11596. 18. Sun, S.; Zeng, H.; Robinson, D. B.; et al. Monodisperse MFe2 O4 (M = Fe, Co, Mn) nanoparticles. J. Am. Chem. Soc. 2004, 126, 273–279. 19. Josephson, L.; Tung, C. H.; Moore, A.; Weissleder, R. High-efficiency intracellular magnetic labeling with novel superparamagnetic-Tat peptide conjugates. Bioconjug. Chem. 1999, 10, 186–191. 20. Portet, D.; Denizot, B.; Rump, E.; Lejeune, J. J.; Jallet, P. Nonpolymeric coatings of iron oxide colloids for biological use as magnetic resonance imaging contrast agents. J. Colloid Interface Sci. 2001, 238, 37–42. 21. Bulte, J. W.; Douglas, T.; Witwer, B.; et al. Magnetodendrimers allow endosomal magnetic labeling and in vivo tracking of stem cells. Nat. Biotechnol. 2001, 19, 1141–1147. 22. De Cuyper, M.; Joniau, M. Mechanistic aspects of the adsorption of phospholipids onto lauric acid stabilized magnetite nanocolloids. Langmuir 1991, 7, 647–652. 23. Nitin, N.; LaConte, L. E.; Zurkiya, O.; Hu, X.; Bao, G. Functionalization and peptide-based delivery of magnetic nanoparticles as an intracellular MRI contrast agent. J. Biol. Inorg. Chem. 2004, 9, 706–712. 24. Bogdanov, A. A.; Martin, C.; Weissleder, R.; Brady, T. J. Trapping of dextran-coated colloids in liposomes by transient binding to aminophospholipid: preparation of ferrosomes. Biochim. Biophys. Acta. 1994, 1193, 212–218. 25. Bulte, J. W.; De Cuyper, M. Magnetoliposomes as contrast agents. Methods Enzymol. 2003, 373, 175–198. 26. Illum, L.; Church, A. E.; Butterworth, M. D.; Arien, A.; Whetstone, J.; Davis, S. S. Development of systems for targeting the regional lymph nodes for diagnostic imaging: in vivo behaviour of colloidal PEG-coated magnetite nanospheres in the rat following interstitial administration. Pharm. Res. 2001, 18, 640–645.
P1: OTA/XYZ P2: ABC c24 JWBS037-Chen January 25, 2011
610
13:31
Printer Name: Yet to Come
MULTIMODAL IMAGING AND THERAPY WITH MAGNETOFLUORESCENT NANOPARTICLES
27. Moffat, B. A.; Reddy, G. R.; McConville, P.; et al. A novel polyacrylamide magnetic nanoparticle contrast agent for molecular imaging using MRI. Mol. Imaging 2003, 2, 324–332. 28. Kellar, K. E.; Fujii, D. K.; Gunther, W. H.; et al. NC100150 Injection, a preparation of optimized iron oxide nanoparticles for positive-contrast MR angiography. J. Magn. Reson. Imaging 2000, 11, 488–494. 29. Kellar, K. E.; Fujii, D. K.; Gunther, W. H.; Briley-Saebo, K.; Spiller, M.; Koenig, S. H. “NC100150”, a preparation of iron oxide nanoparticles ideal for positive-contrast MR angiography. Magma 1999, 8, 207–213. 30. Bulte, J. W.; Douglas, T.; Mann, S.; et al. Magnetoferritin: characterization of a novel superparamagnetic MR contrast agent. J. Magn. Reson. Imaging 1994, 4, 497–505. 31. Wilhelm, C.; Billotey, C.; Roger, J.; Pons, J. N.; Bacri, J. C.; Gazeau, F. Intracellular uptake of anionic superparamagnetic nanoparticles as a function of their surface coating. Biomaterials 2003, 24, 1001–1011. 32. Jung, C. W. Surface properties of superparamagnetic iron oxide MR contrast agents: ferumoxides, ferumoxtran, ferumoxsil. Magn. Reson. Imaging 1995, 13, 675–691. 33. Stark, D. D.; Weissleder, R.; Elizondo, G.; et al. Superparamagnetic iron oxide: clinical application as a contrast agent for MR imaging of the liver. Radiology 1988, 168, 297–301. 34. Weissleder, R.; Elizondo, G.; Stark, D. D.; et al. The diagnosis of splenic lymphoma by MR imaging: value of superparamagnetic iron oxide. AJR. Am. J. Roentgenol. 1989, 152, 175–180. 35. Weissleder, R.; Hahn, P. F.; Stark, D. D.; et al. MR imaging of splenic metastases: ferriteenhanced detection in rats. AJR. Am. J. Roentgenol. 1987, 149, 723–726. 36. Weissleder, R.; Stark, D. D. Magnetic resonance imaging of the liver. Magn. Reson. Q 1989, 5, 97–121. 37. Weissleder, R.; Stark, D. D. Magnetic resonance imaging of liver tumors. Semin. Ultrasound CT MR 1989, 10, 63–77. 38. Weissleder, R.; Stark, D. D.; Compton, C. C.; Wittenberg, J.; Ferrucci, J. T. Ferrite-enhanced MR imaging of hepatic lymphoma: an experimental study in rats. AJR Am. J. Roentgenol. 1987, 149, 1161–1165. 39. Weissleder, R.; Elizondo, G.; Wittenberg, J.; Lee, A. S.; Josephson, L.; Brady, T. J. Ultrasmall superparamagnetic iron oxide: an intravenous contrast agent for assessing lymph nodes with MR imaging. Radiology 1990, 175, 494–498. 40. Weissleder, R.; Elizondo, G.; Wittenberg, J.; Rabito, C. A.; Bengele, H. H.; Josephson, L. Ultrasmall superparamagnetic iron oxide: characterization of a new class of contrast agents for MR imaging. Radiology 1990, 175, 489–493. 41. Weissleder, R.; Stark, D. D.; Engelstad, B. L.; et al. Superparamagnetic iron oxide: pharmacokinetics and toxicity. Am. J. Roentgenol. 1989, 152, 167–173. 42. Harisinghani, M. G.; Barentsz, J.; Hahn, P. F.; et al. Noninvasive detection of clinically occult lymph-node metastases in prostate cancer. N. Engl. J. Med. 2003, 348, 2491–2499. 43. Harisinghani, M. G.; Weissleder, R. Sensitive, noninvasive detection of lymph node metastases. PLoS Med. 2004, 1, e66. 44. Ferrari, M. Cancer nanotechnology: opportunities and challenges. Nat. Rev. Cancer 2005, 5, 161–171. 45. Enochs, W. S.; Harsh, G.; Hochberg, F.; Weissleder, R. Improved delineation of human brain tumors on MR images using a long-circulating, superparamagnetic iron oxide agent. J. Magn. Reson. Imaging 1999, 9, 228–232. 46. Harisinghani, M. G.; Saini, S.; Weissleder, R.; et al. Splenic imaging with ultrasmall superparamagnetic iron oxide ferumoxtran-10 (AMI-7227): preliminary observations. J. Comput. Assist. Tomogr. 2001, 25, 770–776.
P1: OTA/XYZ P2: ABC c24 JWBS037-Chen January 25, 2011
13:31
Printer Name: Yet to Come
REFERENCES
611
47. Saini, S.; Sharma, R.; Baron, R. L.; et al. Multicentre dose-ranging study on the efficacy of USPIO ferumoxtran-10 for liver MR imaging. Clin. Radiol. 2000, 55, 690– 695. 48. Tang, Y.; Kim, M.; Carrasco, D.; Kung, A. L.; Chin, L.; Weissleder, R. In vivo assessment of RAS-dependent maintenance of tumor angiogenesis by real-time magnetic resonance imaging. Cancer Res. 2005, 65, 8324–8330. 49. Weissleder, R.; Lee, A. S.; Khaw, B. A.; Shen, T.; Brady, T. J. Antimyosin-labeled monocrystalline iron oxide allows detection of myocardial infarct: MR antibody imaging. Radiology 1992, 182, 381–385. 50. Kolb, H. C.; Finn, M. G.; Sharpless, K. B. Click chemistry: diverse chemical function from a few good reactions. Angew. Chem. Int. Ed. Engl. 2001, 40, 2004–2021. 51. Devaraj, N. K.; Keliher, E. J.; Thurber, G.; Nahrendorf, M.; Weissleder, R. 18 F labeled nanoparticles for in vivo PET-CT imaging. Bioconjug. Chem. 2009, 20, 397–401. 52. Sun, E. Y.; Josephson, L.; Weissleder, R. “Clickable” nanoparticles for targeted imaging. Mol. Imaging 2006, 5, 122–128. 53. Jaffer, F. A.; Tung, C. H.; Wykrzykowska, J. J.; et al. Molecular imaging of factor XIIIa activity in thrombosis using a novel, near-infrared fluorescent contrast agent that covalently links to thrombi. Circulation 2004, 110, 170–176. 54. Tung, C. H.; Ho, N. H.; Zeng, Q.; et al. Novel factor XIII probes for blood coagulation imaging. Chembiochem 2003, 4, 897–899. 55. Weissleder, R.; Kelly, K.; Sun, E. Y.; Shtatland, T.; Josephson, L. Cell-specific targeting of nanoparticles by multivalent attachment of small molecules. Nat. Biotechnol. 2005, 23, 1418–1423. 56. Kelly, K. A.; Bardeesy, N.; Anbazhagan, R.; et al. Targeted nanoparticles for imaging incipient pancreatic ductal adenocarcinoma. PLoS Med. 2008, 5, e85. 57. Simberg, D.; Duza, T.; Park, J. H.; et al. Biomimetic amplification of nanoparticle homing to tumors. Proc. Natl. Acad. Sci. U.S.A. 2007, 104, 932–936. 58. Lee, H. Y.; Li, Z.; Chen, K.; et al. PET/MRI dual-modality tumor imaging using arginineglycine–aspartic (RGD)-conjugated radiolabeled iron oxide nanoparticles. J. Nucl. Med. 2008, 49, 1371–1379. 59. Hansson, G. K. Inflammation, atherosclerosis, and coronary artery disease. N. Engl. J. Med. 2005, 352, 1685–1695. 60. Hansson, G. K.; Libby, P. The immune response in atherosclerosis: a double-edged sword. Nat. Rev. Immunol. 2006, 6, 508–519. 61. Libby, P. Inflammation in atherosclerosis. Nature 2002, 420, 868–874. 62. Trivedi, R. A.; U-King-Im, J.M.; Graves, M. J.; Kirkpatrick, P. J.; Gillard, J. H. Noninvasive imaging of carotid plaque inflammation. Neurology 2004, 63, 187–188. 63. Trivedi, R. A.; U-King-Im, J. M.; Graves, M. J.; et al. In vivo detection of macrophages in human carotid atheroma: temporal dependence of ultrasmall superparamagnetic particles of iron oxide-enhanced MRI. Stroke 2004, 35, 1631–1635. 64. Jaffer, F. A.; Nahrendorf, M.; Sosnovik, D.; Kelly, K. A.; Aikawa, E.; Weissleder, R. Cellular imaging of inflammation in atherosclerosis using magnetofluorescent nanomaterials. Mol. Imaging. 2006, 5, 85–92. 65. Pande, A. N.; Kohler, R. H.; Aikawa, E.; Weissleder, R.; Jaffer, F. A. Detection of macrophage activity in atherosclerosis in vivo using multichannel, high-resolution laser scanning fluorescence microscopy. J. Biomed. Opt. 2006, 11, 021009. 66. Nahrendorf, M.; Zhang, H.; Hembrador, S.; et al. Nanoparticle PET-CT imaging of macrophages in inflammatory atherosclerosis. Circulation 2008, 117, 379–387.
P1: OTA/XYZ P2: ABC c24 JWBS037-Chen January 25, 2011
612
13:31
Printer Name: Yet to Come
MULTIMODAL IMAGING AND THERAPY WITH MAGNETOFLUORESCENT NANOPARTICLES
67. McCarthy, J. R. Theranostics: agents for diagnosis and therapy. In Molecular Imaging: Principles and Practice. Weissleder, R.; Gambhir, S. S.; Ross, B. D.; Rehemtulla, A. Eds.; People’s Medical Publishing House: Shelton, CT, 2010, 509–522. 68. McCarthy, J. R.; Jaffer, F. A.; Weissleder, R. Imaging and therapy of atherosclerotic lesions with theranostic nanoparticles. In Methods in Bioengineering. Rege, K.; Medintz, I. L. Eds.; Artec House: Boston, 2009, 137–151. 69. McCarthy, J. R.; Jaffer, F. A.; Weissleder, R. A macrophage-targeted theranostic nanoparticle for biomedical applications. Small 2006, 2, 983–987. 70. Sosnovik, D. E.; Nahrendorf, M.; Deliolanis, N.; et al. Fluorescence tomography and magnetic resonance imaging of myocardial macrophage infiltration in infarcted myocardium in vivo. Circulation 2007, 115, 1384–1391. 71. Tsourkas, A.; Shinde-Patil, V. R.; Kelly, K. A.; et al. In vivo imaging of activated endothelium using an anti-VCAM-1 magnetooptical probe. Bioconjug. Chem. 2005, 16, 576–581. 72. Kelly, K. A.; Allport, J. R.; Tsourkas, A.; Shinde-Patil, V. R.; Josephson, L.; Weissleder, R. Detection of vascular adhesion molecule-1 expression using a novel multimodal nanoparticle. Circ. Res. 2005, 96, 327–336. 73. Kelly, K. A.; Nahrendorf, M.; Yu, A. M.; Reynolds, F.; Weissleder, R. In vivo phage display selection yields atherosclerotic plaque targeted peptides for imaging. Mol. Imaging Biol. 2006, 8, 201–207. 74. Nahrendorf, M.; Jaffer, F. A.; Kelly, K. A.; et al. Noninvasive vascular cell adhesion molecule-1 imaging identifies inflammatory activation of cells in atherosclerosis. Circulation 2006, 114, 1504–1511. 75. Thermo Scientific: http://www.piercenet.com/. 76. Molecular Biosciences: http://www.molbio.com/. 77. Quanta Biodesign Limited: http://www.quantabiodesign.com/. 78. Creative PEGWorks: http://www.creativebiochem.com/. 79. Nanocs: http://www.nanocs.com/. 80. Laysan Bio, Inc.: http://www.laysanbio.com/. 81. Kang, H. W.; Josephson, L.; Petrovsky, A.; Weissleder, R.; Bogdanov, A., Jr. Magnetic resonance imaging of inducible E-selectin expression in human endothelial cell culture. Bioconjug. Chem. 2002, 13, 122–127. 82. Reynolds, P. R.; Larkman, D. J.; Haskard, D. O.; et al. Detection of vascular expression of E-selectin in vivo with MR imaging. Radiology 2006, 241, 469–476. 83. Funovics, M.; Montet, X.; Reynolds, F.; Weissleder, R.; Josephson, L. Nanoparticles for the optical imaging of tumor E-selectin. Neoplasia 2005, 7, 904–911. 84. Chen, H. W.; Medley, C. D.; Sefah, K.; et al. Molecular recognition of small-cell lung cancer cells using aptamers. ChemMedChem 2008, 3, 991–1001. 85. Benedetto, S.; Pulito, R.; Crich, S. G.; et al. Quantification of the expression level of integrin receptor alpha(v)beta3 in cell lines and MR imaging with antibody-coated iron oxide particles. Magn. Reson. Med. 2006, 56, 711–716. 86. Montet, X.; Montet-Abou, K.; Reynolds, F.; Weissleder, R.; Josephson, L. Nanoparticle imaging of integrins on tumor cells. Neoplasia 2006, 8, 214–222. 87. Montet, X.; Funovics, M.; Montet-Abou, K.; Weissleder, R.; Josephson, L. Multivalent effects of RGD peptides obtained by nanoparticle display. J. Med. Chem. 2006, 49, 6087–6093. 88. Moore, A.; Medarova, Z.; Potthast, A.; Dai, G. In vivo targeting of underglycosylated MUC-1 tumor antigen using a multimodal imaging probe. Cancer Res. 2004, 64, 1821–1827. 89. Schellenberger, E. A.; Bogdanov, A., Jr.; Hogemann, D.; Tait, J.; Weissleder, R.; Josephson, L. Annexin V-CLIO: a nanoparticle for detecting apoptosis by MRI. Mol. Imaging 2002, 1, 102–107.
P1: OTA/XYZ P2: ABC c24 JWBS037-Chen January 25, 2011
13:31
Printer Name: Yet to Come
REFERENCES
613
90. Schellenberger, E. A.; Bogdanov, A., Jr.; Petrovsky, A.; Ntziachristos, V.; Weissleder, R.; Josephson, L. Optical imaging of apoptosis as a biomarker of tumor response to chemotherapy. Neoplasia 2003, 5, 187–192. 91. Lewin, M.; Carlesso, N.; Tung, C. H.; et al. Tat peptide-derivatized magnetic nanoparticles allow in vivo tracking and recovery of progenitor cells. Nat. Biotechnol. 2000, 18, 410–414. 92. Wunderbaldinger, P.; Josephson, L.; Weissleder, R. Tat peptide directs enhanced clearance and hepatic permeability of magnetic nanoparticles. Bioconjug. Chem. 2002, 13, 264–268. 93. Kelly, K. A.; Setlur, S. R.; Ross, R.; et al. Detection of early prostate cancer using a hepsintargeted imaging agent. Cancer Res. 2008, 68, 2286–2291. 94. Park, J. H.; von Maltzahn, G.; Ruoslahti, E.; Bhatia, S. N.; Sailor, M. J. Micellar hybrid nanoparticles for simultaneous magnetofluorescent imaging and drug delivery. Angew. Chem. Int. Ed. Engl. 2008, 47, 7284–7288. 95. Reddy, G. R.; Bhojani, M. S.; McConville, P.; et al. Vascular targeted nanoparticles for imaging and treatment of brain tumors. Clin. Cancer Res. 2006, 12, 6677–6686. 96. Funovics, M. A.; Kapeller, B.; Hoeller, C.; et al. MR imaging of the her2/neu and 9.2.27 tumor antigens using immunospecific contrast agents. Magn. Reson. Imaging 2004, 22, 843–850. 97. Sakamoto, J. H.; Smith, B. R.; Xie, B.; Rokhlin, S. I.; Lee, S. C.; Ferrari, M. The molecular analysis of breast cancer utilizing targeted nanoparticle based ultrasound contrast agents. Technol. Cancer Res. Treat. 2005, 4, 627–636. 98. Toma, A.; Otsuji, E.; Kuriu, Y.; et al. Monoclonal antibody A7-superparamagnetic iron oxide as contrast agent of MR imaging of rectal carcinoma. Br. J. Cancer 2005, 93, 131–136. 99. DeNardo, S. J.; DeNardo, G. L.; Miers, L. A.; et al. Development of tumor targeting bioprobes ((111)In-chimeric L6 monoclonal antibody nanoparticles) for alternating magnetic field cancer therapy. Clin. Cancer Res. 2005, 11, 7087s–7092s. 100. Montet, X.; Weissleder, R.; Josephson, L. Imaging pancreatic cancer with a peptide–nanoparticle conjugate targeted to normal pancreas. Bioconjug. Chem. 2006, 17, 905–911. 101. Gunn, J.; Wallen, H.; Veiseh, O.; et al. A multimodal targeting nanoparticle for selectively labeling T cells. Small 2008, 4, 712–715.
P1: OTA/XYZ P2: ABC c25 JWBS037-Chen August 29, 2010
10:53
Printer Name: Yet to Come
CHAPTER 25
Gold Nanocages: A Multifunctional Platform for Molecular Optical Imaging and Photothermal Treatment LESLIE AU, CLAIRE M. COBLEY, JINGYI CHEN, and YOUNAN XIA Department of Biomedical Engineering, Washington University, St. Louis, Missouri, USA
Gold nanocages are single-crystalline, hollow structures with porous walls. They are commonly prepared using the galvanic replacement reaction between silver nanocubes and a gold salt precursor. By controlling the amount of gold precursor added, we can tune localized surface plasmon resonance peaks of the resultant gold nanocages into the near-infrared, where the attenuation of light by blood and soft tissue is negligibly low. Calculations based on the discrete dipole approximation method indicate that both the absorption and scattering cross sections of gold nanocages can be up to five orders in magnitude stronger than molecular dyes. In addition, gold nanocages are biocompatible and their surface can easily be modified through the gold-thiolate monolayer chemistry. All these tributes make gold nanocages a novel class of molecular agents for optical bioimaging and photothermal treatment. In this chapter, we present some recent advances in the synthesis and utilization of gold nanocages. Specifically, it will be shown that the optical properties of gold nanocages can be tailored for use as molecular contrast agents in both optical and spectroscopic coherence tomography (OCT and SOCT, respectively), as well as photoacoustic tomography (PAT). Our studies suggest strong enhancement in contrast when gold nanocages were added to tissue phantom for OCT and SOCT. For PAT imaging, we demonstrated that gold nanocages could enhance the optical absorption of vasculature by up to 81%. Additionally, the nanocages were used as tracers for noninvasively mapping the location of sentinel lymph nodes for breast cancer staging. Lastly, when nanocages were modified with tumor-targeting ligands, they were effective in the photothermal ablation of cancer cells.
25.1 INTRODUCTION Metal nanostructures, in particular, those made of silver (Ag) and gold (Au), exhibit a unique property known as localized surface plasmon resonance (LSPR). This optical phenomenon Nanoplatform-Based Molecular Imaging Edited by Xiaoyuan Chen C 2011 John Wiley & Sons, Inc. Copyright
615
P1: OTA/XYZ P2: ABC c25 JWBS037-Chen August 29, 2010
616
10:53
Printer Name: Yet to Come
A MULTIFUNCTIONAL PLATFORM FOR MOLECULAR OPTICAL IMAGING
occurs when the conduction electrons in the metal nanostructure collectively oscillate with the electrical field of light [1]. The accompanying strong extinction (scattering plus absorption) of light has generated great interest in engineering these nanostructures for a variety of applications including optical imaging contrast enhancement and photothermal treatment [1–9]. Gold nanostructures are particularly interesting due to their bioinert nature and strong interaction with the thiolate group, allowing for the attachment of a variety of surface functional groups to enhance circulation in the blood and thus target the malignant cells in a solid tumor [9]. Furthermore, the optical properties of these nanostructures can readily be tailored by controlling their size, shape, composition, and/or interior (hollow vs. solid) [8–11]. For solid gold nanostructures, chemical reduction or thermal decomposition routes have been used to achieve many different shapes, including nanowires [12], nanorods [13–15], nanospheres [16], nanoplates [17–19], and nanocubes [20–22], among others [23–25]. Recently, the galvanic replacement reaction has been demonstrated as a remarkably simple and versatile route to gold hollow nanostructures with tunable LSPR properties [26]. In this chapter, we focus on recent advances from our laboratory with regard to the synthesis and applications of gold nanocages: hollow and porous nanostructures with sizes <100 nm. The nanocages, composed primarily of Au (with some remaining Ag), have optical extinction that can be tuned to the near-infrared region where light scattering from blood and soft tissue is low, making them particularly attractive for in vivo applications [27–29]. We begin by discussing the synthetic protocols and a number of methods we have developed for optimizing the morphology and optical properties. This discussion is followed by highlights of the optical properties of these nanomaterials and finally their applications in techniques such as optical and spectroscopic coherence tomography (OCT and SOCT, respectively), photoacoustic tomography (PAT), and photothermal cancer treatment.
25.2 SYNTHESIS OF GOLD NANOCAGES 25.2.1 Polyol Synthesis of Silver Nanocubes To prepare Au nanocages, we have to first prepare the sacrificial templates—uniform Ag nanocubes. The Ag nanocubes are most commonly synthesized using the polyol method, which involves heating ethylene glycol, AgNO3 , and poly(vinyl pyrrolidone) (PVP) [30]. These three components function as the reducing agent/solvent, metal precursor, and capping agent, respectively. In a typical reaction, ethylene glycol is heated to 150–160 o C in air to generate glycolaldehyde (Eq. (25.1)) [31]: 2HOCH2 CH2 OH + O2 → 2HOCH2 CHO + H2 O
(25.1)
Glycolaldehyde then reduces the Ag+ ions to Ag atoms, which subsequently agglomerate to form nuclei and seeds. As more Ag atoms are produced, the seeds grow into different nanostructures depending on their crystallinity. The role of PVP is to direct the addition of atoms to the {111} facets, due to its stronger binding affinity toward the {100} facets. For example, Ag pentagonal nanorods in the presence of PVP will extend to form nanowires since PVP will bind to the {100} side faces, facilitating the addition of Ag atoms onto the {111} facets at the ends of the wire [32]. Similarly, in the case of the single-crystal Ag cubooctahedrons, PVP will block the addition of atoms onto the {100} side facets,
P1: OTA/XYZ P2: ABC c25 JWBS037-Chen August 29, 2010
10:53
Printer Name: Yet to Come
SYNTHESIS OF GOLD NANOCAGES
(a)
(b)
(c)
(d)
617
FIGURE 25.1 Synthesis of Ag nanocubes based on the polyol reduction coupled with an oxidative etching process. SEM images of Ag nanocubes with various sizes: (a) 30 nm, (b) 50 nm, (c) 90 nm, and (d) 110 nm. (Adapted with permission from Chen et al. [61] and Skrabalak et al. [35]. Copyright © 2005 Wiley and 2007 Nature Publishing Group.)
allowing Ag atoms to add onto the {111} corner facets and promoting their growth into Ag nanocubes with sharp corners. In order to synthesize nanocubes in high yields, it is crucial to control the crystallinity of the initial seeds. In one method developed in our group, we selectively remove twinned seeds by oxidative etching with O2 and Cl− , leaving behind only single-crystal seeds, which grow into Ag nanocubes [33]. Silver nanocubes of different sizes can be routinely produced through this etching method (Fig. 25.1). Most recently, we have demonstrated a much faster method that eliminates the formation of twinned seeds by increasing the reduction rate of Ag+ [34, 35]. This process was achieved by introducing a trace amount (on the ppm level) of Na2 S or NaHS to the reaction, producing Ag2 S nanocrystallites that can catalyze the reduction of additional Ag+ . This method has significantly reduced the overall reaction time from 10–24 hours to less than 20 minutes and is now the most commonly used protocol for producing Ag nanocubes in large quantities. 25.2.2 Galvanic Replacement Reaction with AuCl4 − The galvanic replacement reaction provides a simple and versatile method for generating hollow nanostructures with tunable plasmonic properties, such as Au nanocages [26, 32, 35–38]. When an aqueous suspension of Ag nanocubes with sharp corners is titrated with an aqueous solution of AuCl4 − , the galvanic replacement reaction between these two species occurs immediately, leading to the formation of Au-based nanoboxes and eventually porous
P1: OTA/XYZ P2: ABC c25 JWBS037-Chen August 29, 2010
618
10:53
Printer Name: Yet to Come
A MULTIFUNCTIONAL PLATFORM FOR MOLECULAR OPTICAL IMAGING
nanocages [38]. AuCl4 − oxidizes the sacrificial Ag template to AgCl, which is highly soluble at the boiling temperature used for the reaction (Eq. (25.2)): + − 3Ag(s) + AuCl− 4 (aq) → 3Ag (aq) + Au(s) + 4Cl (aq)
(25.2)
The driving force for this reaction originates from the difference in standard reduction potential for the AuCl4 − /Au pair (0.99 V vs. the standard hydrogen electrode, SHE) and the AgCl/Ag pair (0.22 V). The electrons generated in the oxidation process migrate to the surface of the Ag cubes and reduce AuCl4 − to Au atoms. Gold atoms are able to epitaxially nucleate and grow on the surface of the Ag template since Au and Ag share the same face-centered cubic structure with closely matching lattice constants (4.0786 and 4.0862 ˚ respectively). A, The morphological and compositional changes at various stages of the replacement reaction were monitored using scanning electron microscopy (SEM) and transmission electron microscopy (TEM). After Ag nanocubes with sharp corners shown in Figure 25.2a
(a)
(b)
(c)
(d)
(e)
FIGURE 25.2 Synthesis of Au nanocages that involve Ag nanocubes with sharp corners and AuCl4 − . SEM images of Ag nanocubes (a) titrated with different volumes of 1 mM AuCl4 − solution: (b) 0.05, (c) 0.30, and (d) 0.50 mL. The formation of hollow structures was confirmed by TEM studies on microtomed samples (the inset in panel c). The electron diffraction patterns of the nanocube and nanocage (upper corners in panels a and d) taken perpendicular to the side face. (e) Schematic detailing the major steps involved in the formation of Au nanocages with pores on the side faces. (Adapted with permission from Sun and Xia [38]. Copyright © 2004 American Chemical Society.)
P1: OTA/XYZ P2: ABC c25 JWBS037-Chen August 29, 2010
10:53
Printer Name: Yet to Come
SYNTHESIS OF GOLD NANOCAGES
619
reacted with a small amount of AuCl4 − solution, a pinhole was observed on one of the six faces of each cube, as seen in Figure 25.2b. This pinhole indicated that the replacement reaction was initiated locally at a high-energy site (e.g., surface step, point defect, or hole in the PVP capping layer) [39] rather than over the entire cube surface. While Au was deposited on the surface, the pinhole served as the site for Ag dissolution, facilitating the conversion of the nanocube into a hollow nanobox. In later stages of the reaction, the size of the pinhole was reduced likely through mass diffusion processes or direct deposition of Au near the pinhole [40]. Eventually the pinhole closed, forming seamless nanoboxes with a uniform Au/Ag composition as seen in Figure 25.2c. A TEM image of a microtomed nanobox revealed the hollow interior (Fig. 25.2c, upper inset). When more Au was added, the corners of the nanoboxes became truncated as seen in Figure 25.2d, indicating the start of the dealloying process and the removal of Ag from the Au-Ag alloyed walls. The inset of Figure 25.2d shows the electron diffraction pattern taken perpendicular to a square face of the nanobox and reveals the same symmetry as that of an Ag nanocube as seen in the inset of Figure 25.2a, confirming the epitaxial growth of Au. Figure 25.2e summarizes the morphological changes involved in the galvanic replacement reaction between an individual Ag nanocube with sharp corners and AuCl4 − . Though still highly useful for plasmonic applications, nanocages with randomly distributed and polydispersed pores are not likely to work well for controlled drug release applications. As nanostructures with multiple functionalities are highly desirable, we developed a method to form highly uniform pores confined to the corners by using Ag nanocubes with truncated corners as the sacrificial template. The corners of nanocubes were truncated prior to the galvanic replacement with a thermal annealing process [41]. Figure 25.3a shows SEM and TEM (inset) images of the truncated Ag nanocubes after aging the sample in an ethylene glycol solution containing 1 mM HCl at 160 o C for 5 min in the presence of 0.1 mM PVP (calculated in terms of the repeating unit). The solution contained only a small amount of PVP, thus providing protection for the {100} facets but little for the {111} facets, subsequently leading to the truncation of all corners. Figure 25.3b–d, shows SEM and TEM (inset) images of the nanostructures obtained after the truncated nanocubes reacted with different amounts of AuCl4 − . In contrast to the case of nanocubes with sharp corners, the reaction started simultaneously from all corners of a truncated cube (Fig. 25.3b). As the reaction continued, these initial pits grew into well-defined pores at each corner and the ratio of {111} to {100} facets increased as a result of surface reconstruction (Fig. 25.3c). In the final stage, the nanocages were bounded by {100} facets with large triangular pores at all corners, and the corners grew larger as additional AuCl4 − was added (Fig. 25.3d). Figure 25.3e summarizes the major steps involved in the formation of Au-Ag nanocages with well-controlled pores at the corners.
25.2.3 Galvanic Replacement Reaction with AuCl2 − According to the stoichiometry between Ag and AuCl4 − (Eq. (25.2)), only one Au atom is formed for every three Ag atoms that are oxidized. If AuCl4 − is replaced with a precursor such as AuCl2 − , the stoichiometry will change due to the difference in oxidation number for Au. For the Au(I) precursor, one Au atom will be generated for every Ag atom being oxidized. This change will impact the alloying and dealloying processes and can provide additional flexibility regarding the morphology, wall thickness, and LSPR position of resultant hollow nanostructures.
P1: OTA/XYZ P2: ABC c25 JWBS037-Chen August 29, 2010
620
10:53
Printer Name: Yet to Come
A MULTIFUNCTIONAL PLATFORM FOR MOLECULAR OPTICAL IMAGING
(a)
(b)
(c)
(d)
(e)
FIGURE 25.3 Syntheses of Au nanocages that involve Ag nanocubes with truncated corners and AuCl4 − . SEM and TEM (insets) images of truncated Ag nanocubes (a) and products after the addition of different volumes of 0.1 mM AuCl4 − : (b) 0.6, (c) 1.6, and (d) 3.0 mL. The scale bars in all insets represent 50 nm. (e) Schematic detailing the major steps in the formation of Au nanocages with well-controlled pores at the corners. (Adapted with permission from Chen et al. [41]. Copyright © 2006 American Chemical Society.)
Here we discuss the use of AuCl2 − in place of AuCl4 − in the galvanic replacement reaction with Ag nanocubes [42, 43]. The standard reduction potential of the AuCl2 − /Au pair (1.11 V vs. SHE) is higher than that of the AgCl/Ag; hence Ag nanocubes can also be oxidized by AuCl2 − (Eq. (25.3)) [44]. + − Ag(s) + AuCl− 2 (aq) → Ag (aq) + Au(s) + 2Cl (aq)
(25.3)
In comparison to AuCl4 − , the replacement reaction with AuCl2 − will produce hollow structures with thicker walls as a result of the 1:1 ratio between Ag consumed and Au deposited. These thicker structures are more robust and could sustain longer through the dealloying process, thus allowing for the formation of cubic Au nanoframes in high yields. Figure 25.4a–d, shows the SEM and TEM (insets) images of samples obtained at different stages of the galvanic replacement reaction between Ag nanocubes and AuCl2 − . Ag nanocubes (Fig. 25.4a) were titrated with a NaCl-saturated aqueous solution of AuCl2 − . Au-Ag nanoboxes with truncated corners (Fig. 25.4b) developed via a similar mechanism as described earlier in the chapter. The major differences are that the walls were thicker
P1: OTA/XYZ P2: ABC c25 JWBS037-Chen August 29, 2010
10:53
Printer Name: Yet to Come
SYNTHESIS OF GOLD NANOCAGES
(a)
(b)
(c)
(d)
621
(e)
FIGURE 25.4 Syntheses of Au nanocages that involve Ag nanocubes with sharp corners and AuCl2 − . SEM and TEM (insets) images showing different stages of the galvanic replacement reaction where 50-nm Ag nanocubes (a) were titrated with different volumes of 0.2 mM AuCl2 − : (b) 4.0, (c) 5.0, and (d) 6.0 mL. The scale bars in all insets represent 50 nm. (e) Schematic detailing the major steps in the galvanic replacement reaction between a Ag nanocube and AuCl2 − . (Adapted with permission from Au et al. [42]. Copyright © 2008 Springer.)
and the disappearance of the pinhole occurred much earlier in the reaction with AuCl2 − than with AuCl4 − . Again, this variation can be attributed to the change in stoichiometry, one Au atom depositing for each Ag atom removed. When more AuCl2 − was added, pores formed at the corners and faces (Fig. 25.4c) during the dealloying process. As the titration continued, the pores on the side faces enlarged while the pores at the corners reduced in size, suggesting that the atoms migrated to the more stable {111} facet. When the pores on the sides reached their maximum size, the pores on the corners completely disappeared, resulting in the formation of cubic nanoframes (Fig. 25.4d). Figure 25.4e shows a schematic illustrating the major steps in this reaction. By simply changing the oxidation state of the precursor, this new method provides a route to hollow gold nanostructures with thicker walls, thus allowing improvement for engineering their optical and mechanical properties.
25.2.4 Selective Etching with Fe(NO3 )3 Although the methods described above allow for easy tuning of the LSPR peak to ∼800 nm, it is difficult to shift the peak further into the NIR region while keeping the
P1: OTA/XYZ P2: ABC c25 JWBS037-Chen August 29, 2010
622
10:53
Printer Name: Yet to Come
A MULTIFUNCTIONAL PLATFORM FOR MOLECULAR OPTICAL IMAGING
(a)
(b)
(c)
(d)
(e)
FIGURE 25.5 Syntheses of Au nanocages that involve the dealloying of Au-Ag nanoboxes with Fe(NO3 )3 as an etching agent. TEM and SEM (inset) images of (a) 50-nm Ag nanocubes; (b) AuAg alloy nanoboxes obtained by reacting the nanocubes with 4.0 mL of 0.2 mM AuCl4 − aqueous solution; and (c) nanocages and (d) nanoframes obtained by etching the nanoboxes with 10 and 20 L, respectively, of 50 mM aqueous Fe(NO3 )3 solution. The inset in (d) shows the SEM image obtained at a tilting angle of 45◦ , clearly showing the 3D structure of a nanoframe. The scale bars in all insets represent 50 nm. (e) Schematic illustrating the major steps of the reaction: Ag nanocube was titrated with AuCl4 − , thus forming a Au-Ag alloy nanobox, which was then treated with Fe(NO3 )3 to selectively remove Ag from the wall. (Adapted with permission from Lu et al. [50]. Copyright © 2007 American Chemical Society.)
overall size of the nanocages small, as required for in vivo applications. Since further Ag dealloying also results in Au deposition in galvanic replacement reactions, it is impossible to achieve the ultrathin walls that would be necessary for longer LSPR wavelengths. For this reason, we investigated the use of well-known wet etchants for Ag, such as Fe(NO3 )3 , to remove Ag from the walls of the nanocages without adding Au to the surface [45–51]. Figure 25.5a–d shows the TEM and SEM (insets) images of different stages of the reaction. To begin, 50-nm nanoboxes composed of a thin Au-Ag alloy shell and a partially unreacted Ag in the interior (Fig. 25.5a) were synthesized by stopping the galvanic replacement reaction between Ag nanocubes and AuCl4 − at an early stage. An aqueous solution of Fe(NO3 )3 was introduced into the suspension of Au-Ag alloy nanoboxes, and the remaining Ag in the nanoboxes was dissolved according to Eq. (25.4): Ag(s) + Fe(NO3 )3 (aq) → AgNO3 (aq) + Fe(NO3 )2 (aq)
(25.4)
P1: OTA/XYZ P2: ABC c25 JWBS037-Chen August 29, 2010
10:53
Printer Name: Yet to Come
OPTICAL PROPERTIES OF GOLD NANOCAGES
623
In contrast to the previously described galvanic replacement reaction with AuCl4 − or AuCl2 − , the reaction between the Au-Ag alloy nanoboxes and Fe(NO3 )3 is solely a dealloying process. The nanoboxes developed into nanocages with increasing porosity upon the addition of Fe(NO3 )3 (Fig. 25.5b,c) as increasing amounts of Ag were removed from the walls. The final structures are cubic nanoframes made of pure Au (Fig. 25.5d) with a LSPR beyond 1200 nm. The inset in Figure 25.5d shows the SEM of the product at a 45o tilt angle, making the three-dimensional (3D) morphology of the particle more apparent. As can be seen here, the pores are primarily on the side faces as these {100} facets are more vulnerable to etching. The Au atoms left behind from the Au-Ag alloy dissolution are able to migrate to the more stable edges, creating a frame-like structure. Figure 25.5e shows a schematic detailing all the major steps involved in this procedure.
25.3 OPTICAL PROPERTIES OF GOLD NANOCAGES 25.3.1 Tunable Localized Surface Plasmon Resonance One of the great advantages of the synthesis based on galvanic replacement reaction is its ability to precisely tune the LSPR peak of Au nanocages to a specific wavelength. In this way, the synthesis can be optimized for maximum absorption at the wavelength needed for a specific laser or application. For biological applications, the region of interest is in the near-infrared (800–900 nm), where soft tissue and blood are optically transparent. Figure 25.6 shows that by simply adjusting the amount of AuCl4 − added, the peak can be shifted controllably from ∼450 nm for the 50-nm Ag nanocube template to ∼900 nm when fully transformed into porous Au nanocages. It is easy to monitor the wavelength during
FIGURE 25.6 Spectral tuning for Au nanocages. (Top) Vials containing Au nanocages prepared with 50-nm Ag nanocubes and different volumes of 0.1 mM AuCl4 − aqueous solution: from left to right, 0, 0.3, 0.5, 1.0, 1.5, 2.0, 4.0 and 5.5 mL. (Bottom) The corresponding normalized UV– vis absorbance spectra taken from the Ag nanocubes and Au nanocages. The LSPR peak of the Au nanocages is precisely tunable throughout the visible and near-infrared regions by varying the volume of AuCl4 − solution added. (Reproduced with permission from Skrabalak et al. [35]. Copyright © 2007 Nature Publishing Group.)
P1: OTA/XYZ P2: ABC c25 JWBS037-Chen August 29, 2010
624
10:53
Printer Name: Yet to Come
A MULTIFUNCTIONAL PLATFORM FOR MOLECULAR OPTICAL IMAGING
the reaction with UV–vis spectroscopy, allowing for precise tuning. As discussed above, it is also possible to shift the peak even further into the near-infrared through the use of wet etchants that dealloy and thin the walls.
25.3.2 Calculation of Absorption and Scattering Cross Sections It has been shown that the size, wall thickness, and porosity of the hollow nanostructures have a strong effect on the absorption and scattering cross sections [52–57]. To better engineer Au nanocages for biomedical applications, we calculated the theoretical spectra for a number of different types of cages. Gustav Mie solved Maxwell’s equations to generate an exact formula for optical extinction in 1908; however, this method is only limited to geometries with spherical symmetry and is consequently not suitable for studying Au nanocages [58]. Instead, we use the discrete dipole approximation (DDA), which has been demonstrated to be reliable for calculating the optical properties of arbitrary geometries [54, 59, 60]. The DDA method approximates a nanostructure as a cubic array of polarizable units that interact both with the incident field and with each other. The resulting polarization at each point can be used to calculate the overall optical cross section of the particle at any wavelength, generating simulated spectra. Figure 25.7a,b compares the calculated extinction (Cext ), absorption (Cabs ), and scattering (Csca ) coefficients (note that Cext = Cabs + Csca ) for Au nanoboxes 60 and 40 nm in edge length, respectively, with the wall thickness being 5 nm [61]. The refractive index of bulk
FIGURE 25.7 Extinction, absorption, and scattering spectra calculated using the DDA method for Au nanocages having four different sets of geometric parameters: (a) a Au nanobox of 50 nm in inner edge length and 5 nm in wall thickness; (b) a Au nanobox of 30 nm in inner edge length and 5 nm in wall thickness; (c) the same as in (b) except that the wall thickness is 3 nm; (d) the same as in (b) except that the eight corners are decorated with pores of 5 nm in edge length. (Reproduced with permission from Chen et al. [61]. Copyright © 2005 Wiley.)
P1: OTA/XYZ P2: ABC c25 JWBS037-Chen August 29, 2010
10:53
Printer Name: Yet to Come
ENHANCING OPTICAL IMAGING CONTRAST WITH GOLD NANOCAGES
625
Au was used, and the nanobox was assumed to be surrounded by and completely filled with water. As expected from studies of solid Au particles, absorption dominates the extinction spectra at small sizes (<30 nm), while scattering dominates at larger sizes (>60 nm). These nanoboxes have absorption cross sections at least five orders of magnitude larger than organic chromophores. Figure 25.7c,d shows the influence of reduction in wall thickness and introduction of pores at corners, respectively. Figure 25.7c shows that the extinction peak was red-shifted from 710 nm to 820 nm when the wall thickness of the nanobox was reduced from 5 nm to 3 nm. Interestingly, the magnitudes of both scattering and absorption cross sections only changed slightly. Figure 25.7d shows the calculated spectra of a nanobox similar to that used for Figure 25.7b, except that all eight corners were replaced with holes to form a nanocage. The peak remained roughly at the same position, though a small reduction in the magnitude of the extinction cross section was observed. Further calculations have shown that the extinction coefficient of a 40-nm Au nanocage linearly decreased with the number of holes. The calculated spectra match well with what we have observed experimentally with the solution-phase spectra of nanoboxes; the slight broadening of the peak seen in the experimental data can be attributed to the minor differences in the uniformity of the nanostructures (i.e., size, truncation levels, and wall thickness). These insights into the effect of size, wall thickness, and porosity should allow us to more effectively tune the optical properties of nanocages.
25.4 ENHANCING OPTICAL IMAGING CONTRAST WITH GOLD NANOCAGES 25.4.1 Optical and Spectroscopic Coherence Tomography Optical coherence tomography (OCT) and spectroscopic coherence tomography (SOCT) are emerging as promising diagnostic techniques for noninvasive imaging of biological samples due to their high-resolution capabilities. These techniques are based on Michelson interferometry, which measures the interference signal between the backscattered light of a sample and a reference beam [62]. Consequently, the primary contrast comes from the scattering and absorption of light by tissue, which is typically weak. Gold nanocages can be useful as contrast agents for OCT and SOCT due to their strong scattering and absorption properties. To demonstrate the potential of Au nanocages as a contrast enhancing agent for OCT, 35-nm nanocages with the LSPR peak tuned to 700 nm were embedded in a gelatin phantom containing TiO2 granules to mimic the background scattering in biological tissues [63]. The left portion of the phantom contained only TiO2 , while the right portion was doped with Au nanocages in addition to TiO2 . The OCT and SOCT images were generated using a Ti:Sapphire laser with a central wavelength of 825 nm and a bandwidth of 155 nm. As shown in Figure 25.8a, we could easily distinguish these two portions in the conventional OCT image. Figure 25.8b shows the depth-dependent OCT signal from the portion of the phantom with and without nanocages (solid and dashed line, respectively). It is evident that the demodulated OCT signal decays faster on the portion that was doped with Au nanocages.
P1: OTA/XYZ P2: ABC c25 JWBS037-Chen August 29, 2010
626
10:53
Printer Name: Yet to Come
A MULTIFUNCTIONAL PLATFORM FOR MOLECULAR OPTICAL IMAGING
FIGURE 25.8 OCT and SOCT imaging with Au nanocages. (a) Conventional OCT intensity image of the TiO2 –gelatin phantom displaying the strong attenuation of the light backscattered from the portion that contained Au nanocages. (b) Demodulated OCT intensity decay (solid curve, decay over the region of TiO2 only; dashed curve, decay over the nanocage-doped region). (c) SOCT image of the phantom displaying the centroid of the spatially dependent backscattered spectra. (d) Spatially resolved, depth-dependent spectra of the backscattered light from the phantom. Shorter wavelengths decay faster than longer ones, due to the stronger absorption of the nanocages at shorter wavelengths. (Adapted with permission from Cang et al. [63]. Copyright © 2005 Optical American Society.)
The SOCT image of the phantom shown in Figure 25.8c displays a spatially resolved centroid wavelength. A red-shifted centroid is evident on the right portion of the SOCT image, demonstrating the spectroscopic contrast enhancement by the nanocages. The reason for the red shift arises when the nanocages with LSPR peak at 700 nm absorb the blue side of the broad light source (750–900 nm)—therefore allowing more of the light from the red side of the optical spectrum to be scattered back to the detector. Figure 25.8d clearly shows that the shorter wavelengths of the source spectrum are attenuated more than the longer wavelengths, producing a red shift of the spectrum centroid. The contrast enhancement (or red shift of the centroid) is stronger at deeper depths owing to the increased absorption at shorter wavelengths by the nanocages.
25.4.2 Photoacoustic Tomography PA imaging is an emerging hybrid modality of optical imaging and ultrasonic imaging that can provide high sensitivity, as well as enhanced imaging depth [62]. In PA imaging,
P1: OTA/XYZ P2: ABC c25 JWBS037-Chen August 29, 2010
10:53
Printer Name: Yet to Come
ENHANCING OPTICAL IMAGING CONTRAST WITH GOLD NANOCAGES
627
a short-pulsed laser beam is used to irradiate the tissue and ultrasound is measured to generate an image of its anatomical structures. Different structures in the tissue can absorb light to different extents and the absorbed energy will result in thermoelastic expansion. The resulting pressure rise relaxes with propagation of a wave which lies in the ultrasound frequency range. By analyzing the ultrasonic waves produced by the absorbing structures, it is possible to reconstruct the position, size, shape, and optical properties of the corresponding structures. When Au nanocages are employed in PAT, they provide enhanced contrast due to their strong absorption powers. In this section, we discuss how Au nanocages can be used as a contrast enhancing agent for PAT to image the blood vessels in the brain and as a new class of tracers to locate the sentinel lymph node (SLN) of a rat for breast cancer staging [64, 65]. To demonstrate nanocages as a contrast agent for PAT, the photoacoustic images of the cerebral cortex of a rat were taken, and we saw an 81% enhancement in the optical absorption by comparing the images before (Fig. 25.9a) and after (Fig. 25.9b)
FIGURE 25.9 Demonstration of PAT imaging using Au nanocages as a contrast agent. PAT images of a rat’s cerebral cortex (a) before and (b) after the final injection of nanocages (the highest enhancement point). (c) The differential PAT image of (a) and (b). (d) An open-skull photograph revealing the vasculature of the cerebral cortex. (Adapted with permission from Yang et al. [64]. Copyright © 2007 American Chemical Society.)
P1: OTA/XYZ P2: ABC c25 JWBS037-Chen August 29, 2010
628
10:53
Printer Name: Yet to Come
A MULTIFUNCTIONAL PLATFORM FOR MOLECULAR OPTICAL IMAGING
nanocages injection. Figure 25.9c is the difference of the before and after images, confirming the enhancement achieved with Au nanocages. Figure 25.9d is an open skull photograph of the rat’s cerebral cortex, verifying the vasculature observed in the PA images. This simple study clearly demonstrated that Au nanocages are useful as contrast agents for PAT. The SLN is speculated to be the first draining site of metastatic cancer; hence SLN biopsy has become a standard technique for axillary staging in breast cancer patients. Currently, in clinical facilities, there are two types of tracers utilized to locate the SLN: blue dyes (e.g., isosulfan or methylene blue) and radioactive colloids (e.g., technetium99) [66, 67]. However, both procedures have limitations. The first method relies on the direct visualization of the blue dye, thus requiring a surgical procedure to expose the region near the SLN [68]. The second method requires Geiger counters, which results in low spatial resolution for the noninvasive SLN identification and could expose the patient to damaging radiation. PA imaging provides a new technique to image the SLN with high resolution and large penetration depths, thus allowing for noninvasive biopsy techniques such as fine needle aspiration biopsy [69]. Here we examine the use of Au nanocages as a tracer to provide a high spatial resolution mapping of the SLN using PA imaging [65]. Gold nanocages were intradermally injected into the forepaw pad of a Sprague–Dawley rat. PA images of the axillary region of the rat were taken over a period of time. Figure 25.10a shows a photograph of the axillae with the hair removed, and Figure 25.10b corresponds to the same region with the skin and fatty tissue removed after PA images were recorded. Figure 25.10c is the PA image taken before the injection of Au nanocages, and Figure 25.10d–g shows the PA images taken 5, 59, 150, and 194 min, respectively, after the injection of the nanocages. Before injection of Au nanocages, we cannot discern the location of the SLN, but within 5 min after injection, the SLN was immediately visible. Over time, the nanocages accumulated in the SLN as shown in Figure 25.10h. The peak enhancement occurred around 140 minutes, which was significantly shorter than the 24 hours required for the radioactive colloids. The shorter accumulation time observed for Au nanocages is likely due to their smaller size: the nanocages are 50 nm in edge length while the radioactive colloids are 100–200 nm in diameter. Since the SLN in humans is located 12 ± 5 mm below the surface of the skin, we also demonstrated the deep penetration capability of using PAT to map the SLN with Au nanocages as a contrast agent. Figure 25.11a,b shows PA images of the axillary region before and after the injection of Au nanocages, respectively. Layers of chicken breast were added to mimic imaging through tissue at different depths. Figure 25.11c–e shows PA images of the SLN taken at different imaging depths of 10, 21, and 33 mm, respectively. We see that the SLN can still be discerned at 33 mm deep, much deeper than is necessary for this technique to be useful in humans. The B-scan image in Figure 25.11f shows a cross-sectional view and can serve as a confirmation that the SLN can be distinguished at 33 mm deep: the signal-to-noise ratio was quantified to be 7 dB. Figure 25.11g shows a plot of PA amplitude against depth, revealing the exponential decay of optical energy in tissue. As can be seen from these results, Au nanocages provided strong contrast enhancement in PAT for the noninvasive, in vivo imaging in rat models. Compared to the conventional SLN mapping methods, the nanocage-based PA method allowed for deeper imaging penetration due to the optical absorption in the near-infrared region (vs. visible region), rapid accumulation into lymphatic channels, and good spatial and depth resolution.
P1: OTA/XYZ P2: ABC c25 JWBS037-Chen August 29, 2010
10:53
Printer Name: Yet to Come
ENHANCING OPTICAL IMAGING CONTRAST WITH GOLD NANOCAGES
(a)
(b)
(c)
(d)
(e)
(f)
(g)
629
(h)
FIGURE 25.10 PA SLN time-course mapping images with Au nanocages in a rat model. (a) Photograph of axillary region with hair removed. (b) Photograph with skin and fatty tissue removed after the PA images had been recorded. The inset shows the lymphatic channel of the SLN stained blue from the nanocages. PA images acquired (c) before and after a certain time point following the injection of nanocages: (d) 5 min, (e) 59 min, (f) ∼140 min, and (g) 194 min. (h) The gradual accumulation of nanocages in the SLN over time measured in terms of the amplitude changes of PA signals. Peak accumulation occurred at ∼140 min after the injection. PA signals from the SLN were normalized by those from adjacent blood vessels (the dotted region in (e)) to minimize the ultrasonic focal effect. BV, blood vessels; SLN, sentinel lymph node. (Reproduced with permission from Song et al. [65]. Copyright © 2009 American Chemical Society.)
P1: OTA/XYZ P2: ABC c25 JWBS037-Chen August 29, 2010
630
10:53
Printer Name: Yet to Come
A MULTIFUNCTIONAL PLATFORM FOR MOLECULAR OPTICAL IMAGING
(a)
(b)
(c)
(d)
(e)
(f)
(g)
FIGURE 25.11 Deep penetration capability of PA SLN mapping with Au nanocages in a rat model. PA images (a) before and after the injection of nanocages at a certain imaging depth: (b) 0 mm, (c) 10 mm, (d) 21 mm, and (e) 33 mm. (f) PA B-scan showing the SLN can still be detected 33 mm deep. (g) The amplitude variations of PA signals over imaging depths. BV, blood vessel; SLN, sentinel lymph node. (Reproduced with permission from Song et al. [65]. Copyright © 2009 American Chemical Society.)
P1: OTA/XYZ P2: ABC c25 JWBS037-Chen August 29, 2010
10:53
Printer Name: Yet to Come
PHOTOTHERMAL TREATMENT WITH GOLD NANOCAGES
631
25.5 PHOTOTHERMAL TREATMENT WITH GOLD NANOCAGES 25.5.1 Qualitative Analysis Gold nanocages are also useful for photothermal therapy due to their large absorption cross sections. When nanocages are irradiated with light, the absorbed photons can be converted into phonons, or vibrations in the lattice. This energy will dissipate as heat into its environment. As an initial demonstration, Au nanocages were deposited on a carboncoated copper grid and then exposed to a camera flash [61]. The light exposure from the brief flash was enough to increase the temperature to melt the Au nanocages into solid spheres. Since this experiment was conducted in air, a poor conductor of heat, the heat could not dissipate through the surroundings quickly, thus resulting in the melting of the nanocages. However, in a biological environment, the heat should easily transfer to the surrounding area, which is primarily water, a good conductor of heat. For the photothermal treatment of cancer, the surface of the Au nanocages can be modified with different moieties to actively target cancerous cells. Consequently, the heat generated from the nanocages will transfer to the nearby cancer cells to induce cellular death. The surfaces of Au nanocages are functionalized in a two-step process, shown in Figure 25.12a [61]. In the first step, N-hydroxyl succinimidyl (NHS)-activated polyethylene glycol (PEG) attached to the Au nanocages by breaking its internal disulfide bond and forming a Au–S bond. In the second step, the primary amine of an antibody reacted with the NHS group of the PEGylated nanocage. This functionalized Au nanocage is referred to as an immuno Au nanocage for the rest of this chapter. SK-BR-3 is a well-characterized breast cancer cell line that overexpresses the human epidermal growth receptor 2 (HER2), also known as EGFR-2 and ErbB-2. This cell line was used to demonstrate the photothermal effect of immuno Au nanocages. To target this specific cell line, the Au nanocages were conjugated with monoclonal anti-HER2 antibodies. We found that the immuno Au nanocages effectively attached to the SK-BR-3 cells, whereas in the control experiment, bare Au nanocages did not attach to these cells [70]. After the SK-BR-3 cells had been targeted with immuno Au nanocages, they were irradiated for 5 min at a power density of 1.5 W/cm2 by a femtosecond laser with a central wavelength tuned to the near-infrared peak position of the nanocages [71]. The cells were then stained with calcein-AM and ethidium homodimer 1 (EthD-1) to qualitatively determine the cell viability. Calcein-AM is a colorless dye that fluoresces green upon the enzymatic interaction with live cells. EthD-1 fluoresces red when bonded to the DNA of dead cells. Figure 25.12b,c shows the calcein-AM and EthD-1 staining results, respectively. Both fluorescent images indicate a well-defined circular region of cellular death. The diameter of the cellular death matches the 2-mm diameter of the spot size of the laser, confirming the cell death was caused by and confined to the near-infrared irradiation. Figure 25.12d,e shows the calcium-AM and EthD-1 staining, respectively, for the control experiment without immuno Au nanocages. The cells in the control experiment maintain viability, indicating that the laser itself does not induce cellular death at the used power density.
25.5.2 Quantitative Analysis The photothermal effect of the immuno Au nanocages targeted to SK-BR-3 cells was then quantified using flow cytometry, a useful technique for counting, examining, and sorting
P1: OTA/XYZ P2: ABC c25 JWBS037-Chen August 29, 2010
632
10:53
Printer Name: Yet to Come
A MULTIFUNCTIONAL PLATFORM FOR MOLECULAR OPTICAL IMAGING
(a)
(b)
(c)
(d)
(e)
FIGURE 25.12 Qualitative analysis of the photothermal effect of immuno Au nanocages on SKBR-3 cells. (a) Schematic illustration of the two-step protocol used to functionalize antibodies to the surface of a nanocage. SK-BR-3 breast cancer cells targeted with the immuno Au nanocages were then irradiated by a Ti:Sapphire laser with a spot size of 2 mm in diameter and at a power density of 1.5 W/cm2 for 5 min. The sample showed a well-defined circular zone of dead cells as revealed by (b) calcein-AM assay (where green fluorescence indicates that the cells were alive) and (c) EthD-1 assay (where red fluorescence indicates that the cells were dead). In the control experiment, cells irradiated under the same conditions but without immuno Au nanocages maintained viability, as indicated by (d) calcein-AM fluorescence assay and (e) the lack of intracellular EthD-1 uptake. (Adapted with permission from Chen et al. [61, 71]. Copyright © 2005 and 2007, American Chemical Society.)
cells. We analyzed parameters such as the number of immobilized immuno Au nanocages per cell, the power density of the laser, and the duration of laser exposure [72]. Immuno Au nanocages were incubated with the SK-BR-3 cells and then the unbounded nanocages were removed by washing. The cells were then suspended in solution and analyzed by flow cytometry and inductively coupled mass spectroscopy (ICP-MS) to determine the number of cells in solution and the gold content, respectively. From these values, we calculated there were 400 ± 90 immuno Au nanocages immobilized per cell. Figure 25.13a
P1: OTA/XYZ P2: ABC c25 JWBS037-Chen August 29, 2010
10:53
Printer Name: Yet to Come
PHOTOTHERMAL TREATMENT WITH GOLD NANOCAGES
(a)
(b)
(c)
(d)
633
FIGURE 25.13 Quantitative analysis of the photothermal effect of immuno Au nanocages on SKBR-3 cells. (a) TEM image of a microtomed SK-BR-3 cell with immuno Au nanocages immobilized on the surface of the cell as well as internalized in the cell. (b) Illustration depicting the setup of the photothermal experiments. The laser spot size was 2 mm in diameter and the cellular growth vessel was 6.38 mm in diameter. Thus only 9.8% of the cells were exposed to the laser. Note: Cells are not drawn to scale. Plots of percentage of cellular damage against (c) laser power density for exposure time = 5 min and (d) time of laser exposure at power density = 4.8 W/cm2 . (Adapted with permission from Chen et al. [71]. Copyright © 2007 American Chemical Society.)
shows a microtomed sample of a SK-BR-3 cell with immuno Au nanocages immobilized on the surface of the cell as well as internalized in vesicles within the cell. Figure 25.13b shows an illustration of the photothermal setup. The cells were grown on a 96-well plate that has a diameter of 6.38 mm. The spot size of the laser was 2 mm in diameter, which means only 9.8% of the cells (not drawn to scale) were irradiated by the laser. After the cells had been incubated with immuno Au nanocages, they were irradiated at different laser parameters. The cells were then suspended in solution and subsequently stained with propidium iodide (PI), a common nuclear or chromosomal counterstain that could be used to differentiate between live and damaged cells since PI fluoresces 20–30fold stronger when bound to the DNA of damaged cells compared to unbound PI. We can determine the percentage of damaged cells by analyzing the forward scattering and PI fluorescent signal with flow cytometry. The effect of the power density on cell viability was studied by irradiating the SK-BR-3 cells for 5 min at different laser power densities. Figure 25.13c shows a plot of cellular damage against power density. We found that cells treated with immuno Au nanocages (•) required some power threshold before cellular death was observed. After that point, the cellular death increased linearly with increasing power density. At the highest power density, 6.4 W/cm2 , the cellular damage was 55%, which implied the heat generated transferred to
P1: OTA/XYZ P2: ABC c25 JWBS037-Chen August 29, 2010
634
10:53
Printer Name: Yet to Come
A MULTIFUNCTIONAL PLATFORM FOR MOLECULAR OPTICAL IMAGING
areas outside the spot size of the laser. In the control experiment, cells that were not treated with immuno Au nanocages but were exposed to the same laser (◦) maintained viability. In addition to power density, the effect of time of laser exposure on cell viability was investigated. The cells were irradiated at 4.8 W/cm2 with different laser exposure times. Figure 25.13d shows a plot of cellular damage against time of laser exposure. Cells treated with immuno Au nanocages (•) immediately responded to light exposure while untreated cells (◦) did not. Within 1 minute of laser exposure, the cellular damage was 18%. The percentage of cellular damage continued to increase for the first 5 min. After 5 min, the percentage cellular damage stabilized at 35%. This data suggested that although the nanocages responded immediately to the laser, some time was required for the heat to transfer outside the laser spot size.
25.6 CONCLUSION Gold nanocages provide a promising platform for biomedical research due to their unique optical properties, bioinertness, and facile surface chemistry. They can be routinely prepared via the galvanic replacement reaction between Ag nanocubes and an Au salt precursor. By controlling the size of Ag nanocubes, Au nanocages of different sizes can be produced routinely in high yields. The optical properties of these nanocages can also be tailored, allowing for the peak absorption to be precisely tuned well into the near-infrared region by simply titrating AuCl4 − or AuCl2 − into a solution of Ag nanocubes. Additionally, the selective removal of Ag from the walls of the nanocages with Fe(NO3 )3 produced ultrathin nanoframes with LSPR peaks further into the near-infrared region without compromising their size. Results from both OCT and PAT studies clearly establish that Au nanocages have great potential as contrast agents for optical bioimaging. The noninvasive, in vivo imaging with PAT demonstrates that Au nanocages are useful for enhancing contrast in blood vessels of the brain, as well as improving cancer staging by aiding in noninvasive SLN detection. By converting the incident light into heat, Au nanocages were demonstrated to effectively kill the cancer cells in vitro via the photothermal effect. In summary, Au nanocages can enhance optical contrast for a variety of optical imaging modalities, such as OCT, SOCT, and PAT, as well as provide therapeutic effect through photothermal destruction of cancer cells. By optimizing the reaction conditions, it is also possible to create Au nanocages with uniform pores, a promising structure for drug delivery applications.
REFERENCES 1. Xia, Y.; Halas, N. J. Shape-controlled synthesis and surface plasmonic properties of metallic nanostructures. MRS Bull. 2005, 30, 338–344. 2. Aiken, J. D.; Finke, R. G. A review of modern transition-metal nanoclusters: their synthesis, characterization, and applications in catalysis. J. Mol. Catal. A: Chem. 1999, 145, 1–44. 3. Murphy, C. J.; Sau, T. K.; Gole, A. M.; Orendorff, C. J.; Gao, J.; Gou, L.; Hunyadi, S. E.; Li, T. Anisotropic metal nanopoparticles: synthesis, assembly, and optical applications. J. Phys. Chem. B 2005, 109, 13857–13870. 4. Murray, C. B.; Kagan, C. R.; Bawendi, M. G. Synthesis and characterization of monodisperse nanocrystals and close-packed nanocrystal assemblies. Annu. Rev. Mater. Sci. 2000, 30, 545–610.
P1: OTA/XYZ P2: ABC c25 JWBS037-Chen August 29, 2010
10:53
Printer Name: Yet to Come
REFERENCES
635
5. Shipway, A. N.; Katz, E.; Willner, I. Nanoparticle arrays on surfaces for electronic, optical, and sensor applications. ChemPhysChem 2000, 1, 18–52. 6. Willets, K. A.; Van Duyne, R. P. Localized surface plasmon resonance spectroscopy and sensing. Annu. Rev. Phys. Chem. 2007, 58, 267–297. 7. Xia, Y.; Xiong, Y.; Lim, B.; Skrabalak, S. E. Shape-controlled synthesis of metal nanocrystals: Simple chemistry meets complex physics? Angew. Chem. Int. Ed. 2009, 48, 60–103. 8. El-Sayed, M. A. Some interesting properties of metals confined in time and nanometer space of different shapes. Acc. Chem. Res. 2001, 34, 257–264. 9. Hu, M.; Chen, J.; Li, Z. Y.; Au, L.; Hartland, G. V.; Li, X.; Marquez, M.; Xia, Y. Gold nanostructures: engineering their plasmonic properties for biomedical applications. Chem. Soc. Rev. 2006, 35, 1084–1094. 10. Noguez, C. Surface plasmons on metal nanoparticles: the influence of shape and physical environment. J. Phys. Chem. C 2007, 111, 3806–3819. 11. Wiley, B.; Sun, Y.; Xia, Y. Synthesis of silver nanostructures with controlled shapes and properties. Acc. Chem. Res. 2007, 40, 1067–1076. 12. Xia, Y.; Yang, P.; Sun, Y.; Wu, Y.; Mayers, B.; Gates, B.; Yin, Y.; Kim, F.; Yan, H. Onedimensional nanostructures: synthesis, characterization, and applications. Adv. Mater. 2003, 15, 353–389. 13. Jana, N. R.; Gearheart, L.; Murphy, C. J. Wet chemical synthesis of silver nanorods and nanowires of controllable aspect ratio. Chem. Commun. 2001, 617–618. 14. Jana, N. R.; Gearheart, L.; Murphy, C. J. Wet chemical synthesis of high aspect ratio cylindrical gold nanorods. J. Phys. Chem. B 2001, 105, 4065–4067. 15. Kim, F.; Song, J. H.; Yang, P. Photochemical synthesis of gold nanorods. J. Am. Chem. Soc. 2002, 124, 14316–14317. 16. Jackson, J. B.; Halas, N. J. Silver nanoshells: variations in morphologies and optical properties. J. Phys. Chem. B 2001, 105, 2743–2746. 17. Chen, S. H.; Carrol, D. L. Synthesis and characterization of truncated triangular silver nanoplates. Nano Lett. 2002, 2, 1003–1007. 18. Shao, Y.; Jin, Y.; Dong, S. Synthesis of gold nanoplates by aspartate reduction of gold chloride. Chem. Commun. 2004, 1104–1105. 19. Sun, Y.; Xia, Y. Triangular nanoplates of silver: synthesis, characterization, and use as sacrificial templates for generating triangular nanorings of gold. Adv. Mater. 2003, 15, 695– 699. 20. Huang, C.; Wang, Y.; Chiu, P.; Shih, M.; Meen, T. Electrochemical synthesis of gold nanocubes. Mater. Lett. 2006, 60, 1896–1900. 21. Sun, Y.; Xia, Y. Shape-controlled synthesis of gold and silver nanoparticles. Science 2002, 298, 2176–2179. 22. Xiong, Y.; Chen, J.; Wiley, B.; Xia, Y.; Yin, Y.; Li, Z. Y. Size-dependence of surface plasmon resonance and oxidation for Pd nanocubes synthesized via a seed etching process. Nano Lett. 2005, 5, 1237–1242. 23. Wiley, B.; Xiong, Y.; Li, Z. Y.; Yin, Y.; Xia, Y. Right bipyramids of silver: a new shape derived from single twinned seeds. Nano Lett. 2006, 6, 765–768. 24. Wiley, B. J.; Chen, Y.; McLellan, J. M.; Xiong, Y.; Li, Z. Y.; Ginger, D.; Xia, Y. Synthesis and optical properties of silver nanobars and nanorice. Nano Lett. 2007, 7, 1032– 1036. 25. Zettsu, N.; McLellan, J. M.; Wiley, B.; Yin, Y.; Li, Z. Y.; Xia, Y. Synthesis, stability, and surface plasmonic properties of rhodium multipods, and their use as substrates for surface-enhanced Raman scattering. Angew. Chem. Int. Ed. 2006, 45, 1288–1292.
P1: OTA/XYZ P2: ABC c25 JWBS037-Chen August 29, 2010
636
10:53
Printer Name: Yet to Come
A MULTIFUNCTIONAL PLATFORM FOR MOLECULAR OPTICAL IMAGING
26. Lu, X.; Chen, J.; Skrabalak, S. E.; Xia, Y. Galvanic replacement reaction: a simple and powerful route to hollow and porous metal nanostructures. Proc. IMechE., Part N: JNanoeng. Nanosyst. 2007, 221, 1–16. 27. Skrabalak, S. E.; Au, L.; Lu, X.; Li, X.; Xia, Y. Gold nanocages for cancer detection and treatment. Nanomedicine 2007, 2(5), 657–668. 28. Skrabalak, S. E.; Chen, J.; Au, L.; Lu, X.; Li, X.; Xia, Y. Gold nanocages for biomedical applications. Adv. Mater. 2007, 19, 3177–3184. 29. Skrabalak, S. E.; Chen, J.; Sun, Y.; Lu, X.; Au, L.; Cobley, C. M.; Xia, Y. Gold nanocages: synthesis, properties, and applications. Acc. Chem. Res. 2008, 41, 1587–1595. 30. Fievet, F.; Lagier, J. P.; Figlarz, M. Preparing monodisperse metal powders in micrometer and submicrometer sizes by the polyol process. MRS Bull. 1989, 14, 29–34. 31. Skrabalak, S. E.; Wiley, B. J.; Kim, M.; Formo, E. V.; Xia, Y. On the polyol synthesis of silver nanostructures: glycolaldehyde as a reducing agent. Nano Lett. 2008, 8, 2077–2081. 32. Sun, Y.; Mayers, B.; Herricks, T.; Xia, Y. Polyol synthesis of uniform silver nanowires: a plausible growth mechanism and the supporting evidence. Nano Lett. 2003, 3, 955–960. 33. Wiley, B.; Herricks, T.; Sun, Y.; Xia, Y. Polyol synthesis of silver nanoparticles: use of chloride and oxygen to promote the formation of single-crystal, truncated cubes and tetrahedrons. Nano Lett. 2004, 4, 1733–1739. 34. Siekkinen, A. R.; McLellan, J. M.; Chen, J.; Xia, Y. Rapid synthesis of small silver nanocubes by mediating polyol reduction with a trace amount of sodium sulfide or sodium hydrosulfide. Chem. Phys. Lett. 2006, 432, 491–496. 35. Skrabalak, S. E.; Au, L.; Li, X.; Xia, Y. Facile synthesis of Ag nanocubes and Au nanocages. Nat. Protoc. 2007, 2, 2182–2190. 36. Chen, J.; Wiley, B.; McLellan, J. M.; Xiong, Y.; Li, Z. Y.; Xia, Y. Optical properties of Pd-Ag and Pt-Ag nanoboxes synthesized via galvanic replacement reactions. Nano Lett. 2005, 5, 2058– 2062. 37. Batzill, M.; Koel, B. E. Silver of Pt(100)—room temperature growth and high temperature alloying. Surf. Sci. 2004, 553, 50–60. 38. Sun, Y.; Xia, Y. Mechanistic study on the replacement reaction between silver nanostructures and chloroauric acid in aqueous medium. J. Am. Chem. Soc. 2004, 126, 3892–3901. 39. Wang, Z. L. Transmission electron microscopy of shape-controlled nanocrystals and their assemblies. J. Phys. Chem. B 2000, 104, 1153–1175. 40. Sieradzki, K. Curvature effects in alloy dissolution. J. Electrochem. Soc. 1993, 740, 2868–2872. 41. Chen, J.; McLellan, J.M.; Siekkinen A.; Xiong, Y.; Li, Z.-Y.; Xia, Y. Facile synthesis of gold–silver nanocages with controllable pores on the surface. J. Am. Chem. Soc. 2006, 128, 14776– 14777. 42. Au, L.; Chen, Y.; Zhou, F.; Camargo, P. H. C.; Lim, B.; Li, Z.-Y.; Ginger, D. S.; Xia, Y. Synthesis and optical properties of cubic gold nanoframes. Nano. Res. 2008, 1, 441–449. 43. Au, L.; Lu, X.; Xia, Y. A comparative study of galvanic replacement reactions involving Ag nanocubes and AuCl2− or AuCl4− . Adv. Mater. 2008, 20, 2517–2522. 44. Qian, L.; Yang, X. Preparation and characterization of Ag(Au) bimetallic core–shell nanoparticles with new seed growth method. Colloids Surf. A 2005, 260, 79–85. 45. Choi, Y. U.; Lee, E. C.; Han, K. N. The dissolution behavior of metals from Ag/Cu and Ag/Au alloys in acidic and cyanide solution. Metall. Trans. B 1991, 22, 755–764. 46. Dursun, A.; Pugh, D. V.; Corcoran, S. G. Dealloying of Ag-Au alloys in halide-containing electrolytes. J. Electrochem. Soc. 2003, 150, B355–B360. 47. Martinez, L.; Segarra, M.; Fernandez, M.; Espiell, F. Kinetics of the dissolution of pure silver and silver–gold alloys in nitric acid solution Metall. Trans. B 1993, 24, 827–837.
P1: OTA/XYZ P2: ABC c25 JWBS037-Chen August 29, 2010
10:53
Printer Name: Yet to Come
REFERENCES
637
48. Hunyadi, S. E.; Murphy, C. J. Bimetallic silver-gold nanowires: fabrication and use in surfaceenhanced Raman scattering. J. Mater. Chem. 2006, 16, 3929–3935 49. Xia, Y.; Kim, E.; Whitesides, G. M. Microcontact printing of alkanethiols on silver and its application in microfabrication. J. Electrochem. Soc. 1996, 143, 1070–1079. 50. Lu, X.; Au, L.; McLellan, J.; Li, Z. Y.; Marquez, M.; Xia, Y. Fabrication of cubic nanocages and nanoframes by dealloying Au/Ag alloy nanoboxes with an aqueous etchant based on Fe(NO3 )3 or NH4 OH. Nano Lett. 2007, 7, 1764–1769. 51. Erlenbacher, J.; Aziz, M. J.; Karma, A.; Dimitrov, N.; Sieradzki, K. Evolution of nanoporosity in dealloying. Nature 2001, 410, 450–453. 52. Kreibig, U.; Vollmer, M. Optical Properties of Metal Clusters. Springer: New York, 1995. 53. Fuchs, R. Theory of the optical properties of ionic crystal cubes. Phys. Rev. B 1975, 11, 1732–1740. 54. Haes, A. J.; Haynes, C. L.; McFarland, A. D.; Schatz, G. C.; Van Duyne, R. P.; Zhou, S. Plasmonic materials for surface-enhanced sensing and spectroscopy. MRS Bull. 2005, 30, 368– 375. 55. Kottmann, J. P.; Martin, O. J. F.; Smith, D. R.; Schultz, S. Plasmon resonances of silver nanowires with a nonregular cross section. Phys. Rev. B 2001, 64, 235402. 56. Aizpurua, J.; Garnett, W. B.; Lee, J. R.; Abajo, F. J. G. d.; Brian, K. K.; Mallouk, T. Optical properties of coupled metallic nanorods for field-enhanced spectroscopy. Phys. Rev. B 2005, 71, 235420. 57. Teperik, T. V.; Popov, V. V.; Abajo, F. J. G. d. Radiative decay of plasmons in a metallic nanoshell. Phys. Rev. B 2004, 69, 155402. 58. Mie, G. Beitrage zur optik truber medien, speziell kolloidaler metallosungen. Ann. Phys. 1908, 25, 377–445. 59. Wiley, B.; Im, S. H.; Li, Z. Y.; McLellan, J. M.; Siekkinen, A.; Xia, Y. Maneuvering the surface plasmon resonance of silver nanostructures through shape-controlled synthesis. J. Phys. Chem. B 2006, 110, 15666–15675. 60. Draine, B. T.; Flatau, P. J. Discrete dipole approximation for scattering calculations. J. Opt. Soc. Am. A 1994, 11, 1491–1499. 61. Chen, J.; Wiley, B.; Li, Z. Y.; Cang, H.; Campbell, D.; Saeki, F.; Au, L.; Lee, J.; Li, X.; Xia, Y. Gold nanocages: engineering their structure for biomedical applications. Adv. Mater. 2005, 17, 2255–2261. 62. Wang, L. V.; Wu, H.-I. Biomedical Optics: Principles and Imaging. John Wiley & Sons: Hoboken, NJ, 2007. 63. Cang, H.; Sun, T.; Chen, J.; Wiley, B. J.; Xia, Y.; Li, X. Gold nanocages as contrast agents for spectroscopic optical coherence tomography. Opt. Lett. 2005, 30, 3048–3050. 64. Yang, X.; Skrabalak, S. E.; Li, Z. Y.; Xia, Y.; Wang, L. V. Photoacoustic tomography of a rat cerebral cortex in vivo with Au nanocages as an optical contrast agent. Nano Lett. 2007, 7, 3798– 3802. 65. Song, K. H.; Kim, C.; Cobley, C. M.; Xia, Y.; Wang, L. V. Near-infrared gold nanocages as a new class of tracers for photoacoustic sentinel lymph node mapping on a rat model. Nano Lett. 2009, 9, 183–188. 66. Krag, D. N.; Weaver, D. L.; Alex, J. C.; Fairbank, J. T. Surgical resection and radiolocalization of the sentinel lymph node in breast cancer using a gamma probe. Surg. Oncol. 1993, 2, 335– 339. 67. Morton, D. L.; Wen, D. R.; Wong, J. H.; Economou, J. S.; Cagle, L. A.; Storm, F. K.; Foshag, L. J.; Cochran, A. J. Technical details of intraoperative lymphatic mapping for early stage melanoma. Arch. Surg. 1992, 127, 392–399.
P1: OTA/XYZ P2: ABC c25 JWBS037-Chen August 29, 2010
638
10:53
Printer Name: Yet to Come
A MULTIFUNCTIONAL PLATFORM FOR MOLECULAR OPTICAL IMAGING
68. Nakajima, M.; Takeda, M.; Kobayashi, M.; Suzuki, S.; Ohuchi, N. Nano-sized fluorescent particles as new tracers for sentinel node detection: experimental model for decision of appropriate size and wavelength. Cancer Sci. 2005, 96, 353–356. 69. Oyen, R. H.; Van Poppel, H. P.; Ameye, F. E.; Van de Voorde, W. A.; Baert, A. L.; Baert, L. V. Lymph node staging of localized prostatic carcinoma with CT and CT-guided fine-needle aspiration biopsy: prospective study of 285 patients. Radiology 1994, 190, 315–322. 70. Chen, J.; Saeki, F.; Wiley, B.; Cang, H.; Cobb, M. J.; Li, Z. Y.; Au, L.; Zhang, H.; Kimmey, M. B.; Li, X.; Xia, Y. Gold nanocages: bioconjugation and their potential use as optical imaging contrast agents. Nano Lett. 2005, 5, 473–477. 71. Chen, J.; Wang, D.; Xi, J.; Au, L.; Siekkinen, A.; Warsen, A.; Li, Z. Y.; Zhang, H.; Xia, Y.; Li, X. Immuno gold nanocages with tailored optical properties for targeted photothermal destruction of cancer cells. Nano Lett. 2007, 7, 1318–1322. 72. Au, L.; Zheng, D.; Zhou, F.; Li, Z.-Y.; Li, X.; Xia, Y. A quantitative study on the photothermal effect of immuno gold nanocages targeted to breast cancer cells. ACS Nano 2008, 2, 1645–1652.
P1: OTA/XYZ P2: ABC c26 JWBS037-Chen January 25, 2011
13:33
Printer Name: Yet to Come
CHAPTER 26
Theranostic Applications of Gold Nanoparticles in Cancer PARMESWARAN DIAGARADJANE, PRANSHU MOHINDRA, and SUNIL KRISHNAN Department of Radiation Oncology, University of Texas M. D. Anderson Cancer Center, Houston, Texas, USA
26.1 INTRODUCTION Recent advances in nanotechnology, a multidisciplinary field that has interfaces with chemistry, engineering, and biology, have resulted in the manufacture of a plethora of nanoparticles with different sizes, shapes, core physicochemical properties, and surface modifications that are being investigated for potential medical applications. Nanoparticles typically range in size from a few nanometers to a few hundred nanometers, considerably smaller than human cells, thereby permitting extravasation of smaller nanoparticles out of blood vessels and facilitating interactions with biomolecules at the cellular and molecular level. The most common nanoparticles studied for biomedical applications are liposomes and unior multilamellar vesicles (organic biolipid layers encapsulating imaging and therapeutic payloads), dendrimers (repeatedly branched polymers), quantum dots (metallic core–shell nanoparticles that are intensely fluorescent at specific wavelengths), gold nanoparticles (ranging in shape from spheres and shells to rods and cages), paramagnetic nanoparticles (iron oxide laden particles), and carbon nanotubes. In the arena of cancer research, there has been an explosion of knowledge and research at the genomic and proteomic level that has unraveled clues to the origins and underlying biological underpinnings of cancer. This, in turn, has created opportunities to selectively target critical phenomena unique to the milieu harboring cancer. The focus of this chapter is on gold nanoparticles and their specific applications to cancer diagnosis and treatment (theranostics). A number of features of gold nanoparticles have rendered them particularly attractive to biomedical researchers and account for their popularity in preclinical research leading up to potential clinical translation. The most striking feature is the familiarity of the medical community with gold as a clinically useful therapeutic agent. Indeed, gold has been used in the medical field since 2500 BC, for various ailments such as melancholy, fainting, fevers, syphilis, and arthritis [1, 2]. The most prominent use of gold has been for the treatment of rheumatic arthritis. Since its initial clinical application in the early 20th century, gold Nanoplatform-Based Molecular Imaging Edited by Xiaoyuan Chen C 2011 John Wiley & Sons, Inc. Copyright
639
P1: OTA/XYZ P2: ABC c26 JWBS037-Chen January 25, 2011
640
13:33
Printer Name: Yet to Come
THERANOSTIC APPLICATIONS OF GOLD NANOPARTICLES IN CANCER
formulations have been used to treat a variety of rheumatic diseases including psoriatic arthritis, juvenile arthritis, palindromic rheumatism, and discoid lupus erythematosus [3] and various inflammatory skin disorders such as pemphigus, urticaria, and psoriasis [4]. Treatment of rheumatoid arthritis with a cumulative dose of a little less than 2 g/yr for 10 years without any appreciable toxicity speaks to the good overall tolerance of gold in humans [5]. In addition to its apparent clinical safety and tolerability, gold nanoparticles are relatively easy to synthesize using simple techniques and reagents that are not prohibitively expensive. Furthermore, owing to its physical inertness, gold will not directly interact chemically with biomolecules in humans. To facilitate biological interactions, gold nanoparticles can readily be functionalized via conjugation to biomolecules including peptides, antibodies, radioisotopes, and oligonucleotides. Lastly, the path to clinical translation of some formulations of gold nanoparticles involves their being treated as devices rather than drugs, a feature that could reduce time and expense incurred in transitioning from the bench to the bedside. In spite of the extensive reports on the biocompatibility, safety, and efficacy of gold for the treatment of various medical ailments, its use in the treatment of cancer was not explored until the late 20th century [6]. This chapter focuses on properties of gold nanoparticles, their evolving imaging and therapeutic applications, and methods to quantify and model uptake and distribution of gold nanoparticles within cancers.
26.2 OVERVIEW OF GOLD NANOPARTICLES IN CANCER IMAGING AND THERAPY This section provides an overview of the unique properties of gold nanoparticles that render them capable of theranostic applications, the unique properties of cancers that can be exploited for imaging and therapy, and the practical considerations in clinical investigations involving nanoparticles for cancer imaging and therapy.
26.2.1 Unique Properties of Gold Nanoparticles As with all nanoparticles made of bulk materials, gold nanoparticles also possess properties unique to nanoscale structures—a high surface area to volume ratio wherein the properties of the nanoparticle are largely driven by the properties of the relatively abundant atoms on the surface of the nanoparticle rather than those constituting the bulk of the nanoparticle. The abundance of atoms on the surface of the nanoparticle also increases the contact cross section of the particle for exposure to and interaction with biomolecules at a cellular or molecular level. Furthermore, at the nanoscale fundamental characteristics of a bulk material (i.e., its color, strength, melting point, electric/magnetic/thermal properties) can be altered and fine-tuned without changing the material’s chemical composition. In fact, certain geometries (rods and shells) of gold nanoparticles permit tuning of the optical resonance wavelengths of their surface plasmons to the near-infrared (NIR) region of the spectrum where they absorb light intensely and convert this to heat. From the standpoint of clinical translation of this form of noninvasive heating for cancer therapy, the activation by NIR illumination is highly pertinent because at these wavelengths light penetrates deep within tissue (up to several centimeters).
P1: OTA/XYZ P2: ABC c26 JWBS037-Chen January 25, 2011
13:33
Printer Name: Yet to Come
OVERVIEW OF GOLD NANOPARTICLES IN CANCER IMAGING AND THERAPY
641
26.2.2 Unique Properties of Cancers Tumor or site-specific delivery of a theranostic agent with minimal systemic side effects remains the top priority in real-time clinical applications. The most effective strategy for tumor-specific delivery is to exploit the unique anatomic and pathophysiologic properties of tumors that differentiate them from normal tissues. This has generally been accomplished either via passive accumulation of theranostic agents within tumors or via active targeting of tumor-specific molecules. Passive yet selective sequestration of a theranostic agent within tumors capitalizes on a phenomenon termed enhanced permeability and retention (EPR) effect. Herein, the inherent difference between tumor vasculature and normal tissue vasculature is exploited for greater accumulation and retention of theranostic agents. Typically, when the tumor cells multiply and reach a size of 2–3 mm they outstrip their blood supply—the ever-increasing nutrition and oxygen demands of the growing tumor initiates a process called angiogenesis, where newer channels are created to nourish these cells [7]. In contrast to regular blood vessels in normal tissues, these new vessels are immature, incompletely formed, leaky, and chaotically organized [8, 9]. Characteristically, this tumor neovasculature contains wide interendothelial junctions, abundant transendothelial channels, incomplete or absent basement membranes, and dysfunctional lymphatics. Consequently, a higher proportion of systemically injected macromolecules, nanoparticles, and lipid particles readily extravasate from the bloodstream into the interstitium in tumors rather than in normal tissues. The resulting differential accumulation of nanoparticles in tumor and normal tissues accounts for a higher diagnostic or therapeutic index when passively targeted nanoparticles are used in cancer theranosis. As a broad generalization, nanoparticles smaller than 400 nm in size extravasate into the tumor interstitium without diffusing back into the bloodstream, thereby accumulating progressively in tumor parenchyma over time [10–12]. Apart from passive targeting via the EPR effect, nanoparticles can also be actively targeted to tumors. Functionalizing nanoparticles by conjugating them to tumor-specific docking moieties such as peptides, antibodies, and oligonucleotides allows them to interact specifically with their counterpart biomolecules preferentially overexpressed within tumors. Typical examples of this approach include the conjugation of gold nanoparticles to anti-EGFR (epidermal growth factor receptor) antibody [13, 14], folic acid [15], tumor necrosis factor (TNF) [16] to specifically target the receptors overexpressed in tumors.
26.2.3 Technical Considerations for Clinical Development As with all systemically administered agents (pharmaceutical drugs, imaging contrast agents, drug-delivery systems) in clinical use, key attributes that are intricately linked to their design and deployment are their biodistribution, pharmacokinetics, and target specificity. In the case of gold nanoparticles, their biodistribution is heavily skewed by their robust uptake and clearance by the reticuloendothelial system (RES, primarily comprised of macrophages in the liver, spleen, and lymph nodes). Although very small nanoparticles (< 5.5 nm) evade uptake by the RES and are cleared by the kidney, most metallic nanoparticles accumulate within the liver and spleen [17]. In an effort to create a stealth particle that evades opsonization (process by which molecules such as antibodies and complement bind to surfaces and enhance phagocytosis by macrophages) and RES capture, metallic nanoparticles are routinely decorated with polyethylene glycol (PEG), thereby increasing their circulatory half-life and increasing the probability of their accumulation within and interactions with tumors. This increases the signal-to-noise ratio for tumor imaging probes
P1: OTA/XYZ P2: ABC c26 JWBS037-Chen January 25, 2011
642
13:33
Printer Name: Yet to Come
THERANOSTIC APPLICATIONS OF GOLD NANOPARTICLES IN CANCER
and increases the therapeutic ratio of anticancer treatments. However, it is worth noting that this coating (PEGylation) not only influences the biodistribution of nanoparticles but also influences the pharmacokinetics. Therefore a repeat kinetic assessment is often required after PEGylation to determine the optimal timing of imaging or therapy based on timing of maximal uptake and retention within tumors. Furthermore, without comprehensive pharmacokinetic assessment, it is not easy to discern the difference between the initial rapid nonspecific leakage of nanoparticles into the tumor due to leaky vasculature alone and tumor-specific accumulation of nanoparticles actively targeted to tumors [18]. Lastly, a key component of clinical development of nanoparticles is to determine their toxicity. Unfortunately, this has not been systematically investigated with the rigor and thoroughness one typically encounters with the preclinical evaluation and characterization of pharmaceutical drugs. In addition to the scarcity of such studies, interpretation of the existing data is also complicated by the fact that parameters such as particle purity, heterogeneity of size and structure, imperfections of surface contours, variations in surface decoration/chemistry/charge, and extent of agglomeration contribute significantly to the biodistribution and bioactivity of nanoparticles, which, in turn, impacts assessment of toxicity. Nevertheless, it is evident that gold nanoparticles of shapes and sizes encountered in biological models cross membrane barriers and the larger ones often accumulate within the liver and RES. Whether this confers any increased short-term or long-term risk to patients remains unknown. Also unknown is whether sites of normal vascular breakdown such as within areas of inflammation and/or repair of injury (healing wounds, tissue reorganization following infarction, diabetic retinopathy) are more prone to accumulation of gold nanoparticles and whether this, in turn, poses long-term toxicity concerns. Lastly, in the case of elongated gold nanorods that can resemble asbestos fibers, there is the potential fear that mesothelioma-like lesions may arise if the nanorods encounter the lining membranes of organs. Even though such encounters are more likely with chronic inhalation of these particles (bringing them in proximity to the pleura surrounding the lungs) or intraperitoneal washing with these particles, a recent study demonstrating the development of mesothelioma-like lesions when peritoneal cavities of mice were exposed to multiwalled carbon nanotobes (which tend to be more filamentous and more like asbestos than nanorods) raises this concern. Recognizing the need for systematic toxicity assessment in anticipation of regulatory approval of nanoparticles for clinical use, the National Cancer Institute has created the Nanotechnology Characterization Laboratory that offers characterization of the physical attributes, in vitro biological properties, and in vivo compatibility of nanoparticles.
26.3 THERAPEUTIC APPLICATIONS OF GOLD NANOPARTICLES Although many biodiagnostic applications of gold nanoparticles have been developed since the 1970s [19–23], their use in cancer therapy is a relatively recent development [24]. Outlined below are a few applications of particular interest to cancer biologists and oncologists. 26.3.1 Plasmonic Photothermal Therapy (PPTT) As noted previously, generation of heat via plasmon resonance upon laser illumination is a unique property of certain configurations of gold nanoparticles. This heat can be utilized for photothermal therapy (PTT) of tumors, similar to other ablative methods of
P1: OTA/XYZ P2: ABC c26 JWBS037-Chen January 25, 2011
13:33
Printer Name: Yet to Come
THERAPEUTIC APPLICATIONS OF GOLD NANOPARTICLES
643
inducing tumor coagulation and necrosis. In contrast to radiofrequency ablation and other invasive ablative procedures, PTT using a suitable laser/light source is noninvasive in nature and relatively inexpensive. In the past, PTT was most often achieved using exogenously administered organic dyes such as indocyanine green, naphthalocyanines, and porphyrins coated with transition metals. Upon illumination by a suitable laser/light source, excitation of the molecular electrons results in an increase in their kinetic energy, which, in turn, is converted into heat [25–29]. Similarly, excitation of endogenous chromophores such as hemoglobin and melanin has been used for clinical PTT of vascular malformations such as port-wine stains [30, 31]. However, the efficiency of converting light to heat is very low and rapid photobleaching significantly dampens therapeutic efficacy. The recognition that gold nanostructures are activatable by illumination to generate heat has led to a resurgence of interest in PTT for cancer therapy. Gold nanoparticles tend to be less prone to chemical/ thermal denaturation [28, 29, 32, 33] than organic dyes. In addition, the absorption cross sections of gold are in the range of 10−14 m2 , about a million times more than that of conventional organic dyes. Not only do gold nanoparticles accumulate preferentially in tumors and generate heat efficiently upon illumination but also the heat generated is efficiently transferred to the surrounding tumor due to the excellent metallic conductance of gold. To facilitate maximum penetration of the therapeutic illuminating beam, the surface plasmon resonance of the nanoparticle is tuned to absorb energy in the NIR region where the optical absorption of tissue is minimal and penetration is maximal (the so-called therapeutic window in the electromagnetic spectrum) [34]. Since the generation of heat is due to the surface plasmon resonance of gold, this form of PTT is referred to as plasmonic photothermal therapy (PPTT) [35–37]. In the case of gold nanoshells that were originally synthesized with a dielectric silica core and a thin outer lining (shell) of colloidal gold, this tunability is achieved by altering the core-to-shell ratio [38]. In the case of gold nanorods, this tunability is accomplished by changing the aspect ratio (ratio of length to diameter). The seminal report on the use of gold nanoshells for NIR thermal therapy in mouse tumor models [38] has opened up avenues for several investigations from the clinical perspective. Photothermal ablation was reported on cultured human breast cancer cells incubated with gold nanoshells followed by NIR laser illumination (820 nm, 35 W/cm2 ). Intratumoral injection of gold nanoshells in mouse tumor models, followed by low-dose NIR illumination (820 nm, 4 W/cm2 , 5-mm spot diameter) resulted in a substantial rise in temperature (T = 37.4 ± 6.6 ◦ C), which, in turn, induced irreversible tissue damage within 4–6 min. Controls treated without nanoshells demonstrated significantly lower temperature elevations (T = < 9.1 ± 4.7 ◦ C) on exposure to NIR light (Fig. 26.1). Another interesting aspect of this report was the use of real-time magnetic resonance thermal imaging (MRTI) for online temperature measurement during the laser illumination. MRTI changes corresponded to thermal damage patterns noted on conventional histological evaluation, highlighting its potential utility in real-time feedback control of thermal therapy, both to ensure adequate tumor effect and to minimize collateral normal tissue damage. With the optimized regime of nanoshell and laser dosage, survival studies on a murine colorectal cancer model revealed complete thermoablation of tumors and tumor-free survival of mice for >90 days after treatment [39]. In a subsequent study aimed at clinical translation of this concept, Stern et al. [40] administered gold nanoshells systemically through the tail vein of mice bearing human prostate cancer xenografts prior to photothermal ablation. Using two doses of gold nanoshells (7 L/g and 8.5 L/g of body weight) that accumulated passively in tumors by the EPR effect, they demonstrated that NIR laser illumination for 3 minutes resulted
P1: OTA/XYZ P2: ABC c26 JWBS037-Chen January 25, 2011
644
13:33
Printer Name: Yet to Come
THERANOSTIC APPLICATIONS OF GOLD NANOPARTICLES IN CANCER
(a) 35
35
(b) 35
30
30
35 30
30
25
25
∆T (°C)
25
20 15
20
10
25
15
20
10
15
15 10
6 min
10
5
3 min 1 min
5
0 min
0
0
20
0
1
4 5 2 3 Depth from skin (mm)
6
6 min 3 min 1 min 0 min
0
1
2 3 4 5 Depth from skin (mm)
6
(c)
1 cm FIGURE 26.1 Temperature profiles of (a) nanoshell injected and (b) control tumors as a function of depth, measured using magnetic resonance thermal imaging. (c) Gross pathology after the treatment reveals severe hemorrhage and loss of birefringence in the treatment zone (far left), and the corresponding silver staining (middle) and H&E staining (far right) demonstrating localized necrotic areas induced by the nanoshell-mediated thermal ablation. (Adapted with permission from Hirsch et al. [38].)
in greater therapeutic efficacy with the higher concentration of gold nanoshells (93% tumor necrosis and regression with an average temperature rise to 65.4 ◦ C). In similar experiments using an orthotopic canine model, the feasibility of using optical fiber-based NIR illumination for thermoablation of intracranial tumors was demonstrated [41]. Gold nanoshells were delivered systemically by intravenous infusion and allowed to passively accumulate (∼24 h) in the intracranial tumors and a 3.5-W average, 3-minute laser dose at 808 nm selectively elevated the temperature of tumors to 65.8 ± 4.1 ◦ C. Identical laser doses applied to normal white and grey matter on the contralateral side of the brain yielded sublethal temperatures of 48.6 ± 1.1◦ C. The postmortem histopathology of the treated brain sections confirmed the selective and localized thermal ablation induced by the gold nanoshell thermal therapy. 26.3.2 Radiation Dose Enhancement Clinical radiation therapy relies on ionizing radiation beams that interact physically with tissues in its path to deposit varying doses of its incident energy within different tissues.
P1: OTA/XYZ P2: ABC c26 JWBS037-Chen January 25, 2011
13:33
Printer Name: Yet to Come
THERAPEUTIC APPLICATIONS OF GOLD NANOPARTICLES
645
For kilovolt radiation energies (typically 100 kV to 1 MV), the predominant phenomenon mediating energy deposition is photoelectric absorption, wherein the incident photon transfers its entire energy to the inner shell electron of an atom. Depending on the absorbed energy, the electron is either excited to a higher electronic orbit level or knocked out of the atom. However, for megavolt radiation energies (typically 1–10 MV), the predominant interaction within tissues is Compton scattering, wherein the incident photon knocks out an electron from the outermost electronic orbit of an atom. Whereas the photoelectric effect is highly dependent on the atomic number (Z) of tissues (dose deposited is directly proportional to Z 3-4 ), Compton scattering is independent of Z. Since the electron density of tissues has a significant influence on dose deposition, one method to modify dose to tumors is to artificially alter their electron density such that they are distinctly different from the surrounding normal tissues. Indeed, if sufficient quantities of high-Z material are introduced into tumors, considerable radiation dose enhancement is achievable, particularly immediately adjacent to the scattering surface of the high-Z material. Several studies have demonstrated the concept of dose escalation in cell monolayers in the presence of high-Z materials like iodine, barium, and gadolinium. However, the clinical utility of this effect will only be realized if adequate concentrations of high-Z materials can be delivered to and maintained within solid tumors at the time of radiation therapy. With gold being biologically inert, arguably nontoxic, relatively affordable, and capable of preferentially accumulating within tumors, it was natural to consider radiation dose enhancement via intratumoral gold nanoparticle incorporation. Initial in vitro and in vivo studies were performed using commercially available gold particles 1.5–3.0 m in diameter [42]. A dose enhancement factor of 1.42 was measured for 1% gold particle solutions irradiated with 200-kVp X-rays. When rodent and human cells were irradiated in the presence of 1% gold nanoparticles, the dose enhancement factor averaged 1.43 (range 1.36–1.54). Intratumoral delivery of gold particles resulted in a modest radiosensitization but extremely heterogeneous distribution. Subsequent in vivo experiments were performed with much smaller gold nanoparticles measuring 1.9 nm in diameter. A single intravenous injection of gold nanoparticles (2.7 g Au/kg) in tumor-bearing mice resulted in ∼7 mg Au/g accumulation in tumor tissues with tumor-to-normal tissue gold concentration ratio of ∼8:1 during the X-ray radiation. A significant dose enhancement was also reported with a lower concentration (1.35 g Au/kg) of gold nanoparticles [43]. The 1-year survival of mice treated with gold nanoparticles alone (no radiation) was 0%, whereas irradiated mice had 1-year survivals of 86%, 50%, and 20%, respectively, for those treated with the higher concentration of gold nanoparticles, the lower concentration of gold nanoparticles, and no nanoparticles. To estimate the range of radiation dose enhancement that is potentially achievable in clinical settings using typical radiation energies, these in vivo tumor concentrations were used in Monte Carlo calculations using 140-kVp X-rays, 4- and 6-MV photon beams, and 192 Ir gamma rays [44]. This mathematical modeling confirmed that, using 140-kVp X-rays, a dose enhancement factor of 2 was achievable within tumors with a gold concentration of 7 mg Au/g tumor. Under similar conditions a dose enhancement factor of 0.01–0.07 was achievable with 4- and 6-MV photon beams and 0.30 with 192 Ir gamma rays. While these macroscopic dose enhancement factor estimates assume a uniform distribution of nanoparticles within tumors and do not account for relative position of cells (and their DNA—for biological consequences, the final target of electrons expelled from their orbits)
P1: OTA/XYZ P2: ABC c26 JWBS037-Chen January 25, 2011
646
13:33
Printer Name: Yet to Come
THERANOSTIC APPLICATIONS OF GOLD NANOPARTICLES IN CANCER
in correlation to nanoparticles, sophisticated event-by-event Monte Carlo codes modeling individual electron tracks may provide greater prediction of dose at a nano- or microscale and greater clinical and biological correlation. Nevertheless, theoretical estimation does confirm that biologically meaningful dose enhancement is achievable with low-energy X-rays and 192 Ir gamma rays used in clinical practice.
26.3.3 Hyperthermic Radiosensitization As noted above, gold nanoparticles can be employed to improve the efficacy of clinical radiation therapy via tumor-specific dose enhancement. In addition to such physical dose enhancement, biological dose enhancement may be achieved by targeting key processes involved in intrinsic resistance of cancer cells to ionizing radiation [45]. One such mechanism is intratumoral hypoxia (inadequate oxygenation) secondary to disorderly and immature blood vessels found within tumors [8, 9]. One mechanism of targeting hypoxia is mild temperature hyperthermia. Unfortunately, despite multiple randomized trials having demonstrated improved response rates and survival rates when patients with locally advanced malignancies are treated with locoregional hyperthermia and radiotherapy compared to radiotherapy alone, hyperthermic radiosensitization is underutilized in routine clinical practice. This is largely attributable to the invasive means of achieving and maintaining hyperthermia and the lack of good thermal dosimetry [46]. Gold nanoparticle mediated hyperthermia is therefore a simple and elegant solution to this problem. Herein, hyperthermia can be achieved noninvasively using NIR illumination of tumors that have passively accumulated gold nanoshells and MRTI can be used to monitor hyperthermia noninvasively. In the initial preclinical demonstration of this concept, ∼8 × 108 nanoshells/g body weight of mice was systemically injected 24 hours before generating mild hyperthermia levels (∼41 ◦ C) for ∼20 minutes using an 808-nm diode laser. Direct thermocouple measurements of temperature were confirmed using real-time noninvasive MRTI (Fig. 26.2a). A single 10-Gy dose of radiation (125-kV X-rays) immediately after the hyperthermia resulted in approximately two-fold increase (Fig. 26.2b) in tumor growth delay (time taken for tumors to double in volume compared to untreated mice) when compared to the animals treated with radiation alone [47]. Furthermore, this sensitization of tumors to radiation was demonstrated to be due to both an early increase in blood flow to the core of tumors (imaged by dynamic contrast enhanced magnetic resonance imaging) immediately after hyperthermia and a subsequent disruption of vasculature creating large swaths of necrosis within the tumor (visualized on hematoxylin/eosin-stained slides and immunofluorescence staining for vascular number and architecture) (Fig. 26.2c). Furthermore, this focal vascular disruption was attributable to vasculature-centered focal temperature elevations due to perivascular sequestration of gold nanoshells that are too large to diffuse freely into tumor interstitium but large enough to leak out of tumor vasculature. These heterogeneous temperature profiles of tumors with focal hotspots of temperature elevation near tumor vasculature created by perivascularly sequestered nanoshells and the gradual fall-off in temperature away from these areas are a unique feature of nanoshell-mediated hyperthermia. Conceivably, with the use of gold nanorods that have higher concentrations of gold than nanoshells, are small enough to penetrate deeper into tumors, and capable of greater intratumoral gold content increases, both physical radiation dose enhancement and hyperthermic radiosensitization can be integrated to improve the therapeutic index of radiation therapy.
P1: OTA/XYZ P2: ABC c26 JWBS037-Chen January 25, 2011
13:33
Printer Name: Yet to Come
THERAPEUTIC APPLICATIONS OF GOLD NANOPARTICLES
647
FIGURE 26.2 (a) In vivo magnetic resonance thermal images of tumor tissues during near-infrared laser illumination. The color bar represents the increase in temperature over baseline (∼30 ◦ C). (b) Normalized tumor volume plot of control, hyperthermia, radiation, and thermoradiotherapy groups showing the mean ± SE values at different time periods after each treatment. Time to tumor doubling was significantly longer in the thermoradiotherapy group than in the radiation alone group (p < 0.005). (c) Immunofluorescence staining of hypoxic cells (pimonidazole, green) and vascular perfusion (Hoechst 33342, blue) demonstrates a structured architectural pattern with regions of perfusion separated by areas of hypoxia in the radiation group (upper panel, left). In contrast, the thermoradiotherapy group demonstrates a distorted architectural pattern with patchy hypoxic regions with minimal vascular perfusion (upper panel, right). Taken together with the corresponding hematoxylin and eosin stained images of tumors from these groups demonstrating extensive necrosis in the thermoradiotherapy group (lower panel) and electron microscopic images of gold nanoshell sequestration in the perivascular space, this is suggestive of vascular disruption in the combined treatment group. (Adapted with permission from Diagaradjane, et al. [47].)
26.3.4 Active Targeting of Gold Nanoparticles to Tumors The therapeutic efficacy of gold nanoparticles can further be enhanced by active targeting of these nanoparticles to cell surface receptors that are specifically overexpressed in tumors when compared to surrounding normal tissues. Active targeting of gold nanoparticles significantly increases their retention in the tumor interstitium, thereby widening the tumor-to-normal tissue ratio [48]. Early bioconjugation procedures involved citrate capping of gold nanospheres to bind antibodies by electrostatic interactions. More recently, several bioconjugation strategies such as poly(styrenesulfonate) bridging layers, chemical binding
P1: OTA/XYZ P2: ABC c26 JWBS037-Chen January 25, 2011
648
13:33
Printer Name: Yet to Come
THERANOSTIC APPLICATIONS OF GOLD NANOPARTICLES IN CANCER
of functional groups of antibodies to the metal surface, and use of antibody-PEG linker complexes to bind to the gold surface have been adopted to increase the strength of the conjugation [24, 49]. Interestingly, these conjugation protocols do not significantly affect the inherent optical characteristics of the gold nanoparticles. Sokolov et al. [13] reported that 12-nm gold nanoparticles conjugated to an anti-EGFR antibody resulted in a small red shift of ∼6 nm in the surface plasmon resonance with a 10% reduction in transmission. The estimation of 5 × 104 conjugates per cell correlated with the reported 2 × 104 to 2 × 105 EGFR receptors per cell and demonstrated the selectivity of these actively targeted gold nanoparticles [50]. Subsequently, Loo et al. [24] conjugated gold nanoshells to herceptin (an antibody targeting EGFR-2) to demonstrate that immunotargeted nanoshells can provide scattering contrast for imaging while also exhibiting sufficient absorption to enable effective photothermal therapy in breast cancer cells. Folic acid conjugated gold nanoparticles have been shown to have higher selectivity for cancer cells with increased folic acid requirements [39, 51–53]. Paciotti et al. [16] have used colloidal gold nanoparticles (∼33 nm) as a vehicle to deliver TNF selectively to tumor. El-Sayed et al. [14] evaluated the ability of anti-EGFR antibody conjugated gold nanoparticles (∼35 nm) to discriminate between cancerous and noncancerous oral epithelial cell lines. The anti-EGFR antibody conjugates demonstrated a specific and homogeneous binding to the surface of the cancerous cells with 600% greater affinity than to the noncancerous cells, suggesting that these conjugates can be used in molecular biosensor techniques for the diagnosis of oral epithelial cancer cells [14]. Furthermore, anti-EGFR antibody conjugated nanorods bound to the human oral cancer cells were found to exhibit sharp, polarized, and enhanced Raman spectrum. The enhancement in the Raman spectrum observed due to the specificity of the conjugated nanorods for the cancer cells could potentially be used as diagnostic signatures for cancer cells. Although several studies have reported on the preclinical benefit of conjugated gold nanoparticles in cancer therapy, such bioconjugates have also been used in nononcologic applications. Laser illumination of antibody-conjugated gold nanorods targeting the extracellular parasite Toxoplasma gondii has provided a new paradigm for the treatment of toxoplasmosis. Another interesting application of actively targeted nanoparticles is the use of peptide conjugated nanoparticles (∼10 nm) to activate macrophages. Bone marrow derived macrophages recognize the gold nanoparticles that are conjugated to biologically relevant peptides (viz., amyloid growth inhibitory peptide and sweet arrow peptide), while they do not recognize peptides or nanoparticles alone [56]. The above-mentioned studies using gold nanoparticles of different compositions, sizes, and shapes for active targeting of cancers have demonstrated the feasibility and promise of customized imaging and therapy at a cellular level. Extensive studies are in progress to demonstrate the feasibility of using actively targeted gold nanoparticles in vivo.
26.4 ESTIMATION OF GOLD ACCUMULATION The estimation of gold accumulation in tumors after passive or active delivery is critical in optimizing treatment efficacy and for clinical translation. Currently, there are two methods to achieve this—neutron activation analysis and inductively coupled plasma mass spectrometry—both of which require the tissue of interest to be harvested and processed to estimate the gold content. An alternative and relatively new noninvasive method is diffuse optical spectroscopy. This section describes each of these techniques and the tissue preparation processes for the estimation of the gold content.
P1: OTA/XYZ P2: ABC c26 JWBS037-Chen January 25, 2011
13:33
Printer Name: Yet to Come
ESTIMATION OF GOLD ACCUMULATION
649
26.4.1 Neutron Activation Analysis (NAA) NAA is the “gold standard” analytical technique used for the qualitative and quantitative multielement analysis of elements in samples. The sensitivity of this technique, on the order of parts per billion or better, is superior to that attainable by other methods. NAA is significantly different from other spectroscopic analytical techniques in that it is based not on electronic transitions but on nuclear transitions. When a neutron interacts with the target nucleus via a nonelastic collision, a compound nucleus is formed in the excited state, which almost instantaneously decays to a more stable configuration through emission of one or more characteristic prompt gamma rays. In many cases, this new configuration yields a radioactive nucleus, which also decays by emission of one or more characteristic delayed gamma rays, but at a much slower rate according to the unique half-life of the radioactive nucleus. Typically, elemental analysis involves delayed gamma-ray NAA, where the measurements are made following radioactive decay rather than during radiation [57]. The basic essentials required to carry out an analysis of samples by NAA are a source of neutrons, instrumentation suitable for detecting gamma rays, and knowledge of the reactions that occur when neutrons interact with target nuclei. Typically, freshly collected tissue samples are lyophilized and weighed into precleaned polyethylene irradiation vials. In order to minimize the solid angle uncertainties in irradiation and counting positions, the physical size of the control and samples are maintained the same. The sample and control vials are placed into a suitable irradiation facility and bombarded with neutrons (n · cm−2 · s−1 ) and the gamma-ray spectroscopy is carried out on all the irradiated materials after a delay of 4–8 days to allow for decay of matrix activity. The amount of gold in tissues can be calibrated using measured standard values [58]. 26.4.2 Inductively Coupled Plasma Mass Spectrometry (ICP-MS) ICP-MS is a relatively new and powerful tool that is used to detect trace or ultratrace amounts of elements with a high degree of specificity and sensitivity. The heart of ICP-MS is the inductively coupled plasma ion source, which is operated at temperatures of 7000 K. A plasma or gas consisting of ions, electrons, and neutral particles, formed from argon gas, is utilized to atomize and ionize the elements in the sample matrix. Due to the extremely high temperature, all molecules in the sample are broken into their component atoms. The generated ions are identified using their mass-to-charge ratio by passing the generated ions into a high vacuum analyzer through a series of apertures. The intensity of the peak in the mass spectrum is proportional to the amount of the elemental isotope from the original sample. Several research studies have reported on the use of ICP-MS to measure the amount of gold present in tissues [59, 60]. More recently, this technique has been used to detect the homing of gold nanorods into tumors [61]. In general, the samples for ICP-MS analysis are frozen, lyophilized, and dissolved in aqua regia (nitric acid + 37% hydrochloric acid) for several hours to dissolve gold particles. The dissolved gold particles are suitably diluted with nitric acid and analyzed using ICP-MS to estimate the gold content. 26.4.3 Diffuse Optical Spectroscopy Although NAA and ICP-MS techniques are used for the accurate measurement of gold in extracted tissues ex vivo, a technique for noninvasive real-time measurement of gold nanoparticle content in the tissue is often needed to guide the treatment strategy in clinical
P1: OTA/XYZ P2: ABC c26 JWBS037-Chen January 25, 2011
650
13:33
Printer Name: Yet to Come
THERANOSTIC APPLICATIONS OF GOLD NANOPARTICLES IN CANCER
settings. Diffuse optical spectroscopy (DOS) offers such an option for quantifying gold content in tissues by building on its prior use for in vivo quantification of the hemoglobin concentration and blood oxygen saturation, the amount of scattering, water content, and melanin content in tissues [23, 62]. In vivo DOS, or point spectroscopy, uses fiberoptic probes coupled to both a source and a spectrometer to transmit light of a desired wavelength range to the tissue surface. As the photons enter the tissue, a portion of the light is absorbed by the tissue chromophores, some of it is reflected (or scattered) from the surface, and some of it passes through the tissue unperturbed (transmitted). During its path in the tissue, some photons will be scattered by the different structures within the tissues and return to the tissue surface and emerge for detection. The reflection that is caused on the tissue surface is specular reflection and the one that is caused from within the tissue structure is called diffuse reflection. Since only a portion of the diffuse reflected photons are returned to the detector, the intensity of the emerging light is attenuated. The changes in the intensity of the reflected light over a spectral range provide information about the relative changes of specific structures within the tissue. When compared to the visible region (450–700 nm), light in the NIR region (700–1000 nm) penetrates deeper into the tissues, due to the lack of light-absorbing chromophores in this region. The major and important tissue chromophore that absorbs light in the NIR region is hemoglobin. Hemoglobin exists both in an oxygenated state (HbO2 ) and a deoxygenated state (Hb), with each of these forms exhibiting its own typical characteristic absorption spectrum. These unique characteristics are used to determine the tissue oxygenation (HbO2 ) and blood volume (as total hemoglobin, tHb = HbO2 + Hb) using DOS. Since the gold nanoparticles exhibit a distinct reflectance spectral signature in the NIR region, DOS can be used to measure the concentration of gold nanoparticles in tissues in vivo. Zaman et al. [63] first reported such in vivo measurements in a mouse tumor model using DOS, and compared their findings with standard NAA analysis. DOS could measure in vivo gold concentrations in living mice with an accuracy of within 12% [63]. These results represent a technique capable of real-time in vivo measurement of gold nanoparticles in tissues, which opens up the possibility of monitoring the dynamics of nanoparticle accumulation in tissues.
26.5 IMAGING OF SPATIAL DISTRIBUTION OF INTRATUMORAL GOLD ACCUMULATION Advances in optical imaging have contributed significantly to the field of biology and medicine at both microscopic and macroscopic levels. Macroscopic imaging techniques are widely being used to evaluate the gross differences in a relatively large tissue area to discriminate normal from diseased conditions. Endoscopy-based imaging techniques such as high-magnification endoscopy (confocal endoscopy and endocytoscopy), narrowband imaging (NBI), autofluorescence, and optical computed tomography have shown promise in early clinical applications such as identification of Barrett’s esophagus, esophageal neoplasia, early malignant bronchial lesions, and colorectal polyps [64]. These macroscopic and microscopic techniques predominantly utilize the differences in the native optical characteristics that arise due to the variation in vascular architecture, cellular morphology, and the alterations in native chromophores, to differentiate normal from diseased tissues. The image contrasts can be further enhanced by introducing tumor- or site-specific exogenous contrast agents, which in most cases are fluorescent molecules. The effective utilization
P1: OTA/XYZ P2: ABC c26 JWBS037-Chen January 25, 2011
13:33
Printer Name: Yet to Come
IMAGING OF SPATIAL DISTRIBUTION OF INTRATUMORAL GOLD ACCUMULATION
651
of these imaging techniques to image or identify therapeutic agents like gold nanoshells or nanorods in the tumor tissues could potentially provide valuable information about the distribution of these particles, which in turn can be used to model biological effects, explain biological outcomes observed, and/or refine imaging/treatment techniques. In this section, the principles of macroscopic narrowband imaging and microscopic two-photon confocal microscopy techniques are described in the context of their utilization for imaging the distribution of gold nanoshells or nanorods in tumor tissues. 26.5.1 Narrowband Imaging (NBI) NBI is an optical filter technology that involves the illumination and collection of a narrow wavelength band of light from the target tissue. In one technique of NBI (we call this method A), white light from a broadband light source is passed through a collimated lens assembly (CLA) and a bandpass filter (BPF) to generate narrow discrete wavelength bands (blue and green), which are used individually to illuminate tissues, and reflected light from the tissues is passed through the CLA and collected using a monochromatic detector, usually a charge coupled device (CCD). The collected monochromatic images are integrated to form a contrast enhanced composite narrowband image. Alternatively, a broadband white light source is used to illuminate the tissues and the reflected light is collected by a CCD through filters of narrow wavelength bands corresponding to the chromophores of interest (method B). In both these techniques the narrowband wavelength selections are usually dependent on the target of interest, which in most cases using biological tissue samples is hemoglobin that has absorption maxima at 413 nm and 580 nm. The reflected signal at either of these wavelengths is compared with the reflected signal from a narrow wavelength band at which the absorption due to hemoglobin is less. For example, the image pattern acquired using the blue and green illumination method (method A) will capitalize on the difference in the absorption characteristics of blood at both these wavelength bands of interest. Also, as the light penetration in tissues is directly dependent on the wavelength, with longer wavelengths permitting deeper light penetration, the image from the blue narrowband will provide information about the superficial capillary networks, while the green narrowband image will provide details of the deeper subepithelial blood vessels. The combination of these two monochromatic images offers a contrast enhanced image of the tissue surface with fine details about the vascular distribution. Recently, Puvanakrishnan et al. [65] have used the second imaging method (method B) to identify the pattern of gold nanoshell distribution in the tumor tissues extracted from mice 24 hours after the injection of gold nanoshells. Two narrowband wavelengths in the visible (570 nm) and NIR (810 nm) regions corresponding to the hemoglobin and gold nanoshell absorption were optimized for good contrast enhanced narrowband images. The visible images of the tumor tissues from control and gold nanoshell-injected mice demonstrated similar contrast characteristics indicating similar vascular distribution in these mice. However, the NIR images from the mouse injected with the gold nanoshells demonstrated not only enhanced contrast when compared to the control mouse but also a markedly nonhomogeneous dispersion of nanoparticles within the tumor [65]. The enhanced visualization of blood vessels and gold nanoshells using a macroscopic imaging technique like NBI might provide useful information to understand the extent of tumor angiogenesis and the geographical heterogeneity of nanoparticle distribution within biological systems. Given the high photothermal efficiency of gold nanoshells, NBI followed by NIR irradiation may be used as a combined imaging and photothermal therapy platform for both identifying and
P1: OTA/XYZ P2: ABC c26 JWBS037-Chen January 25, 2011
652
13:33
Printer Name: Yet to Come
THERANOSTIC APPLICATIONS OF GOLD NANOPARTICLES IN CANCER
ablating intact tumors and residual disease at resection margins after surgery. Indeed, NBI is particularly suited for incorporation into the surgical arena because it is noninvasive, uses simple portable instrumentation with video-rate imaging capability, and is a noncontact technique compatible with use in a sterile operating field. 26.5.2 Two-Photon Induced Photoluminescence (TPIP) Imaging The concept of two-photon excitation was predicted theoretically by Maria G¨oppert-Mayer [66]. In 1931 when she stated that photons of lesser energy together can cause an excitation that is normally produced by the absorption of a single photon of higher energy. The probability of the near-simultaneous absorption of two photons is extremely low. Therefore a high flux of excitation photons is typically required, usually a femtosecond laser. Since its first demonstration by Webb and colleagues [67], two-photon microscopy has been extensively explored for various imaging applications and it has evolved as a technique of choice for fluorescence microscopy in thick tissues and in live animals. The concept is based on the idea that two or more photons of low energy can excite a fluorophore, resulting in the emission of a fluorescence photon, typically at a higher energy than either of the two excitation photons. TPIP is a serial process involving sequential absorption of photons and emission from the recombination of electrons in the spband and holes in the dbands [68]. Although two-photon excitation is widely used to image fluorescence markers in tissues, TPIP has also been described for metal surfaces including those of gold nanoparticles. In general, the two-photon luminescence signal from tissues is relatively weak but can be amplified by several orders of magnitude when produced from roughened metal substrates due to the resonant coupling with localized surface plasmons. Under plasmon resonance conditions, the quantum efficiency of single photon luminescence from gold nanorods has been demonstrated to be enhanced by a factor of 1 million [69]. Wang et al. [70] further evaluated single GNRs using two-photon excitation laser scanning microscopy and determined that their TPIP signal brightness was about 60 times higher than that of single molecules of rhodamine. Lastly, in vivo imaging of BALB/c mice ears demonstrated feasibility of imaging TPIP signals of flowing nanorods, which are significantly different from the autofluorescence signals [70]. Using anti-EGFR antibody conjugated gold nanorods, Durr et al. [71] demonstrated that the TPIP intensity from gold-nanorod-labeled cancer cells was three orders of magnitude brighter than the two-photon autofluorescence (TPAF) emission intensity from unlabeled cancer cells. Furthermore, TPIP imaging of nanorod-labeled cells required 64 times less power than TPAF imaging of unlabeled cells in order to achieve the similar collected intensity. Given the quadratic dependence of emission intensity on the incident power, this observation implies that, for equal excitation powers, TPIP imaging of nanorod-labeled cancer cells can generate more than 4000 times larger emission signals than TPAF imaging of unlabeled cells [71]. The brightness of TPIP from gold nanoshells, gold nanorods, and fluorescent beads was also characterized. Nanoshells were noted to be as bright as nanorods and 140 times brighter than fluorescent beads [72]. The potential of gold nanoshells for multiphoton microscopy in bulk tissue was demonstrated in bulk tumor tissues represented by a mouse tumor extracted 24 hours after intravenous injection of gold nanoshells (∼12 × 108 ) and immediately after fluorescein injection. TPIP imaging exhibited an intense photoluminescence signal from the tumor tissues, whereas the control animals (injected with saline) did not show any significant photoluminescence signal. Three-dimensional volume visualization demonstrated the ability of TPIP imaging of nanoshells up to a
P1: OTA/XYZ P2: ABC c26 JWBS037-Chen January 25, 2011
13:33
Printer Name: Yet to Come
CONCLUSION
653
depth of 130 m in bulk tumor. The distribution of gold nanoshells in the tumor tissues was evaluated by simultaneous imaging of gold nanoshells and the tumor vasculature labeled with fluorescein-conjugated dextran. Using appropriate bandpass filters, prominent nanoshell luminescence coregistered with the fluorescein fluorescence was recorded. The three-dimensional volumetric visualization of TPIP images demonstrated that passive accumulation of gold nanoshells in tumors predictably results in close proximity of nanoshells to tumor blood vessels. With these promising results, the next step toward potential clinical translation is the development of miniaturized imaging tools, including compact microscopes for handheld imaging, endoscopes for insertion into hollow tissue cavities, and microendoscopes for minimally invasive real-time clinical imaging in solid tumors. For instance, Bao et al. [73] have developed a fast handheld two-photon fiber optic fluorescence endoscope with a field of view of 475 × 475 m and depth perception of 250 m with the lateral and axial resolution of 1 and 14.5 m, respectively [73]. In theory, this 3D in vivo cellular imaging endoscope capable of revealing 3D surface and subsurface histological characteristics could obviate the need for a biopsy. Similarly, recent advances in optical technologies such as gradient index (GRIN) lenses offer further opportunities to miniaturize imaging instrumentation [74]. GRIN lenses are rod-shaped lenses with a concentration gradient of doping atoms such as thallium, lithium, or silver such that their refractive indices decrease continuously along the radial coordinate. The main advantages of the GRIN lenses over conventional lenses are the small diameters down to 0.2 mm (maximum 2 mm) and the flat surfaces that allow easy assembly of complex optical systems. A more recent report on the development of a miniaturized two-photon fluorescence endoscope using GRIN lenses demonstrated that this technique is highly flexible and controllable in terms of time acquisition, resolution, and magnification and yet capable of discerning and resolving microscopic cellular details [75].
26.6 CONCLUSION The promise of nanotechnology in translating the knowledge about cancer from the postgenomics era into tangible clinical diagnostic and therapeutic gains requires continued interdisciplinary collaboration between teams of biomedical/optical/manufacturing engineers, chemists, biologists, and clinicians. In particular, novel nanoparticle constructs and newer techniques for their targeted delivery specifically to tumors are likely to be wedded to improvements in technologies that facilitate noninvasive visualization, quantification, and activation of these nanoparticles for improved imaging and therapy of cancers. Among this new arsenal of nanoparticles, gold nanoparticles are likely to figure prominently because of their biological inertness, ease of manufacture and bioconjugation, presumed lack of toxicity, relative affordability, and the familiarity with use of systemically administered gold in clinical medicine. However, while gold nanoparticles have some inherent logistical and potential clinical advantages, their primary disadvantage is their inability to be readily imaged in deep tissues. Consequently, their greatest value is likely to be in surface imaging/therapy applications (oral mucosa, skin, intraoperative fields) and endoscopic applications (esophageal, head and neck, colorectal, bronchial, cervical lesions). While the optimal theranostic anticancer agent is a multifunctional nanoparticle capable of simultaneously sensing, imaging, and treating tumors, such an ideal nanoparticle remains to be fabricated, optimized, and systematically evaluated. Nevertheless, in recent years there has
P1: OTA/XYZ P2: ABC c26 JWBS037-Chen January 25, 2011
654
13:33
Printer Name: Yet to Come
THERANOSTIC APPLICATIONS OF GOLD NANOPARTICLES IN CANCER
been a figurative “gold rush” in the identification of newer applications of more sensitive, flexible, and multifunctional gold nanoparticles for thernostic applications in cancer. This enthusiasm, the promising early outcomes, and the continued crosstalk between multiple disciplines are likely to herald a new era of nanotechnology applications in oncology.
REFERENCES 1. Higby, G. J. Gold in medicine: a review of its use in the West before 1900. Gold Bull. 1982, 15(4), 130–140. 2. Parish, R. V. Biologically-active gold(III) complexes. Met. Based Drugs 1999, 6(4-5), 271– 276. 3. Champion, G. D.; Graham, G. G.; et al. The gold complexes. Baillieres Clin. Rheumatol. 1990, 4(3), 491–534. 4. Thomas, R. E.; Papandrea, R. A. Treatment of psoriasis with topical auranofin. Med. J. Aust. 1993, 158(10), 720. 5. Abrams, M. J.; Murrer, B. A. Metal compounds in therapy and diagnosis. Science 1993, 261(5122), 725–730. 6. Regulla, D. F.; Hieber, L. B.; et al. Physical and biological interface dose effects in tissue due to X-ray-induced release of secondary radiation from metallic gold surfaces. Radiat. Res. 1998, 150(1), 92–100. 7. Folkman, J. Angiogenesis in cancer, vascular, rheumatoid and other disease. Nat. Med. 1995, 1(1), 27–31. 8. Baluk, P.; Morikawa, S.; et al. Abnormalities of basement membrane on blood vessels and endothelial sprouts in tumors. Am. J. Pathol. 2003, 163(5), 1801–1815. 9. McDonald, D. M.; Choyke, P. L. Imaging of angiogenesis: from microscope to clinic. Nat. Med. 2003, 9(6), 713–725. 10. Matsumura, Y.; Maeda, H. A new concept for macromolecular therapeutics in cancer chemotherapy: mechanism of tumoritropic accumulation of proteins and the antitumor agent smancs. Cancer Res. 1986, 46(12 Pt 1), 6387–6392. 11. Ishida, O.; Maruyama, K.; et al. Size-dependent extravasation and interstitial localization of polyethyleneglycol liposomes in solid tumor-bearing mice. Int. J. Pharm. 1999, 190(1), 49–56. 12. Maeda, H. The enhanced permeability and retention (EPR) effect in tumor vasculature: the key role of tumor-selective macromolecular drug targeting. Adv. Enzyme Regul. 2001, 41, 189– 207. 13. Sokolov, K.; Follen, M.; et al. Real-time vital optical imaging of precancer using anti-epidermal growth factor receptor antibodies conjugated to gold nanoparticles. Cancer Res. 2003, 63(9), 1999–2004. 14. El-Sayed, I. H.; Huang, X.; et al. Surface plasmon resonance scattering and absorption of antiEGFR antibody conjugated gold nanoparticles in cancer diagnostics: applications in oral cancer. Nano Lett. 2005, 5(5), 829–834. 15. Dixit, V.; Van den Bossche, J.; et al. Synthesis and grafting of thioctic acid-PEG-folate conjugates onto Au nanoparticles for selective targeting of folate receptor-positive tumor cells. Bioconjug. Chem. 2006, 17(3), 603–609. 16. Paciotti, G. F.; Myer, L.; et al. Colloidal gold: a novel nanoparticle vector for tumor directed drug delivery. Drug Deliv. 2004, 11(3), 169–183. 17. Choi, H. S.; Liu, W.; et al. Renal clearance of quantum dots. Nat. Biotechnol. 2007, 25(10), 1165–1170.
P1: OTA/XYZ P2: ABC c26 JWBS037-Chen January 25, 2011
13:33
Printer Name: Yet to Come
REFERENCES
655
18. Diagaradjane, P.; Orenstein-Cardona, J. M.; et al. Imaging epidermal growth factor receptor expression in vivo: pharmacokinetic and biodistribution characterization of a bioconjugated quantum dot nanoprobe. Clin. Cancer Res. 2008, 14(3), 731–741. 19. Elghanian, R.; Storhoff, J. J.; et al. Selective colorimetric detection of polynucleotides based on the distance-dependent optical properties of gold nanoparticles. Science 1997, 277(5329), 1078–1081. 20. Bendayan, M. Tech.Sight. Worth its weight in gold. Science 2001, 291(5507), 1363–1365. 21. Hirsch, L. R.; Jackson, J. B.; et al. A whole blood immunoassay using gold nanoshells. Anal. Chem. 2003, 75(10), 2377–2381. 22. Nam, J. M.; Thaxton, C. S.; et al. Nanoparticle-based bio-bar codes for the ultrasensitive detection of proteins. Science 2003, 301(5641), 1884–1886. 23. Zhao, X.; Hilliard, L. R.; et al. A rapid bioassay for single bacterial cell quantitation using bioconjugated nanoparticles. Proc. Natl. Acad. Sci. U.S.A. 2004, 101(42), 15027–15032. 24. Loo, C.; Lowery, A.; et al. Immunotargeted nanoshells for integrated cancer imaging and therapy. Nano Lett. 2005, 5(4), 709–711. 25. Anderson, R. R.; Parrish, J. A. Microvasculature can be selectively damaged using dye lasers: a basic theory and experimental evidence in human skin. Lasers Surg. Med. 1981, 1(3), 263–276. 26. Anderson, R. R.; Parrish, J. A. Selective photothermolysis: precise microsurgery by selective absorption of pulsed radiation. Science 1983, 220(4596), 524–527. 27. Jori, G.; Spikes, J. D. Photothermal sensitizers: possible use in tumor therapy. J. Photochem. Photobiol. B. 1990, 6(1-2), 93–101. 28. Chen, W. R.; Adams, R. L.; et al. Chromophore-enhanced in vivo tumor cell destruction using an 808-nm diode laser. Cancer Lett. 1995, 94(2), 125–131. 29. Chen, W. R.; Adams, R. L.; et al. Chromophore-enhanced laser–tumor tissue photothermal interaction using an 808-nm diode laser. Cancer Lett. 1995, 88(1), 15–19. 30. Anvari, B.; Milner, T. E.; et al. Selective cooling of biological tissues: application for thermally mediated therapeutic procedures. Phys. Med. Biol. 1995, 40(2), 241–252. 31. Nelson, J. S.; Milner, T. E.; et al. Dynamic epidermal cooling during pulsed laser treatment of port-wine stain. A new methodology with preliminary clinical evaluation. Arch. Dermatol. 1995, 131(6), 695–700. 32. Landsman, M. L.; Kwant, G.; et al. Light-absorbing properties, stability, and spectral stabilization of indocyanine green. J. Appl. Physiol. 1976, 40(4), 575–583. 33. Chen, C.; Parangi, S.; et al. A strategy to discover circulating angiogenesis inhibitors generated by human tumors. Cancer Res. 1995, 55(19), 4230–4233. 34. Weissleder, R. A clearer vision for in vivo imaging. Nat. Biotechnol. 2001, 19(4), 316–317. 35. Pitsillides, C. M.; Joe, E. K.; et al. Selective cell targeting with light-absorbing microparticles and nanoparticles. Biophys. J. 2003, 84(6), 4023–4032. 36. Zharov, V. P. Far-field photothermal microscopy beyond the diffraction limit. Opt. Lett. 2003, 28(15), 1314–1316. 37. Zharov, V. P.; Galitovskaya, E. N.; et al. Synergistic enhancement of selective nanophotothermolysis with gold nanoclusters: potential for cancer therapy. Lasers Surg. Med. 2005, 37(3), 219–226. 38. Hirsch, L. R.; Stafford, R. J.; et al. Nanoshell-mediated near-infrared thermal therapy of tumors under magnetic resonance guidance. Proc. Natl. Acad. Sci. U.S.A. 2003, 100(23), 13549–13554. 39. O’Neal, D. P.; Hirsch, L. R.; et al. Photo-thermal tumor ablation in mice using near infraredabsorbing nanoparticles. Cancer Lett. 2004, 209(2), 171–176. 40. Stern, J. M.; Stanfield, J.; et al. Selective prostate cancer thermal ablation with laser activated gold nanoshells. J. Urol. 2008, 179(2), 748–753.
P1: OTA/XYZ P2: ABC c26 JWBS037-Chen January 25, 2011
656
13:33
Printer Name: Yet to Come
THERANOSTIC APPLICATIONS OF GOLD NANOPARTICLES IN CANCER
41. Schwartz, J. A.; Shetty, A. M.; et al. Feasibility study of particle-assisted laser ablation of brain tumors in orthotopic canine model. Cancer Res. 2009, 69(4), 1659–1667. 42. Herold, D. M.; Das, I. J.; et al. Gold microspheres: a selective technique for producing biologically effective dose enhancement. Int. J. Radiat. Biol. 2000, 76(10), 1357–1364. 43. Hainfeld, J. F.; Slatkin, D. N.; et al. The use of gold nanoparticles to enhance radiotherapy in mice. Phys. Med. Biol. 2004, 49(18): N309–315. 44. Cho, S. H. Estimation of tumour dose enhancement due to gold nanoparticles during typical radiation treatments: a preliminary Monte Carlo study. Phys. Med. Biol. 2005, 50(15), N163–173. 45. Moeller, B. J.; Richardson, R. A.; et al. Hypoxia and radiotherapy: opportunities for improved outcomes in cancer treatment. Cancer Metastasis Rev. 2007, 26(2), 241–248. 46. Moros, E. G.; Corry, P. M.; et al. Thermoradiotherapy is underutilized for the treatment of cancer. Med. Phys. 2007, 34(1), 1–4. 47. Diagaradjane, P.; Shetty, A.; et al. Modulation of in vivo tumor radiation response via gold nanoshell-mediated vascular-focused hyperthermia: characterizing an integrated antihypoxic and localized vascular disrupting targeting strategy. Nano Lett. 2008, 8(5), 1492–1500. 48. Iyer, A. K.; Khaled, G.; et al. Exploiting the enhanced permeability and retention effect for tumor targeting. Drug Discov. Today 2006, 11(17-18), 812–818. 49. Ai, H.; Fang, M.; et al. Electrostatic layer-by-layer nanoassembly on biological microtemplates: platelets. Biomacromolecules 2003, 3(3), 560–564. 50. Carpenter, G. Receptors for epidermal growth factor and other polypeptide mitogens. Annu. Rev. Biochem. 1987, 56, 881–914. 51. Leamon, C. P.; Low, P. S. Folate-mediated targeting: from diagnostics to drug and gene delivery. Drug Discov. Today 2001, 6(1), 44–51. 52. Nayak, S.; Lee, H.; et al. Folate-mediated cell targeting and cytotoxicity using thermoresponsive microgels. J. Am. Chem. Soc. 2004, 126(33), 10258–10259. 53. Kam, N. W.; O’Connell, M.; et al. Carbon nanotubes as multifunctional biological transporters and near-infrared agents for selective cancer cell destruction. Proc. Natl. Acad. Sci. U.S.A. 2005, 102(33), 11600–11605. 54. El-Sayed, I.; Huang, X.; et al. Effect of plasmonic gold nanoparticles on benign and malignant cellular autofluorescence: a novel probe for fluorescence based detection of cancer. Technol. Cancer Res. Treat. 2007, 6(5), 403–412. 55. Pissuwan, D.; Valenzuela, S. M.; et al. A golden bullet? Selective targeting of Toxoplasma gondii tachyzoites using antibody-functionalized gold nanorods. Nano Lett. 2007, 7(12), 3808– 3812. 56. Bastus, N. G.; Sanchez-Tillo, E.; et al. Peptides conjugated to gold nanoparticles induce macrophage activation. Mol. Immunol. 2009, 46(4), 743–748. 57. De Soete, D.; Gijbels, R.; et al. Neutron Activation Analysis. John Wiley & Sons: Hoboken, NJ, 1972. 58. James, W. D.; Hirsch, L. R.; et al. Application of INAA to the build-up and clearance of gold nanoshells in clinical studies in mice. J. Radioanal. Nucl. Chem. 2007, 271(2), 455–459. 59. Hillyer, J. F.; Albrecht, R. M. Correlative instrumental neutron activation analysis, light microscopy, transmission electron microscopy, and X-ray microanalysis for qualitative and quantitative detection of colloidal gold spheres in biological specimens. Microsc. Microanal. 1998, 4(5), 481–490. 60. De Jong, W. H.; Hagens, W. I.; et al. Particle size-dependent organ distribution of gold nanoparticles after intravenous administration. Biomaterials 2008, 29(12), 1912–1919. 61. von Maltzahn, G.; Park, J. H.; et al. Computationally guided photothermal tumor therapy using long-circulating gold nanorod antennas. Cancer Res. 2009, 69(9), 3892–3900.
P1: OTA/XYZ P2: ABC c26 JWBS037-Chen January 25, 2011
13:33
Printer Name: Yet to Come
REFERENCES
657
62. Zonios, G.; Perelman, L. T.; et al. Diffuse reflectance spectroscopy of human adenomatous colon polyps in vivo. Appl. Opt. 1999, 38(31), 6628–6637. 63. Zaman, R. T.; Diagaradjane, P.; et al. In vivo detection of gold nanoshells in tumors using diffuse optical spectroscopy. IEEE. J. Selected Topics Quantum Electronics 2007, 13(6), 1715–1720. 64. Curvers, W. L.; Singh, R.; et al. Endoscopic tri-modal imaging for detection of early neoplasia in Barrett’s oesophagus: a multi-centre feasibility study using high-resolution endoscopy, autofluorescence imaging and narrow band imaging incorporated in one endoscopy system. Gut 2008, 57(2), 167–172. 65. Puvanakrishnan, P.; Park, J.; et al. Near-infrared narrow-band imaging of gold/silica nanoshells in tumors. J. Biomed. Opt. 2009, 14(2), 024044. ¨ 66. G¨oppert-Mayer, M. Uber elementarakte mit zwei quantunspr¨ungen. Ann. Phys. 1931, 9, 273–294. 67. Denk, W.; Strickler, J. H.; et al. Two-photon laser scanning fluorescence microscopy. Science 1990, 248(4951), 73–76. 68. Boyd, G. T.; Yu, Z. H.; et al. Photoinduced luminescence from the noble-metals and its enhancement on roughened surfaces. Phys. Rev. B 1986, 33(12), 7923–7936. 69. Mohamed, M. B.; Volkov, V.; et al. The “lightning” gold nanorods: fluorescence enhancement of over a million compared to the gold metal. Chem. Phys. Lett. 2000, 317(6), 517–523. 70. Wang, H.; Huff, T. B.; et al. In vitro and in vivo two-photon luminescence imaging of single gold nanorods. Proc. Natl. Acad. Sci. U.S.A. 2005, 102(44), 15752–15756. 71. Durr, N. J.; Larson, T.; et al. Two-photon luminescence imaging of cancer cells using molecularly targeted gold nanorods. Nano Lett. 2007, 7(4), 941–945. 72. Park, J.; Estrada, A.; et al. Two-photon-induced photoluminescence imaging of tumors using near-infrared excited gold nanoshells. Opt. Express 2008, 16(3), 1590–1599. 73. Bao, H.; Allen, J.; et al. Fast handheld two-photon fluorescence microendoscope with a 475 micron x 475 micron field of view for in vivo imaging. Opt. Lett. 2008, 33(12), 1333–1335. 74. Konig, K.; Ehlers, A.; et al. Clinical two-photon microendoscopy. Microsc. Res. Tech. 2007, 70(5), 398–402. 75. Schwarz, M.; Riemann, I.; et al. New developments in two photon endoscopy. In: Tearney GT, editor. Endoscopic miscroscopy. Proc. of. SPIE. 2009, 7172: 717204.
P1: OTA/XYZ P2: ABC c27 JWBS037-Chen August 31, 2010
16:1
Printer Name: Yet to Come
CHAPTER 27
Gold Nanorods as Theranostic Agents ALEXANDER WEI, QINGSHAN WEI, and ALEXEI P. LEONOV Department of Chemistry, Purdue University, West Lafayette, Indiana, USA
27.1 INTRODUCTION Recent advances in nanomaterials engineering have enabled the synthesis of anisotropic gold nanoparticles with strong optical absorption or scattering at visible and near-infrared (NIR) wavelengths [1]. Gold nanoparticles (d < 100 nm) with plasmon resonances in the NIR region are presently under active investigation as contrast agents for optical imaging modalities and nonlinear optical microscopies [2]. The NIR spectrum between 750 and 1300 nm provides a “biological window” for optical imaging, as shorter wavelengths are extinguished by hemoglobin or other endogenous pigments, and longer wavelengths are strongly attenuated by water [3]. Anisotropic gold nanoparticles are biochemically inert and have found clinical applications as colloidal adjuvants for in vivo radiotherapies [4–6] and also for animal studies for rheumatoid arthritis [7, 8] and are thus ideally suited as contrast agents for in vivo biological imaging. Gold nanorods (GNRs) are especially attractive due to their strong NIR absorption cross sections, the tunability of their plasmon resonances, and their scalable and highly reproducible synthesis. GNRs can be prepared with lengths below 100 nm and have narrower linewidths than spheroidal Au nanoparticles at comparable resonance frequencies. With respect to tunability, GNRs are readily prepared in micellar surfactant solutions, with numerous modifications introduced to establish fine control over aspect ratio, scalability and uniformity, optical and colloidal stability, and absorption and scattering properties. The latter define the types of optical imaging modalities that can be empowered by GNRs as contrast agents. The efficient NIR absorption of GNRs can enhance biomedical optical imaging modalities capable of relatively deep penetration into tissues, such as optical coherence tomography (OCT) and photoacoustic tomography (PAT), and can also support several types of nonlinear optical microscopies for imaging with low background and subcellular resolution. In particular, GNRs have large two-photon absorption cross sections in the NIR and can be excited with ultrafast laser pulses to produce a two-photon excited luminescence (TPL) that can be detected at the single-particle level [9–11]. Multiphoton imaging of GNRs can
Nanoplatform-Based Molecular Imaging Edited by Xiaoyuan Chen C 2011 John Wiley & Sons, Inc. Copyright
659
P1: OTA/XYZ P2: ABC c27 JWBS037-Chen August 31, 2010
660
16:1
Printer Name: Yet to Come
GOLD NANORODS AS THERANOSTIC AGENTS
be performed under confocal conditions with submicron resolution and has been employed in vitro for visualizing cell uptake [12] and for the targeted labeling of tumor cells [13, 14] and bacterial spores [15], and even for noninvasive in vivo imaging of GNRs passing through subcutaneous blood vessels [9]. Plasmon-resonant GNRs are more than just passive imaging agents, as most of the absorbed NIR photons are actually converted into heat [16].When exposed to high photon densities, GNRs can mediate intense photothermal effects and inflict localized injury on nearby cells and tissues. Such effects have inspired new concepts in nanomedicine, in which therapeutic effects are integrated with diagnostic imaging, now popularly referred to as theranostics [17]. Hyperthermic effects from GNRs can induce necrosis in nearby cells, with a kill radius depending on particle concentration and the dosimetry of NIR irradiation. At the microscopic level, photothermally excited GNRs may be able to raise the local temperature by many hundreds of degrees, resulting in superheating and cavitation [18, 19]. In vivo demonstrations of GNRs as theranostic agents are now emerging and serve as important milestones en route to clinical applications. However, much depends on the ability to produce uniform GNRs on a batch scale and also robust methods of surface functionalization, to enable their passage through preclinical evaluations. In this chapter we present some recent developments in the chemistry and photophysical properties of GNRs, their application toward biological imaging, and their potential for photothermally activated therapies. While much attention has been focused on the synthesis and structure–function relationships of GNRs, the process chemistry supporting their biological applications such as detoxification and surface functionalization is extremely important and also merits discussion. Experimental studies on the synthesis and functionalization of GNRs are presented in detail, with practical insights for avoiding common pitfalls. Examples of GNRs as imaging or photothermal agents are also presented in a biological context, to illustrate the scope and potential of these highly promising nanomaterials for biomedical use.
27.2 PHYSICAL PROPERTIES OF GNRs 27.2.1 Optical Properties The optical resonances of GNRs are intimately associated with an electrodynamic phenomenon known as surface plasmons, which embody the collective oscillations of conduction electrons when excited by light. In brief, colloidal Au nanoparticles much smaller than the optical wavelength (diameter ) will exhibit localized surface plasmon resonances, which are characterized by their remarkably strong extinction at specific electromagnetic frequencies. The plasmons of isotropic Au particles can be essentially described by Mie theory as dipolar modes that are sensitive to particle size, shape, material composition, and the local dielectric environment [20]. However, anisotropic particles such as GNRs are able to support at least two resonance modes—a longitudinal resonance (LR) along the GNR axis and a transverse resonance (TR) normal to that axis. The dominant LR mode shifts strongly into the NIR with increasing aspect ratio, whereas the weaker TR mode shifts only slightly toward shorter wavelengths (Fig. 27.1). LR wavelengths (LR ) have been calculated in water using modified Mie theory under quasistatic conditions and suggest that GNRs with aspect ratios ranging between 4 and 9 are optically active within the NIR “biological window” between 800 and 1300 nm, respectively [21].
P1: OTA/XYZ P2: ABC c27 JWBS037-Chen August 31, 2010
16:1
Printer Name: Yet to Come
PHYSICAL PROPERTIES OF GNRs
(a)
661
(b) Longitudinal mode (λLR)
Transverse mode (λTR)
Au
Wavelength (nm)
FIGURE 27.1 (a) Gold nanorod (GNR) with longitudinal and transverse plasmon resonances (LR and TR). (b) Simulations of LR in GNRs with different aspect ratios (R), based on modified Mie theory [21]. (Reproduced with permission of American Chemical Society.)
El-Sayed and co-workers have put forth a simple semiempirical relationship between LR and GNR aspect ratio R [22]: LR = (53.71R − 42.29)εm + 495.14
(27.1)
where ε m is the permittivity of the dielectric medium. Equation (27.1) suggests a linear shift in LR with increasing R but makes some simplifying assumptions about the GNR shape and surface dielectric, so it is valid only to the first degree of approximation. Nevertheless, the correlation matches qualitatively with Figure 27.1 at low values of R and is useful for estimating the effect of R and ε m on the LR mode. The tunability of the LSPRs is best appreciated by the range of colors produced by GNRs as a function of aspect ratio R. Shorter GNRs appear blue while longer GNRs appear red, indicating the minimal extinction at those wavelengths (Fig. 27.2a). The colors are the consequence of strong extinction of orange and green wavelengths, which are on opposite sides of the color wheel to blue and red and correspond, respectively, to LR (for short GNRs) and TR (for long GNRs; Fig. 27.2b). Plasmon-resonant GNRs below 100 nm in length typically produce extinction profiles that are dominated by absorption, but their scattering intensities are sufficient for visualization by dark-field microscopy. The scattering cross sections of GNRs do not increase commensurately with length [23]; on the other hand, an increase in diameter strengthens the scattering response but compromises the GNR aspect ratio, shifting LR to the visible range [24]. In addition to absorption and scattering, Au nanoparticles are capable of producing detectable photoluminescence (PL), particularly with pulsed laser excitation. Metals are better known for their ability to quench fluorescence by back-electron transfer, so PL is not commonly considered to be significant. In fact, PL can be generated from nanostructured or nanoparticulate Au including GNRs and can be amplified by resonant coupling with plasmons [9–11, 25–27]. Excitation of GNRs using continuous-wave laser irradiation has been found to produce PL with em >TR , and plasmons were determined to amplify the PL quantum efficiency by more than 106 times relative to bulk Au [28].
P1: OTA/XYZ P2: ABC c27 JWBS037-Chen August 31, 2010
662
16:1
Printer Name: Yet to Come
GOLD NANORODS AS THERANOSTIC AGENTS
FIGURE 27.2 (a) Optical extinction spectra of GNRs with different aspect ratios; (b) color wheel with reference to LR (and TR ) for GNR samples (labeled a–e) [2]. (Reproduced with permission of Wiley Interscience.)
27.2.2 Nonlinear Optical Properties GNRs also exhibit several nonlinear optical properties when excited near LR , particularly when using ultrashort laser pulses. Two-photon excited luminescence (TPL) [9–11], hyperRayleigh scattering [29, 30], and second harmonic generation [31, 32] can all be generated using NIR excitation with very low autofluorescence, which is highly appealing for the imaging of biological samples [33]. TPL is initiated by the absorption of two NIR photons to generate a photoexcited state, followed shortly afterwards by electron–hole recombination to produce photoemission [25, 34]. The two-photon absorption cross section in GNRs can be amplified by the coupling of localized plasmon modes and incident excitation [9, 11] and is on the order of 2000 GM, intermediate between that of typical dye molecules (102 –103 GM) and semiconductor quantum dots (104 GM). Other Au nanostructures have also been found to exhibit TPL activity, but GNRs also exhibit polarization-dependent excitation, aligned with the longitudinal plasmon mode. Wang et al. [9] have shown the TPL intensity of GNRs to be quadratically dependent on excitation power, as expected for a two-photon process (Fig. 27.3). The emission spectra are broad and depolarized and contain peaks associated with interband transitions near the X and L symmetry points of the Brillouin zone, corresponding respectively to the {001} and {111} lattice planes in fcc Au [34]. The TPL of individual submicron GNRs have also been examined by SNOM by Imura et al. [10, 35] and shown to depend strongly on the local density of states with maximum PL enhancement at the tips where the electromagnetic field is strongest. Biological applications of TPL imaging will be further discussed in Section 27.5.1. GNRs have also been applied toward bioanalytical and imaging applications based on surface-enhanced Raman scattering (SERS) [36–39]. Raman vibrational modes of chemical species adsorbed on the surfaces of plasmon-resonant nanoparticles can be amplified by orders of magnitude via coupling with local electromagnetic fields, such that the resulting emissions are highly dependent on the plasmon resonance modes of the underlying substrate. GNRs excited by an NIR laser source have been shown to generate SERS signals
P1: OTA/XYZ P2: ABC c27 JWBS037-Chen August 31, 2010
16:1
Printer Name: Yet to Come
663
PHYSICAL PROPERTIES OF GNRs
1.5
(b)
(a) 1.2
Increasing IPL: slope = 1.97
0.6 0.4
0.4
1.2
0.9
0.2 0.0 400
Decreasing 0.6
0.0 500
600
700
800
900
IPL: slope = 2.17
2.7
Wavelength (nm) (c)
Log(IPL(a.u.))
0.8
IPL(a.u.)
Absorbance (a.u.)
0.8
2.8
2.9
3.0
Log(Pin(µW))
4
Intensity (a.u.)
3
2
1 λex=780 nm
λex=730 nm
λex=830 nm
0 400 500 600 Wavelength (nm)
400 500 600 Wavelength (nm)
400 500 600 Wavelength (nm)
FIGURE 27.3 TPL generated from GNRs, using a femtosecond-pulsed Ti:sapphire laser [9]. (a) Excitation intensities superimposed onto GNR absorption spectrum; (b) quadratic dependence of PL intensity with excitation power; (c) TPL emission spectra from GNRs in aqueous solution, excited at 730, 780, and 830 nm, respectively.
with enhancement factors as high as 109 but the signal intensities are very sensitive to aspect ratio, with differences of up to 100-fold between resonant and nonresonant conditions. 27.2.3 Photothermal Properties GNRs and other Au nanoparticles are highly efficient converters of light energy into heat, making them promising agents for targeted photothermal effects, particularly in conjunction with imaging applications. Plasmon-resonant heating is due to nonradiative relaxation of the optically excited conduction electrons, whose energies are transferred to the metal lattice on the picosecond timescale via electron–phonon coupling [40]. A simple relationship describing the temperature change at the nanoparticle surface has been suggested: T =
E abs mcp
(27.2)
where Eabs is the absorbed photon energy, m represents the mass of the nanoparticle, and cp is the heat capacity of Au [41]. The power density requirement for photothermal heating
P1: OTA/XYZ P2: ABC c27 JWBS037-Chen August 31, 2010
664
16:1
Printer Name: Yet to Come
GOLD NANORODS AS THERANOSTIC AGENTS
is quite low; for example, a 5-nm Au sphere can experience a 15-K increase in surface temperature when saturated at plasmon resonance by a 20-mW laser [42]. The photothermal energy can be dissipated into the surrounding medium as heat and also through photoacoustic effects such as cavitation, depending on the rate of heat exchange. Govorov and co-workers have described a simple heat transfer model to calculate the temperature distribution around a spherical Au nanoparticle [43]: T (r ) =
VNP Q 4k0r
(27.3)
where V NP is the nanoparticle volume, Q is photothermal heat, k0 is the thermal conductivity of the medium, and r is the distance from the nanoparticle surface. This model suggests a thermal gradient with a 1/r dependence; for instance, in the case above of a 5-nm Au sphere in water, the surface temperature rise at saturation is 15 K but drops to 3 K just 10 nm from the surface. However, heat capacity increases with size (or more precisely, the ratio of volume to surface area), and the photothermal relaxation time goes up quadratically with particle diameter [44]. Therefore larger nanoparticles should generate more intense photothermal effects. In vivo photothermolysis should be performed with NIR radiation, due to its greater penetration depth into biological tissue [13]. GNRs below 100 nm are arguably more efficient photothermal transducers than most other Au nanostructures per unit volume, based on their optimal size, relatively narrow LPR linewidths (FWHM ca. 100 nm), and their large absorption cross sections. Chen and co-workers irradiated aqueous GNR suspensions at plasmon resonance with a NIR laser power of 20 mW and measured a bulk temperature increase of 26 ◦ C in 5 min, whereas GNRs embedded in polyurethane increased the local temperature to over 100 ◦ C within 1 min [16]. GNRs have subsequently been employed as actively targeted photothermal agents against tumor cells [13, 14, 45–49], macrophages [50], and parasitic pathogens [51, 52], and also passively targeted to primary tumor sites in mice [110, 111] (see Section 27.5.2). The surfaces of GNRs and other plasmon-resonant nanoparticles may reach temperatures ranging from hundreds to thousands of degrees, and an insufficient rate of heat dissipation can lead to cavitation effects [19, l3]. This process is initiated when the superheated medium reaches the kinetic spinodal, causing the expansion of transient microbubbles on the nanosecond time scale. Microbubbles can expand by as much as several microns during their short lifespan on the order of microseconds, and their sudden collapse causes the release of shockwaves and other forms of acoustic emission. The threshold fluence for photoacoustic damage can be remarkably low if ultrashort laser pulses are used: for example, Lin and co-workers have demonstrated the necrosis of lymphocytes labeled with Au nanoparticles by “optoporation” with just a single 20-ns laser pulse of 0.35 J/cm2 [18]. Therefore some attention should be paid toward the mechanisms of photothermolysis, to be further discussed in Section 27.5.2.
27.3 SYNTHESIS OF GNRs NIR-resonant GNRs of well-defined size are most often prepared by a seeded growth approach in aqueous solutions of cetyltrimethylammonium bromide (CTAB), a micellar surfactant [54]. This practical and popular method can be finely tuned by changes in
P1: OTA/XYZ P2: ABC c27 JWBS037-Chen August 31, 2010
16:1
Printer Name: Yet to Come
SYNTHESIS OF GNRs
665
FIGURE 27.4 (a) TEM image of dumbbell-shaped GNRs, freshly isolated after seeded growth; (b) oblate GNRs were obtained upon standing in solution for 24 hours [56]. (Reproduced with permission of the American Chemical Society.)
stoichiometry [55] and by introducing chemical additives [56] for further control over growth kinetics. In particular, the presence of Ag ion has been found to be critical for producing uniform GNRs in high yield [57, 58] and is thought to selectively passivate high-energy facets of the GNRs during anisotropic growth [59]. The synthesis of GNRs in micellar CTAB solutions is also scalable and has been performed on a gram scale with narrow size dispersity and excellent shape control [60, 61]. One problem often encountered while preparing NIR-absorbing GNRs is a gradual shift in LPR wavelength from NIR to visible wavelengths, over a period of hours to days. The drift in optical absorption is due to continued growth in the transverse dimension, as established by careful analysis using transmission electron microscopy (TEM, Fig. 27.4) [56]. The “fattening” of GNRs can take place even after precipitation and resuspension in fresh solution, and appears to depend on residual Au ions associated with CTAB micelles, which eventually displace the passivating Ag along the sides of the GNRs. The transverse (possibly isotropic) growth process can be quenched by adding Na2 S to reduce residual Au ions just after the initial longitudinal growth phase, producing NIR-resonant GNRs that remain optically stable for many months. We provide a detailed procedure for the synthesis of sulfide-treated GNRs in the following section. 27.3.1 Materials and Methods Deionized (DI) water was obtained using an ultrafiltration system with a measured resistivity above 18 M·cm, and passed through a 0.22-m filter to remove particulate matter. Seed solutions were prepared using high-purity CTAB (99+%), gold chloride (HAuCl4 ), and sodium borohydride (NaBH4 ); growth solutions were prepared using CTAB, HAuCl4 , silver nitrate (AgNO3 ), and ascorbic acid.
Preparation of Seed Particles In a typical experiment, 5 mL of a 0.2-M CTAB solution was combined with 5 mL of a 0.5-mM HAuCl4 solution at room temperature. Next, 0.6 mL of a 10-mM NaBH4 solution was cooled to 0 ◦ C in an ice bath for 10 min,
P1: OTA/XYZ P2: ABC c27 JWBS037-Chen August 31, 2010
666
16:1
Printer Name: Yet to Come
GOLD NANORODS AS THERANOSTIC AGENTS
then injected into the reaction mixture [57]. The solution, which turned dark brown, was vigorously mixed with magnetic stirring for 2 min. A portion of this solution was used within 2 h.
Seeded Growth of GNRs The following solutions were prepared in advance using DI water [56, 58]: 0.2 M CTAB (SigmaUltra, >99%), 200 mL; 1 mM gold chloride (HAuCl4 ), 200 mL; 4 mM silver nitrate (AgNO3 ), 9.6 mL; 78.8 mM ascorbic acid, 2.8 mL; and 4 mM sodium sulfide (Na2 S), 100 mL. The GNR growth medium was prepared by combining the first four solutions at room temperature. This mixture was then treated with 0.48 mL of a freshly prepared solution of Au nanoparticle seeds (see above), swirled by hand, and allowed to stand at room temperature between 15 and 50 min. The reaction mixture turned from colorless to a deep purplish-brown after 15 min, indicating the presence of GNRs (see Section 27.3.2 for additional details). Prolonging the reaction time increased the overall size of the GNRs, but often reduced the aspect ratio. The reaction was quenched by treatment with the Na2 S solution, yielding an approximately 500-mL suspension of GNRs with an absorption maximum near 800 nm and an optical density (O.D.) close to 1. The GNR suspension was subjected to centrifugation at 24,000g for 20 min using a fixed-angle rotor, and the CTAB-saturated supernatant was removed by decantation. The residual GNRs could be resuspended in 20–25 mL water to yield a highly concentrated sample (O.D. ca. 16.8, based on 1/20 dilution). 27.3.2 Notes The source and purity of CTAB is now known to have an influential role on GNR growth [62], and deviations in growth kinetics have recently been traced to the presence of contaminating iodide [63]. Some immediate signs of a poor quality control include the lack of an induction period and the appearance of a strongly tinted solution, attributed to an accelerated rate of gold chloride reduction. Subsequent TEM and optical analysis typically reveals poor control over particle shape, with broad absorption in the far-red region. Other sources of variability include (1) the quality of the seed particles and (2) the cleanliness of the glassware, which should be cleaned with aqua regia and rinsed with filtered DI water prior to use. Seed particles prepared by the method above are typically below 4 nm and formed as single crystals, according to TEM analysis [59]. However, the average particle size is age dependent and can increase due to Ostwald ripening. Therefore it is recommended to use freshly prepared nanoparticles for GNR growth. LPR peaks typically appear at NIR wavelengths (max > 800 nm) 15–20 min after seed addition and reach their maximum absorbance intensities after 50–60 min. However, the resonance peaks often shift toward shorter wavelengths during this period due to a reduction in aspect ratio, as the growth becomes less anisotropic. Furthermore, GNR formation can be quite sensitive to both nucleation and growth conditions, so it is not uncommon for the LPR peak to vary from batch to batch. GNR batches prepared and isolated under apparently identical conditions can differ in their absorbance maxima by as much as 30 nm. The addition of Na2 S is not absolutely required but is recommended in the case of incomplete reduction of Au ions, to prevent a shift in GNR absorbance maxima from NIR to visible wavelengths. Simply isolating GNRs from the reaction solution by centrifugation is not sufficient to arrest growth, as significant amounts of gold ion are still associated with the CTAB layers. Quenching the reaction with millimolar concentrations of Na2 S prevents further lateral growth by reducing residual gold ions, both in solution and in association with
P1: OTA/XYZ P2: ABC c27 JWBS037-Chen August 31, 2010
16:1
Printer Name: Yet to Come
SURFACE CHEMISTRY OF GNRs
667
CTAB. It should be noted that higher concentrations of Na2 S can have the opposite effect on GNR plasmon resonance (i.e., cause a significant shift toward longer wavelengths), which may be of further use for NIR imaging and excitation. These acute shifts have been attributed to the formation of metal sulfides (particularly Ag2 S) on the GNR surface, which modulates the surface dielectric [56, 64].
27.4 SURFACE CHEMISTRY OF GNRs With respect to GNR coatings and functionalization, numerous surface conjugation methods have been introduced to support biological studies in a laboratory setting (see Fig. 27.5 for some examples). However, such surface coatings must meet at least several criteria to be considered eligible for in vivo clinical studies: (1) dispersion stability in blood and other physiological fluids, (2) sufficiently long circulation lifetimes to allow efficient delivery to the region of interest, (3) resistance against nonspecific cell uptake and protein adsorption, (4) ability to support ligand functionalization for site-selective targeting and/or cell uptake, and (5) low cytotoxicity and inflammatory response. Meeting these criteria also implies that coated GNRs should be robust against chemical degradation while under biological exposure, to avoid compromising their ability to meet the criteria above. Consequently, the surface chemistry of GNRs requires at least as much attention and development as their synthesis, if they are to be useful in clinically relevant settings. 27.4.1 Detoxification of GNRs The potential of GNRs for practical biomedical applications is challenged by the presence of CTAB, the micellar surfactant used in GNR synthesis (Section 27.3). CTAB has a poor biocompatibility profile, with in vitro toxicological studies yielding IC50 values in the low micromolar range [71–73]. CTAB-coated GNRs are also susceptible to nonspecific uptake, even at very low surfactant levels [12]. These issues demand a rigorous purification of GNR formulations prior to biomedical testing, but efforts to remove CTAB from GNRs are very often met with a loss of dispersion stability. One approach that can be performed on a laboratory scale is to exchange CTAB with chemisorptive surfactants within the confines of an ion-exchange resin, which provides a physical support to prevent GNRs from flocculation [74]. CTAB exchange can also be performed on a larger batch scale in the absence of resin, simply by applying sodium polystyrenesulfonate (PSS, 70 kDa) as an adsorbent and detergent [75]. This is somewhat counterintuitive, given the precedent of using polyelectrolytes to stabilize GNRs by electrostatic adsorption (Fig. 27.5a). However, it has recently been made clear that PSS-stabilized GNRs can slowly release a cytotoxic substance over a period of weeks, most likely a persistent PSS–CTAB complex [75]. Fortunately, CTAB-laden PSS can be exchanged with fresh polyelectrolyte to produce biocompatible, CTAB-depleted GNRs with no significant cytotoxicity up to 85 g/mL (Fig. 27.6). These purification methods set the stage for further development of surfacefunctionalized GNRs, without concern for contamination by CTAB. 27.4.2 Surface Functionalization of GNRs The examples described in Figure 27.5 illustrate some of the many useful bioconjugation methods that can be applied to an appropriately modified GNR surface. Most of the
P1: OTA/XYZ P2: ABC c27 JWBS037-Chen August 31, 2010
668
16:1
Printer Name: Yet to Come
GOLD NANORODS AS THERANOSTIC AGENTS
FIGURE 27.5 Some surface functionalization and bioconjugation methods applied toward GNRs [2]. (a) Electrostatic adsorption onto polyelectrolyte (PE)-coated GNRs [65, 66]; (b) covalent attachment via carbodiimide coupling [67]; (c) “click” bioconjugation [68, 69]; (d) chemisorption using thiols [15, 70]; and (e) chemisorption using in situ dithiocarbamate (DTC) formation [12–14]. (Reproduced with permission of Wiley-Interscience.)
conventional bioconjugation methods (and also many “bio-orthogonal” crosslinking schemes) have been compiled in an updated monograph by Hermanson [76], and so will not be discussed here. Instead, we briefly review methods for introducing chemisorptive ligands onto metal surfaces, and for encapsulating GNRs within chemically modifiable shells.
Chemisorption Surface modification by chemisorption offers an alternative route to covalent surface conjugation and has become extremely popular due to the simplicity and
P1: OTA/XYZ P2: ABC c27 JWBS037-Chen August 31, 2010
16:1
Printer Name: Yet to Come
SURFACE CHEMISTRY OF GNRs
669
70-kDa PSS CTAB-coated GNRs
further exchange PSS-CTAB coating
“CTAB-free” GNRs
Percent control
100 80 60 40 20 0 0.1
1 10 Dose (μg/mL)
100
FIGURE 27.6 (Left) Graphical scheme for removing CTAB from GNRs using PSS exchange [75]. (Right) Cytotoxicity profile (using KB cells) of PSS-coated GNRs contaminated with CTAB (), and after further exchange with unadulterated PSS (). (Reproduced with permission from the American Chemical Society.)
ease with which it can be applied under laboratory settings. Thiols continue to be one of the most widely used functional groups [77] and have been used to functionalize GNRs with oligopeptides and proteins (typically via cysteine residues) [15, 70, 78], oligonucleotides [79, 80], and various other ligands. Thiolated ligands have been found to adsorb preferentially at the tips of GNRs resulting in anisotropic surface functionalization [81], which has enabled the development of optical biosensors based on LPR shifts [82–86]. Despite its utility, thiol-based chemisorption has long been known to provide limited stability when substrates are exposed to physiological conditions [87, 88]. A number of reports have since demonstrated that thiols are displaced by competing surfactants via surface exchange [89–91] and can desorb at appreciable rates under oxidative conditions [92, 93]. Multivalent ligands may not fare any better: a recent study has shown that the oxidative stability of chemisorbed dihydrolipoic acid is significantly worse than that of monovalent thiols [94]. These degradative pathways compromise the integrity of thiolbased coatings, with negative consequences for in vivo applications. More robust alternatives to thiols exist and are being actively investigated for the functionalization of GNRs and other nanoparticles. One class of ligands with particular promise is based on dithiocarbamates (DTCs, NCS2 − ), which can be formed in situ by the condensation of alkylamines with CS2 under moderately basic conditions [95, 96]. DTC chemisorption comes with several benefits: (1) it retains the simplicity of chemical self-assembly and can be performed in water; (2) it expands the chemical toolbox by enabling the conjugation of amines onto Au surfaces; and (3) once adsorbed, the DTC ligands are resistant to displacement by competing surfactants and to ambient oxidation, much more so than their thiolated counterparts [95]. In situ DTC chemisorption has been demonstrated on GNRs using amine-terminated polyethyleneglycol (PEG) [12] and also with diamine-functionalized PEG, with conjugation to targeting ligands such as folic acid (Fig. 27.5e) [14, 97].
Core–Shell Formation In addition to chemisorption, it is possible to functionalize metal nanoparticles by first encapsulating them inside a polymeric or inorganic shell. With respect to the former, several examples of emulsion polymerization have been reported [98–101]; in the latter case, Au nanoparticles encapsulated in silica shells (Au@SiO2 ) have been synthesized via a seeded variant of the well-known St¨ober method [102–104]. CTAB-stabilized GNRs have also been prepared by application of the St¨ober process [105, 106]; by increasing the amount of CTAB, it is even possible to generate a mesoporous
P1: OTA/XYZ P2: ABC c27 JWBS037-Chen August 31, 2010
670
16:1
Printer Name: Yet to Come
GOLD NANORODS AS THERANOSTIC AGENTS
SiO2 shell [107]. Methods for functionalizing silica are of course well known and present another opportunity for endowing GNRs with biomolecular functionality. An example of silica-coated GNRs as SERS labels has recently been reported [39]. 27.4.3 Material and Methods Sodium PSS (M w 70 kDa) was dispersed in DI water as a 10 mg/mL (1% w/v) solution with the aid of a sonicating bath (15 min) and stirred for 30 min prior to use. A 73-g/mL PSS solution was prepared by diluting 51 L of a 1% PSS solution in 7 mL of aqueous NaOH at pH 10. Purifications were performed with a stirred ultrafiltration cell using cellulose membrane filters with a nominal molecular weight limit (NMWL) of 100 kDa. In situ DTC formation was performed using CS2 distilled from CaH2 and commercially available amine-terminated PEG derivatives. All solutions were freshly prepared before use.
GNR Detoxification A concentrated suspension of CTAB-stabilized GNRs (8 mL, max = 784 nm, O.D. 16.8, 586 g/mL) were combined with 8 mL of chloroform and agitated for 1 min with a vortex mixer to produce an emulsion. The phases were separated by centrifugation at 1000g for 4 min, and the aqueous phase was carefully removed and treated with 0.45 mL of a 1% w/v solution of 70-kDa PSS. Additional CTAB was removed by washing the aqueous GNR suspension three more times with 8 mL of chloroform, every 3 h. The PSS-treated GNR suspension was diluted with fresh water to 200 mL, then concentrated to 2 mL using a stirred ultrafiltration cell (Millipore, Model 8010) with a regenerated cellulose membrane (NMWL 100 kDa). The GNRs were subjected to two more cycles of ultrafiltration, with starting volumes of 200 mL and final volumes of 2–7 mL. The final dialyzed suspensions of PSS-treated GNRs could be diluted with phosphate buffered saline (PBS) solution (pH 7.4) without loss of dispersion stability and were stable at room temperature for at least 30 days. The PSS-treated GNRs (7 mL, O.D. 16.6) were subjected to centrifugation at 6000g for 5 min to strip the CTAB-contaminated PSS coating from the GNR surface. The supernatant was decanted and the GNRs were redispersed in 7 mL of a 73 g/mL solution of 70-kDa PSS. The centrifugation–redispersion cycle was repeated twice more to yield PSScoated GNR suspensions depleted of CTAB. These GNRs did not exhibit any significant cytotoxicity, based on a standard MTT cell viability assay (Fig. 27.6) [75]. Amine-Functionalized GNRs by In Situ DTC Formation (One-Pot Procedure) In this procedure, the molarity of the chemisorptive DTC ligand is determined by the efficiency of in situ DTC formation in water. An aqueous suspension of PSS-stabilized GNRs at O.D. 1 (3 mL) was treated while stirring with a 10-mM solution of O,O -bis(2aminoethyl)octadecaethylene glycol (Fluka) adjusted by base titration to pH 9.5 (1 mL), followed by a saturated (28 mM) aqueous solution of CS2 (0.1 mL). The mixture was stirred for 12 h, then subjected to membrane dialysis with a molecular weight cutoff (MWCO) of 6000–8000 for 2 h in 200 mL water. The amine-functionalized GNRs can then be subjected to standard bioconjugation procedures. Amine-Functionalized GNRs by In Situ DTC Formation (Two-Pot Procedure) In this procedure, DTC formation is presumed to be quantitative, with CS2 as the limiting agent. A methanolic solution of O,O -bis(2-aminoethyl)octadecaethylene glycol (58 mM, 190 L) was treated with a methanolic solution of CS2 (110 mM, 10 L), then agitated with
P1: OTA/XYZ P2: ABC c27 JWBS037-Chen August 31, 2010
16:1
Printer Name: Yet to Come
GNRs AS THERANOSTIC AGENTS
671
a vortex mixer and allowed to stand for 1 h to yield a 5.5-mM solution of the corresponding mono-DTC derivative and excess diamine. This was diluted 100× with aqueous NaOH at pH 10 to produce a 55-M solution of the mono-DTC derivative (20 mL). An aqueous suspension of PSS-stabilized GNRs (O.D. 14.4, 21 mL) was adjusted to pH 10 by dropwise addition of 1 M NaOH, then treated with the mono-DTC derivative (55 M, 110 L). The suspension was agitated with a vortex mixer and allowed to stand overnight, then subjected to membrane dialysis as described above. 27.4.4 Notes The dispersion stability of PSS-treated GNRs is dependent on the amount of residual CTAB. For instance, the CTAB-depleted and essentially nontoxic GNRs described in Section 27.4.1 have a dispersion half-life of approximately 24 h in PBS solution when dispersed with PSS, and much less than that in serum. Therefore PSS is most beneficial as a detergent for removing CTAB and should be replaced with an alternative peptizing agent. Substitution with PEG-DTC or PEG-thiol has proved to be useful for in vitro studies involving cell uptake [12] and also for in vivo GNR biodistribution studies [108–111]. For short-term in vitro and in vivo studies, it is possible to use freshly prepared PSS-coated GNRs for surface functionalization and subsequent targeted delivery, if care is taken to remove excess CTAB and desorbed PSS–CTAB prior to administration. However, the long-term stability of polyelectrolyte-coated GNRs cannot be assumed. The one-pot surface functionalization of GNRs with amine-terminated oligoethyleneglycol chains is convenient; however, the in situ DTC formation in water is not necessarily quantitative. DTC formation generally proceeds well in polar solvents, but its efficiency can depend on several factors, including concentration and basicity. Methanol and ethanol appear to be ideal media for efficient DTC formation and can be employed in the twostep procedure, which is cleaner and may provide better control over surface coverage. In particular, the surface ligand density can be adjusted as a simple function of reagent concentration.
27.5 GNRs AS THERANOSTIC AGENTS The optical properties of plasmon-resonant GNRs (described in Section 27.2) are ideally suited toward biological imaging using NIR wavelengths. GNRs have several advantages over conventional NIR dyes: their linear and nonlinear optical cross sections are many times larger than organic chromophores, and they are far less vulnerable to photodegradation. Furthermore, GNRs are capable of producing intense photothermal effects not typically observed with organic dyes, permitting one to consider synergies between optical imaging and localized hyperthermia. This coupling has spawned a global effort to develop multifunctional agents for “theranostics,” which has emerged as a key paradigm within the rapidly growing field of nanomedicine [13, 17]. 27.5.1 GNRs as Optical Contrast Agents GNRs have been employed in a variety of biophotonic imaging applications, both for microscopic cellular imaging and for modalities capable of deeper imaging within biological
P1: OTA/XYZ P2: ABC c27 JWBS037-Chen August 31, 2010
672
16:1
Printer Name: Yet to Come
GOLD NANORODS AS THERANOSTIC AGENTS
tissues. Here we will provide select examples, with sufficient coverage for a balanced overview of the current state of the field.
Dark-Field Microscopy: GNRs and other plasmon-resonant scatterers are easily visualized with dark-field microscopy, a popular tool for imaging biological samples. GNRs are able to support dual plasmon modes in the visible and NIR range (TR and LR, respectively). The latter can be expected to produce a polarization-dependent response [112], although this feature appears to have been little exploited for biological imaging. GNRs have been used in dark-field microscopy for targeted cancer cell imaging: for example, antibody-labeled GNRs were used to label malignant carcinoma cell lines via their cellsurface EFGRs, with selectivity over normal human keratinocytes [45]. In this case, the TR mode of GNRs produced an easily detectable orange-red scattering with white light excitation. The same mode has been used to track the targeted nuclear delivery of GNRs conjugated with transferring or cell-penetrating peptides [66, 68]. In fact, the TR scattering from GNRs is several times less intense than LR scattering at NIR wavelengths, but still sufficient to be monitored by dark-field microscopy. Dark-field microscopy with white-light illumination can also support multiplex labeling strategies, as demonstrated by the simultaneous detection of GNRs with different aspect ratios, targeted toward separate cell-surface biomarkers on human breast epithelial cells [86]. Dark-field GNR imaging can even be used to measure tissue properties: longer GNRs (R ∼15) embedded in a cardiac fibroblast network could track local deformations induced by mechanical stress [113]. Multiphoton Confocal Microscopy GNRs have been found to exhibit appreciable luminescence under both linear and nonlinear excitation conditions (Sections 27.2.1 and 27.2.2) [9, 28, 114, 115]. However, nonlinear excitation modalities such as TPL are advantageous for biological imaging for several reasons (1) NIR illumination has greater transmittivity through biological structures and can therefore achieve greater penetration depth than visible light; (2) multiphoton excitation produces much lower autofluorescence and greater signal-to-noise; and (3) nonlinear optical signal intensities are highly power dependent, which increases 3D spatial resolution and minimizes collateral photodamage. The utility of GNRs as TPL contrast agents for biomedical imaging has been demonstrated both in vitro and in vivo [9]. A seminal in vivo TPL imaging study was conducted by Cheng and co-workers, who injected a dilute solution of GNRs into the bloodstream of an anesthetized mouse, followed some minutes later by the detection of TPL signals produced by GNRs passing through ear blood vessels (Fig. 27.7). Subsequent in vivo TPL studies with PEG-conjugated GNRs have revealed that such particles can remain in the bloodstream with a half-life of several hours (L. Tong and J.-X. Cheng, personal communication). A three-dimensional TPL imaging modality has also been developed for tissues using GNRs as contrast agents, with penetration depths up to 75 m in a tissue phantom [115]. In vitro TPL imaging has been used to track the trajectory and eventual fate of individual GNRs incubated with KB cells [12, 13]. Unfunctionalized (CTAB-stabilized) GNRs were internalized by KB cells within a few hours and observed to migrate toward the nucleus with a bidirectional motion, suggestive of cotransport with endosomes along microtubules [12]. In contrast, GNRs coated with PEG chains (anchored by in situ DTC formation) were not taken up by KB cells, as characterized by the near-absence of TPL signals. Folate-conjugated GNRs have also been targeted to tumor cell surfaces overexpressing the high-affinity folate receptor and observed by TPL imaging to accumulate on the outer cell membrane for many hours, prior to their receptor-mediated endocytosis and delivery to
P1: OTA/XYZ P2: ABC c27 JWBS037-Chen August 31, 2010
16:1
Printer Name: Yet to Come
GNRs AS THERANOSTIC AGENTS
673
FIGURE 27.7 Still-frame TPL image of several GNRs (indicated by arrows) passing through a mouse ear blood vessel, several minutes after a tail vein injection [9]. Blood vessel walls in transmission overlay enhanced for clarity.
the perinuclear region [14]. This image-guided delivery of GNRs exemplifies the potential for synergy between imaging and therapeutic effect: not only do the GNRs identify cancer cells, but the TPL imaging can also be used to time the delivery of NIR dosage for maximum photothermal damage (Section 27.5.2).
Optical Coherence Tomography OCT is a noninvasive optical imaging modality and analogous in several respects to ultrasound imaging, except that reflections of NIR light are detected rather than sound. OCT is capable of 2–3-mm depth penetration, with axial resolution on the order of 10 m and lateral resolution in the low micron range [116]. OCT is primarily used in clinical opthalmology, but recent technological advances have made it possible to image nontransparent tissues, extending its application toward a broader range of medical specialties. However, unlike other noninvasive imaging modalities such as ultrasound, magnetic resonance imaging (MRI), or X-ray computed tomography (CT), OCT can image cellular and even subcellular structures, with 10–25 times greater spatial resolution than that produced by ultrasound imaging, and up to 100 times better than MRI or CT [117]. OCT typically generates images based on morphology-dependent scattering but can also produce images by differential absorption contrast (spectroscopic mode) or by differences in absorption/scattering profiles. These spectroscopic modes can benefit enormously from NIR-active contrast agents such as GNRs, whose optical resonances have relatively narrow linewidths for sharp spectral distinction. In this regard, it must also be noted that the GNR optical response is dominated by its absorption; its scattering cross section is small relative to other types of plasmon-resonant nanoparticles. Although GNRs have been employed in conventional backscattering OCT, a very high concentration is needed to produce detectable contrast [118]. GNRs are much better suited to support OCT modalities based on differential absorption or backscattering albedo (the ratio of backscattering to total extinction), which have the advantage of producing contrast in tissues with intrinsically high scattering cross sections. An OCT modality based on low backscattering albedo has been demonstrated with GNRs in
P1: OTA/XYZ P2: ABC c27 JWBS037-Chen August 31, 2010
674
16:1
Printer Name: Yet to Come
GOLD NANORODS AS THERANOSTIC AGENTS
FIGURE 27.8 Spectroscopic OCT image depicting the GNR density (green) in an excised human breast carcinoma [121]. Structural OCT image is presented in red. (Reproduced with permission from the Royal Chemical Society.)
highly backscattering tissue phantoms, with an estimated detection limit as low as 30 ppm [119]. The narrow absorption linewidths of GNRs can also be used to enhance spectroscopic OCT and were recently investigated as contrast agents in an excised sample of human breast invasive ductal carcinoma (Fig. 27.8) [120, 121]. The polarization-dependent extinction of GNRs has not yet been exploited for OCT imaging but is anticipated to provide further enhancements to absorption-based modalities, particularly for polarization-sensitive OCT.
Photoacoustic Tomography GNRs can also generate optoacoustic contrast for PAT, another emerging imaging modality with analogy to ultrasound [122, 123]. Pulsed NIR laser irradiation above a threshold power results in photoinduced cavitation effects (see below), and the low diffusion of propagating acoustic waves allows a depth resolution of up to several centimeters in biological tissue [124]. Images are then obtained by the reconstruction of the optical energy absorption distributed across an array of acoustic transducers. The advantage of PAT over pure optical or ultrasonic imaging technique relies on the combined merits of optical irradiation and acoustic detection. Laser excitation can provide micron-level spatial resolution (versus the millimeter resolution of ultrasound waves), whereas the low diffusion of the propagating acoustic waves has a major advantage over reflected optical signals, which suffer from scattering by biological tissue [124]. The depth resolution of PAT can be several centimeters in biological tissue [125], whereas the depth penetration of purely optical imaging is a few millimeters [126]. GNR-enhanced PAT imaging has been demonstrated in vivo using nude mice and could produce variations in signal intensity based on concentration differences as low as 1.25 pM (Fig. 27.9) [122]. Another recent in vivo example illustrates the utility of GNRs as contrast agents for sentinel lymph node (SLN) mapping by PAT [127]. Some other examples involving GNR-enhanced PAT include quantitative flow analysis in biological tissues [128] and an ex vivo analysis of distribution kinetics of drug delivery systems [129]. 27.5.2 GNRs as Photothermal Agents As previously discussed in Section 27.2.3, plasmon-resonant nanoparticles such as GNRs are highly efficient at converting light energy into heat, making them promising agents for the targeted photothermolysis of cells and tissues [130]. Two types of photothermal
P1: OTA/XYZ P2: ABC c27 JWBS037-Chen August 31, 2010
16:1
Printer Name: Yet to Come
GNRs AS THERANOSTIC AGENTS
(a)
675
(b)
FIGURE 27.9 Photoacoustic tomography of a nude mouse, (a) before and (b) after injection of GNRs [122]. (Reproduced with permission from the American Chemical Society.)
effects have been established: at the macroscopic level, the dissipated heat can induce a nonspecific, hyperthermic response in nearby cells, typically resulting in necrosis [131, 132]. The hyperthermic effects are localized and can be controlled by the dosimetry of NIR irradiation. At the microscopic level, photothermally excited nanoparticles can raise the temperature of its immediate environment by hundreds to thousands of degrees, resulting in superheating and cavitation [133, 134]. The photoacoustic effects are generated on a timescale of nano- to microseconds [53] and can deliver a mechanical form of cell injury if the GNRs are directly attached [135, 136]. GNRs have been exploited for the photothermolysis of tumor cells and microbial pathogens, in vitro as well as in vivo. Two recent examples of the latter involve the delivery of PEG-coated GNRs to vascularized tumor sites implanted within athymic nude mice, via a passive uptake pathway commonly referred to as enhanced permeabilization and retention (EPR) [110, 111]. GNRs were administered by tail-vein injection, followed some time later (at least 24 h postinjection) by irradiation with an NIR laser. The quantity of GNRs injected in each case was relatively high, on the order of 104 –1012 particles (up to 20 mg Au/kg). Photothermal therapy (PTT) was based on a single dose of laser irradiation at power densities of 2–4 W/cm 2 for 3–10 min (fluences of 10–18 W·min/cm 2 ). A dramatic reduction in tumor volume was observed in both studies, as well as significant tumor resorption and enhanced survival. In the first study involving squamous cell carcinoma xenografts [110], tumor resorption was observed in approximately 25% of mice over a 13-day period (N = 7); in a second study involving explanted melanoma cells [111], all mice treated by GNR-mediated PTT survived over a 50-day period (N = 4) whereas control mice continued to experience tumor progression and were euthanized within 2–3 weeks. It is worth mentioning that the latter study was complemented by a biodistribution analysis of PEG-coated GNRs, with a blood circulation half-life of approximately 17 h and a 72-h postinjection tumor uptake of 7% ID/g [111]. The high dosage of GNRs used was also sufficient to generate X-ray contrast, permitting some enhancement for CT imaging. Overall, the results of these studies should be interpreted objectively: while the GNR concentrations may be high, they also clearly demonstrate the exciting potential of GNR-mediated PTT for combating cancer. Incorporation of active targeting mechanisms can be anticipated to reduce GNR loadings in the development of effective PTT protocols for clinical cancer treatment. While the issue of selective tumor delivery is of primary importance for future theranostic applications, attention must also be paid toward the mechanisms of photoinduced cell injury. In particular, necrosis is often assumed to be the result of hyperthermia, for which a few
P1: OTA/XYZ P2: ABC c27 JWBS037-Chen August 31, 2010
676
16:1
Printer Name: Yet to Come
GOLD NANORODS AS THERANOSTIC AGENTS
(a)
before
(b)
after
(c)
FIGURE 27.10 (a, b) TPL signals from folate-conjugated GNRs (red) residing on the membranes of KB cells, before and after a 1-min exposure to a scanning NIR laser (12 J/cm2 ). (c) GNR-induced membrane perforation followed by blebbing, visualized by ethidium bromide (red) and a fluorescent dye indicating free intracellular Ca2+ (green) [14]. (Reproduced with permission from Wiley-VCH.)
degrees is sufficient to cause cell and tissue malfunction [131, 132, 137, 138]. However, this is not necessarily the case at the single-cell level. A recent study by Tong et al. [14] has shown that GNRs can mediate the “optoporation” of tumor cell membranes and induce necrosis in KB cells by an intracellular influx of Ca ions, accompanied by a dramatic membrane blebbing response (Fig. 27.10). The GNRs are thought to disrupt intracellular homeostasis by generating local cavitation effects as discussed above [53, 133–136], which can inflict damage on the cell membrane and permit the rapid influx of Ca ions, followed by degradation of the actin cytoskeleton. Other mechanisms may also contribute toward cell death and are likely to involve a biochemical imbalance caused by the loss of membrane integrity.
27.6 OUTLOOK The future of GNRs as a platform for nanomedicine is most certainly bright, literally as well as figuratively. The many examples presented here amply demonstrate the feasibility of employing GNRs as multifunctional agents for biomedical diagnostics and image-guided therapies. Optical microscopies and biomedical imaging modalities based on linear and nonlinear optical processes can be integrated with GNR-based theranostics in cells and tissues with a high level of synergy; in particular, emerging in vivo imaging modalities such as OCT and PAT are making significant strides and are now taking full advantage of GNRs as NIR-active contrast agents. While recent in vivo studies underscore the exciting potential of GNRs as photothermal agents for cancer treatment, preclinical evaluation of functionalized GNRs remains an outstanding issue. Colloidal Au nanoparticles have been employed as adjuvants in clinical radiotherapies for many years (particularly for brachytherapy) [4–6, 139], but functionalized GNRs must be regarded as novel combination products and thus require the same pharmacokinetic testing as any other novel chemical entity. Surface chemistry will continue to play a critical role in the development of clinically useful nanomaterials and may prove to be the limiting factor for meeting regulatory standards. For this reason, the successful deployment of GNRs for disease imaging and treatment will require a high level of cooperation between chemists, physicists, biomedical engineers, and clinicians to ensure the successful translation of promising research discoveries to meaningful patient outcomes.
P1: OTA/XYZ P2: ABC c27 JWBS037-Chen August 31, 2010
16:1
Printer Name: Yet to Come
REFERENCES
677
ACKNOWLEDGMENTS The authors gratefully acknowledge financial support from the National Institutes of Health (EB-001777) and an active exchange with the Nanomaterials Characterization Laboratory (SAIC–Frederick), supported under NCI contract N01-CO-12400.
REFERENCES 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26.
Xia, Y.; Halas, N. J. MRS Bull. 2005, 30, 338–343. Tong, L.; Wei, Q.; Wei, A.; Cheng, J.-X., Photochem. Photobiol. 2009, 85, 21–32. Weissleder, R. Nature Biotechnol. 2001, 19, 316–317. Root, S. W.; Andrews, G. A.; Kniseley, R. M.; Tyor, M. P. Cancer 1954, 7, 856–866. Moses, C.; Kent, E.; Boatman, J. B. Cancer 1955, 8, 417–423. Rubin, P.; Levitt, S. H. J. Nucl. Med. 1964, 5, 581–594. Durkes, T. E. Probl. Vet. Med. 1992, 4, 207–211. Tsai, C. Y.; Shiau, A. L.; Chen, S. Y.; Chen, Y. H.; Cheng, P. C.; Chang, M. Y.; Chen, D. H.; Chou, C. H.; Wang, C. R.; Wu, C. L. Arthritis Rheum. 2007, 56, 544–554. Wang, H.; Huff, T. B.; Zweifel, D. A.; He, W.; Low, P. S.; Wei, A.; Cheng, J.-X., Proc. Natl. Acad. Sci. USA 2005, 102, 15752–15756. Imura, K.; Nagahara, T.; Okamoto, H. J. Phys. Chem. B 2005, 109, 13214–13220. Bouhelier, A.; Bachelot, R.; Lerondel, G.; Kostcheev, S.; Royer, P.; Wiederrecht, G. P. Phys. Rev. Lett 2005, 95, 267405. Huff, T. B.; Hansen, M. N.; Zhao, Y.; Cheng, J.-X., Wei, A. Langmuir 2007, 23, 1596– 1599. Huff, T. B.; Tong, L.; Zhao, Y.; Hansen, M. N.; Cheng, J.-X., Wei, A. Nanomedicine 2007, 2, 125–132. Tong, L.; Zhao, Y.; Huff, T. B.; Hansen, M. N.; Wei, A.; Cheng, J.-X., Adv. Mater. 2007, 19, 3136–3141. He, W.; Henne, W. A.; Wei, Q.; Zhao, Y.; Doorneweerd, D. D.; Cheng, J.-X., Low, P. S.; Wei, A. Nano Res. 2008, 1, 450–456. Chou, C.-H.; Chen, C.-D.; Wang, C. R. C. J. Phys. Chem. B 2005, 109, 11135–11138. Cuenca, A. G.; Jiang, H.; Hochwald, S. N.; Delano, M.; Cance, W. G.; Grobmyer, S. R. Cancer 2006, 107, 459–466. Pitsillides, C. M.; Joe, E. K.; Wei, X.; Anderson, R. R.; Lin, C. P. Biophys. J. 2003, 84, 4023–4032. Kotaidis, V.; Plech, A. Appl. Phys. Lett. 2005, 87, 213102. Kreibig, U.; Vollmer, M. Optical Properties of Metal Clusters. Springer: New York, 1995. van der Zande, B.; B¨ohmer, M. R.; Fokkink, L. G. J.; Sch¨onenberger, C. Langmuir 2000, 16, 451–458. Link, S.; Mohamed, M. B.; El-Sayed, M. A. J. Phys. Chem. B 1999, 103, 3073–3077. Jain, P. K.; Lee, K. S.; El-Sayed, I. H.; El-Sayed, M. A. J. Phys. Chem. B 2006, 110, 7238–7248. Ni, W.; Kou, X.; Yang, Z.; Wang, J. ACS Nano 2008, 2, 677–686. Mooradian, A. Phys. Rev. Lett. 1969, 22, 185–187. Varnavski, O. P.; Mohamed, M. B.; El-Sayed, M. A.; Goodson, T. J. Phys. Chem. B 2003, 107, 3101–3104.
P1: OTA/XYZ P2: ABC c27 JWBS037-Chen August 31, 2010
678
16:1
Printer Name: Yet to Come
GOLD NANORODS AS THERANOSTIC AGENTS
27. Dulkeith, E.; Niedereichholz, T.; Klar, T. A.; Feldmann, J.; von Plessen, G.; Gittins, D. I.; Mayya, K. S.; Caruso, F. Phys. Rev. B 2004, 70, 205424. 28. Mohamed, M. B.; Volkov, V.; Link, S.; El-Sayed, M. A. Chem. Phys. Lett. 2000, 317, 517– 523. 29. Nappa, J.; Revillod, G.; Abid, J.-P., Russier-Antoine, I.; Jonin, C.; Benichou, E.; Girault, H. H.; Brevet, P. F. Faraday Discuss. 2004, 125, 145–156. 30. Darbha, G. K.; Rai, U. S.; Singh, A. K.; Ray, P. C. Chem. Eur. J. 2008, 14, 3896–3903. 31. Hubert, C.; Billot, L.; Adam, P. M.; Bachelot, R.; Royer, P.; Grand, J.; Gindre, D.; Dorkenoo, K. D.; Fort, A. Appl. Phys. Lett. 2007, 90, 181105. 32. Celebrano, M.; Zavelani-Rossi, M.; Polli, D.; Cerullo, G.; Biagioni, P.; Finazzi, M.; Duo, L.; Labardi, M.; Allegrini, M.; Grand, J.; Adam, P. M. J. Microsc. 2008, 229, 233–239. 33. Helmchen, F.; Denk, W. Nat. Methods 2005, 2, 932–940. 34. Boyd, G. T.; Yu, Z. H.; Shen, Y. R. Phys. Rev. B 1986, 33, 7923–7936. 35. Imura, K.; Nagahara, T.; Okamoto, H. J. Am. Chem. Soc. 2004, 126, 12730–12731. 36. Orendorff, C. J.; Gole, A.; Sau, T. K.; Murphy, C. J. Anal. Chem. 2005, 77, 3261–3266. 37. Nikoobakht, B.; Wang, J.; El-Sayed, M. A. Chem. Phys. Lett. 2002, 366, 17–23. 38. Kawamura, G.; Yang, Y.; Nogami, M. Appl. Phys. Lett. 2007, 90, 261908. 39. Wang, C.; Chen, Y.; Wang, T.; Ma, Z.; Su, Z. Adv. Funct. Mater. 2008, 18, 355–361. 40. Link, S.; El-Sayed, M. A. Int. Rev. Phys. Chem. 2000, 19, 409–453. 41. Burda, C.; Chen, X.; Narayanan, R.; El-Sayed, M. A. Chem. Rev. 2005, 105, 1025–1102. 42. Boyer, D.; Tamarat, P.; Maali, A.; Lounis, B.; Orrit, M. Science 2002, 297, 1160–1163. 43. Richardson, H. H.; Hickman, Z. N.; Govorov, A. O.; Thomas, A. C.; Zhang, W.; Kordesch, M. E. Nano Lett. 2006, 6, 783–788. 44. Hu, M.; Hartland, G. V. J. Phys. Chem. B 2002, 106, 7029–7033. 45. Huang, X.; El-Sayed, I. H.; Qian, W.; El-Sayed, M. A. J. Am. Chem. Soc. 2006, 128, 2115–2120. 46. Takahashi, H.; Niidome, T.; Nariai, A.; Niidome, Y.; Yamada, S. Chem. Lett. 2006, 35, 500– 501. 47. Huang, Y.-F., Sefah, K.; Bamrungsap, S.; Chang, H.-T., Tan, W. Langmuir 2008, 24, 11860–11865. 48. Black, K. C.; Kirkpatrick, N. D.; Troutman, T. S.; Xu, L.; Vagner, J.; Gillies, R. J.; Barton, J. K.; Utzinger, U.; Romanowski, M. Mol. Imaging 2008, 7, 50–57. 49. Li, J. L.; Day, D.; Gu, M. Adv. Mater. 2008, 20, 3866–3871. 50. Pissuwan, D.; Valenzuela, S.; Killingsworth, M.; Xu, X.; Cortie, M. J. Nanopart. Res. 2007, 9, 1109–1124. 51. Pissuwan, D.; Valenzuela, S. M.; Miller, C. M.; Cortie, M. B. Nano Lett. 2007, 7, 3808–3812. 52. Norman, R. S.; Stone, J. W.; Gole, A.; Murphy, C. J.; Sabo-Attwood, T. L. Nano Lett. 2008, 8, 302–306. 53. Vogel, A.; Noack, J.; H¨uttmann, G.; Paltauf, G. Appl. Phys. B 2005, 81, 1015–1047. 54. Jana, N. R.; Gearheart, L.; Murphy, C. J. Adv. Mater. 2001, 13, 1389–1393. 55. Murphy, C. J.; Gole, A. M.; Hunyadi, S. E.; Orendorff, C. J. Inorg. Chem. 2006, 45, 7544–7554. 56. Zweifel, D. A.; Wei, A. Chem. Mater. 2005, 17, 4256–4261. 57. Nikoobakht, B.; El-Sayed, M. A. Chem. Mater. 2003, 15, 1957–1962. 58. Sau, T. K.; Murphy, C. J. Langmuir 2004, 20, 6414–6420. 59. Liu, M.; Guyot-Sionnest, P. J. Phys. Chem. B 2005, 109, 22192–22200. 60. Jana, N. R. Small 2005, 1, 875–882. 61. Khanal, B. P.; Zubarev, E. R. J. Am. Chem. Soc. 2008, 130, 12634–12635.
P1: OTA/XYZ P2: ABC c27 JWBS037-Chen August 31, 2010
16:1
Printer Name: Yet to Come
REFERENCES
62. 63. 64. 65. 66. 67. 68. 69. 70. 71. 72. 73. 74. 75. 76. 77. 78. 79. 80. 81. 82. 83. 84. 85. 86. 87. 88. 89. 90. 91. 92. 93. 94. 95.
679
Smith, D. K.; Korgel, B. A. Langmuir 2008, 24, 644–649. Smith, D. K.; Miller, N. R.; Korgel, B. A. Langmuir 2009, 25, 9518–9524. Liu, M. Z.; Guyot-Sionnest, P. J. Mater. Chem. 2006, 16, 3942–3945. Gole, A.; Murphy, C. J. Chem. Mater. 2005, 17, 1325–1330. Ding, H.; Yong, K.-T., Roy, I.; Pudavar, H. E.; Law, W. C.; Bergey, E. J.; Prasad, P. N. J. Phys. Chem. C 2007, 111, 12552–12557. Gole, A.; Murphy, C. J. Langmuir 2005, 21, 10756–10762. Oyelere, A. K.; Chen, P. C.; Huang, X.; El-Sayed, I. H.; El-Sayed, M. A. Bioconjug. Chem. 2007, 18, 1490–1497. Gole, A.; Murphy, C. J. Langmuir 2008, 24, 266–272. Liao, H.; Hafner, J. H. Chem. Mater. 2005, 17, 4636–4641. Cortesi, E.; Esposito, E.; Menegatti, E.; Gambari, R.; Nastruzzi, C. Int. J. Pharm. 1996, 139, 69–78. Mirska, D.; Schirmer, K.; Funari, S.; Langner, A.; Dobner, B.; Brezesinski, G. Colloids Surf. B 2005, 40, 51–59. Takahashi, H.; Niidome, Y.; Niidome, T.; Kaneko, K.; Kawasaki, H.; Yamada, H. Langmuir 2006, 22, 2–5. Dai, Q.; Coutts, J.; Zou, J. Q. Huo, Chem. Commun. 2008, 2858–2860. Leonov, A. P.; Zheng, J.; Clogston, J. D.; Stern, S. T.; Patri, A. K.; Wei, A. ACS Nano 2008, 2, 2481–2488. Hermanson, G. T. Bioconjugate Techniques, 2nd ed.; Academic Press: San Diego, 2008. Love, J. C.; Estroff, L. A.; Kriebel, J. K.; Nuzzo, R. G.; Whitesides, G. M. Chem. Rev. 2005, 105, 1103–1170. Tong, L.; Cheng, J.-X., Nanomedicine 2009, 4, 265–276. Dujardin, E.; Mann, S.; Hsin, L.-B.; Wang, C. R. C. Chem. Commun. 2001, 1264–1265. Pan, B.; Cui, D.; Ozkan, C.; Xu, P.; Huang, T.; Li, Q.; Chen, H.; Liu, F.; Gao, F.; He, R. J. Phys. Chem. C 2007, 111, 12572–12576. Caswell, K. K.; Wilson, J. N.; Bunz, U. H. F.; Murphy, C. J. J. Am. Chem. Soc. 2003, 125, 13914–13915. Chang, J.-Y.; Wu, H.; Chen, H.; Ling, Y.-C.; Tan, W. Chem. Commun. 2005, 1092–1094. Shibu Joseph, S. T.; Ipe, B. I.; Pramod, P.; Thomas, K. G. J. Phys. Chem. B 2006, 110, 150–157. Yu, C.; Irudayaraj, J. Anal. Chem. 2007, 79, 572–579. Wang, C.; Irudayaraj, J. Small 2008, 4, 2204–2208. Yu, C.; Nakshatri, H.; Irudayaraj, J. Nano Lett. 2007, 7, 2300–2306. Mrksich, M.; Dike, L. E.; Tien, J.; Ingber, D. E.; Whitesides, G. M. Exp. Cell Res. 1997, 235, 305–313. Flynn, N. T.; Tran, T. N.; Cima, M. J.; Langer, R. Langmuir 2003, 19, 10909–10915. Demers, L. M.; Mirkin, C. A.; Mucic, R. C.; Reynolds, R. A. III; Letsinger, R. L.; Elghanian, R.; Viswanadham, G. Anal. Chem. 2000, 72, 5535–5541. Castelino, K.; Kannan, B.; Majumdar, A. Langmuir 2005, 21, 1956–1961. Hong, R.; Han, G.; Fernandez, J. M.; Kim, B.-J.; Forbes, N. S.; Rotello, V. M. J. Am. Chem. Soc. 2006, 128, 1078–1079. Schoenfisch, M. H.; Pemberton, J. E. J. Am. Chem. Soc. 1998, 120, 4502–4513. Dasog, M.; Scott, R. W. J. Langmuir 2007, 23, 3381–3387. Hou, W.; Dasog, M.; Scott, R. W. J. Langmuir 2009, 25, 12954–12961. Zhao, Y.; P´erez-Segarra, W.; Shi, Q.; Wei, A. J. Am. Chem. Soc. 2005, 127, 7328–7329.
P1: OTA/XYZ P2: ABC c27 JWBS037-Chen August 31, 2010
680
16:1
Printer Name: Yet to Come
GOLD NANORODS AS THERANOSTIC AGENTS
96. Zhu, H.; Coleman, D. M.; Dehen, C. J.; Geisler, I. M.; Zemlyanov, D.; Chmielewski, J.; Simpson, G. J.; Wei, A. Langmuir 2008, 24, 8660–8666. 97. Wei, Q.; Song, H.-M., Leonov, A. P.; Hale, J. B.; Oh, D.; Ong, Q. K.; Ritchie, K. P.; Wei, A. J. Am. Chem. Soc. 2009, 131, 9728–9734. 98. Hasegawa, M.; Arai, K.; Saito, S. J. Polym. Sci. A 1987, 25, 3117–3125. 99. Van Herk, A. M., Polymeric Dispersions: Principles and Applications; Asua, J. M., Ed.; Kluwer Academic: Boston, 1997; Vol. 335, pp. 435–450. 100. Quaroni, L.; Chumanov, G. J. Am. Chem. Soc. 1999, 121, 10642–10643. 101. Clark, H. A.; Campagnola, P. J.; Wuskell, J. P.; Lewis, A.; Loew, L. M. J. Am. Chem. Soc. 2000, 122, 10234–10235. 102. Liz-Marz´an, L. M.; Giersig, M.; Mulvaney, P. Chem. Commun. 1996, 731–732. 103. Liz-Marz´an, L. M.; Giersig, M.; Mulvaney, P. Langmuir 1996, 12, 4329–4335. 104. Graf, C.; van Blaaderen, A. Langmuir 2002, 18, 524–534. 105. Pastoriza-Santos, I.; P´erez-Juste, J.; Liz-Marz´an, L. M. Chem. Mater. 2006, 18, 2465–2467. 106. Wang, C.; Ma, Z.; Su, Z. Nanotechnology 2006, 17, 1819–1824. 107. Gorelikov, I.; Matsuura, N. Nano Lett. 2008, 8, 369–373. 108. Niidome, T.; Yamagata, M.; Okamoto, Y.; Akiyama, Y.; Takahashi, H.; Kawano, T.; Katayama, Y.; Niidome, Y. J. Control. Release 2006, 35, 500–501. 109. Niidome, T.; Akiyama, Y.; Shimoda, K.; Kawano, T.; Mori, T.; Katayama, Y.; Niidome, Y. Small 2008, 4, 1001–1007. 110. Dickerson, E. B.; Dreaden, E. C.; Huang, X.; El-Sayed, I. H.; Chu, H.; Pushpanketh, S.; McDonald, J. F.; El-Sayed, M. A. Cancer Lett. 2008, 269, 57–66. 111. von Maltzahn, G.; Park, J.-H.; Agrawal, A.; Bandaru, N. K.; Das, S. K.; Sailor, M. J.; Bhatia, S. N. Cancer Res. 2009, 69, 3892–3900. 112. S¨onnichsen, C.; Alivisatos, A. P. Nano Lett. 2005, 5, 301–304. 113. Stone, E. W.; Sisco, P. N.; Goldsmith, E. C.; Baxter, S. C.; Murphy, C. J. Nano Lett. 2007, 7, 116–119. 114. Eustis, S.; El-Sayed, M. A. J. Phys. Chem. B 2005, 109, 16350–16356. 115. Durr, N. J.; Larson, T.; Smith, D. K.; Korgel, B. A.; Sokolov, K.; Ben-Yakar, A. Nano Lett. 2007, 7, 941–945. 116. Zysk, A. M.; Nguyen, F. T.; Oldenburg, A. L.; Marks, D. L.; Boppart, S. A. J. Biomed. Opt. 2008, 12, 051403. 117. Fujimoto, J. G. Nat. Biotechnol. 2003, 21, 1361–1367. 118. Troutman, T. S.; Barton, J. K.; Romanowski, M. Opt. Lett. 2007, 32, 1438–1440. 119. Oldenburg, A. L.; Hansen, M. N.; Zweifel, D. A.; Wei, A.; Boppart, S. A. Opt. Express 2006, 14, 6724–6738. 120. Oldenburg, A. L.; Hansen, M. N.; Wei, A.; Boppart, S. A. Proc. SPIE 2008, 6867, 68670E. 121. Oldenburg, A. L.; Hansen, M. N.; Ralston, T. S.; Wei, A.; Boppart, S. A. J. Mater. Chem. 2009, 19, 6407–6411. 122. Eghtedari, M.; Oraevsky, A.; Copland, J. A.; Kotov, N.; Conjusteau, A.; Motamedi, M. Nano Lett. 2007, 7, 1914–1918. 123. Song, K. H.; Kim, C.; Maslov, K.; Wang, L. V. Eur. J. Radiology 2009, 70, 227–231. 124. Wang, X.; Pang, Y.; Ku, G.; Xie, X.; Stoica, G.; Wang, L. V. Nat. Biotechnol. 2003, 21, 803–806. 125. Ku, G.; Wang, L. V. Opt. Lett. 2005, 30, 507–509. 126. Zhang, H. F.; Maslov, K.; Stoica, G.; Wang, L. V. Nat. Biotechnol. 2006, 24, 848–851. 127. Song, K. H.; Kim, C.; Maslov, K.; Wang, L. V. Eur. J. Radiol. 2009, 70, 227–231.
P1: OTA/XYZ P2: ABC c27 JWBS037-Chen August 31, 2010
16:1
Printer Name: Yet to Come
REFERENCES
681
128. Liao, C.-K.; Huang, S.-W.; Wei, C.-W.; Li, P.-C. J. Biomed. Opt. 2007, 12, 064006. 129. Chamberland, D. L.; Agarwal, A.; Kotov, N.; Fowlkes, J. B.; Carson, P. L.; Wang, X. Nanotechnology 2008, 19, 095101. 130. Anderson, R. R.; Parrish, J. A. Science 1983, 220, 524–527. 131. Bowler, K. Symposium of the Society of Experimental Biology; Bowler, K., Fuller, B. J., Eds.; Society for Experimental Biology: Cambridge, UK 1987, pp. 157–186. 132. Falk, M. H.; Issels, R. D. Int. J. Hyperthermia 2001, 17, 1–18. 133. Lin, C. P.; Kelly, N. W. Appl. Phys. Lett. 1998, 72, 2800–2802. 134. Lin, C. P.; Kelly, N. W.; Sibayan, S. A. B.; Latina, M. A.; Anderson, R. R. IEEE J. Sel. Top. Quant. 1999, 5, 963–968. 135. Pitsillides, C. M.; Joe, E. K.; Wei, X.; Anderson, R. R.; Lin, C. P. Biophys. J. 2003, 84, 4023–4032. 136. Zharov, V. P.; Galitovskaya, E. N.; Johnson, C.; Kelly, T. Lasers Surg. Med. 2005, 37, 219–226. 137. Coakley, W. T. Symposium of the Society Experimental Biology; Bowler, K., Fuller, B. J., Eds.; Society for Experimental Biology: Cambridge, UK 1987, pp. 187–211. 138. Wust, P.; Hildebrandt, B.; Sreenivasa, G.; Rau, B.; Gellerman, J.; Riess, H.; Felix, R.; Schlag, P. M. Lancet Oncology 2002, 3, 487–497. 139. Seal, S. H.; Crosignani, S.; Valvassori, G.; Nickson, J. J.; Agostino, D. Am. J. Obstet. Gynecol. 1958, 75, 1027–1033.
P1: OTA/XYZ P2: ABC c28 JWBS037-Chen February 1, 2011
0:19
Printer Name: Yet to Come
CHAPTER 28
Theranostic Applications of Gold Core–Shell Structured Nanoparticles WEI LU, MARITES P. MELANCON, and CHUN LI Department of Experimental Diagnostic Imaging, University of Texas M. D. Anderson Cancer Center, Houston, Texas, USA
Gold core–shell nanostructures have unique optical properties due to strong and tunable surface plasmon resonance, which is a collective oscillation of conduction-band electrons within the structures induced by the oscillating dipole of a resonant wavelength of light. By varying the core composition, core radius, and shell thickness, the optical resonances of core–shell nanostructures—including silica-core gold nanoshells, hollow gold nanoshells, and hollow gold nanocages—can be continuously tuned through wavelengths ranging from the ultraviolet region to the infrared region. These novel nanoparticles have been shown to be useful for a variety of potential applications in biomedicine, including cancer molecular optical imaging, controlled drug delivery, and photothermal ablation therapy. In this chapter, we provide an up-to-date summary of the synthesis, characterization, and preclinical evaluation of gold core–shell nanostructures. We also touch upon the development of multifunctional gold core–shell nanoparticles for theranostic applications (i.e., simultaneous treatment and diagnosis) for cancer.
28.1 INTRODUCTION Theranostics are agents that have both a treatment and a diagnostic function. The theranostic approach tailors treatment to the individual patient and has the potential to offer improved therapeutic efficacy, reduced side effects, and improved cost effectiveness. Nanotechnology as an enabling technology has attracted much attention in recent years. In particular, multifunctional nanomaterials incorporating both diagnostic and therapeutic functions have been created and evaluated for cancer theranostic applications. Nanomaterials that interact with light provide a unique opportunity [1]. These photocoupling agents can not only report the molecular-specific signatures of cancer with high sensitivity and high spatial resolution but also mediate photochemical, photothermal, and photomechanical responses that can be harnessed to kill cancer cells. Nanoplatform-Based Molecular Imaging Edited by Xiaoyuan Chen C 2011 John Wiley & Sons, Inc. Copyright
683
P1: OTA/XYZ P2: ABC c28 JWBS037-Chen February 1, 2011
684
0:19
Printer Name: Yet to Come
THERANOSTIC APPLICATIONS OF GOLD CORE–SHELL STRUCTURED NANOPARTICLES
Nanostructures made of noble metal, especially gold and silver, exhibit a unique and tunable optical property, termed surface plasmon resonance (SPR) [2, 3]. SPR is a collective oscillation of conduction-band electrons within the structures induced by the oscillating dipole of a resonant wavelength of light. For solid spherical particles, the resonance peaks appear at approximately 520 nm for gold and 400 nm for silver, and the peak varies slightly depending on the size of the particle and the embedding medium. The SPR peaks of nanostructures can be tuned from the visible to the near-infrared (NIR) region by controlling the size, shape (e.g., nanorods), and structure (e.g., core–shell structured nanoparticles). These unique optical properties combined with excellent biocompatibility of gold make gold nanostructures well suited for various biomedical applications [4–9]. Gold core–shell nanostructures consist of a dielectric or semiconducting core surrounded by an ultrathin gold shell. Core particles of different morphologies—such as rods, wires, tubes, rings, and cubes—can be coated with a thin shell to obtain core–shell structured nanoparticles with the desired shape [10]. These nanostructures possess photochemical properties different from those of their single-component counterparts of the same size. These nanostructures have been shown to be useful for a variety of potential applications in biomedicine, including cancer molecular optical imaging, controlled drug delivery, and photothermal ablation therapy. In this chapter, we provide an up-to-date review of the synthesis, characterization, and preclinical evaluation of gold core–shell structured gold nanoparticles, with emphasis on the development of multifunctional gold nanoparticles for cancer theranostic applications.
28.2 SYNTHESIS, CHARACTERIZATION, AND TUNING 28.2.1 Synthesis and Characterization Gold core–shell structured nanoparticles are generally synthesized by a template method, that is, the core nanoparticle is coated with a gold shell by reduction of chloroauric acid (HAuCl4 ) onto the surface of the core. Au3+ can readily be reduced to Au(0) because the reduction potential of the AuCl4 − /Au pair is much higher than that of the standard hydrogen electrode (SHE): − − ◦ AuCl− 4 + 3e = Au(s) + 4Cl + 0.93 E (V)
(28.1)
Gold–gold sulfide nanoshells consisting of an Au2 S dielectric core surrounded by a gold shell (Au2 S@Au) were the first nanoshells studied [11, 12]. The nanoparticles were prepared by mixing HAuCl4 and sodium sulfide (Na2 S) with the size not more than 40 nm. Au2 S@Au nanoshells with different core and shell thicknesses were synthesized by varying the feed ratio between HAuCl4 and Na2 S. It was possible to shift the SPR peak to longer wavelengths, that is, from ∼520 nm to ∼900 nm, by varying the size of the nanoshells. However, this synthesis did not permit accurate control over the core and shell dimensions [13, 14]. Moreover, a large fraction of gold colloid was formed as a by-product of this synthesis, generating an additional absorption peak at ∼520 nm [13]. In order to permit control of the core and shell dimensions independently, other materials, such as semiconductors and dielectrics, are used as core particles. In particular, silica and polystyrene are commonly used as the core. Oldenburg et al. [14, 15] grew monodisperse silica nanoparticles via the St¨ober [14, 15] process as the dielectric cores. Organosilane molecules (3-aminopropyltriethoxysilane) were then adsorbed onto these nanoparticles.
P1: OTA/XYZ P2: ABC c28 JWBS037-Chen February 1, 2011
0:19
Printer Name: Yet to Come
SYNTHESIS, CHARACTERIZATION, AND TUNING
685
The organosilane molecules bonded to the surface of the silica nanoparticles, extending their amine groups outward as a new termination of the nanoparticle surface. After the silanecoated silica particles were isolated from residual reactants, a solution of very small gold colloid (1–2 nm in diameter) was added [16]. The gold nanoseeds bonded covalently to the organosilane linkage molecules via the amine group [17]. A subsequent reduction of an aged mixture of chloroauric acid and potassium carbonate by a solution of sodium borohydride (NaBH4 ), in which the gold-decorated silica nanoparticles were used as nucleation sites for the reduction, resulted in an increasing coverage of gold on the nanoparticle surface. The progression in metal nanoshell growth occurred within a few seconds during the reduction. Initially, the seed colloids increased in size as the reduction progressed. Then, the seed colloids began to coalesce on the nanoparticle surface, until finally the apparent formation of a continuous metallic nanoshell on the dielectric nanoparticle surface could be observed [14]. Recently, second-generation gold core–shell structured nanoparticles consisting of only a thin gold wall with a hollow interior were fabricated. These hollow gold nanospheres (HAuNS) were synthesized using a reductive metal as a template to form the core. By reducing the HAuCl4 , the template core is consumed, and the resulting metal ions diffuse into bulk solution, leaving a hollow interior. Liang et al. [18] first reported the formation of HAuNS with tunable interior-cavity sizes using cobalt (Co) nanoparticles as a reductive template. The preparation of Co nanoparticles was carried out according to the method reported by Kobayashi et al. [19]. In brief, CoCl2 solution was added into deaerated aqueous solution and reduced by NaBH4 in the presence of citric acid as stabilizer. The pink color of Co2+ solution changed to dark brown as Co colloids were formed. Hydrogen was evolved while NaBH4 was oxidized, a process that continued for several minutes. When the gas evolution ceased, the Co nanoparticle colloidal solution was transferred to stirred HAuCl4 aqueous solution. Since the reduction potential of the AuCl4 − /Au redox couple (0.935 V vs. SHE) is much higher than that of the Co2+ /Co redox couple (−0.377 V vs. SHE), AuCl4 − was reduced to Au atoms as soon as Co nanoparticles were added into the chloroauric acid solution: 2+ + 8Cl− 3Co + 2AuCl− 4 = 2Au + 3Co
(28.2)
The reduced Au atoms nucleated, grew into very small particles, and eventually evolved into a thin shell around the Co nanoparticles. The shells presumably had an incomplete porous structure at an early stage when Co2+ and AuCl4 − were able to continuously diffuse across the shell in reverse directions [18, 20]. When AuCl4 − was completely consumed, the remaining Co “core” was continually oxidized by H+ in aqueous solution (the reduction potential of the H+ /H2 redox couple is 0.0 vs. SHE). Because the reduced gold atoms were largely confined to the vicinity of sacrificial Co template’s outer surface, the size of the resulting HAuNS was governed by the initial size of the Co nanoparticles. The synthetic procedure is depicted in Figure 28.1 [21].
FIGURE 28.1 Schematic drawings of hollow gold nanosphere (HAuNS) synthesis.
P1: OTA/XYZ P2: ABC c28 JWBS037-Chen February 1, 2011
686
0:19
Printer Name: Yet to Come
THERANOSTIC APPLICATIONS OF GOLD CORE–SHELL STRUCTURED NANOPARTICLES
Sun et al. [22, 23] described the galvanic replacement reaction between Ag nanoparticles and an aqueous HAuCl4 solution as a simple and convenient route to synthesize metal nanostructures with hollow interiors and highly crystalline walls. The replacement transformed Ag nanocubes into Au nanoboxes and nanocages (nanoboxes with porous walls). The synthesis of Ag nanocubes was built upon the polyol process, in which a polyol (e.g., ethylene glycol) served as both solvent and reducing agent [24]. Generally, AgNO3 was reduced by ethylene glycol at an elevated temperature to generate Ag nanocrystals (seeds). In the presence of poly(vinylpyrrolidone), which selectively bound to the (100) surface of the Ag crystals, the additional Ag atoms grew preferentially at the (111) facets on the ends to form well-defined and controllable shapes [23, 25, 26]. This preference led to the formation of the sharp corners of the Ag crystals, which subsequently generated nanocubes ranging from 30 to 200 nm in edge length and with various degrees of corner truncation [24]. The driving force for the replacement reaction was the higher standard reduction potential of the AuCl4 − /Au redox couple than the AgCl/Ag redox couple (+0.22 V vs. SHE) [27]. The chemical reaction involved in the galvanic replacement was the following [28]: + − 3Ag(s) + AuCl− 4 (aq) = Au(s) + 3Ag (aq) + 4Cl (aq)
(28.3)
Xia and colleagues fully investigated the mechanism of this replacement [29, 30]. At the early stage, the silver nanocubes bounded by (100) facets reacted with the HAuCl4 to form small holes on a specific face. As the replacement reaction continued, Au atoms resulting from the above reaction were epitaxially deposited on the surface of the silver nanocube to generate a thin shell. The Ag atoms could also diffuse into the gold shell, leading to the formation of a closed box made of Au-Ag alloy. The (111) facets at the corners of the nanoboxes were dealloyed and further etched when more HAuCl4 solution was added. As consequence, a hole formed at each corner of the nanobox [24]. As alloying and dealloying proceeded, the Ag nanocube template was gradually dissolved, leading to a series of nanocages with increasing porosities. In later stages of dealloying, small holes could also be formed on (100) facets, resulting in breakage of the nanocage into small pieces [24]. Suzuki and Kawaguchi described an interesting method for the synthesis of core–shell structured gold nanoparticles in which the gold layer was fixed between a rigid core and a thermosensitive shell by using thermosensitive core–shell polymeric nanoparticles as templates [31]. To obtain the template nanoparticles, N-isopropylacrylamide (NIPAM), glycidyl methacrylate (GMA), and methylenebisacrylamide (crosslinker) were copolymerized in a soap-free aqueous medium using azobis(amidinopropane) dihydrochloride as an initiator. Because NIPAM and GMA had different reactivity, GMA tended to be consumed faster than NIPAM. Therefore the interior of the core particle was rich in poly(GMA), while the exterior was rich in thermosensitive poly(NIPAM) chains. After the introduction of amino groups to the surface of the core by addition of 2-aminoethanethiol, the gold nanoseeds were subsequently formed at the interface between the poly(GMA) core and the poly(NIPAM) shell. Additional electroless gold plating was carried out to control the thickness of the gold nanoshells. The absorption peaks of the resulting core–shell structured gold nanoparticles were in the range of 500–800 nm. 28.2.2 Tuning Surface Plasmon Resonance SPR is an optical phenomenon arising from the interaction between an electromagnetic wave and the conduction electrons in a metal [32]. When exposed to light, the conduction
P1: OTA/XYZ P2: ABC c28 JWBS037-Chen February 1, 2011
0:19
Printer Name: Yet to Come
SYNTHESIS, CHARACTERIZATION, AND TUNING
687
electrons in a metal nanostructure are driven by the electric field to collectively oscillate at a resonance frequency relative to the lattice of positive ions. At this resonance frequency, the incident light is absorbed by the nanostructure. Some of these absorbed photons are released with the same frequency in all directions, a process known as scattering. At the same time, a fraction of these photons are converted into phonons (vibrations of the lattice), a process known as absorption. In general, the SPR peak of a metal nanostructure should include both scattering and absorption components [30]. The optical properties of gold nanospheres are quantified in terms of their calculated absorption and scattering efficiency (Qabs and Qsca ) and their optical resonance wavelength (max ). The efficiencies Qi for the interaction of radiation with a scattering sphere of radius r are cross sections Ci normalized to the particle cross section, r2 , where i stands for extinction (i = ext), absorption (i = abs), or scattering (i = sca); thus [33] Qi =
Ci r 2
(28.4)
Energy conservation requires that Q ext = Q sca + Q abs
(28.5)
Cext = Csca + Cabs
(28.6)
or
According to the Mie theory, for a metal nanosphere with particle size much smaller than the wavelength of light (quasistatic and dipole limit), the nanoparticle’s extinction cross section (Cext ) is given by [2, 34, 35] 3/2
Cext =
⑀i 24 2 r 3 ⑀m · (⑀r + ⑀m )2 + ⑀i2
(28.7)
where ⑀ r and ⑀ i are, respectively, the real and imaginary components of the dielectric function of the metal, ⑀ m is the external environment dielectric function, r is the radius of the particle, and is a factor related to the eccentricity of the particle. From this equation, one can reasonably predict the position and shape of SPR for spherical and spheroidal metal particles. The optical extinction thus has a band maximum at the resonance condition roughly given by ⑀r + ⑀m = 0 for the noble metals such as Au, Ag, and Cu [36], provided that the imaginary component of the dielectric constant is small or weakly dependent on the frequency [37]. Whereas the real part of the dielectric component of the metal determines the wavelength position of the resonance, the imaginary component determines the bandwidth. Therefore the plasmon band is centered around a wavelength of 520 nm for small (<20 nm) Au nanocrystals dispersed in water (refractive index) 1/2 ⑀m = 1.33 [38]. For nanospheres of gold, Qabs and Qsca are calculated on the basis of the Mie theory for homogeneous spheres [34]. Because absorption and scattering together comprise the optical extinction by the particle, the Mie total extinction and scattering efficiency Qext and
P1: OTA/XYZ P2: ABC c28 JWBS037-Chen February 1, 2011
688
0:19
Printer Name: Yet to Come
THERANOSTIC APPLICATIONS OF GOLD CORE–SHELL STRUCTURED NANOPARTICLES
Qsca for a homogeneous sphere are expressed as infinite series [33]:
Q ext =
∞ 2 (2n + 1)Re(an + bn ) x 2 n=1
(28.8)
Q sca =
∞ 2 (2n + 1)(|an |2 + |bn |2 ) x 2 n=1
(28.9)
Q abs = Q ext − Q sca
(28.10)
where n is the refractive index of the sphere, an and bn are the Mie coefficients of the scattered field, and Re is the real part of the Mie coefficients. x is the size parameter given as x = 2n mr/, where nm is the refractive index of the surrounding medium. Numerical calculations of the Mie series are performed at discrete points in the wavelength range from 300 to 800 nm. The plasmon resonance of spherical solid gold nanoparticles varies only weakly with the particle size, shifting to longer wavelength as the particle size increases, and is also sensitive to the dielectric environment of the nanoparticle [32]. In contrast to what is observed with solid gold nanoparticles, the optical properties of gold core–shell nanostructures depend on the size and thickness of the metallic shell layer. The plasmon resonance frequency is determined by the relative sizes of the inner radius (r1 ) and outer radius (r2 ) of the shell [12]. Calculations of the optical absorption and scattering efficiencies of silica–gold nanoshells are performed by using a computer code employing Mie scattering for concentric sphere geometry developed by Ivan Charamisinau et al. [39]. The required parameters for the code are the value of the core and shell radii, r1 and r2 , and the complex refractive indices for the core, shell, and the surrounding medium, nc , ns , and nm , respectively. Jain et al. [33] used Mie theory and the discrete dipole approximation method to calculate absorption and scattering efficiencies and optical resonance wavelengths for solid gold nanospheres and silica–gold nanoshells. The calculated spectra clearly reflected the well-known dependence of nanoparticle optical properties (i.e., the resonance wavelength), the extinction cross section, and the ratio of scattering to absorption on the nanoparticle dimensions. When the size of the gold nanospheres was increased from 20 to 80 nm, the magnitude of extinction and the relative contribution of scattering to the extinction rapidly increased. Gold nanospheres (∼40 nm) showed an absorption cross section five orders higher than that of conventional absorbing dyes. The magnitude of light scattering by 80-nm gold nanospheres was five orders higher than the light emission from strongly fluorescing dyes. The plasmon wavelength maximum of nanospheres increased slightly from ∼520 to 550 nm when the size increases from 20 to 80 nm. Gold nanoshells were found to have extinction cross sections similar to and even higher than those of the nanospheres [30]. However, the resonance wavelength of gold nanoshells can be increased rapidly either by increasing the nanoshell size or by increasing the ratio of the core thickness to the shell radius. The relative contribution of scattering to the extinction can be increased rapidly by increasing the nanoshell size or decreasing the ratio of the core thickness to the shell radius [33]. HAuNS and nanocages, although they have different shape, are expected to have an absorption cross section of the same order of magnitude since they have similar composition, size, and shell thickness [20].
P1: OTA/XYZ P2: ABC c28 JWBS037-Chen February 1, 2011
0:19
Printer Name: Yet to Come
MOLECULAR OPTICAL IMAGING
689
Prodan et al. [40] and Halas [32] used “plasmon hybridization” theory to explain the remarkable tunability of the optical resonance of the nanoshells. The highly geometrydependent plasmon response can be seen as an interaction or “hybridization” between the essentially fixed-frequency plasmon response of a solid nanosphere and the plasmon response of a nanocavity. The sphere and cavity plasmons are electromagnetic excitations that induce surface charges at the inner and outer interfaces of the metal shell. Because of the finite thickness of the shell layer, the sphere and cavity plasmons interact with each other. The strength of the interaction between the sphere and cavity plasmons is controlled by the thickness of the metal shell layer. This interaction results in the splitting of the plasmon resonances into two new resonances: the lower-energy symmetric, or “bonding,” plasmon and the higher-energy antisymmetric, or “anti-bonding,” plasmon [40]. Only the lower-energy plasmon interacts strongly with an incident optical field. That is the reason why the nanoshell plasmon shifts to the lower energies (longer wavelengths) as the shell thickness is reduced. The theory of plasmon hybridization provided a general principle to guide the design of metallic nanostructures and prediction of their plasmon response qualitatively and quantitatively [32]. The construction of tunable nanoshells was achieved experimentally in a silica-cored gold nanoshell system [14]. When the diameter of the silica core was fixed at 120 nm, the plasmon resonance of the nanoparticles varied from ∼730 to ∼1020 nm as the core radius–shell thickness ratio was varied between 3 and 12 nm [14, 41]. Another group showed that the SPR peak could also be controlled by adjusting the interior-cavity sizes of HAuNS using Co as the displacement template. In this case, Co colloidal particles reduced AuCl4 − to form the gold shell outside. The shell formation was an inward growth process, in which the thickness of the gold shell increased inward as the replacement between Co nanoparticles and HAuCl4 continued. Therefore the outer diameters were controlled by the diameter of Co nanoparticles, whereas interior-cavity sizes were controlled by changing the stoichiometric ratio of HAuCl4 over the reducing agents [18]. For example, the SPR peak of HAuNS increased from 526 to 628 nm as the interior-cavity size was increased from 0 to 40 nm while the exterior size was fixed at ∼60 nm [18]. Schwartzberg et al. [20] modified the synthetic method and expanded the tunable peak of the surface plasmon band absorption to between 550 and 820 nm at a smaller exterior size of around 40 nm [20]. Similarly, the SPR peaks of gold nanocages can be tuned by controlling the degree of galvanic replacement reaction to generate walls of different thicknesses and porosities. In reported studies, as alloying and dealloying proceeded by titrating 30-nm silver nanocubes with different volumes of HAuCl4 solution, the positions of SPR peaks could be continuously adjusted from the visible light (ca. 400 nm for silver nanocubes) to the NIR light region up to 1200 nm [29, 30, 42].
28.3 MOLECULAR OPTICAL IMAGING 28.3.1 Optical Coherence Tomography Noninvasive in vivo imaging with light photons represents a powerful avenue for extracting relevant biological information. Whereas light in the visible range is routinely used for intravital microscopy [43], imaging of deeper tissues (>500 m) requires the use of NIR light. Hemoglobin and water, the major absorbers of visible and infrared light, respectively,
P1: OTA/XYZ P2: ABC c28 JWBS037-Chen February 1, 2011
690
0:19
Printer Name: Yet to Come
THERANOSTIC APPLICATIONS OF GOLD CORE–SHELL STRUCTURED NANOPARTICLES
have their lowest absorption coefficient in the NIR region around 650–900 nm [44]. Therefore NIR light permits deep optical penetration into biological samples [45]. Unique among the numerous diagnostic systems under investigation that employ NIR optical probes to enhance contrast between target and nontarget tissues, optical coherence tomography (OCT) offers extremely high resolution (typically 10–15 m), real-time data acquisition, and optical sectioning through biological tissues. The OCT technique is based on the Michelson interferometer, which measures the interference signal between light backscattered from the sample and a reference beam [1]. It detects the reflections of a low-coherence light source directed into a tissue and determines at what depth the reflections occurred. By employing a heterodyne optical detection scheme, OCT is able to detect very faint reflections relative to the incident power delivered to the tissue. In OCT imaging, out-of-focus light is strongly rejected owing to the coherence gating inherent to the technique. This permits imaging deeper tissues than can be imaged using alternative methods such as reflectance confocal microscopy, for which the out-of-focus rejection achievable is far less [1]. OCT is used clinically in ophthalmology and in imaging of aortae [46]. Core–shell structured gold nanoparticles have been shown to be highly efficient contrast agents for OCT owing to enhanced backscattering of laser radiation in the NIR region [47]. The backscattering efficiencies (Q back ) can be calculated as follows [48]:
Q back
1 = 2 x
∞ 2 n (2n + 1)(−1) (an − bn )
(28.11)
n=1
where n is the refractive index of the sphere, an and bn are the Mie coefficients of the scattered field, x is the size parameter given as x = 2n mr/, and is the wavelength in the surrounding medium. By varying the core radius and shell thickness, one can adjust the backscattering maximum for a given wavelength and obtain particles with preset optical properties [1]. This enables much higher contrast in OCT images acquired with gold nanoshells than in OCT images acquired with polystyrene nanospheres or saline solution. Significant contrast enhancement was observed in OCT images of tumors from mice injected with pegylated, silica-cored gold nanoshells [49]. In vivo OCT imaging of rabbit skin demonstrated significant contrast of the borders between the areas containing and not containing silica-cored gold nanoshells [47]. Topical application of gold nanoshells resulted in an increase in the intensity of the OCT signal in superficial layers of the skin (epidermis, superficial dermis) and an increase in the intensity of the OCT signal contrast between superficial and deep dermis and between hair follicles and glands. Numerical Monte Carlo simulations confirmed the possibility of contrasting skin layer boundaries and constituent structures by the application of gold nanoshells [47]. Adler et al. [50] proposed a new contrast-enhanced imaging technique based on phasesensitive OCT. Gold nanoshells were detected by inducing photothermal modulation with an 808-nm laser diode. Absorption of NIR light by the nanoshells resulted in a change in the optical path length, corresponding to phase modulation. This modulation was measured using a swept-source OCT phase microscope and detected at the modulation frequency using Fourier transform methods. High signal-to-noise ratio was achieved, indicating that the use of molecularly targeted nanoshell contrast agents with photothermal-modulated OCT imaging may enable highly sensitive and specific detection of diseases such as cancer [50].
P1: OTA/XYZ P2: ABC c28 JWBS037-Chen February 1, 2011
0:19
Printer Name: Yet to Come
MOLECULAR OPTICAL IMAGING
691
Compared with silica-cored gold nanoshells, gold nanocages can be prepared with much smaller dimensions (on the scale of 40 nm) while maintaining strong optical resonance peaks in the NIR region (800–1200 nm) [51]. Cang et al. [52] described gold nanocages as a new class of potential contrast agent for spectroscopic OCT. In conventional OCT, both the envelope of interference fringes and the peak intensity of the envelope are measured for imaging. In contrast, in spectroscopic OCT, a technique for revealing the wavelength-dependent backscattering at a given position, the interference fringes rather than the envelope are measured. Spatially resolved spectra can be obtained by short-time Fourier transform of the interferogram [30, 52]. A clear spectroscopic OCT image was acquired of a nanocage-doped phantom made of gelatin embedded with TiO2 granules to mimic the background scattering of typical biological tissues [52]. The depth-dependent OCT backscattered signal intensity at a single wavelength could readily be obtained from spectroscopic OCT analysis. 28.3.2 Photoacoustic Tomography Photoacoustic tomography (PAT) is a hybrid technology that images the internal distribution of optical energy deposition in biological tissues on the basis of the detection of laser-induced ultrasonic waves (photoacoustic or optoacoustic waves) [53, 54]. PAT provides higher spatial resolution in deep biological tissues than traditional optical imaging because ultrasonic scattering is two orders of magnitude less than optical scattering in such tissues [55, 56]. This lower scattering significantly reduces the signal intensity of background scattering and improves the image quality. PAT has been applied successfully to the visualization of different structures in biological tissues and has been especially useful in imaging the cerebral cortex of small animals [57, 58]. PAT is also suitable for monitoring the circulation of exogenous optical contrast agents with high sensitivity and specificity [45, 59]. Core–shell structured gold nanoparticles with optical resonance precisely tunable at the NIR region can generate greater photoacoustic wave signal and thus greater contrast than endogenous chromophores. Moreover, gold nanoshells are not susceptible to photobleaching, a problem commonly associated with the use of other organic dyes. Guillon et al. [60] investigated the acoustic vibration of gold nanoshells in colloidal solution by using a time-resolved pump-probe technique. The oscillation amplitude was significantly stronger and its period was considerably longer than those in pure gold nanospheres of the same size, which was ascribed to fundamental breathing vibration of the gold nanoshells. The damping time of the oscillations for gold nanoshells was much shorter than that for pure gold nanospheres, suggesting a faster energy transfer from nanoshells to the surrounding medium. Such distinct signatures allowed unambiguous identification of nanoshell acoustic vibration and separation of their contribution from that of other entities possibly present in a colloidal solution [60]. PAT had high spatial resolution and satisfactory sensitivity in rat brain in vivo when pegylated silica-cored gold nanoshells with a tunable absorption spectrum in the NIR region were used [45]. The images captured after three sequential administrations of nanoshells demonstrated a gradual enhancement of the optical absorption in the brain vessels, by up to 63% in total enhancement. Yang et al. [61] used pegylated gold nanocages for PAT of rat cerebral cortex in vivo. After three sequential administrations of nanocages, a gradual enhancement of the optical absorption in the cerebral cortex, by up to 81% in total enhancement, was observed. The advantages of Au nanocages over silica-cored Au nanoshells include their more compact size (<50 nm for Au nanocages compared to
P1: OTA/XYZ P2: ABC c28 JWBS037-Chen February 1, 2011
692
0:19
Printer Name: Yet to Come
THERANOSTIC APPLICATIONS OF GOLD CORE–SHELL STRUCTURED NANOPARTICLES
>100 nm for Au nanoshells) and larger optical absorption cross sections (3.6 × 10−14 vs. 2.7 × 10−14 m2 ) [61]. PAT using gold nanostructures targeted to molecular biomarkers associated with various diseases may become a viable tool for molecular interrogation of these diseases. In addition to being used for PAT of brain vasculature, NIR Au nanocages were recently used for imaging of lymph nodes. In a rat model, Song et al. [62] noninvasively identified a sentinel lymph node with enhanced contrast and good spatial resolution using Au nanocages as the contrast agent. In sentinel lymph node mapping, Au nanocages could be detected as deep as ∼33 mm. Compared to the conventional sentinel lymph node mapping methods, the Au nanocages-based PAT method offers several advantages: (1) noninvasive mapping, (2) peak optical absorption in the NIR region, allowing deeper imaging penetration than in the visible range, (3) rapid drainage of the contrast agent into lymphatic channels, (4) accumulation of Au nanocages in excess of the original concentration, generating stronger photoacoustic signals (high signal-to-noise ratio), and (5) spatial resolution with an axial resolution of 15 m and a lateral resolution of 45 m in its focal zone [54, 62, 63].
28.4 PHOTOTHERMAL THERAPY IN CANCER 28.4.1 Photothermal Effect When the incident light frequency matches the intrinsic electron oscillation frequency of a gold core–shell nanostructure, light is absorbed, resulting in SPR absorption. SPR absorption in Au nanoparticles is followed by energy relaxation through nonradioactive decay channels. This results in an increase in kinetic energy, leading to overheating of the local environment around the light-absorbing species. This phenomenon of rapid conversion of absorbed light into heat is known as the photothermal effect [3, 64, 65]. Using femtosecond transient absorption spectroscopy, Link and El-Sayed [66] showed that the photoexcitation of metal nanostructures results in the formation of a heated electron gas that subsequently cools rapidly—within ∼1 picosecond—by exchanging energy with the nanoparticle lattice. This is followed by phonon–phonon interactions in which the nanoparticle lattice cools rapidly by exchanging energy with the surrounding medium within ∼100 ps [66]. This mechanism has also been confirmed in gold nanocages using ultrafast laser spectroscopy or time-resolved spectroscopy as well as vibrational spectroscopy [67, 68]. The photothermal effect in gold nanostructures is highly efficient. Hot electron temperatures of several thousand kelvins are easily reached in the nanoparticles, even with laser excitation powers as low as 100 nJ, and lattice temperatures on the order of a few tens of degrees can be achieved [66]. The absorption cross section of gold nanoparticles is several orders of magnitude stronger than that of the strongest-absorbing rhodamine 6G dye molecules [33, 69]. Conventional NIR dyes like indocyanine green possess an absorption cross section of ∼1.66 × 10−20 m2 at wavelengths of ∼800 nm, while silica-cored gold nanoshells possess absorption cross sections on the order of 3.8 × 10−14 m2 . Thus such nanoshells are over 1 million-fold more likely than the comparable dye to encounter an absorbing event and convert that light into thermal energy [7]. The photothermal heating of gold nanostructures is much different from conventional thermal heating. At a much lower temperature (∼523 K) than their corresponding melting points (1337 K), gold nanocages become nanospheres; in contrast, gold nanocages maintain
P1: OTA/XYZ P2: ABC c28 JWBS037-Chen February 1, 2011
0:19
Printer Name: Yet to Come
PHOTOTHERMAL THERAPY IN CANCER
693
their shapes under ultrafast laser pulse excitation with very high temperature jumps. In one experiment, gold nanocages remained intact at 17 J per pulse, corresponding to a lattice temperature around 1100 ± 100 K [30]. This stability is due to the rapid heat dissipation in solution. 28.4.2 Photothermal Therapy Photothermal therapy (PTT) is a therapeutic modality in which photothermal agents are employed to achieve selective heating of the local environment [70–73]. The heat produced can lead to localized temperatures far above the threshold temperature (∼330 K) that causes irreversible cell death [30]. This method is very appealing because it is a physical treatment and therefore associated with fewer side effects compared with conventional cancer treatment strategies such as radiotherapy and chemotherapy. PTT can also be performed repeatedly without the accumulation of toxic side effects. Mild increases in temperatures can also boost the efficacy of radiotherapy and chemotherapy [74]. Gold nanostructures are ideally suited as photothermal coupling agents. In addition to enhanced absorption cross sections of the gold nanostructures that ensure effective laser therapy at relatively low laser output energies, the SPR absorption band of the core–shell structured nanoparticles is readily tunable in the NIR region, which makes it possible to deliver thermal dose to deeper tissues. Metal nanostructures also display much higher photostability than organic dyes, and their surface can readily be modified with receptor ligands to direct nanoparticles to the target sites, allowing selective thermal ablation of targeted tumor cells. The effectiveness of PTT was studied in human prostate cancer cells in vitro with laseractivated silica-cored gold nanoshells in which the nanoshells were taken up by cells via nonspecific phagocytosis or pinocytosis. It was shown that a ratio of at least 5000 nanoshells per human prostate cancer cell was critical for achieving efficient cell killing effect [75–77]. The efficacy of PTT can be significantly improved with targeted nanoparticles. Loo et al. [1, 79] and Lowery et al. [78] demonstrated targeted photothermal destruction of human breast cancer cells in vitro using immunoconjugated nanoshells directed against human epidermal growth factor receptor 2 (HER2). The anti-HER2 antibodies were bound to a polyethylene glycol (PEG) linker, orthopyridyl-disulfide-PEG, (OPSS)-PEG-NHS, through an activated ester, N-hydroxysuccinimide (NHS), which in turn was attached to the surface of silica-cored gold nanoshells through the sulfur-containing OPSS group located at the distal end of the PEG linker. The immunoconjugated nanoshells mediated effective PTT in vitro when breast cancer cells were irradiated with an NIR laser (820 nm) at 35 W/cm2 for 7 min [1, 78]. In another study, treatment with HER2-targeted nanoshells, but not with nontargeted nanoshells, followed by exposure to laser light induced cell death of medulloblastoma Daoy.2 cells, which overexpress HER2 receptors. Combined treatment with HER2-targeted nanoshells and NIR laser was not effective against dermal fibroblasts, which do not express HER2 [80]. In yet another study, gold nanocages with a relatively small size (e.g., ∼45 nm in edge length) were conjugated with anti-HER2 antibodies. Theoretical calculations showed that the nanocage had an absorption cross section of 3.48 × 10−14 m2 , similar to that of silica-cored nanoshells (3.8 × 10−14 m2 ) [7, 81]. Preliminary results showed that the immunonanocages strongly absorbed light in the NIR region, with an output power threshold of 1.5 W/cm2 required to induce thermal destruction in the cancer cells. In the range of 1.5–4.7 W/cm2 , the percentage of damaged cells increased linearly
P1: OTA/XYZ P2: ABC c28 JWBS037-Chen February 1, 2011
694
0:19
Printer Name: Yet to Come
THERANOSTIC APPLICATIONS OF GOLD CORE–SHELL STRUCTURED NANOPARTICLES
with irradiation power density [81]. Bernardi et al. [80] conjugated an antibody against interleukin-13 receptor-alpha 2 (IL13R␣2), an antigen that is frequently overexpressed in gliomas, to gold–silica nanoshells. They showed that these immunonanoshells were capable of inducing cell death in two high-grade glioma cell lines (U373 and U87) that express IL13R␣2, but not in A431 epidermoid carcinoma cells, which do not express significant levels of IL13R␣2 [80]. We used 40-nm-diameter HAuNS as an efficient photothermal coupling agent. We carefully designed the size of our agent because the particle size is of utmost importance for efficient cellular uptake [82]. Particles larger than 100 nm may not readily enter cells and may interrupt some cellular functions. Similarly, particles that are too small, less than 20 nm, may not be retained by cells. Thus nanostructures with sizes between 20 and 100 nm may be favorable [20]. Moreover, the small size of HAuNS is advantageous when they are used for targeted delivery [21, 83]. In our studies, C225, a monoclonal antibody directed at epidermal growth factor receptor (EGFR), was covalently attached to HAuNS (C225HAuNS). When an aqueous solution of C225-HAuNS (7.3 × 1010 nanoshells/mL) was exposed to an NIR laser at an output power of 8 W/cm2 , the temperature of the solutions ◦ increased 16.5 C after ∼4 min of light exposure. Without HAuNS, little temperature change was observed at the same laser power and duration of exposure [83]. Our in vitro results showed EGFR-mediated selective uptake of C225-HAuNS in EGFR-positive A431 tumor cells but no uptake of IgG-HAuNS control. Irradiation of A431 cells treated with C225-HAuNS plus NIR laser resulted in selective destruction of these cells. In contrast, treatment with C225-HAuNS alone, laser alone, or IgG-HAuNS plus laser did not result in any observable effect on cell viability (Fig. 28.2) [83]. The efficacy of photothermal ablation mediated by locally injected nanoshells has also been assessed in several in vivo studies [7, 84]. Magnetic resonance thermal imaging of mice with solid tumors directly injected with pegylated silica-cored gold nanoshells revealed that exposure to low doses of NIR light (820 nm, 4 W/cm2 ) resulted in average maximum temperatures capable of inducing irreversible tissue damage (T = 37.4 ± 6.6 ◦ C) within 4–6 min [7]. Controls treated without nanoshells demonstrated significantly lower average temperatures on exposure to NIR light (T < 10 ◦ C). Histological evaluations showed that the tissues heated to above the thermal damage threshold displayed coagulation, cell shrinkage, and loss of nuclear staining, which were indicators of irreversible thermal damage. However, little or no damage was found in surrounding tissue. This study also provided information about the relationships between nanoshell dosages, light intensity, and duration of illumination and the ultimate thermal profile and resultant tissue damage [7]. In another study, the nanoshells were applied for PTT of spontaneous tumors, including cutaneous squamous cell carcinoma of the external ear canal, squamous cell carcinoma of the oral mucosa, basal cell cutaneous tumor of the external ear canal, and malignant melanoma of the oral mucosa. In all cases, effective optical destruction of cancer cells was demonstrated by local injection of gold nanoshells followed by continuous-wave semiconductor laser irradiation at a wavelength of ∼800 nm [84]. Importantly, direct injection of nanoshells into the tumor mass may not be feasible when lesions become inaccessible to intratumoral injection and when uneven intratumoral distribution of nanoparticles owing to intratumoral distribution compromises the effectiveness of treatment. In such cases, it is necessary to administer gold nanostructures systemically. O’Neal et al. [85] intravenously injected silica-cored gold nanoshells (∼130 nm in average diameter) conjugated with PEG into mice bearing murine colon carcinoma and investigated the tumor accumulation of the nanoshells and the efficacy of subsequent
P1: OTA/XYZ P2: ABC c28 JWBS037-Chen February 1, 2011
0:19
Printer Name: Yet to Come
PHOTOTHERMAL THERAPY IN CANCER
695
FIGURE 28.2 Selective photothermal ablation of A431 cells by C225-HAuNS. (a) Cell viability after various treatments. Cells retained normal morphology with no apparent death observed (stained green with calcein AM) when cells were not treated or treated with C225-HAuNS alone, NIR laser alone, or nontargeted IgG-HAuNS plus NIR laser. In contrast, most cells were dead after treatment with C225-HAuNS plus NIR laser. Dead cells were labeled red with EthD-1. Magnification, 40×. (b) Images of untreated viable cells and dead cells treated with C225-HAuNS and NIR laser at higher magnification (400×). The dead cells showed rounded morphology (asterisk) and membrane damage as indicated by positive staining with EthD-1 (arrow, red). DIC, differential interference contrast.
photothermal ablation. These PEG-conjugated silica-cored gold nanoshells accumulate in tumors via a passive mechanism referred to as the “enhanced permeability and retention” effect [86–89]. O’Neal et al. [85] found that this extravasation of agent into the abnormal and leaky vasculature, termed passive targeting, resulted in complete regression of tumors within 10 days following treatment with nanoshells and NIR light (808 nm, 4 W/cm2 , 3 min), and the mice appeared healthy and tumor free >90 days later. Survival times for mice with the nanoshell treatment in this study were significantly improved compared to survival times for untreated mice or those receiving laser treatment alone [85]. In a similar study, Stern et al. [90] evaluated the antitumor efficacy of systemic delivery of the pegylated silica-cored gold nanoshells in a subcutaneous prostate cancer model. Differences in temperature changes between the treated and untreated groups offered indirect evidence of nanoshell accumulation in tumor. After 3 min of laser activation, temperatures up to 65.4 ◦ C (a 35 ◦ C increase from baseline) were achieved in the treated group. However, despite an average 14 ◦ C increase in temperature, no tumor ablation effect was observed in the control mice treated only with NIR irradiation [90]. In passive targeting, particle size is a critical characteristic. It was shown that smaller gold nanoparticles (20–40 nm) coated with PEG had longer blood circulation time than larger gold nanoparticles (80 nm). The 20-nm gold nanoparticles exhibited the lowest uptake by
P1: OTA/XYZ P2: ABC c28 JWBS037-Chen February 1, 2011
696
0:19
Printer Name: Yet to Come
THERANOSTIC APPLICATIONS OF GOLD CORE–SHELL STRUCTURED NANOPARTICLES
reticuloendothelial cells and the slowest clearance from the body [91]. Moreover, smaller particles have a better chance of extravasating from blood vessels into extravascular fluid space. In certain tumors, such as glioma and ovarian cancer, pore cutoff size was between 7 and 100 nm [86, 92]. Therefore, to improve targeted delivery of gold nanostructures, it is desirable that the size of these nanoparticles be less than 80 nm. Toward this end, HAuNS, with a diameter of ∼40 nm, have been identified as a promising photothermal mediator. Attachment of targeting molecules that possess high affinity for unique molecular signatures found in malignant cells has been found to improve the residence time of the agent in tumors. Melancon et al. [83] used C225-HAuNS as the immunonanoshells targeting EGFR overexpressing A431 human squamous carcinoma in mice. Radiotracer counting for tumor biodistribution showed that C225-HAuNS had a higher uptake value in the A431 tumors than did IgG-conjugated HAuNS (control) at 24 h after injection, representing a 48% increase for targeted HAuNS. There was a greater than threefold increase in the number of nanoshells and/or nanoshell aggregates per field observed under dark-field microscope in the perivascular area of the tumor in mice injected with C225-HAuNS than in mice injected with IgG-HAuNS, suggesting that the interaction between C225-HAuNS and EGFR may have facilitated extravasation of the nanoshells into the interstitial space [83]. In addition to being conjugated with an antibody, HAuNS have also been conjugated with a small-molecular-weight peptide, [Nle4 ,D-Phe7 ]␣-MSH (NDP-MSH), as a targeting moiety through a PEG linker [21]. NDP-MSH is a potent agonist of melanocortin type-1 receptor, which is overexpressed in many melanoma cells [93], and binds to the receptor with high affinity (IC50 = 0.21 nM) [94]. The tumor uptake of NDP-MSH-conjugated pegylated HAuNS was significantly higher than that of control pegylated HAuNS without peptide conjugation (12.6 ± 3.1%ID/g versus 4.3 ± 1.2%ID/g) at 4 h after intravenous injection in B16/F10 melanoma-bearing nude mice [21]. This suggested that the targeted HAuNS had enhanced extravasation from tumor blood vessels and dispersion into the tumor matrix. Furthermore, the PTT effect of the nanoparticles was confirmed both histologically, by examination of excised tissue, and functionally, by [18 F]fluorodeoxyglucose positron emission tomography (Fig. 28.3). The successful active targeting of NDP-MSH-conjugated HAuNS to melanoma suggested their potential application in targeted PTT for melanoma [21].
28.5 NEAR-INFRARED LIGHT-TRIGGERED DRUG DELIVERY Gold nanostructures with high photothermal coupling efficiency may be incorporated into hydrogels and liposomes to produce controlled-drug-release systems that can be induced by NIR light [95, 96]. Sershen et al. [95] developed composite hydrogels formed from N-isopropylacrylamideco-acrylamide (NIPAAm-co-AAm) and NIR-absorbing silica-cored gold nanoshells. The incorporation of the nanoshells transformed a thermoresponsive polymer into a photothermally responsive nanoshell–polymer composite. NIPAAm-co-AAm formed hydrogel materials with a lower critical solution temperature ranging from 32 ◦ C to 60 ◦ C, a temperature at which the hydrogel material underwent a dramatic phase change [97]. When the thermoresponsive polymer was heated owing to absorption of light at 832 nm by nanoshells embedded in the hydrogel, the polymer underwent a reversible decrease in volume. This induced a reversible and repeatable light-driven collapse of the composite with a weight change of 90% after illumination at 1.8 W/cm2 . The degree of shrinkage of the hydrogel was controlled by the laser fluence at 808 nm as well as by the concentration of
P1: OTA/XYZ P2: ABC c28 JWBS037-Chen February 1, 2011
0:19
Printer Name: Yet to Come
NEAR-INFRARED LIGHT-TRIGGERED DRUG DELIVERY
697
FIGURE 28.3 In vivo photothermal ablation (PTA) effect of the targeted NDP-MSH-PEG-HAuNS inducing selective destruction of B16/F10 melanoma in nude mice. (a) [18 F]FDG PET imaging shows significantly reduced metabolic activity of the tumors after PTA therapy only in mice pretreated with NDP-MSH-PEG-HAuNS. [18 F]FDG PET study was conducted before (0 h) and 24 h after laser irradiation, which was commenced 4 h after intravenous injection of NDP-MSH-PEG-HAuNS, PEGHAuNS, or saline control. For each mouse, one flank of tumors was randomly selected for laser treatment (0.5 W/cm2 at 808 nm for 1 min). T; tumors; arrowheads, tumors irradiated with NIR light. (b) Histological assessment of tumor necrosis. The whole tumor from each mouse 24 h after NIR irradiation was removed, fixed, and stained with H&E (bar = 500 m). Representative images show that in the group treated with NDP-MSH-PEG-HAuNS plus laser, most of the tumor cells were necrotic, indicated by pyknosis (arrows), karyolysis (arrowheads), cytoplasmic acidophilia, and degradation and corruption of the extracellular matrix of the tumor (asterisks). Only a small fraction of necrotic zone was found in the tumors treated with PEG-HAuNS plus laser, laser alone, and tumor without any treatment. Bar = 50 m.
SiO2 -Au nanoshells [98]. Modulated drug delivery profiles for methylene blue, insulin, and lysozyme were demonstrated by irradiating the drug-loaded nanoshell-composite hydrogels, which showed that drug release was dependent on the molecular weight of the therapeutic molecules. The system demonstrated the ability to switch between a passive, diffusional rate of drug delivery and an active, accelerated rate of drug delivery in response to NIR irradiation. This system has the potential to provide a simple, minimally invasive
P1: OTA/XYZ P2: ABC c28 JWBS037-Chen February 1, 2011
698
0:19
Printer Name: Yet to Come
THERANOSTIC APPLICATIONS OF GOLD CORE–SHELL STRUCTURED NANOPARTICLES
alternative to serial injections when repetitive dosing is required. Additionally, alteration of the irradiation parameters, hydrogel composition, and loading methods can match the drug release profile to the requirements of an individual treatment [98]. Incorporation of Au-Au2 S nanoshells into NIPAAm-co-AAm hydrogel has also been reported [96]. The extinction spectra of the composite were dictated by the nanoshells, while the phase transition characteristics of a NIPAAm-co-AAm copolymer with a lower critical solution temperature of 40 ◦ C were maintained in the composite. Significantly enhanced drug release from composite hydrogels was achieved in response to irradiation by pulsed Nd:YAG laser light at 1064 nm. When the temperature of the copolymer exceeded the lower critical solution temperature, the hydrogel collapsed, causing a burst release of embedded methylene blue, ovalbumin, and bovine serum albumin. Additionally, nanoshellcomposite hydrogels can have multiple pulsatile delivery of protein in response to repeated NIR irradiation [96]. Besides being incorporated into hydrogels, gold nanoshells can be incorporated into liposomes and used to produce hyperthermia in cancer cells [99]. In one study, the liposomes were made of cationic PEG-lipid conjugates, which included a hydrophobic lipid anchor of dipalmitoylphosphatidylethanolamine, a hydrophilic spacer of PEG, and a cationic head group [100]. Encapsulating the silica-cored gold nanoshells, the liposomes were used as delivery vehicles to shuttle the nanoshells into human MCF-7 mammary carcinoma cells via endocytosis, and then cells were exposed to NIR light peaked at 785 nm. Liposomal delivery enhanced the intracellular bioavailability of gold nanoshells and was able to induce a higher degree of cell death in vitro more effectively than free-standing gold nanoshells. Liposomes facilitated tropism and uptake by endocytosis due to the similarity of their constituents and that of the cellular membrane. Following uptake, the liposomes were eventually destroyed and released their contents, gold nanoshells, into the cytoplasm. Wu et al. [101] synthesized small HAuNS that were either encapsulated within liposomes (by an interdigitation-fusion method) or tethered to the liposome membrane with a Au-SHPEG-lipid linker. HAuNS with a maximum absorption at 820 nm were then coated with 750-Da PEG-thiol to enhance particle stability and were concentrated by ultracentrifugation. The diameter of the HAuNS was 33 ± 13 nm with shell thickness of 3.4 ± 0.9 nm. HAuNS were encapsulated within dipalmitoylphosphatidylcholine liposomes together with a fluorescent dye, 6-carboxyfluorescein, used as a soluble model drug. It was proved that the pulsed NIR light absorbed by HAuNS triggered the near-instantaneous release of liposome contents, providing precise spatial and temporal control. The laser-heated HAuNS acted as optically triggered nano-“sonicators” to temporally disrupt the lipid membrane. HAuNS tethered to, encapsulated within, or suspended freely outside liposomes all induced liposome disruption; however, tethered HAuNS achieved the highest release efficacy owing to their proximity to the lipid membrane. With this new NIR-activated release, disease cells can be synergistically targeted by combining drug-carrying particles (liposomes) and energy-absorbing particles (HAuNS). Continued irradiation of the HAuNS can induce localized hyperthermia or permeabilize cell membranes, both of which can facilitate the cellular uptake of large macromolecules, including proteins and DNA [101].
28.6 MULTIFUNCTIONAL CORE–SHELL STRUCTURED GOLD NANOPARTICLES AS THERANOSTIC AGENTS One of the promises of nanotechnology is its ability to provide multiple functions that small molecular compounds cannot. In particular, integrating diagnostic and therapeutic
P1: OTA/XYZ P2: ABC c28 JWBS037-Chen February 1, 2011
0:19
Printer Name: Yet to Come
MULTIFUNCTIONAL CORE–SHELL STRUCTURED GOLD NANOPARTICLES
699
capabilities into a single entity allows tailoring nanomedicine to individual patients on the basis of successful delivery and detection of nanoparticles at the diseased sites. Magnetic resonance imaging (MRI) is widely used in designing nanoparticles for theranostic applications because of its high spatial and temporal resolution, lack of ionizing radiation, and relatively high detection sensitivity compared to computed tomography. Superparamagnetic iron oxide (SPIO) nanoparticles have a high transverse (or spin–spin) relaxivity (R2) that results in negative contrast on T2-weighted images. To date, several SPIO-based MRI contrast agents have been approved for human use or are being studied in clinical trials [102]. Therefore nanostructures combining the magnetic properties of SPIO and the optical properties of gold nanoshells are expected to provide a novel platform for simultaneous diagnostic imaging and photothermal treatment of cancer. Ji et al. [103] described the synthesis, characterization, and use of hybrid nanoparticles with a SPIO-silica core and a gold nanoshell. These nanoparticles were prepared by first synthesizing silica-coated SPIO (Fe2 O3 ) nanoparticles according to the St¨ober process [15]. The SPIO nanoparticles (average diameter, 10 nm) stabilized with oleic acid in water were coated with amorphous silica via the sol-gel process. The thickness of the silica sphere could be tuned from 2 to 100 nm simply by changing the concentration of the sol-gel precursor, tetraethylorthosilicate [104]. The shell of silica-coated SPIO particles was functionalized with amine groups by treatment with NH4 OH and 3-aminopropyltrimethoxysilane. Gold nanocrystal seeds (2–3 nm) were then attached to the amino groups on the silica sphere by reduction of HAuCl4 with tetrakis(hydroxymethyl)phosphonium chloride [16]. Because the gold nanoseeds had net negative surface charges, they firmly attached to the amino groups on the silica sphere, which were positively charged at acidic pH. Finally, the attached gold nanoseeds were used to nucleate the growth of a gold layer on the silica surface to form a gold nanoshell [103]. The resulting multifunctional nanoparticles, SPIO@silica@Au, displayed superparamagnetic characteristics as well as a significant absorbance in the NIR region of the electromagnetic spectrum. They exhibited high transverse relaxivities, R2, and a large R2/R1 ratio, and therefore they could be imaged by MRI to obtain T2-weighted images. Moreover, the SPIO@silica@Au nanoshells showed efficient photothermal effects when exposed to NIR light. It is expected that the efficacy of dual-function nanoshellmediated PTT may be significantly enhanced because such nanoshells permit real-time in vivo MRI imaging of the distribution of the nanoparticles before, during, and after PTT. In addition to silica coating, paramagnetic iron oxide (Fe3 O4 ) nanoparticles were also embedded in a polymer layer, which in turn served as an interface for the introduction of the gold shell [105]. The polymer layer was formed through copolymerization of methacrylic acid and acrylamide [106]. Because the surface of the resulting intermediate Fe3 O4 @polymer nanoparticles was negatively charged, the nanoparticles were coated with another thin layer of positively charged polyelectrolyte, poly(allylamine hydrochloride), through electrostatic interaction. Poly(allylamine hydrochloride) provided a positively charged surface for electrostatic attachment of gold nanoseeds. Finally, gold shells were formed upon reduction of chloroauric acid in the presence of gold nanoseed-coated Fe3 O4 @polymer nanoparticles. The resulting Fe3 O4 @polymer@Au core–shell structured gold nanoparticles, 273±17.9 nm in diameter, showed both a strong magnetization (saturation magnetization, M s of Fe3 O4 @polymer@Au was 61.0 emu·g−1 compared with that of Fe3 O4 core 75.6 emu·g−1 ) and optical absorption in the NIR region [105]. In an alternative design, Kim et al. [107] described the assembly of Fe3 O4 (magnetite) nanoparticles and gold seed nanoparticles onto amino-modified silica spheres, followed by the growth of gold shells around the silica spheres. Thus amine-terminated silica nanoparticles were coated with 2-bromo-2-methylpropionic acid-stabilized 7-nm Fe3 O4
P1: OTA/XYZ P2: ABC c28 JWBS037-Chen February 1, 2011
700
0:19
Printer Name: Yet to Come
THERANOSTIC APPLICATIONS OF GOLD CORE–SHELL STRUCTURED NANOPARTICLES
nanoparticles [108, 109] through a direct nucleophilic substitution reaction between the bromo groups in Fe3 O4 nanoparticles and the amino groups in silica nanoparticles. Gold seed nanoparticles of 1–3 nm were attached to the residual amino groups of the Fe3 O4 coated silica spheres. Finally, a complete 15-nm-thick gold shell with embedded Fe3 O4 nanoparticles was formed around the silica spheres [107]. These dual-function core–shell structured gold nanoparticles were further tagged with anti-HER2 antibodies. In vitro study demonstrated targeted imaging and selective ablation of breast cancer cells overexpressing HER2 receptors using the antibody-coated silica-Fe3 O4 @Au nanoparticles [107]. Paramagnetic iron oxide nanoparticles may also be encapsulated in yolk–shell gold nanostructures using silver-coated Fe3 O4 nanoparticles as templates [110]. Specifically, monodisperse Fe3 O4 nanoparticles 9–11 nm in diameter, first deposited with gold layers with thickness of 2–3 nm, were sequentially coated with silver for a subsequent replacement reaction [111]. The yolk–shell nanostructures were finally formed by displacing silver layers with gold layers while HAuCl4 was reduced to gold by silver. The resonance peak of the resulting Fe3 O4 @Au yolk–shell nanostructure could be tuned to the NIR region by controlling the thickness of the outer gold layer, which in turn was determined by the thickness of the silver layer. The novel nanostructures possessed both NIR-absorbing (and scattering) optical characteristics and magnetic properties for T2-weighted MRI. Multifunctional gold nanoparticles were also designed for photothermally controlled drug delivery and MRI contrast enhancement using biodegradable polymer and metal to form multilayer half-shell nanoparticles [112]. In this study, rhodamine as a model drug was encapsulated within biocompatible and biodegradable poly(lactic-co-glycolic acid) (PLGA) copolymer nanoparticles. Manganese and gold layers were deposited on these nanoparticles. Since the physical deposition method yielded half-shells, the drug was released from the open half of the shell. The gold layer provided a photothermal conversion effect under NIR light irradiation, causing temperature elevation, which accelerated the release rate of rhodamine from the PLGA nanoparticles to about twice the release rate seen without NIR irradiation. The manganese layer provided contrast enhancement in T2-weighted MR images.
28.7 PERSPECTIVES Gold core–shell nanostructures have a fascinating optical property referred to as the SPR. Depending on the size and composition of the core and shell of the nanostructures, they can be designed to either absorb or scatter light over much of the visible and infrared regions of the electromagnetic spectrum, including the NIR region, where penetration of light through tissue is maximal. As these particles are readily conjugated to antibodies and other biomolecules, they are effective substrates for mediating a variety of potential applications in biomedicine, including molecular optical imaging, laser-triggered controlled drug delivery, and photothermal ablation therapy. (See Table 28.1). One can envision a myriad of potential applications in theranostics for cancer. It is expected that these gold core–shell structured nanoparticles would enable acquisition of in vivo imaging data using OCT, PAT, and MRI in preoperative, intraoperative, and postoperative settings as well as image-guided surgery and treatment, such as PTT and NIR-triggered drug treatment. To date, most research work gold core–shell structured nanoparticles stands at the proof-of-concept stage. Detailed preclinical studies should be conducted, especially in pharmacokinetics and in vivo tumor targeting and therapeutic effect. Furthermore, the metabolism of the gold nanoparticles as well as acute and long-term toxicity needs to
P1: OTA/XYZ P2: ABC c28 JWBS037-Chen February 1, 2011
0:19
Printer Name: Yet to Come
PERSPECTIVES
701
TABLE 28.1 Summary of Theranostic Application of Core–Shelled Gold Nanostructures Nanostructure
Characteristics
Silica-cored gold nanoshell
Size: 150-nm diameter silica core and 25-nm-thick gold shell Absorbance: 850–950 nm Size: 291-nm diameter silica core and 25-nm-thick gold shell Absorbance:1310 nm Size: 200-nm diameter silica core and 20-nm-thick gold shell Absorbance: 800 nm Size: 119-nm diameter silica core and 12-nm-thick gold shell Absorbance: 800 nm Surface: PEG Size: 125-nm diameter silica core and 10-12-nm-thick gold shell Absorbance: 800 nm Surface: PEG Size: 110-nm diameter silica core and 10-nm-thick gold shell Absorbance: 815 nm Surface: anti-HER2 antibody Surface: anti-HER2 antibody; anti-IL13R␣2 antibody Size: 110-nm diameter silica core and 10-nm-thick gold shell Absorbance: 815-nm Size: 110-nm diameter silica core and 10-nm-thick gold shell Absorbance: 815 nm Surface: PEG Size: 110-nm diameter silica core and 10-nm-thick gold shell Absorbance: 815 nm Surface: PEG
Application Optical coherence tomography
Reference
Rabbit; applied on the skin
47
In vitro
48
In vitro
1
CT-26 murine colon carcinomabearing mouse; IV injection
49
Photoacoustic tomography
Rat; IV injection
45
Photothermal therapy
PC-3 and E4-2 human prostate cancer cell
75–77
SKBr3 breast cancer cell Medulloblastoma cell line Daoy.2; U373 and U87 glioma cell Rat; subcutaneous injection
1, 78, 79
Canine transmissible venereal tumor bearing mouse; Intratumor injection CT26.WT murine colon carcinoma bearing mouse; IV injection
80
84
7
85
(Continued )
P1: OTA/XYZ P2: ABC c28 JWBS037-Chen February 1, 2011
702
0:19
Printer Name: Yet to Come
THERANOSTIC APPLICATIONS OF GOLD CORE–SHELL STRUCTURED NANOPARTICLES
TABLE 28.1 (Continued ) Nanostructure
Characteristics Size: 110-nm diameter silica core and 10-nm-thick gold shell Absorbance: 815 nm Surface: PEG Size: 100-nm diameter silica core and 7-nm-thick gold shell Absorbance: 832 nm Size: 119-nm diameter silica core and 14-nm-thick gold shell Absorbance: 804 nm Absorbance: 785 nm
Gold nanocage
Hollow gold nanosphere
Au–Au2 S nanoshell
Hollow gold nanosphere
Size: 36.7-nm edge length and 3.3-nm wall thickness Absorbance: 800 nm Size: ∼50-nm edge length Absorbance: 820 nm Surface: PEG Size: ∼50-nm edge length Absorbance: 735 nm Surface: PVP Size: 45-nm edge length and 3.5-nm wall thickness Absorbance: 810 nm Surface: anti-HER2 antibody Size: 30.4-nm diameter Surface: C225 monoclonal antibody Absorbance: 808 nm Size: 43.5-nm diameter Surface: melanocyte stimulating hormone analog Absorbance: 808 nm Size: 37-nm diameter silica core and 4-nm-thick gold shell Size: 33-nm diameter silica core and 3.4-nm-thick gold shell Absorbance: 820 nm
Application
Reference
PC-3 bearing mouse; IV injection
90
Nanoshells@poly (NIPAAm-coAAm) hydrogel Laser: 832 nm Nanoshells@poly (NIPAAm-coAAm) hydrogels Laser: 808nm Nanoshells@ liposomes Laser: 785 nm In vitro
95
51, 52
Rat; IV injection
61
Rat; left forepaw pad injection
62
Photothermal Therapy
SKBr3 breast cancer cells
81
Photothermal Therapy
In vitro A431 human squamous carcinoma
83
B16/F10 murine melanoma bearing mouse; IV injection
21
Nanoshells@poly (NIPAAm-coAAm) hydrogel Laser: 1064 nm Nanoshells@ liposomes Laser: 820 nm
96
NIR lighttriggered drug delivery
Optical coherence tomography Photoacoustic tomography
NIR lighttriggered drug delivery
98
99
101
P1: OTA/XYZ P2: ABC c28 JWBS037-Chen February 1, 2011
0:19
Printer Name: Yet to Come
REFERENCES
703
TABLE 28.1 (Continued ) Nanostructure
Characteristics
SPIO@silica@Au
Size: ∼80 nm Surface: Multifunction PEG Absorbance: 600–900 nm
Fe3 O4 @polymer@Au Size: 273 nm
Silica@Fe3 O4 @Au
Hollow gold nanoshell (Fe3 O4 @Au)
PLGA-Mn-Au multilayer half-shell nanoparticles
Absorbance: from 700 nm to NIR region Surface: anti-HER2/neu antibody Absorbance: ∼800 nm Surface: anti-HER2/neu antibody Absorbance: ∼800 nm Shape: half-shell
Application In vitro photothermal effect Magnetic resonance relaxation In vitro NIR absorption spectra magnetization In vitro SKBR3 cell line T2-weighted MR imaging photothermal ablation In vitro SKBR3 cell line T2-weighted MR imaging photothermal ablation In vitro T2-weighted MR imaging; photothermal effect; controlled drug release
Reference 103
105
107
110
112
be examined thoroughly, although preliminary studies suggest that gold nanoparticles are nontoxic. Based on the standard of new drug development, gold core–shell nanostructures should meet the safety, effectiveness, and quality control standards. ACKNOWLEDGMENTS We thank Stephanie Deming for editing the manuscript. This work was supported in part by a grant from the National Institutes of Health (R01 CA119387), a Seed Grant through the Alliance for NanoHealth by the Department of Army Telemedicine and Advanced Technology Research Center (W81XWH-07-2-0101), and an Odyssey Fellowship (to MPM). The Odyssey Fellowship is supported by the Odyssey Program and the Cockrell Foundation for Scientific Achievement at The University of Texas M. D. Anderson Cancer Center. REFERENCES 1. Loo, C.; et al. Nanoshell-enabled photonics-based imaging and therapy of cancer. Technol. Cancer Res. Treat. 2004, 3, 33–40. 2. Jain, P. K.; Huang, X.; El-Sayed, I. H.; El-Sayad, M. A. Review of some interesting surface plasmon resonance-enhanced properties of noble metal nanoparticles and their applications to biosystems. Plasmonics 2007, 2, 107–118.
P1: OTA/XYZ P2: ABC c28 JWBS037-Chen February 1, 2011
704
0:19
Printer Name: Yet to Come
THERANOSTIC APPLICATIONS OF GOLD CORE–SHELL STRUCTURED NANOPARTICLES
3. Huang, X. H.; Jain, P. K.; El-Sayed, I. H.; El-Sayed, M. A. Plasmonic photothermal therapy (PPTT) using gold nanoparticles. Lasers Med. Sci. 2008, 23, 217–228. 4. Hirsch, L. R.; Jackson, J. B.; Lee, A.; Halas, N. J.; West, J. A whole blood immunoassay using gold nanoshells. Anal. Chem. 2003, 75, 2377–2381. 5. West, J. L.; Halas, N.; Sershen, S. R. Optically-responsive nanoshell composites. Abstr. Papers Am. Chem. Soc. 2003, 225, U522–U522. 6. Gobin, A. M.; Lee, M. H.; Drezek, R. A.; Halas, N. J.; West, J. L. Nanoshells for combined cancer therapy and imaging in vivo. Clin. Cancer Res. 2005, 11, 9095s–9095s. 7. Hirsch, L. R.; et al. Nanoshell-mediated near-infrared thermal therapy of tumors under magnetic resonance guidance. Proc. Nat. Acad. Sci. U.S. Am. 2003, 100, 13549–13554. 8. West, J. L.; Halas, N. J. Engineered nanomaterials for biophotonics applications: improving sensing, imaging, and therapeutics. Annu. Rev. Biomed. Eng. 2003, 5, 285–292. 9. Lal, S.; Clare, S. E.; Halas, N. J. Nanoshell-enabled photothermal cancer therapy: impending clinical impact. Acc. Chem. Res. 2008, 41, 1842–1851. 10. Kalele, S.; Gosavi, S. W.; Urban, J.; Kulkarni, S. K. Nanoshell particles: synthesis, properties and applications. Curr. Sci. 2006, 91, 1038–1052. 11. Zhou, H. S.; Honma, I.; Komiyama, H.; Haus, J. W. Controlled synthesis and quantum-size effect in gold-coated nanoparticles. Phys. Rev. B 1994, 50, 12052–12056. 12. Averitt, R. D.; Sarkar, D.; Halas, N. J. Plasmon resonance shifts of Au-coated Au2S nanoshells: insight into multicomponent nanoparticle growth. Phys. Rev. Lett. 1997, 78, 4217–4220. 13. Hirsch, L. R.; et al. Metal nanoshells. Ann. Biomed. Eng. 2006, 34, 15–22. 14. Oldenburg, S. J.; Averitt, R. D.; Westcott, S. L.; Halas, N. J. Nanoengineering of optical resonances. Chem. Phy. Lett. 1998, 288, 243–247. 15. St¨ober, W.; Fink, A.; Bohn, E. Controlled growth of monodisperse silica spheres in the micron size range. J. Colloid Interface Sci. 1968, 26, 62–69. 16. Duff, D. G.; Baiker, A.; Edwards, P. P. A new hydrosol of gold clusters .1. Formation and particle-size variation. Langmuir 1993, 9, 2301–2309. 17. Leff, D. V.; Brandt, L.; Heath, J. R. Synthesis and characterization of hydrophobic, organicallysoluble gold nanocrystals functionalized with primary amines. Langmuir 1996, 12, 4723–4730. 18. Liang, H. P.; Wan, L. J.; Bai, C. L.; Jiang, L. Gold hollow nanospheres: tunable surface plasmon resonance controlled by interior-cavity sizes. J. Phys. Chem. B 2005, 109, 7795–7800. 19. Kobayashi, Y.; Horie, M.; Konno, M.; Rodriguez-Gonzalez, B.; Liz-Marzan, L. M. Preparation and properties of silica-coated cobalt nanoparticles. J. Phys. Chem. B 2003, 107, 7420–7425. 20. Schwartzberg, A. M.; Olson, T. Y.; Talley, C. E.; Zhang, J. Z. Synthesis, characterization, and tunable optical properties of hollow gold nanospheres. J. Phys. Chem. B 2006, 110, 19935– 19944. 21. Lu, W.; et al. Targeted photothermal ablation of murine melanomas with melanocyte-stimulating hormone analog-conjugated hollow gold nanospheres. Clin. Cancer Res. 2009, 15, 876– 886. 22. Sun, Y. G.; Mayers, B. T.; Xia, Y. N. Template-engaged replacement reaction: a one-step approach to the large-scale synthesis of metal nanostructures with hollow interiors. Nano Lett. 2002, 2, 481–485. 23. Sun, Y. G.; Xia, Y. N. Shape-controlled synthesis of gold and silver nanoparticles. Science 2002, 298, 2176–2179. 24. Chen, J. Y.; et al. Gold nanocages: engineering their structure for biomedical applications. Adv. Mater. 2005, 17, 2255–2261. 25. Sun, Y. G.; Mayers, B.; Herricks, T.; Xia, Y. N. Polyol synthesis of uniform silver nanowires: a plausible growth mechanism and the supporting evidence. Nano Lett. 2003, 3, 955–960.
P1: OTA/XYZ P2: ABC c28 JWBS037-Chen February 1, 2011
0:19
Printer Name: Yet to Come
REFERENCES
705
26. Sun, Y. G.; Gates, B.; Mayers, B.; Xia, Y. N. Crystalline silver nanowires by soft solution processing. Nano Lett. 2002, 2, 165–168. 27. Au, L.; Lu, X. M.; Xia, Y. N. A comparative study of galvanic replacement reactions involving Ag nanocubes and AuC12 − or AuCl4 − . Adv. Mater. 2008, 20, 2517. 28. Sun, Y. G.; Xia, Y. N. Mechanistic study on the replacement reaction between silver nanostructures and chloroauric acid in aqueous medium. J. Am. Chem. Soc. 2004, 126, 3892–3901. 29. Sun, Y. G.; Xia, Y. A. Alloying and dealloying processes involved in the preparation of metal nanoshells through a galvanic replacement reaction. Nano Lett. 2003, 3, 1569–1572. 30. Hu, M.; et al. Gold nanostructures: engineering their plasmonic properties for biomedical applications. Chem. Soc. Rev. 2006, 35, 1084–1094. 31. Suzuki, D.; Kawaguchi, H. Gold nanoparticle localization at the core surface by using thermosensitive core–shell particles as a template. Langmuir 2005, 21, 12016–12024. 32. Halas, N. Playing with plasmons. Tuning the optical resonant properties of metallic nanoshells. MRS Bull. 2005, 30, 362–367. 33. Jain, P. K.; Lee, K. S.; El-Sayed, I. H.; El-Sayed, M. A. Calculated absorption and scattering properties of gold nanoparticles of different size, shape, and composition: applications in biological imaging and biomedicine. J. Phys. Chem. B 2006, 110, 7238–7248. 34. Mie, G. Beitr¨age zur Optik tr¨uber Medien, speziell kolloidaler Metall¨osungen. Ann. Phys. 1908, 25, 377–445. 35. Schwartzberg, A. M.; Zhang, J. Z. Novel optical properties and emerging applications of metal nanostructures. J. Phys. Chem. C 2008, 112, 10323–10337. 36. Lee, K. S.; El-Sayed, M. A. Gold and silver nanoparticles in sensing and imaging: sensitivity of plasmon response to size, shape, and metal composition. J. Phys. Chem. B 2006, 110, 19220–19225. 37. Link, S.; El.-Sayed, M. A. Optical properties and ultrafast dynamics of metallic nanocrystals. Annu. Rev. Phys. Chem. 2003, 54, 331–366. 38. Link, S.; El-Sayed, M. A. Size and temperature dependence of the plasmon absorption of colloidal gold nanoparticles. J. Phys. Chem. B 1999, 103, 4212–4217. 39. Charamisinau, I.; et al. Semiconductor laser insert with uniform illumination for use in photodynamic therapy. Appl. Optics 2004, 44, 5055–5068. 40. Prodan, E.; Radloff, C.; Halas, N. J.; Nordlander, P. A hybridization model for the plasmon response of complex nanostructures. Science 2003, 302, 419–422. 41. Oldenburg, S. J.; Jackson, J. B.; Westcott, S. L.; Halas, N. J. Infrared extinction properties of gold nanoshells. Appl. Phys. Lett. 1999, 75, 2897–2899. 42. Skrabalak, S. E.; Au, L.; Li, X. D.; Xia, Y. Facile synthesis of Ag nanocubes and Au nanocages. Nat. Protocols 2007, 2, 2182–2190. 43. Fukumura, D.; et al. Tumor induction of VEGF promoter activity in stromal cells. Cell 1998, 94, 715–725. 44. Weissleder, R. A clearer vision for in vivo imaging. Nat. Biotechnol. 2001, 19, 316–317. 45. Wang, Y. W.; et al. Photoacoustic tomography of a nanoshell contrast agent in the in vivo rat brain. Nano Lett. 2004, 4, 1689–1692. 46. Fujimoto, J. G.; et al. Optical biopsy and imaging using optical coherence tomography. Nat. Med. 1995, 1, 970–972. 47. Zagaynova, E. V.; et al. Contrasting properties of gold nanoparticles for optical coherence tomography: phantom, in vivo studies and Monte Carlo simulation. Phys. Med. Biol. 2008, 53, 4995–5009. 48. Agrawal, A.; et al. Quantitative evaluation of optical coherence tomography signal enhancement with gold nanoshells. J. Biomed. Optics 2006, 11, 041121.
P1: OTA/XYZ P2: ABC c28 JWBS037-Chen February 1, 2011
706
0:19
Printer Name: Yet to Come
THERANOSTIC APPLICATIONS OF GOLD CORE–SHELL STRUCTURED NANOPARTICLES
49. Gobin, A. M.; et al. Near-infrared resonant nanoshells for combined optical imaging and photothermal cancer therapy. Nano Lett. 2007, 7, 1929–1934. 50. Adler, D. C.; Huang, S. W.; Huber, R.; Fujimoto, J. G. Photothermal detection of gold nanoparticles using phase-sensitive optical coherence tomography. Optics Express 2008, 16, 4376–4393. 51. Chen, J.; et al. Gold nanocages: bioconjugation and their potential use as optical imaging contrast agents. Nano Lett. 2005, 5, 473–477. 52. Cang, H.; et al. Gold nanocages as contrast agents for spectroscopic optical coherence tomography. Optics Lett. 2005, 30, 3048–3050. 53. Zhang, H. F.; Maslov, K.; Wang, L. H. V. In vivo imaging of subcutaneous structures using functional photoacoustic microscopy. Nat. Protocols 2007, 2, 797–804. 54. Zhang, H. F.; Maslov, K.; Stoica, G.; Wang, L.H.V. Functional photoacoustic microscopy for high-resolution and noninvasive in vivo imaging. Nat. Biotechnol. 2006, 24, 848–851. 55. Oleary, M. A.; Boas, D. A.; Chance, B.; Yodh, A. G. Experimental images of heterogeneous turbid media by frequency-domain diffusing-photon tomography. Optics Lett. 1995, 20, 426–428. 56. Boas, D. A.; et al. Imaging the body with diffuse optical tomography. IEEE Signal Processing Mag. 2001, 18, 57–75. 57. Wang, X. D.; et al. Noninvasive laser-induced photoacoustic tomography for structural and functional in vivo imaging of the brain. Nat. Biotechnol. 2003, 21, 803–806. 58. Hoelen, C. G. A.; de Mul, F. F. M.; Pongers, R.; Dekker, A. Three-dimensional photoacoustic imaging of blood vessels in tissue. Optics Lett. 1998, 23, 648–650. 59. Wang, X. D.; et al. Noninvasive photoacoustic angiography of animal brains in vivo with near-infrared light and an optical contrast agent. Optics Lett. 2004, 29, 730–732. 60. Guillon, C.; et al. Coherent acoustic vibration of metal nanoshells. Nano Lett. 2007, 7, 138–142. 61. Yang, X. M.; Skrabalak, S. E.; Li, Z. Y.; Xia, Y. N.; Wang, L. H. V. Photoacoustic tomography of a rat cerebral cortex in vivo with au nanocages as an optical contrast agent. Nano Lett. 2007, 7, 3798–3802. 62. Song, K. H.; Kim, C.; Cobley, C. M.; Xia, Y.; Wang, L. V. Near-infrared gold nanocages as a new class of tracers for photoacoustic sentinel lymph node mapping on a rat model. Nano Lett. 2009, 9, 183–188. 63. Maslov, K.; Stoica, G.; Wang, L. V. H. In vivo dark-field reflection-mode photoacoustic microscopy. Optics Lett. 2005, 30, 625–627. 64. El-Sayed, M. A. Some interesting properties of metals confined in time and nanometer space of different shapes. Acc. Chem. Res. 2001, 34, 257–264. 65. Jain, P. K.; El-Sayed, I. H.; El-Sayed, M. A. Au nanoparticles target cancer. Nano Today 2007, 2, 18–29. 66. Link, S.; El-Sayed, M. A. Shape and size dependence of radiative, non-radiative and photothermal properties of gold nanocrystals. Int. Rev. Phys. Chem. 2000, 19, 409–453. 67. Hu, M.; et al. Ultrafast laser studies of the photothermal properties of gold nanocages. J. Phys. Chem. B 2006, 110, 1520–1524. 68. Hartland, G. V. Measurements of the material properties of metal nanoparticles by time-resolved spectroscopy. Phys. Chem. Chem. Phys. 2004, 6, 5263–5274. 69. Du, H.; Fuh, R. C. A.; Li, J. Z.; Corkan, L. A.; Lindsey, J. S. PhotochemCAD: a computer-aided design and research tool in photochemistry. Photochem. Photobiol. 1998, 68, 141–142. 70. Anderson, R. R.; Parrish, J. A. Selective photothermolysis—precise microsurgery by selective absorption of pulsed radiation. Science 1983, 220, 524–527. 71. Jori, G.; Spikes, J. D. Photothermal sensitizers—possible use in tumor-therapy. J. Photochem. Photobiol. B–Biol. 1990, 6, 93–101. 72. Camerin, M.; et al. Photothermal sensitisation as a novel therapeutic approach for tumours: studies at the cellular and animal level. Eur. J. Cancer 2005, 41, 1203–1212.
P1: OTA/XYZ P2: ABC c28 JWBS037-Chen February 1, 2011
0:19
Printer Name: Yet to Come
REFERENCES
707
73. Soncin, M.; Busetti, A.; Fusi, F.; Jori, G.; Rodgers, M.A.J. Irradiation of amelanotic melanoma cells with 532 nm high peak power pulsed laser radiation in the presence of the photothermal sensitizer Cu(II)-hematoporphyrin: a new approach to cell photoinactivation. Photochem. Photobiol. 1999, 69, 708–712. 74. Moyer, H. R.; Delman, K. A. The role of hyperthermia in optimizing tumor response to regional therapy. Int. J. Hyperthermia 2008, 24, 251–261. 75. Stern, J. M.; Hsieh, J. T.; Park, S.; Lotan, Y.; Cadeddu, J. A. Successful gold nanoshell assisted laser ablation of prostate carcinoma in vitro. J. Urol. 2007, 177, 52–52. 76. Stern, J. M.; Hsieh, J. T.; Park, S.; Qiu, J.; Cadeddu, J. Gold nanoshell assisted laser ablation of a prostate cancer cell line. J. Endourol. 2006, 20, A193. 77. Stern, J. M.; et al. Efficacy of laser-activated gold nanoshells in ablating prostate cancer cells in vitro. J. Endourol. 2007, 21, 939–943. 78. Lowery, A. R.; Gobin, A. M.; Day, E. S.; Halas, N. J.; West, J. L. Immunonanoshells for targeted photothermal ablation of tumor cells. Int. J. Nanomed. 2006, 1, 149–154. 79. Loo, C.; Lowery, A.; Halas, N.; West, J.; Drezek, R. Immunotargeted nanoshells for integrated cancer imaging and therapy. Nano Lett. 2005, 5, 709–711. 80. Bernardi, R. J.; Lowery, A. R.; Thompson, P. A.; Blaney, S. M.; West, J. L. Immunonanoshells for targeted photothermal ablation in medulloblastoma and glioma: an in vitro evaluation using human cell lines. J. Neuro-Oncol. 2008, 86, 165–172. 81. Chen, J. Y.; et al. Immuno gold nanocages with tailored optical properties for targeted photothermal destruction of cancer cells. Nano Lett. 2007, 7, 1318–1322. 82. Chithrani, B. D.; Ghazani, A. A.; Chan, W.C.W. Determining the size and shape dependence of gold nanoparticle uptake into mammalian cells. Nano Lett. 2006, 6, 662–668. 83. Melancon, M. P.; et al. In vitro and in vivo targeting of hollow gold nanoshells directed at epidermal growth factor receptor for photothermal ablation therapy. Mol. Cancer Ther. 2008, 7, 1730–1739. 84. Maksimova, I. L.; et al. Near-infrared laser photothermal therapy of cancer by using gold nanoparticles: Computer simulations and experiment. Med. Laser Application 2007, 22, 199–206. 85. O’Neal, D. P.; Hirsch, L. R.; Halas, N. J.; Payne, J. D.; West, J. L. Photo-thermal tumor ablation in mice using near infrared-absorbing nanoparticles. Cancer Lett. 2004, 209, 171–176. 86. Hobbs, S. K.; et al. Regulation of transport pathways in tumor vessels: role of tumor type and microenvironment. Proc. Nat. Acad. Sci. U.S.A. 1998, 95, 4607–4612. 87. Maeda, H.; Sawa, T.; Konno, T. Mechanism of tumor-targeted delivery of macromolecular drugs, including the EPR effect in solid tumor and clinical overview of the prototype polymeric drug SMANCS. J. Control. Release 2001, 74, 47–61. 88. Maeda, H.; Wu, J.; Sawa, T.; Matsumura, Y.; Hori, K. Tumor vascular permeability and the EPR effect in macromolecular therapeutics: a review. J. Control. Release 2000, 65, 271–284. 89. Moghimi, S. M.; Hunter, A. C.; Murray, J. C. Long-circulating and target-specific nanoparticles: theory to practice. Pharmacol. Rev. 2001, 53, 283–318. 90. Stern, J. M.; Stanfield, J.; Kabbani, W.; Hsieh, J. T.; Cadeddu, J. A. Selective prostate cancer thermal ablation with laser activated gold nanoshells. J. Urol. 2008, 179, 748–753. 91. Zhang, G.; et al. Influence of anchoring ligands and particle size on the colloidal stability and in vivo biodistribution of polyethylene glycol-coated gold nanoparticles in tumor-xenografted mice. Biomaterials 2009, 30, 1928–1936. 92. Yuan, F.; et al. Vascular permeability in a human tumor xenograft: molecular size dependence and cutoff size. Cancer Res. 1995, 55, 3752–3756. 93. Siegrist, W.; et al. Characterization of receptors for alpha-melanocyte-stimulating hormone on human melanoma cells. Cancer Res. 1989, 49, 6352–6358.
P1: OTA/XYZ P2: ABC c28 JWBS037-Chen February 1, 2011
708
0:19
Printer Name: Yet to Come
THERANOSTIC APPLICATIONS OF GOLD CORE–SHELL STRUCTURED NANOPARTICLES
94. Chen, J.; Cheng, Z.; Hoffman, T. J.; Jurisson, S. S.; Quinn, T. P. Melanoma-targeting properties of (99m )technetium-labeled cyclic alpha-melanocyte-stimulating hormone peptide analogues. Cancer Res. 2000, 60, 5649–5658. 95. Sershen, S. R.; Westcott, S. L.; West, J. L.; Halas, N. J. An opto-mechanical nanoshell-polymer composite. App. Phys. B–Lasers Optics 2001, 73, 379–381. 96. Sershen, S. R.; Westcott, S. L.; Halas, N. J.; West, J. L. Temperature-sensitive polymer–nanoshell composites for photothermally modulated drug delivery. J. Biomed. Mater. Res. 2000, 51, 293–298. 97. Priest, J. H.; Murray, S. L.; Nelson, R. J.; Hoffman, A. S. Lower critical solution temperatures of aqueous copolymers of N-isopropylacrylamide and other N-substituted acrylamides. ACS Symp. Ser. 1987, 350, 255–264. 98. Bikram, M.; Gobin, A. M.; Whitmire, R. E.; West, J. L. Temperature-sensitive hydrogels with SiO2 -Au nanoshells for controlled drug delivery. J. Control. Release 2007, 123, 219–227. 99. Kasili, P. M.; Vo-Dinh, T. Liposome encapsulated gold nanoshells for nanophototherapy induced hyperthermia. Int. J. Nanotechnol. 2005, 2, 397–410. 100. Chen, T.; et al. Distal cationic poly(ethylene glycol) lipid conjugates in large unilamellar vesicles prepared by extrusion enhance liposomal cellular uptake. J. Liposome Res. 2004, 14, 155–173. 101. Wu, G.; et al. Remotely triggered liposome release by near-infrared light absorption via hollow gold nanoshells. J. Am. Chem. Soc. 2008, 130, 8175–8177. 102. Harisinghani, M. G.; et al. Noninvasive detection of clinically occult lymph-node metastases in prostate cancer. N. Engl. J. Med. 2003, 348, 2491–U2495. 103. Ji, X. J.; et al. Bifunctional gold nanoshells with a superparamagnetic iron oxide–silica core suitable for both MR imaging and photothermal therapy. J. Phys. Chem. C 2007, 111, 6245–6251. 104. Lu, Y.; Yin, Y. D.; Mayers, B. T.; Xia, Y. N. Modifying the surface properties of superparamagnetic iron oxide nanoparticles through a sol-gel approach. Nano Lett. 2002, 2, 183–186. 105. Wang, L. Y.; Bai, J. W.; Li, Y. J.; Huang, Y. Multifunctional nanoparticles displaying magnetization and near-IR absorption. Angew. Chem.-Int. Ed. 2008, 47, 2439–2442. 106. Bai, F.; Yang, X. L.; Li, R.; Huang, B.; Huang, W. Q. Monodisperse hydrophilic polymer microspheres having carboxylic acid groups prepared by distillation precipitation polymerization. Polymer 2006, 47, 5775–5784. 107. Kim, J.; et al. Designed fabrication of multifunctional magnetic gold nanoshells and their application to magnetic resonance imaging and photothermal therapy. Angew. Chem. Int. Ed. Engl. 2006, 45, 7754–7758. 108. Park, J.; et al. Ultra-large-scale syntheses of monodisperse nanocrystals. Nat. Mater. 2004, 3, 891–895. 109. Wang, Y.; Teng, X. W.; Wang, J. S.; Yang, H. Solvent-free atom transfer radical polymerization in the synthesis of Fe2 O3 @polystyrene core–shell nanoparticles. Nano Lett. 2003, 3, 789–793. 110. Lim, Y. T.; Cho, M. Y.; Kim, J. K.; Hwangbo, S.; Chung, B. H. Plasmonic magnetic nanostructure for bimodal imaging and photonic-based therapy of cancer cells. Chembiochem 2007, 8, 2204–2209. 111. Wang, L. Y.; et al. Monodispersed core–shell Fe3 O4 @Au nanoparticles. J. Phys. Chem. B 2005, 109, 21593–21601. 112. Park, H.; et al. Multifunctional nanoparticles for photothermally controlled drug delivery and magnetic resonance imaging enhancement. Small 2008, 4, 192–196.
P1: OTA/XYZ P2: ABC c29 JWBS037-Chen February 1, 2011
4:29
Printer Name: Yet to Come
CHAPTER 29
Magnetic Nanoparticle Carrier for Targeted Drug Delivery: Perspective, Outlook, and Design R. D. K. MISRA Center for Structural and Functional Materials, University of Louisiana at Lafayette, Lafayette, Louisiana, USA
29.1 INTRODUCTION The primary challenge in the delivery of a drug to a tumor site is to target the anticancer drug specifically into and around tumors at concentrations that will decrease their growth and/or viability. Excellent vehicles to achieve targeted drug delivery are magnetic nanocarriers. An approach to accomplish this objective is to fabricate a novel temperature and pH-responsive magnetic nanocarrier that combines tumor targeting and controlled release. In addition to controlled release, these carriers simultaneously offer the possibility of imaging the delivery process by magnetic resonance imaging (MRI). The novel aspect of this approach is the combination of thermal and pH sensitivity of the drug-containing shell with magnetic properties of the core in a single unit. Furthermore, application of an external magnetic field will heat the nanoparticles and thus nanoparticle-containing cells. The targeted heat will aid in the killing of tumor cells.
29.2 BACKGROUND AND SIGNIFICANCE OF NANOPARTICLE DRUG CARRIER Metastasis of malignant cells in the human body is a cancer-specific complication that poses serious risks in addition to the effect of the original tumor. The routes that are generally adopted to target the tumor cells include surgical resection, radiation, and chemotherapy. New therapeutic drugs are being developed [1], and a significant effort is being devoted to improve the present noninvasive methods such as photodynamic therapy [2] and application of hyperthermia [3].
Nanoplatform-Based Molecular Imaging Edited by Xiaoyuan Chen C 2011 John Wiley & Sons, Inc. Copyright
709
P1: OTA/XYZ P2: ABC c29 JWBS037-Chen February 1, 2011
710
4:29
Printer Name: Yet to Come
MAGNATIC NANOPARTICLE CARRIER FOR TARGETED DRUG DELIVERY
The basic challenge in drug delivery is the transfer of drugs to the targeted site at the appropriate time [4]. Chemotherapeutic drugs are cytotoxic to healthy cells and there is limited control on the rate of drug release. Site-specific delivery of drug has the potential to increase the effectiveness of drugs, enhance local concentrations, and thus minimize undesirable side effects and toxicity to nontumor cells. One approach to accomplish these characteristics is magnetic field-induced targeted drug delivery. It was recently shown that magnetic nanocrystals coated with biocompatible polymer loaded with anticancer agents via linking to functional groups are viable drug vesicles [5, 6]. Drug-loaded nanoparticle carrier can be injected into the bloodstream and transported to the desired site using a high gradient magnetic field. An advantage of utilizing magnetic nanoparticles is the use of an external localized magnetic field gradient to attract the particles to the desired site, hold them at the desired location until the therapy is completed, followed by removal of the magnetic field. To achieve effective tumor-specific drug delivery, it is important that the vesicles distinguish tumor cells from healthy cells either morphologically or physiologically. Because of the rapid growth rate of cancerous cells, they have a high metabolic rate and the intracellular pH is low [7, 8]. Therefore one of the strategies to deliver drugs to tumor cells is to conjugate drug vesicles with nutrients required for a specific tumor, so that tumor cells will associate with the drug carrier at a significantly faster rate than healthy cells. Furthermore, the drug carrier is designed to release its drug cargo in response to the slightly endosomal acidic pH. To enhance the specificity of tumor targeting, receptor-specific groups can be conjugated to drug carriers. The folate receptor targets are considered appropriate for tumor-selective drug delivery for a number of reasons. Folic acid is stable, poorly immunogenic, and can preferentially target tumor cells because the folate receptor is overexpressed on the surface of tumor cells [7]. The receptor binds folate to nourish the rapidly dividing tumor cells [8–10]. Once the drug carrier has reached the desired site, the drug can be released directly into the bloodstream via enzymatic biodegradation, by changes in temperature or pH, or by direct cell uptake of biofunctionalized nanoparticles [11]. Drug release in response to a specific stimulus (i.e., triggered release) is an essential feature of effective targeted delivery systems. A few examples based on magnetic nanoparticles have been reported. A recent study designed drug-loaded mesoporous silica nanorods capped with superparamagnetic iron oxide nanoparticles through disulfide bonds, which can be transported to a specific site with an external magnet and release the drug upon reducing the disulfide bonds [12]. A simpler approach was the use of magnetic hydrogel, where it was shown that drug release can be turned ‘on and off’ by an external magnet [13]. A magnetic drug-targeting delivery system with triggered release is a new and effective delivery system and should constitute the prime focus for controlled drug delivery. The combination of localized transport of magnetic nanoparticles by an external magnetic field and tumor targeting will greatly contribute to the development of a new tumor-targeted controlled drug delivery system. Stimuli-responsive polymers are a unique class of polymers that respond to changes in environmental conditions, notably pH, temperature, and electric field [14, 15]. Poly(Nisopropylacrylamide) (PNIPAAm) homopolymer is a typical example of stimuli-responsive polymer. It is characterized by lower critical solution temperature (LCST) in aqueous solution such that its volume and shape undergo changes in a reversible manner in response to small changes in temperature in the vicinity of LCST. Additionally, the enzymatic degradation of PNIPAAm can be promoted by grafting with water-soluble and biodegradable
P1: OTA/XYZ P2: ABC c29 JWBS037-Chen February 1, 2011
4:29
Printer Name: Yet to Come
T > LCST, low pH
DOX
DOX
DOX
Magnetic core loaded with drug and encapsulated with thermosensitive polymer shell
DOX DOX
T < LCST, high pH
DOX
Magnetic core
711
DOX
DOX
BACKGROUND AND SIGNIFICANCE OF NANOPARTICLE DRUG CARRIER
Controlled release of drug
FIGURE 29.1 Schematic of “smart” polymer response with temperature and pH.
dextran. Furthermore, PNIPAAm with a molecular weight of less than 40,000 Da can be cleared or excreted by the physiological system [16, 17]. By combining the stimuli-responsive or “smart” behavior of a polymer with the magnetic properties of nanoparticles and the tumor-targeting characteristic of folic acid, the system can serve as an effective drug delivery system with targeted accumulation and controlled drug release. For instance, a thermoresponsive polymer having lower critical solution temperature (LCST) can be “tuned” as desired by varying hydrophilic or hydrophobic comonomer content. The phase transition is shown schematically in Figures 29.1 and 29.2. In a similar manner, these polymers can be made sensitive to changes in pH [18, 19]. Thus these smart polymers are expected to play an important role in drug delivery because they will not only dictate where the drug is to be delivered, but also when and how the drug is released. In summary, the drug-loaded, polymer-coated magnetic drug carrier containing targetspecific payload will first be transported to the desired site by an external magnetic field. Upon removal of the magnetic field, the magnetic nanocarrier will detect tumor cells because of its tumor-recognizing function. Then the encapsulated drugs will be released when the particles are taken up by the cells because of change in pH from physiological pH (7.4) to endosomal pH (5.3). The release rate of drug can be additionally modulated using a magnetic field, which provides an external means to accelerate the release of drugs [5].
100
Transmittance (%)
80 60 40
Collapse T > LCST or low pH
Swollen, T < LCST or high pH
20 0
LCST 26 28 30 32 34 36 38 40 42 44 46 Temperature (°C)
FIGURE 29.2 LCST of 5 wt% smart polymer in phosphate buffer solution pH 7.4 [6].
P1: OTA/XYZ P2: ABC c29 JWBS037-Chen February 1, 2011
712
4:29
Printer Name: Yet to Come
MAGNATIC NANOPARTICLE CARRIER FOR TARGETED DRUG DELIVERY
During exposure to the magnetic field, the nanoparticle carrier will oscillate, alternately creating compressive and tensile forces. This in turn will act as a pump, providing an increased release of drug. Furthermore, the oscillating relaxation of the particles’ magnetic moments will locally heat the nanoparticle carrier. The increase in temperature will shrink the shell and release the encapsulated drugs inside the tumor cells. In principle, both the release period and the amount of drug released can be controlled by the intensity and duration of the magnetic field. It is expected that the carrier will stably retain drug in the biological body until drug release is induced. Also, the locally generated heat will have therapeutic effects and assist in destroying tumor cells. Thus an ideal approach is to tailor-make graft copolymers, consisting of biodegradable polysaccharides and stimuli-responsive grafts, as building blocks for the new dual-stimuliresponsive (pH and temperature sensitive) polymeric system. This is in contrast to most polymers that respond to a single stimulus. After the release of drug, the metabolism of iron oxide crystals and their susceptibility to degradation will leach Fe2+ and circulate in the bloodstream as micronutrient [20]. It has been shown that iron uptake decreased to 22% in a few days. Thus the metabolic system cleans up the nanoparticles, implying that the damage to healthy tissues is not a serious concern.
29.3 RECENT STUDIES Routes have been developed to synthesize magnetic nanocrystals for biomedical [5, 6] and antimicrobial applications [21, 22]. They include a room-temperature reverse micelle process [23–25] and a high-temperature decomposition process to make nanoparticles [6]. In the reverse micelle process, water-in-oil microemulsion is used. These are nanosized droplets of water sustained in a hydrocarbon bulk phase using surfactants such as bis(2ethylhexyl) sodium sulfosuccinate. Figure 29.3 is a generic flow sheet illustrating the steps in the chemical synthesis of magnetite or ferrite nanoparticles carried out in the AOT reverse micelle system. In the high-temperature decomposition process, the reduction of Fe(III) salt to an Fe(II) intermediate occurs, followed by the decomposition of the intermediate at high temperature (260 ◦ C). These methods enabled modulation of physicochemical parameters of magnetic nanoparticles. Figure 29.4 illustrates remarkably uniform-sized monodisperse nanoparticles as imaged using transmission electron microscopy. Uniformity of particles is an important requirement for drug delivery because it enhances the probability of magnetic capture. At room temperature, they were characterized by superparamagnetic behavior (absence of hysteresis, immeasurable coercivity and remanence, Fig. 29.5) [23–25]. Superparamagnetism assures that particles do not retain magnetism upon withdrawal of the external magnetic field. Additionally, the superparamagnetic nanoparticles were capable of generating impressive levels of heating at lower magnetic field strength. The heat generated expressed in terms of the specific absorption rate was observed to be ∼210 W/g at a low field strength of ∼15 kA/m. This is adequate for magnetic particle hyperthermia, to induce the changes in temperature required for the collapse of thermosensitive polymer, and encourage release of drugs. A series of exploratory studies [5, 6, 26] were conducted by Misra’s group, where magnetic nanocrystals were encapsulated with polymethylacrylic acid (PMAA), polyvinyl alcohol (PVA), or polyethylene glycol (PEG), and an anticancer agent (doxorubicin) was
P1: OTA/XYZ P2: ABC c29 JWBS037-Chen February 1, 2011
4:29
Printer Name: Yet to Come
RECENT STUDIES
713
FIGURE 29.3 Schematic of reverse micelle process developed in the PI’s laboratory for synthesis of magnetic nanoparticles (magnetite and ferrites). The surfactant refers to bis(2-ethylhexyl) sodium sulfosuccinate (AOT). Reverse micelle is a colloidal aggregate of amphilic surfactant molecules such that hydrophilic ends are in contact. It includes two microemulsions (oil phase microemulsion and aqueous phase microemulsion). The reaction takes place inside nanoreactors.
tethered to the ends of the polymer chains. These synthetic polymers degrade by hydrolytic cleavage of their backbone and achieve sustained release of drug [4, 27, 28]. A comparison of magnetic properties of as-synthesized and polymer encapsulated magnetic nanoparticles confirmed that magnetic nanocrystals retained their magnetic properties (Table 29.1) [5]. Subsequently, the study was extended from the viewpoint of selective
FIGURE 29.4 Representative transmission electron micrographs for magnetic (magnetite, Fe3 O4 ) nanoparticles (a) without and (b) with encapsulation of thermosensitive polymer.
P1: OTA/XYZ P2: ABC c29 JWBS037-Chen February 1, 2011
714
4:29
Printer Name: Yet to Come
MAGNATIC NANOPARTICLE CARRIER FOR TARGETED DRUG DELIVERY
FIGURE 29.5 Variation of magnetization with applied field at room temperature showing superparamagnetic nature of magnetic nanoparticles [23–25].
tumor-targeted drug delivery, where folic acid activated with dicyclohexyl carbodiimide (DCC) was conjugated to the surface of the polymer encapsulated magnetic nanocrystals. An analysis of characteristic adsorption bands in the Fourier transform infrared (FTIR) spectra (Fig. 29.6) of magnetite (Fe3 O4 ), polymer (PEG), and folic acid confirmed the successful chemical synthesis of folic acid-conjugated polymer encapsulated magnetic nanoparticles. Interestingly, the drug release response of magnetic nanocrystals functionalized with PEG and conjugated with folic acid–doxorubicin studied by UV–visible spectrometer in the presence of magnetic field of ∼1000 Oe was characterized by an initial rapid drug release followed by a controlled release (Fig. 29.7). The concentration of drug released was quantified by fluorescence or VIS spectrometery [3] (Fig. 29.8). The effect of magnetic stimulation on the accelerated release of drugs was proposed to be a consequence of repeated pulsatile mechanical deformation generating compression and decompression of the matrix [6]. A similar behavior was observed for polymethacrylic acid (PMMA) encapsulated nanoparticles. The release of drug occurred by desorption, diffusion, polymer degradation, or some combination of both. Electron microscopy studies
TABLE 29.1 Comparison of Magnetic Properties of As-Synthesized and Polymer-Encapsulated Magnetic Nanocrystalsa Saturation Magnetization (emu/g) Sample Magnetic nanocrystals Polymer-coated magnetic nanocrystals a Normalized
Coercivity (kOe)
Remanence (emu/g)
300 K
2K
300 K
2K
300 K
65.4 64.9
0.40 0.53
0 0
8.8 7.5
0 0
with respect to the magnetic content. Source: Zhang et al. [5].
P1: OTA/XYZ P2: ABC c29 JWBS037-Chen February 1, 2011
4:29
Printer Name: Yet to Come
RECENT STUDIES
d 3435
3003 2958
1426 1459 1523
1093
2853 2924
c Transmittance (%)
1627
442 630 588
29592844 2924
3543 3324 3416
715
1411 1640 1484 1605 1694
b
951 1726 1426 1097 1637 1523
3001 2954 2852 2924
343 5
442 630 588
a 1631 1426 152 3
3001 2954
1097
2852 2924
343 5
442 630 588
4000
3500
3000
2500
2000
1500
1000
500
Wavenumber (cm –1)
FIGURE 29.6 Fourier transform infrared (FTIR) spectra of (a) as-synthesized Fe3 O4 nanoparticles, (b) PEG encapsulated Fe3 O4 nanoparticles (Fe3 O4 /PEG = 1:2, weight ratio), (c) folic acid (pure), and (d) Fe3 O4 /PEG (1:2)-folic acid/DCC (1:1; mole ratio) [26].
suggested that the polymer was porous and the pores appeared to have coalesced, implying that erosion occurred during the delivery process. Having proved the success of polymer encapsulated magnetic nanocrystals as a viable drug carrier, the study was recently extended to explore the potential of a novel stimuli-responsive polymer to develop the drug delivery system based on stimuli-responsive polymer. Temperature- and pH-responsive dextran grafted poly(N-isopropylacrylamide)
100 % Cumulative drug release
90 80 70 60 50 40 30 20 10 0
0
5
10 15 20 25 30 35 40 45 50 Time (h)
FIGURE 29.7 Drug release response of folate receptor activated polyethylene glycol functionalized magnetite nanoparticles. Bar indicates the maximum range over which the values are observed [26].
P1: OTA/XYZ P2: ABC c29 JWBS037-Chen February 1, 2011
716
4:29
Printer Name: Yet to Come
MAGNATIC NANOPARTICLE CARRIER FOR TARGETED DRUG DELIVERY
FIGURE 29.8 Absorption (left) and fluorescence spectrum (right) of doxorubicin in water. The extinction coefficient of 17250 (OD/g) and specific fluorescence (Exmax = 500 and Emmax = 585) allow specific and sensitive detection of less than 5 g/mL. Fluorescence measurements can achieve detection of <50 ng doxorubicin.
copolymerized with dimethylacrylamide (dextran-g-poly(NIAAm-co-DMAAm)) using the steps summarized in Figure 29.9 [6] was synthesized. The 1 H-NMR spectra of poly(NIPAAm-co-DMAAm) and dextran-g-poly(NIPAAm-coDMAAm) smart polymer shown in Figure 29.10 confirmed the chemical composition of the smart polymer. Characteristic peaks from the different components indicating dextran (backbone) and the poly(NIPAAm-co-DMAAm) polymer (grafted chain) were identified. Peaks a–e result from NIPAAm and DMAAm of the poly(NIPAAm-co-DMAAm) polymer, while the peaks f–l originate from dextran. Using the standard procedure of calculation of characteristic peak integration of DMAAm and NIPAAm in 1 H-NMR spectra, molecular weight and average number of grafts in the dextran-g-poly(NIAAm-co-DMAAm) smart polymer were estimated to be about 79,000 DPa and 7.6, respectively. Furthermore, the molecular weight of poly(NIAAm-co-DMAAm) was significantly less than 10,000 DPa and can be cleared or excreted by the physiological system. Thermal gravimetric analysis (TGA) of drug-loaded magnetite encapsulated with dextran-grafted poly(NIAAm-co-DMAAm) indicated mass loss of 40 wt% (Fig. 29.11) on heating from room temperature to 600 ◦ C, implying that the drug carrier consisted of 60 wt% of Fe3 O4 nanoparticles and 40 wt% of the organic substance including drug and smart polymer. This polymer system exhibited lower critical solution temperature (LCST) of 37–39 o C (Fig. 29.2) in a phosphate buffered saline (PBS) solution, which is representative of a phase-transition behavior. For the stimuli-responsive polymer, the drug release is expected to be controlled by sensitivity to pH and temperature. If the polymer is highly sensitive, an abrupt increase in release is anticipated to be observed. However, release also depends on the degradability of the polymer and the removal of polymer particles by the physiological system. Thus magnetic nanoparticles possess characteristics suitable for controlled drug release using an external magnetic field to bring the drug carrier to a specific location. LCST can be fine-tuned to be above the body temperature by incorporating comonomer units
P1: OTA/XYZ P2: ABC c29 JWBS037-Chen February 1, 2011
4:29
Printer Name: Yet to Come
717
RECENT STUDIES
Step 1: Synthesis of the poly(NIPAAm-co-DMAAm)-COOCH3 O H2C n
CH C
H2C
+
m
O
O
CH2 CH
n
C
O
C
CH3
CH
N
NH CH3
CH3
CH2 CH
CH3O C CH2CH2S
Methanol, 50°C, 24h C
N
HSCH2CH2COOCH3
CH
CH
H
NH
CH3
CH3
CH3
CH3
CH3
m
O
Step 2: Transformation of -COOCH3 group into -NHNH2 group O
O
CH3O C CH2CH2S
CH2 CH C
n
CH2 CH
O
C
N CH3
m
H
+
H2N NH2H2 O
Methanol Refluxing 5h
H2N NH C CH2CH2S
CH2 CH C
O
C
m
CH3
CH CH3
CH3
Step 3: Preparation of dextran with 4-nitrophenyl chloroformate O
O O
HO HO
O O
OH O HO HO
+ CI C O
O
HO HO
DMAP NO2
DMSO/Pyridine
OH O
OH O HO
O2N
O
O OH O
O C O
Dextran
Activated Dextran
Step 4: Synthesis of dextran grafted with poly(NIPAAm-co-DMAAm) O O
O CH ) ( CH2 CH ) SCH2CH2O C HN NH2 m n CO CO NH CH CH CH3 CH3 CH3 CH3
H ( CH2
HO HO
+
OH O HO
O2N
O
O OH O
O C O
Activated Dextran
O O HO HO DMSO, 48h Room temperature
O HO O CH ) SCH2CH2O C HN NH C n CO
OH O
O
O
CH ) ( CH2 m CO NH CH CH CH3 CH3 CH3 CH3
H ( CH2
H
O
NH CH3
CH3
CH
CH3
CH2 CH
N
NH CH3
n
O
OH O
FIGURE 29.9 Steps in the synthesis of dextran grafted with poly(NIPAAm-co-DMAAm) [6].
P1: OTA/XYZ P2: ABC c29 JWBS037-Chen February 1, 2011
718
4:29
Printer Name: Yet to Come
MAGNATIC NANOPARTICLE CARRIER FOR TARGETED DRUG DELIVERY
(a)
(b)
5.0
4.5
4.0
FIGURE 29.10
3.5
1
3.0
2.5
2.0
1.5
1.0
ppm
5.5
5.0
4.5
4.0
3.5
3.0
2.5
2.0
1.5
1.0
ppm
H-NMR spectrum of the dextran-g-poly(NIPAAm-co-DMAAm) smart polymer.
in the thermosensitive polymer. The contribution of magnetic field and stimuli-response mechanisms can therefore be utilized for controlled release of drug into targeted cells. The above outlined perspective further allows investigation of [1] the physicochemical aspects of the nanoparticle drug carrier that govern potential therapeutic utility; [2] structural control of graft polymers, such as changing graft length and number, will elucidate hydration–dehydration transition of thermoresponsive grafts; tailor-made graft copolymers will serve as building blocks for designing multistimuli-receptive hydrogels, by balancing hydrophobic/hydrophilic interactions in the polymer chain; [3] the response of a nanoparticle delivery system to changes in pH, temperature, and magnetic field intensity will provide control of drug release that can be adjusted in response to external or internal signals; [4] the kinetic analysis of drug release will provide an understanding of drug release in response to changing stimuli; and [5] the examination of nanoparticle uptake and release of (fluorescing) drugs using a confocal scanning laser microscope will provide a valuable tool to assess the efficacy of drug delivery.
Weight Remaining (%)
100
3.55 mg
90
80
40%, 1.44 mg
70
60
2.13 mg 100
200
300
400
500
600
700
800
Temperature (deg)
FIGURE 29.11 Thermogravimetric analysis of smart polymer encapsulated nanocarrier.
P1: OTA/XYZ P2: ABC c29 JWBS037-Chen February 1, 2011
4:29
Printer Name: Yet to Come
NANOPARTICLE DRUG CARRIER: SYNTHESIS
719
29.4 NANOPARTICLE DRUG CARRIER: SYNTHESIS 29.4.1 Synthesis of Magnetic Nanoparticles Reverse micelle technique is an appropriate technique to prepare magnetite (Fe3 O4 ) nanoparticles. This approach offers substantial control over the size and size distribution of particles because the reaction takes place inside nanoreactors, which are of similar size [5, 23–25]. In the reverse micelle method, two microemulsion systems are prepared. The first system consists of an oil-phase microemulsion containing isooctane and surfactant diiso-octylsulphoccinate (AOT) and the second is an aqueous phase emulsion consisting of isooctane and surfactant AOT with the reactant salts (hydrated iron sulfate). Microemulsion system I typically consists of 2 mL of 30% NH4 OH + 2.4 mL of water + 66 mL of 0.50 M AOT-isooctane. Microemulsion system II contains 0.576 g of FeSO4 · 7H2 O dissolved in 8 mL of water + 66 mL of AOT-isooctane. Prior to use both emulsion systems are sonicated for 10 min. In microemulsion system I, NH4 OH is the precipitating agent. The two microemulsions are subjected to rapid mechanical stirring for 75 min at a temperature of 50 o C. Atomic force microscope (AFM) studies indicated that at this temperature particles with reduced roughness and high saturation magnetization are obtained [23]. The iron hydroxide is precipitated within the water phase of reverse micelles and oxidized to magnetite. The precipitation of Fe3 O4 occurs according to the following reaction: 3FeSO4 · 7H2 O + 6NH4 OH + 12 O2 → Fe3 O4 + 3(NH4 )2 SO4 + 24H2 O
(29.1)
After rapid mechanical stirring, methanol is added to the resulting mixture, to extract the surfactant and the organic solvent. The resulting liquid is separated and the magnetite product is centrifuged with more methanol. The resulting solid product is washed a number of times with 50% methanol and acetone mixture and distilled water, and dried in an oven at 90 o C for 30 min. 29.4.2 Surface Functionalization of Magnetite Nanoparticles and Conjugation with Drug The synthesized magnetite (Fe3 O4 ) nanoparticles are hydrophobic and require surface functionalization to make them suitable for magnetic drug targeting. The Fe3 O4 nanoparticles are surface modified with bifunctional methyl 3-mercaptopropionate (HSCH2 CH2 COOCH3 ) to chemically bond to the surface of the magnetite nanoparticles via Fe–S covalent bonds [27]. 3-Mercaptopropionate is also useful for the synthesis of stimuli-responsive polymer. To conjugate with drug, the OCH3 functional group is converted to NHNH2 functional group by hydrazinolysis reaction because the NHNH2 functional group facilitates subsequent conjugation with doxorubicin, an antitumor drug. This group is more stable than the OCH3 group. Hydrazinolysis is required because hydrazide end groups ( NHNH2 ) that provide hydrazone linkage with drug (doxorubicin) are acid-labile linkers. Hydrazones of this kind are stable at normal physiological pH 7, but the slightly acidic (pH = 5–5.5) compartment, characteristic for endosomes/lysosomes of the cancer cells, hydrolyzes the hydrazone link, releasing the biologically active doxorubicin [28–33]. The hydrazinolysis reaction is confirmed by examining the product using FTIR. The presence of the characteristic absorption band at 1655 cm−1 corresponds to the N C bond and its intensity provides evidence of the hydrolytic cleavage of hydrazone. A typical procedure to accomplish the
P1: OTA/XYZ P2: ABC c29 JWBS037-Chen February 1, 2011
720
4:29
Printer Name: Yet to Come
MAGNATIC NANOPARTICLE CARRIER FOR TARGETED DRUG DELIVERY
conjugation involves dispersion of hydrophobic Fe3 O4 nanoparticles in diphenyl ether to form a colloidal solution. Then methyl 3-mercaptopropionate is added to the suspension and refluxed at ∼260 o C for 1 h. Subsequently, the solution is cooled and hydrazine monohydrate (N2 H4 •H2 O) is added dropwise to the solution, while continuously stirring the solution for 2 h. The resulting nanoparticles are then separated by centrifugation, washed several times with methanol, and dried at ∼50 ◦ C for 24 h. This procedure functionalizes the hydrophilic Fe3 O4 nanoparticles, with NHNH2 groups [34]. Subsequently, the antitumor drug (e.g., doxorubicin (DOX)) is conjugated to the surface of functionalized magnetite nanoparticles by linking the hydrazone bond to the C O group of DOX according to the chemical reaction illustrated below:
+
Doxorubicin
Functionalized magnetite
Drug-loaded magnetite
29.4.3 Encapsulation of Drug-Loaded Magnetic Nanoparticles with Smart Polymer In the selection of thermosensitive polymer; there are two kinds of important temperatureresponsive materials: liposomes and poly(N-isopropylacrylamide) (PNIPAAm)-based polymers. Liposomes are concentric bilayered vesicle structures made of amphiphilic phospholipids, typically surrounding an aqueous core. They range in size from 0.025 to 10 m in diameter. The size and morphology of liposomes are regulated by the method of preparation and composition. Upon heating, phospholipid bilayers exhibit an endothermic transition at a specific temperature (T c ) that is below its melting point. Below T c , the phospholipid bilayer exists in a gel state and above T c the bilayer exists in a liquid-crystalline state. In the last two decades, liposome-mediated delivery of therapeutic agents has progressed considerably making the transition from the laboratory to the clinic [35–37]. However, one of the problems encountered with targeted liposomes is that the synthetic liposomes’ preparation requires a complicated process and is associated with high operating costs, while the natural liposomes do not have a fixed phase transformation temperature. This has limited their application and popularization in the clinic. Thus it is appropriate to explore inexpensive synthetic polymers that are temperature sensitive. The important responsive polymers in this class are poly(N-isopropylacrylamide) (PNIPAAm), poly(N-vinylcaprolactam) (PVCL), polyethylene glycol (PEG) or polyethylene oxide (PEO), and polypropylene oxide (PPO) [18]. PNIPAAm homopolymer is a thermosensitive polymer. It is hydrated and extended and inhibits transport of solute below the LCST, while at the same time it is hydrophobic and collapses above the LCST. In other words, it undergoes a sharp coil–globule transition
P1: OTA/XYZ P2: ABC c29 JWBS037-Chen February 1, 2011
4:29
Printer Name: Yet to Come
OUTCOME OF APPROACH AND CONCLUSION
721
in water at the LCST, changing from a hydrophilic state below this temperature to a hydrophobic state above it (Figs. 29.1 and 29.2). This is because, at temperatures below the LCST, the predominantly intermolecular hydrogen bonding between the PNIPAAm chains and water molecules result in the PNIPAAm chains adopting a water-soluble coiled structure. At a temperature above the LCST, intermolecular hydrogen bonding between C O and N H groups in the PNIPAAm chains causes the chains to become compact and collapse, reducing their water solubility [38]. This “smart” behavior is very attractive and conducive for controlled drug delivery and biomedical applications [39]. To combine biodegradable and stimuli-responsive encapsulation polymer with tumor recognition of the nanocarrier to efficiently deliver the anticancer drug into the tumor cells requires a new synthetic approach involving the use of a biodegradable polymer, dextran, a natural polysaccharide. The introduction of a biodegradable lining helps the encapsulated polymer to degrade under physiological conditions without harmful effect or significant changes in the hydration–dehydration of thermoresponsive polymer. For instance, dextran-grafted poly(N-isopropylacrylamide-co-N,Ndimethylacrylamide)—dextran-g-poly(NIPAAm-co-DMAAm)s— that is derived from Nisopropylacrylamide (NIPAAm) includes the advantage of bio- and cytocompatibility and exhibits “enzymatic degradation” upon temperature increase [16, 40]. A large decrease in viscosity is anticipated to be observed above the LCST, because of steric hindrance of enzymatic accessibility by grafted chains. Furthermore, PNIPAAm with molecular weight of less than 40,000 Da is easily excreted. Thus by tailor grafting copolymers, consisting of biodegradable polysaccharides and stimuli-responsive grafts, building blocks for new dualstimuli-responsive (pH and temperature sensitive) polymeric systems are created, while most stimuli-responsive polymers are designed to respond to a single stimulus. For tumorrecognizing characteristics, folic acid conjugated dextran-g-poly(NIPAAm-co-DMAAm) can be used for tumor-targeted delivery. The approach presented here involves encapsulation of drug-loaded magnetic nanoparticles with the thermosensitive polymer with an LCST of 40–41 o C to form a core–shell structure. This step involves synthesis of the smart polymer—dextran- grafted poly(NIPAAmco-DMAAm)—followed by encapsulation of drug-loaded magnetic nanoparticles with the smart polymer and finally conjugation with folic acid [6]. A schematic of the different steps involved in the synthesis of core–shell nanocarrier for targeted drug delivery is summarized in Figure 29.12.
29.5 OUTCOME OF APPROACH AND CONCLUSION The perspective and approach described here for the nanoparticle drug carrier is designed to investigate the following: 1. The structural and physicochemical aspects of the nanoparticle drug carrier that govern the potential therapeutic utility. 2. Building block of multistimuli-receptive hydrogels via structural control of graft polymers such as length and number that modulate hydration–dehydration transition (LCST). The controlled solubility of polymer in water is of significant interest in applications that require smart materials, such as sensors for physiological conditions and actuators for enzymatic activity and cell patterning.
P1: OTA/XYZ P2: ABC c29 JWBS037-Chen February 1, 2011
722
4:29
Printer Name: Yet to Come
MAGNATIC NANOPARTICLE CARRIER FOR TARGETED DRUG DELIVERY
FIGURE 29.12 Schematic illustration of steps involved in the synthesis of core–shell magnetic carrier.
3. The response of the nanoparticle drug delivery device to changes in pH, temperature, and magnetic field will provide future directions for the control of drug release from tunable devices that respond to external or internal signals. Additionally, the empirical kinetic data will provide details in the factors that govern drug release rate from stimuli-responsive gels. 4. Insight of nanoparticle uptake into cells in response to changes in pH, temperature, and magnetic fields using confocal laser scanning microscope will be valuable to observe delivery and longevity of internalized drugs.
REFERENCES 1. Gu, H.; Yu, K.; Yang, Z.; Chang, C. K.; Yu, B. Synthesis and cellular uptake of porphyrin decorated iron oxide nanoparticles: a potential candidate for bimodal anticancer therapy. Chem. Commun. 2005, 34, 4270. 2. Detty, M. R.; Gibson, S. L.; Wagner, S. J. Current clinical and preclinical photosensitizers for use in photodynamic therapy. J. Med. Chem. 2004, 47, 3897. 3. Jordan, A.; Scholz, R.; Wust, P.; Fahling, H.; Felix, R. Magnetic fluid hyperthermia (MFH): cancer treatment with AC magnetic field induced excitation of biocompatible superparamagnetic nanoparticles. J. Magn. Magn. Mater. 1999, 201, 413. 4. Orive, G.; Hern´andez, R. M.; Gasc´on, A. R.; Dom´ınguez-Gil, A.; Pedraz, J. L. Drug delivery in biotechnology: present and future. Curr. Opin. Biotechol. 2003, 14, 659.
P1: OTA/XYZ P2: ABC c29 JWBS037-Chen February 1, 2011
4:29
Printer Name: Yet to Come
REFERENCES
723
5. Zhang, J. L.; Srivastava, R. S.; Misra, R. D. K. Core–shell magnetite nanoparticles surface encapsulated with smart stimuli-responsive polymer: synthesis, characterization and LCST of viable drug-targeting delivery system. Langmuir 2007, 23, 6342. 6. Rana, S.; Gallo, A.; Srivastava, R. S.; Misra, R. D. K. On the suitability of nanocrystalline ferrites as a magnetic carrier for drug delivery: Functionalization, conjugation, and drug release kinetics. Acta Biomater. 2007, 3, 233. 7. Weitman, S. D.; Lark, R. H.; Coney, L. R.; Fort, D. W.; Frasca, V.; Zurawski, V. R.; Kamen, B. A. Distribution of the folate receptor Gp38 in normal and malignant cell lines and tissues. Cancer Res. 1992, 52, 3396. 8. Antony, A. C. Folate receptors. Ann. Rev. Nutr. 1996, 16, 501. 9. Lu, Y.; Low, P. S. Folate-mediated delivery of macromolecular anticancer therapeutic agents. Adv. Drug Deliv. Rev. 2002, 54, 675. 10. Maziarz, K. M.; Monaco, H. L.; Shen, F.; Ratnam, M. Complete mapping of divergent amino acids responsible for differential ligand binding of folate receptors alpha and beta. J. Biol. Chem. 1999, 274, 11086. 11. Grief, A. D.; Richardson, G. Mathematical modelling of magnetically targeted drug delivery. J. Magn. Magn. Mater. 2005, 293, 455. 12. Giri, S.; Trewyn, B. G.; Stellmaker, M. P.; Lin, V.S.-Y. Stimuli-responsive controlled-release delivery system based on mesoporous silica nanorods capped with magnetic nanoparticles. Angew. Chem. Int. Ed. 2005, 44, 5038. 13. Liu, T. Y.; Hu, S. H.; Liu, T. Y.; Liu, D. M.; Chen, S. Y. Magnetic-sensitive behavior of intelligent ferrogels for controlled release of drug. Langmuir 2006, 22, 5974. 14. Soppimath, K. S.; Aminabhavi, T. M.; Dave, A. M.; Kumbar, S. G.; Rudzinski, W. E. Stimulusresponsive “smart” hydrogels as novel drug delivery systems. Drug Dev. Indust. Pharm. 2002, 28, 957. 15. Qiu, Y.; Park, K. Enviroment-sensitive hydrogels for drug delivery. Adv. Drug Deliv. Rev. 2001, 53, 321. 16. Huh, K. M.; Kumashiro, Y.; Ooya, T.; and Yui, N. New synthetic route for dextran graft copolymers containing thermo-responsive polymers. Polym. J. 2001, 33, 108. 17. Grinberg, V. Y.; Gronberg, N. V.; Usov, A. I.; Shusharina, N. P.; Khokhlov, A. R.; Kruif, K. G. de. Thermodynamics of conformational ordering of -carrageenan in KCl solutions using high-sensitivity differential scanning calorimetry. Biomacromolecules 2001, 2, 874. 18. Alarc´on, C. D. I. H.; Pennadam, S.; Alexander, C. Stimuli responsive polymers for biomedical applications. Chem. Soc. Rev. 2005, 34, 276. 19. Wu, J.; Su, Z. G.; Ma, G. H. A thermo- and pH-sensitive hydrogel composed of quaternized chitosan/glycerophosphate. Int. J. Pharm. 2006, 315, 1. 20. Rana, S.; Rawat, J.; Sorensson, M. M.; Misra, R. D. K. Antimicrobial function of Nd3+ -doped anatase titania-coated nickel ferrite composite nanoparticles: a biomaterial system. Acta Biomater. 2006, 2, 421. 21. Rana, S.; Rawat, J.; Misra, R. D. K. Anti-microbial active composite nanoparticles with magnetic core and photocatalytic shell: TiO2-NiFe2O4 bio-material system. Acta Biomater. 2005, 1, 691. 22. Nathani, H.; Gubbala, S.; Misra, R. D. K. Magnetic behavior of nanocrystalline nickel ferrite: Part I. The effect of surface roughness. Mater. Sci. Eng. B 2005, 121, 126. 23. Misra, R. D. K.; Gubbala, S.; Kale, A.; Egelhoff, W. F. Jr.; A comparison of the magnetic characteristics of nanocrystalline nickel, zinc, and manganese ferrites synthesized by reverse micelle technique. Mater. Sci. Eng. B 2004, 111, 164.
P1: OTA/XYZ P2: ABC c29 JWBS037-Chen February 1, 2011
724
4:29
Printer Name: Yet to Come
MAGNATIC NANOPARTICLE CARRIER FOR TARGETED DRUG DELIVERY
24. Gubbala, S.; Nathani, H.; Koizol, K.; Misra, R. D. K. Magnetic properties of nanocrystalline Ni–Zn, Zn–Mn, and Ni–Mn ferrites synthesized by reverse micelle technique. Physica B 2004, 348, 317. 25. Kale, A.; Gubbala, S.; Misra, R. D. K. Magnetic behavior of nanocrystalline nickel ferrite synthesized by the reverse micelle technique. J. Magn. Magn. Mater. 2004, 277, 350. 26. Zhang, J. L.; Rana, S.; Srivastava, R. S.; Misra, R. D. K. On the chemical synthesis and drug delivery response of folate-receptor-activated, polyethylene glycol-functionalized magnetite nanoparticles. Acta Biomater. 2008, 4, 40. 27. Gupta, A. K.; Wells, S. Surface-modified superparamagnetic nanoparticles for drug delivery: preparation, characterization, and cytotoxity studies. IEEE Trans. Nanobiosci. 2005, 41, 4137. 28. Etrych, T.; Jel´ınkov´a, M.; ?´ıhov´a, B.; Ulbrich, K. New HPMA copolymers containing doxorubicin bound via pH-sensitive linkage: synthesis and preliminary in vitro and in vivo biological properties. J. Control. Release 2001, 73, 89. 29. Christie, R. J.; Grainger, D. W. Design strategies to improve soluble macromolecular delivery constructs. Adv. Drug Deliv. Rev. 2003, 55, 421. 30. Willner, D.; Trail, P. A.; Hofstead, S. J.; King, H. D.; Lasch, S. J.; Braslawsky, G. R.; Greenfield, R. S.; Kaneko, T.; Firestone, R. A. (6-Maleimidocaproyl)hydrazone of doxorubicin—a new derivative for the preparation of immunoconjugates of doxorubicin. Bioconjug. Chem. 1993, 4, 521. 31. Firestone, R. A.; Willner, D.; Hofstead, S. J.; King, H. D.; Kaneko, T.; Braslawsky, G. R.; Greenfield, R. S.; Trail, P. A.; Lasch, S. J.; Henderson, A. J.; Casazza, A. M.; Hellstrom, I.; Hellstrom, K. E. Synthesis and antitumor activity of the immunoconjugate BR96-Dox. J. Control. Release 1996, 39, 251–259. 32. Trail, P. A.; Willner, D.; Knipe, J.; Henderson, A. J.; Lasch, S. J.; Zoeckler, M. E.; TrailSmith, M. D.; Doyle, T. W.; King, H. D.; Casazza, A. M.; Braslawsky, G. R.; Brown, J.; Hofstead, S. J.; Greenfield, R. S.; Firestone, R. A.; Mosure, K.; Kadow, K. F.; Yang, M. B.; Hellstrom, K. E.; Hellstrom, I. Effect of linker variation on the stability, potency, and efficacy of carcinoma-reactive BR64-doxorubicin immunoconjugates. Cancer Res. 1997, 57, 100. 33. King, H. D.; Yurgaitis, D.; Willner, D.; Firestone, R. A.; Yang, M. B.; Lasch, S. J.; Hellstrom, K. E.; Trail, P. A. Monoclonal antibody conjugates of doxorubicin prepared with branched linkers: a novel method for increasing the potency of doxorubicin immunoconjugates. Bioconjug. Chem. 1999, 10, 279. 34. Zhang, J.; Misra, R. D. K. Magnetic drug-targeting carrier encapsulated with thermosensitive smart polymer: Core–shell nanoparticle carrier and drug release response. Acta Biomater. 2007, 3, 838. ´ Moreau, P.; Leroux, J-C. On the characterization of pH35. Roux, E.; Lafleur, M.; Lataste, E.; sensitive liposome/polymer complexes. Biomacromolecules 2003, 4, 240. 36. H¨afeli, U. O. Magnetically modulated therapeutic systems. Int. J. Pharm. 2004, 277, 19. 37. Kono, K. Thermosensitive polymer-modified liposomes. Adv. Drug Deliv. Rev. 2001, 53, 307. 38. You, Y. Z.; Hong, C. Y.; Pan, C. Y.; Wang, P. H. Synthesis of a dendritic core–shell nanostructure with a temperature-sensitive shell. Adv. Mater. 2004, 16, 1953. 39. Needham, D.; Dewhirst, M. W. The development and testing of a new temperature-sensitive drug delivery system for the treatment of solid tumors. Adv. Drug Deliv. Rev. 2001, 53, 285. 40. Lemarchand, C.; Gref, R.; Couvreur, P. Polysaccharide-decorated nanoparticles. Eur. J. Pharm. Biopharm. 2004, 58, 327.
P1: OTA/XYZ P2: ABC c30 JWBS037-Chen January 25, 2011
13:43
Printer Name: Yet to Come
CHAPTER 30
Perfluorocarbon Nanoparticles: A Multidimensional Platform for Targeted Image-Guided Drug Delivery GREGORY M. LANZA∗ Department of Medicine, Washington University Medical School, St. Louis, Missouri, USA
SHELTON D. CARUTHERS† Department of Medicine, Washington University Medical School, St. Louis, Missouri, and Philips Healthcare, Andover, Massachusetts, USA
ANNE H. SCHMIEDER, PATRICK M. WINTER, TILLMANN CYRUS, and SAMUEL A. WICKLINE∗ Department of Medicine, Washington University Medical School, St. Louis, Missouri, USA
The expanding front of nanomedicine offers diagnostic and therapeutic tools to approach entrenched medical problems from new perspectives. Although these emergent techniques will augment the clinical armamentarium for treatment and monitoring of established pathology, the greatest opportunity may lie in the detection, quantification, and personalized treatment of early, even asymptomatic disease. Many nanotechnologies are being developed to address this opportunity, including perfluorocarbon (PFC) nanoparticles. PFC nanoparticles are a versatile platform technology with imaging capability across a broad spectrum of clinically relevant modalities, which may be used in concert with targeted drug delivery. This chapter delves into the in vivo drug delivery data developed with targeted PFC nanoparticles in cancer and cardiovascular disease.
30.1 INTRODUCTION Molecular imaging and targeted drug delivery are intertwined key opportunities emerging from the translation of nanotechnology into medicine, sometimes referred to as ∗ SAW † SDC
and GML are scientific co-founders and minority stockholders of Kereos, Inc. is an employee of Philips Healthcare
Nanoplatform-Based Molecular Imaging Edited by Xiaoyuan Chen C 2011 John Wiley & Sons, Inc. Copyright
725
P1: OTA/XYZ P2: ABC c30 JWBS037-Chen January 25, 2011
726
13:43
Printer Name: Yet to Come
PERFLUOROCARBON NANOPARTICLES
nanomedicine. The concept of molecular imaging with nanoparticles is an expansion of the nuclear medicine field of immunoscintigraphy to other diagnostic modalities to achieve noninvasive visualization or characterization of in vivo biological processes. In comparison with typical radiotracer technologies, nanoparticles offer considerably higher image resolution using commercial clinical magnetic resonance, ultrasound, and computed tomography scanners, but larger pharmaceutical dosages are required to offset lower sensitivities of detection. The higher dosages and greater complexity of targeted nanoparticle constructs unfortunately increase the development and regulatory challenges that must be addressed to prove platform safety, stability, and manufacturing control needed for investigational new drug (IND) approval to initiate a Phase I human clinical trial. For some strictly diagnostic nanotechnologies, the research and manufacturing expenses may outweigh the clinical value of its use or extend the time for development beyond its window of market opportunity. However, versatile theranostic platform technologies, that is, those with both diagnostic and drug delivery potential, are expected to provide a significantly improved value proposition, which help to justify the significant financial investment required for clinical translation.
30.2 PERFLUOROCARBON NANOPARTICLES: AN OVERVIEW Perfluorocarbon nanoparticles are a multifunctional platform technology with versatile imaging and drug delivery potential demonstrated in a variety of preclinical animal models (Fig. 30.1). They are comprised of a phospholipid monolayer encapsulated around a perfluorocarbon core (98% v/v). Perfluorocarbon nanoparticles differ distinctly from typical oil-based emulsions by virtue of the physicochemical properties of fluorine, the most electronegative of all elements. Fluorine has a high ionization potential and very low polarizability [1]. Larger than hydrogen, fluorine creates bulkier, stiffer compounds
FIGURE 30.1 Schematic diagram illustrating the versatility of the multifunctional perfluorocarbon nanoparticle platform.
P1: OTA/XYZ P2: ABC c30 JWBS037-Chen January 25, 2011
13:43
Printer Name: Yet to Come
PERFLUOROCARBON NANOPARTICLES FOR MOLECULAR IMAGING
727
that typically adopt a helical conformation. The C F bond is chemically and thermally stable and essentially biologically inert due to an encroachment barrier established by the dense electron cloud of fluorine in the chain [1]. Unusually, perfluorocarbons are extremely hydrophobic and lipophobic simultaneously! The biocompatibility of liquid fluorocarbons is well documented. Even at large doses, most fluorocarbons are innocuous and physiologically inactive. No toxicity, carcinogenicity, mutagenicity, or teratogenic effects have been reported for pure fluorocarbons within the 460–520 MW range. PFCs have tissue half-life residencies ranging from 4 days for perfluorooctylbromide up to 65 days for perfluorotripropylamine, and are not metabolized, but rather become slowly reintroduced from clearance tissues into the circulation in dissolved form by lipid carriers and ultimately expelled through the lungs [1]. The nominal size of perfluorocarbon nanoparticles can be manipulated by chemical processing techniques, the composition of the surfactant, and the PFC-to-surfactant ratio, but typical PFC nanoparticle size for molecular imaging is approximately 250 nm, which confines their distribution to the vasculature during the targeting phase. Unfortunately, few biochemical markers are pathognomonic for a specific disease and most are expressed in various densities by multiple cell types or within the interstitial matrix of numerous tissues. Steric vascular constraint eliminates potential interactions of homed PFC nanoparticles with extravascular sites of target epitope expression. Although vascular targeting precludes investigation of many important pathological sites of interest, it represents a logical first step in the translation of nanomedicine from the bench to the clinic.
30.3 PERFLUOROCARBON NANOPARTICLES FOR MOLECULAR IMAGING Perfluorochemicals have varied medical applications [2–7], but the inherent potential for imaging was first studied by Mattrey and colleagues for blood pool diagnostic applications with ultrasound (US), magnetic resonance (MR), and computed tomography (CT) [8–11]. A distillation of this body of work reveals that the effectiveness of PFC emulsions for blood pool US and CT imaging was limited by a relatively weak contrast effect versus microbubble or iodinated contrasts, respectively. For MRI, the PFC nanoparticles provided excellent luminal contrast for gastrointestinal studies, but the added benefit to the patient was offset by higher material costs relative to simpler, inexpensive barium-based agents. 30.3.1 Ultrasound Molecular Imaging Functionalizing PFC nanoparticles with homing ligands, for example, monoclonal antibodies, peptides, or peptidomimetics, coupled to the outer lipid surface was first demonstrated for fibrin targeting in dogs with clinical linear phased array transducers (7.5 MHz), where the acoustic conspicuity of acute arterial thrombus was markedly enhanced by the bound agent [12]. Importantly, circulating unbound PFC particles remained essentially invisible, allowing the targeted clot to be appreciated without concomitant background contrast interference [13]. The marked improvement in ultrasound contrast appreciated when PFC nanoparticles were bound versus free resulted from an increase in the specular reflection of thrombus from the acoustic impedance mismatch [14]. This acoustic mismatch results from a combination of a slower speed of sound in PFC (∼700 m/s) versus water (∼1450 m/s) and
P1: OTA/XYZ P2: ABC c30 JWBS037-Chen January 25, 2011
728
13:43
Printer Name: Yet to Come
PERFLUOROCARBON NANOPARTICLES
an increased density (e.g., ∼2.0 g/mL PFC vs. 1.0 g/mL H2 O @STP) [15–17]. Homed PFC nanoparticles form an effective layer over the surface, which can be mathematically represented as an acoustic transmission line model [18, 19]. Unlike contemporary ultrasound contrast agents of the time, which lasted for seconds and were poorly conducted through the pulmonary vasculature, the minute-sized highly stable PFC nanoparticles had prolonged circulation times during which fibrin epitopes were well saturated by the targeted agent. While for some particles, the nanosizing of material provides new physical properties, for many biological applications, the reduced size and resultant prolonged increased circulation has been one of the major advantages. Numerous other advances with the acoustic imaging of PFC nanoparticles have been published over the last decade, including their use in highfrequency intravascular applications [20–22], the induction of temperature shifts to enhance targeted contrast [23], and the particular sensitivity with novel ultrasound receivers, called information theoretical detectors [24, 25]. 30.3.2 Magnetic Resonance Molecular Imaging Another key advantage offered by nanoparticles, particularly for paramagnetic MR imaging, is their inherent increase in surface-to-volume ratio. Iron oxide nanoparticles, the first nanoparticles used for MR imaging, demonstrated excellent negative (i.e., dark) tissue contrast due to strong magnetic susceptibility artifacts [26–29]. Unfortunately, persistent circulation of these superparamagnetic metal oxides required delays of 24 h or more between the time of injection and subsequent imaging. Important advances in MR pulse sequences and image postprocessing techniques now reverse the dark contrast appearance [30–34], but the undesirable delays between treatment and imaging as well as dipole bloom artifacts persist. The prevailing dogma of the 1990s within the MR community was that targeted paramagnetic contrast agents, unlike iron oxides, would not provide adequate signal amplification for molecular imaging. However, the synthesis of lipid-based paramagnetic nanoparticles, polymerized liposomes [35] and PFC nanoparticles [36–39], facilitated the easy incorporation and presentation of large payloads of gadolinium in the outer surface, which resulted in bright, paramagnetic contrast when bound and concentrated at the target site. Although the first example of integrin-targeted polymerized paramagnetic liposomes required 24 h for adequate signal development in the Vx2 model [35], numerous examples of integrin-targeted PFC nanoparticles in multiple cancer and cardiovascular models have demonstrated robust, stable MR molecular imaging in 2 h or less [37, 40–42]. Moreover, paramagnetic PFC nanoparticles have sub-100 pM/voxel detection sensitivity using clinically available MR scanners and coils [43, 44]. Uniquely, the fluorine dense core of perfluorocarbon nanoparticles provides a second element for MR spectroscopy and imaging, which affords targeting confirmation, quantification, and multispectral benefits [38, 45, 46]. Fluorine has a high gyromagnetic ratio relative to the proton signal (i.e., 83%) and negligible background in the body. Building on the early perfluorocarbon emulsion work with MR fluorine spectroscopy and imaging [47–53], recent investigators have studied the use of PFC particles to image macrophages [54], dendritic T cells [55, 56], and stem cells [57, 58]. The development of clinical 1 H and 19 F imaging is rapidly progressing, and today 1 H and 19 F images can be acquired and displayed simultaneously, minimizing scanning time, ensuring proper registration, and providing unprecedented quantification of MR imaging results [59–63].
P1: OTA/XYZ P2: ABC c30 JWBS037-Chen January 25, 2011
13:43
Printer Name: Yet to Come
THERANOSTIC APPLICATIONS OF PERFLUOROCARBON NANOPARTICLES IN CANCER
729
30.4 THERANOSTIC APPLICATIONS OF PERFLUOROCARBON NANOPARTICLES IN CANCER 30.4.1 Angiogenesis Molecular Imaging Neovessel formation (i.e., angiogenesis) is an important biosignature of cancer. One molecular signature, ␣v 3 integrin, has garnered prominent attention for angiogenic targeting applications because it is expressed on the luminal surface of activated endothelial cells but not on mature quiescent cells. The ␣v 3 integrin, a heterodimeric transmembrane glycoprotein, is expressed by numerous cell types including endothelial cells [64, 65], macrophages [66], platelets [67], lymphocytes [67], smooth muscle cells [68], and tumor cells [69, 70]. Fortunately, the steric constraint of perfluorocarbon nanoparticles to the vasculature precludes significant interaction with nonendothelial integrin expressing cells, which greatly enhances neovascular target specificity [71, 72]. The ␣v 3 -integrin-targeted paramagnetic nanoparticles sensitively detect histologicalcorroborated angiogenic endothelium at 1.5 T in New Zealand white rabbits bearing Vx2 tumors (<1.0 cm) implanted into the hind limb 12 days previously [40]. In vivo competition studies in that report demonstrated that the ligand-directed homing was specific for the ␣v 3 -targeted nanoparticles. Moreover, the T1-weighted images obtained with ␣v 3 targeted nanoparticles differentiated the growing tumors from the inflammatory remnants of cancer, which were entirely infiltrated with macrophages and other white cells [40]. These rejected tumors were not differentiated from developing cancers using standard T2weighted MR imaging, nor would 18 F deoxyglucose PET imaging be expected to discern the high oxygen utilization of macrophage metabolism from proliferating Vx2 tumor cell demand. This simple example illustrates how MR molecular imaging may provide fundamental data that support medical decisions to biopsy, treat, or observe anomalous pathology or to interrogate early tumor response to chemotherapy or radiation treatment. In a challenging follow-up study in athymic mice, MR signal enhancement from the targeted angiogenic vasculature of minute xenograft melanoma tumors (ATCC C32, <40 mm3 ) was detected within 0.5 h following intravenous (IV) administration of ␣v 3 -integrin-targeted PFC particles, and this contrast increased in strength over the next 2 h [73]. In this xenograft model, like the Vx2 rabbit model, in vivo competition studies demonstrated the high specificity of ligand-directed targeting, which was further corroborated by ␣v 3 -targeted bimodal particles (fluorescent and MR) and immunofluorescent microscopy.
30.4.2 Angiogenesis Image-Guided Drug Delivery Antiangiogenesis therapy in conjunction with established chemotherapy or radiation therapy has become well established for lung, colon, and breast cancer [74–76]. However, the optimal effectiveness of antiangiogenic pretreatment is achieved only in a limited subset of patients; moreover, the clinical timing of the treatment and its duration are fixed and not individualized to response. Given the projected annual expense (>$100,000) of these therapies, the ever rising constraints on healthcare expense, and the potential adverse effects associated with the treatments (i.e., hypertension, proteinuria, hemorrhage, thrombosis, wound healing complications, and gastrointestinal perforation) [77–80], a compelling need exists to better risk stratify candidates for treatment. PFC nanotechnology offers an approach to better select and monitor treatment in the most eligible patients as well as to deliver the targeted antiangiogenic therapy. In addition to ultrahigh payloads of
P1: OTA/XYZ P2: ABC c30 JWBS037-Chen January 25, 2011
730
13:43
Printer Name: Yet to Come
PERFLUOROCARBON NANOPARTICLES
paramagnetic chelates for diagnostic MR imaging, ␣v 3 -targeted nanoparticles can incorporate therapeutic agents for targeted drug delivery [81–89]. Such dual function agents may be referred to as “theranostics.” Chemotherapeutics, such as paclitaxel, doxorubicin, rapamycin, and fumagillin, as well as toxic peptides, such as mellitin, and thrombolytic enzymes, have been incorporated into the surfactant of PFC nanoparticles. 30.4.3 Antiangiogenic Therapy with Integrin-Targeted Fumagillin Nanoparticles in Rabbits Fumagillin is a mycotoxin produced by Aspergillus fumigatus, which suppresses angiogenesis by inhibition of methionine aminopeptidase 2 (MetAP2) [90, 91]. MetAP2 is one of two methionine aminopeptidase forms in eukaryotes responsible for cleavage of the NH2 -terminal methionine residue from nascent proteins [92]. Although both MetAP1 and MetAP2 have common activity and substrates, only MetAP2 is upregulated during cellular proliferation [93]. MetAp2 has greater efficiency (1000-fold) catalyzing methionine removal from certain proteins, such as glyceraldehydes-3-phosphate, which is an important intermediate in glycolysis and gluconeogenesis [94, 95]. TNP-470 is a water-soluble functional analog of fumagillin that also alkylates His231 near the enzymatic center of MetAP2 to selectively inhibit proliferating endothelial cells (i.e., angiogenesis) and a subset of tumor cells, with little effect on nonendothelial cell types [90, 96]. The antitumor efficacy of TNP-470 was demonstrated in a wide variety of cancer models in rodents [97–100] and in human clinical trials [101–105], but at dosages required for therapeutic effects, TNP470 elicited sudden, moderately severe symptoms of neurotoxicity including weakness, nystagmus, diplopia, and ataxia [101, 102, 105]. Inclusion of fumagillin into the surfactant of the ␣v 3 -targeted fumagillin nanoparticles enables the delivery of drug into proliferating endothelial cells via “contact facilitated transport,” which is promoted by the ligand-based tethering of the nanoparticle to the target cell surface and the spontaneous passive exchange of the lipid surfactant components with similar membrane lipids through hemifusion complexes [81, 88, 89]. The antiangiogenic effectiveness of ␣v 3 -targeted fumagillin nanoparticles was studied in the syngeneic Vx2 adenocarcinoma rabbit model using a minimal fraction of the TNP-470 dosages previously used in preclinical animal models [87]. Vx2 tumor volume was reduced by one-half to twothirds among rabbits receiving ␣v 3 -targeted fumagillin nanoparticles when compared to animals given nontargeted fumagillin nanoparticles, ␣v 3 -targeted nanoparticles without drug, or saline. Vx2 tumor volume did not differ between animals receiving nontargeted fumagillin nanoparticles, ␣v 3 -targeted nanoparticles without drug, or saline alone. MR molecular imaging of angiogenesis with ␣v 3 -targeted paramagnetic nanoparticles in Vx2 tumor control rabbits on day 16 revealed a predominantly peripheral distribution of neovascularity accounting for 7.2% of the tumor rim volume, whereas the neovasculature in the core regions of the control tumors was very sparse (1.7%). The ␣v 3 -targeted fumagillin nanoparticles decreased neovascular signal enhancement in the rim to 2.8%, with no impact on the central contrast. Three-dimensional reconstruction and mapping of the contrast enhanced voxels over a chain model of the tumor surface demonstrated a coherent asymmetric peripheral distribution characterized by dense neovessel regions interspersed with a finer, reticular pattern (Fig. 30.2). MR imaging was corroborated by fluorescence microscopy, which demonstrated that ␣v 3 -targeted rhodamine nanoparticles were distributed to the periphery of the tumor constrained by the vasculature where they colocalized with FITC-lectin staining consistent with previous findings. Rhodamine particles were rarely
P1: OTA/XYZ P2: ABC c30 JWBS037-Chen January 25, 2011
13:43
Printer Name: Yet to Come
THERANOSTIC APPLICATIONS OF PERFLUOROCARBON NANOPARTICLES IN CANCER
731
FIGURE 30.2 Three-dimensional reconstruction of angiogenesis proliferative patterns at 3 hours postinjection obtained in the Vx2 adenocarcinoma model. Tumor was implanted into the popliteal fossa of rabbits receiving control and ␣v 3 -targeted fumagillin nanoparticles. Yellow contrast enhanced overlaid on 2D baseline slice. 3D reconstruction presents a gray volume-rendered tumor with blue contrast enhanced pixel overlay. Black arrows point to contrast enhanced regions of tumor, which must be adjacent to a source of existing vasculature. In this example, angiogenesis is derived from the spouting of vessels within the adjacent muscle fascia. Reproduced with permission from Winter et al. [87].)
appreciated within the mid or central regions of the tumor, which is consistent with both the paucity of neovasculature in these localities as well as the inability of nanoparticles to extravasate due to their size.
30.4.4 Antiangiogenic Theranostics in Mice Another biomarker for angiogenesis in cancer is the ␣5 1 integrin, which was studied using PFC paramagnetic nanoparticles in the MDA-MB-435 xenograft mouse model [85]. The ␣5 1 integrin, like ␣v 3 integrin, is an important adhesion molecule, which regulates endothelial cell migration and survival during neovascularization [106]. The ␣5 1 -integrin is poorly expressed on normal quiescent blood vessels, but its expression is induced on tumor blood vessels and in response to angiogenic factors, including basic fibroblast growth factor, interleukin-8, tumor necrosis factor-␣, and the angiomatrix protein Del-1 [107]. Similar to ␣v 3 integrin, ␣5 1 integrin regulates human endothelial cell vacuolation and lumen formation in three-dimensional (3D) fibrin matrices, and these morphogenic events are completely inhibited by the simultaneous addition of anti-␣v 3 -integrin and anti-␣5 -integrin antibodies [108]. While the role of ␣v 3 integrin is well documented in aggressive melanoma and breast cancer metastasis, ␣5 1 integrin is frequently expressed by low malignant potential tumors in addition to aggressive carcinomas (e.g., in ovarian cancers). MR theranostic agents directed against a spectrum of biomarkers will likely be required to properly
P1: OTA/XYZ P2: ABC c30 JWBS037-Chen January 25, 2011
732
13:43
Printer Name: Yet to Come
PERFLUOROCARBON NANOPARTICLES
characterize, segment, and treat the widely heterogeneous population of tumors and associated neovasculature. The potential utilities of ␣5 1 -(RGD peptide) and ␣5 1 (␣v 3 )-(peptidomimetic) PFC nanoparticles were compared in the MDA 435 xenograft mouse model to ␣v 3 (peptidomimetic) PFC nanoparticles as MR theranostics to characterize and map tumor angiogenesis and response to antiangiogenic therapy [85]. In the first in vivo phase of the project, the effectiveness of homing to the ␣5 1 integrin versus the ␣v 3 integrin was studied in tumor bearing mice (14d). The ␣5 1 -targeted paramagnetic nanoparticle contrast in the xenograft model was sparse (1.8 ± 0.4% of the tumor volume), predominantly in the rim (90%) yet markedly greater versus the irrelevant control, which enhanced only 0.2 ± 0.1% of the tumor volume. High-resolution three-dimensional neovascular maps depicted the peripheral asymmetric distribution of ␣5 1 -integrin neovascular expression withnegligible coherent contrast enhancement in the tumor interior; large regions of the tumor surface were devoid of signal enhancement. The percentage of surface neovasculature was notably less prominent than the results previously obtained with the Vx2 tumor model. Although routine immunohistological staining of tumors revealed the prevalence of ␣5 1 -integrin expression within both vascular and extravascular tissues, ␣5 1 -integrintargeted rhodamine nanoparticles were distributed in the tumor periphery, constrained to the vasculature, and colocalized with endothelial FITC-lectin (Fig. 30.3). Within the tumor core, prominent fluorescent signal from microvessel-bound lectin but not the ␣5 1 -integrintargeted rhodamine nanoparticles was observed. The pattern of ␣5 1 -targeted paramagnetic nanoparticle enhancement observed in the reconstructed images paralleled these microscopic observations. In the second phase of the study,mice bearing MDA-MB-435 xenografts received four serial treatments of either ␣v 3 -targeted fumagillin nanoparticles, ␣5 1 (␣v 3 )-targeted fumagillin nanoparticles, or saline control via tail vein injection. On day 22 postimplantation, MR T1-weighted images obtained with ␣5 1 (␣v 3 )-targeted paramagnetic nanoparticles showed that ␣5 1 (␣v 3 )-targeted fumagillin nanoparticles elicited a greater decrease (p < 0.05) in tumor neovasculature relative to the saline control than the ␣v 3 -targeted fumagillin nanoparticles. High-resolution 3D neovascular mapping further illustrated that the decrease in angiogenesis by treatment with ␣5 1 (␣v 3 )-targeted fumagillin nanoparticles manifested morphologically as smaller, contracted islands of neovasculature with a diminished appearance of speckled angiogenic signal (Fig. 30.4). This effect was validated by histological analysis of the decrease in angiogenesis. While ␣5 1 (␣v 3 )-targeted fumagillin nanoparticles decreased the already sparse neovasculature of the MDA-MB-435 xenograft to negligible levels, no differences in tumor volume were appreciated between the control (202 ± 55 mm3 ) and fumagillin-treated mice (199 ± 71 mm3 ). Similarly, no difference in tumor proliferation, as assessed microscopically by BrdU staining, was noted between the saline and ␣5 1 (␣v 3 )-targeted fumagillin treatment group. The effectiveness of targeted antiangiogenic therapies logically requires adequate expression of the biomarker and a dependence of the tumor growth on neovasculature expansion. In contradistinction to the Vx2 tumor model, the MDA-MB-435 tumor model had a very sparse neovasculature and further reduction of angiogenesis to minimal levels elicited no change in tumor volume. These early data suggest that angiogenic characterization of small tumors may be prognostic of subsequent response to antiangiogenics. Development and validation of an angiogenesis index to better stratify patients into medical strategies involving anti-VEGF therapy would help ensure that risks and costs of antiangiogenic agents are associated with the patients most likely to benefit from the treatment.
P1: OTA/XYZ P2: ABC c30 JWBS037-Chen January 25, 2011
13:43
Printer Name: Yet to Come
(a)
(b)
(c)
FIGURE 30.3 Histological examination of ␣5 1 -integrin expression and nanoparticle localization. (a) Fluorescent microscopy images (20×) of conventional immunohistochemistry from tumor border stained for ␣5 1 integrin (green) and endothelium (red), with nuclei counterstained blue (DAPI). There is prevalent expression of ␣5 1 -integrin throughout the tissue section, both within and outside the vasculature. (b,c) Images of 30-m tumor sections from mice intravenously injected with integrintargeted nanoparticles (red) and endothelium-avid lectin (green). (b) The merged images confirmed that both particles are constrained to the vasculature and are colocalized within the angiogenic vasculature. Note that no ␣5 1 integrin expressed abundantly in the extravascular regions of the tumor or by other cells were associated with ␣5 1 -targeted nanoparticles. (c) ␣5 1 -targeted nanoparticles (red) were absent from the mature vasculature in the tumor core despite marked staining with lectin (green), consistent with the MR findings. (White scale bar = 200 m) (Reproduced with permission from Schmieder et al. [85].)
P1: OTA/XYZ P2: ABC c30 JWBS037-Chen January 25, 2011
734
13:43
Printer Name: Yet to Come
PERFLUOROCARBON NANOPARTICLES
Tumor Volume Enhanced (%)
2.5
NS
2
1.5
1
0.5
0
ανβ3α5β1(ανβ3)Fumagillin NP Fumagillin NP
Saline
3D MR Reconstruction of Sparse Angiogenesis in MDA-435 Tumor
1cm
Saline
α5β1 (ανβ3)-targeted Fumagillin NP
FIGURE 30.4 Assessment of the antiangiogenic response to integrin-targeted fumagillin nanoparticles with ␣5 1 (␣v 3 )-targeted MR contrast agent. (Top) The extent of neovascularity was quantified by calculating the amount of signal enhancement in the tumor periphery. The ␣5 1 (␣v 3 )-targeted fumagillin nanoparticles reduced peripheral tumor neovascularity relative to control (p < 0.05, n = 5). ␣v 3 -targeted fumagillin nanoparticles had no significant effect on angiogenesis, compared to control. (Bottom) The effect of ␣5 1 (␣v 3 )-targeted fumagillin nanoparticles on tumor neovascular morphology is clearly apparent on 3D reconstructions of MR signal enhancement. Tumor volume is outlined in gray and contrast enhanced pixels are shown in blue. There was no effect of tumor growth by targeted fumagillin despite almost complete elimination of angiogenesis, suggesting that cancers with sparse neovasculatures may not be responsive to purely antiangiogenic formulations. (Reproduced with permission from Schmieder et al. [85].)
30.4.5 Anticancer Therapy with Cytolytic Peptides and Perfluorocarbon Nanoparticles Host defense peptides are an interesting class of small amphipathic peptides (10–50 a.a.) found in most eukaryotic cells that serve diverse functions stemming from their antibiotic, anticancer, and anti-inflammatory properties [109–113]. These peptides rapidly associate
P1: OTA/XYZ P2: ABC c30 JWBS037-Chen January 25, 2011
13:43
Printer Name: Yet to Come
NANOMEDICINE APPLICATIONS IN CARDIOVASCULAR DISEASE
735
with phospholipid cell membranes creating structural defects (e.g., pores) and induce lethal biochemical changes [114]. Melittin is a cytolytic peptide derived from honeybee venom that nonspecifically attacks all lipid membranes, leading to significant toxicity when injected intravenously. Cytolytic peptides have shown promise in preclinical cancer studies, conjugated to homing peptides and coupled to iron oxide particles [115–118]. PFC nanoparticles provide a stable in vivo mellitin delivery platform for advanced or preclinical tumors with minimal off-target toxicity. PFC nanoparticle–melittin complexes, referred to as “NanoBees,” have cell-selective homing potential, which is trackable and quantifiable with MR 19 F imaging or spectroscopy [119, 120]. Melittin nanoparticles were successfully demonstrated in three rodent cancer models. In the xenograft MDA-435 model, mice implanted with tumor cells in the footpad were injected intravenously with NanoBees every 3 days between 7 and 22 days [120]. Ultrasound measurements of tumor volume produced 25% growth inhibition relative to controls. In syngeneic B16-F10 mouse melanoma tumors implanted subcutaneously in C57BL/6 mice, a dramatic reduction of melanoma growth was observed after four intravenous injections of melittin nanoparticles every other day starting at day 5. The final tumor weight on day 14 was 80% lower for the treated mice compared to saline controls and or mice given nanoparticles alone without melittin. Histology revealed decreased blood vessels (CD31 staining), decreased proliferating cells, and significantly increased areas of necrosis in the melittin nanoparticle treated tumors (Fig. 30.5). In the K14-HPV16 mouse model, a genetically engineered model of squamous carcinoma associated with abundant angiogenesis during the precancerous dysplastic and in the invasive squamous cancer stage, integrin-targeted NanoBees elicited a quantitative 80% reduction in dermal papillae, which is the anatomical site of the genetically driven dysplastic changes in this model. The addition of free melittin to cells in vitro results in necrosis, characterized by LDH release due to the disruption of cell membranes and cytochrome c release following mitochondria damage. Cells treated with targeted melittin nanoparticles in vitro released cytochrome c from their mitochondria but only minimal LDH was measured, a pattern suggestive of increased apoptosis. Even at high concentrations, melittin delivered to the cell surface via PFC nanoparticles drove cells into apoptotic and not necrotic cell death. This example illustrates the ability of nanoparticle-based drug delivery to influence the therapeutic mechanisms of action.
30.5 NANOMEDICINE APPLICATIONS IN CARDIOVASCULAR DISEASE Atherosclerotic plaque progresses from an early atheromatous lesion to a thin-capped vulnerable plaque through aggressive inflammatory and immune responses, comprising macrophage infiltration with necrotic core enlargement, neovascular expansion of the vasa vasorum, and intraplaque hemorrhage [121–123]. Increased plaque angiogenesis, driven by hypoxia [124], proangiogenic growth factors [125], and oxidative stress [126] portends unstable vascular disease [121, 122]. Angiogenesis is correlated with plaque ruptures [121] and is associated with the morphological features of vulnerable atheroma including macrophage infiltrated fibrous caps [121], lipid-rich cores [127], and thin-cap shoulders [121]. Reverse cholesterol transport via HDL serves to eliminate cholesterol from the plaque and other body tissues, particularly adipose depots, but large amounts of cholesterol are retained in the vascular wall as droplets of the cholesterol ester. The cholesterol burden is further exacerbated by the additional load imparted from erythrocyte membranes deposited
P1: OTA/XYZ P2: ABC c30 JWBS037-Chen January 25, 2011
736
13:43
Printer Name: Yet to Come
PERFLUOROCARBON NANOPARTICLES
(b) 3000
Saline
2500
Saline control
Melittin NP
Nanoparticles
H and E
2000 melittin nanoparticles 8mg/kg
1500 1000
dead
**
500 0 7
9
11
14
day post implant
Melittin NP
CD31
Saline control
BrdU
tumor volume (cu. mm)
(a)
100 µm
FIGURE 30.5 Therapeutic efficacy of melittin nanoparticles in syngeneic B16F10 mouse melanoma tumors. (a) Graph showing the increase in tumor volume of B16F10 melanoma tumors during the course of treatment with melittin nanoparticles (8.5 mg/kg) or controls (saline or nanoparticles alone, n = 5 each group). Note the dramatic difference in tumor volume at day 14 after four doses of melittin nanoparticles. Data are represented as mean ± SD. **p < 0.01. (b) Histological assessment of B16-F10 melanoma tumors excised at day 14. Note the extensive nonproliferating dead areas in the treated tumors along with the markedly decreased vascularity. (Reproduced with permission from Soman et al. [120].)
through intraplaque hemorrhage, which arises from the disruption of fragile angiogenic vessels [123, 128]. Based on the pathobiology of atherosclerotic plaque development and the preponderance of data from experimental models and human pathological samples, assessments of plaque neovasculature may serve as a molecular imaging biomarker of atherosclerotic severity and cardiovascular disease risk, as well as a target to passivate plaque progression. The American Heart Association and American College of Cardiology have endorsed the use of HMG-CoA reductase inhibitors for cardiovascular risk reduction achieved by lowering LDL cholesterol [129], but their utility may exceed the primary lipid-lowering effect. The pleiotropic benefits of statins have been attributed to antioxidant effects [130], diminished leukocyte–endothelial cell adhesion [131], attenuated macrophage activation and cytokine release [132], increased endothelial nitric oxide activity [133], or atherosclerotic plaque stabilization [134]. Pathological data from excised carotid arteries of patients treated with statins have revealed a reduction in microvascular density, which was suggested to account for the additional benefit of statins [135]. While clear direct evidence of decreased intraplaque angiogenesis due to statins and subsequent clinical improvement
P1: OTA/XYZ P2: ABC c30 JWBS037-Chen January 25, 2011
13:43
Printer Name: Yet to Come
NANOMEDICINE APPLICATIONS IN CARDIOVASCULAR DISEASE
737
needs further investigation, a normalized vasculature achieved through neovascular pruning may be less prone to intraplaque hemorrhage and promote plaque stabilization [123, 136, 137]. Moreover, assuming that the impact of antineovascular therapy demonstrated in cancer is predictive of effects in atherosclerosis, antiangiogenesis may be used to normalize the plaque microenvironment, to restore pressure gradients across capillary walls, to increase drug penetration, and to improve plaque oxygenation [138, 139]. We have shown that ␣v 3 -targeted paramagnetic nanoparticles can be used to quantify angiogenesis in atherosclerosis [41] and with the incorporation of fumagillin, this theranostic agent delivered an acute antiangiogenic effect [82] that was demonstrated by MR molecular imaging and confirmed directly by microscopic counts of neovessels in the treated and control aortas. However, the effectiveness of antiangiogenesis therapy is transient and must be combined with a complementary treatment, such as the FDA approved combined use of Bevacizumab® and 5-fluorouracil in GI cancer. In a series of in vivo experiments performed in hyperlipidemic rabbits, the antiangiogenic pharmacodynamic effects of a single ␣v 3 -targeted fumagillin nanoparticle dosage was defined, which led to a second study to sustain the antiangiogenic effects of ␣v 3 -targeted fumagillin nanoparticles further with oral atorvastatin, an HMG-CoA reductase inhibitor [86].
30.5.1 Antiangiogenic Effect After Treatment with ␣ v 3 -Integrin-Targeted Fumagillin Nanoparticles The hypothesis tested in study I was whether a single pulsed dose of ␣v 3 -targeted fumagillin nanoparticles could persistently suppress angiogenesis in the aortic wall. The duration of the antiangiogenic response was quantified with MR molecular imaging of the neovasculature. Hyperlipidemic rabbits were administered ␣v 3 -targeted, paramagnetic nanoparticles with or without fumagillin at baseline and were reimaged with ␣v 3 -targeted, paramagnetic nanoparticles (no drug) to assess angiogenesis response weekly for 1 month. Half of the animals were continued on a high cholesterol chow while the remainder were switched to a normal rabbit diet at baseline. At the start of study I, the average percent signal enhancement from ␣v 3 -targeted paramagnetic nanoparticles ranged between 20% and 25% for the six treatment groups (p > 0.05), which was consistent with two prior studies [41, 82]. (Fig. 30.6). The ␣v 3 targeted paramagnetic contrast was diffusely distributed across and within slices and the enhancement calculated represented the average of all aortic voxels rather than a thresholdsegmented subset. Targeted characterization of ␣v 3 -integrin expression revealed that MR neovascular signal enhancement did not differ between fat-fed animals switched to standard chow and those continuing on the cholesterol-enriched diet over the 4-week study regardless of treatment. On week 1 post-treatment, neovascular contrast among animals receiving ␣v 3 -targeted fumagillin nanoparticles decreased approximately 50% to 75% relative to the control animals (p < 0.05). The decreased neovascular signal persisted through week 3 and was not different from controls by week 4. Neither the duration nor magnitude of antiangiogenic response was improved significantly by increasing the fumagillin dose from 30 g/mL to 90 g/mL over the 4-week trial, although the higher dose exhibited a trend toward longer effectiveness when compared to its own baseline angiogenesis signal.
P1: OTA/XYZ P2: ABC c30 JWBS037-Chen January 25, 2011
738
13:43
Printer Name: Yet to Come
PERFLUOROCARBON NANOPARTICLES
Signal Enhancement (%)
30 25 Controls 20 15 Normal Chow 10 Cholesterol-rich
5 Fumagillin
0 0
1
2 Week of Study
3
4
Signal Enhancement (%)
30 25
Controls
20 15 10 Fumagillin 90 µg/kg
5 Fumagillin 30 µg/kg 0 0
1
2 Week of Study
3
4
FIGURE 30.6 Aortic signal enhancement up to 4 weeks post-treatment with targeted fumagillin nanoparticles. (Top) Fumagillin treatment at week 0 reduced angiogenesis compared to untreated controls, but withdrawing the high-cholesterol feed had no effect. (Bottom) Fumagillin treatment at 30 vs. 90 g/kg produced identical responses (*p < 0.05). (Reproduced with permission from Winter et al. [86].)
30.5.2 Antiangiogenic Synergism of ␣ v 3 -Targeted Fumagillin Nanoparticles and Atorvastatin In study II, the potential of HMG-CoA reductase therapy (i.e., atorvastatin) to sustain the acute antiangiogenic response of ␣v 3 -targeted fumagillin nanoparticles was evaluated. Aortic neovasculature signal was followed for 8 weeks with serial MR molecular imaging at baseline and on weeks 1, 2, 4, 6, and 8 in hyperlipidemic rabbits maintained on a cholesterol-enriched diet, with a portion of the animals receiving feed supplemented with atorvastatin throughout the 8-week interval. The ␣v 3 -targeted fumagillin treatment was given either once (baseline only) or twice (baseline and week 4). Nearly identical to the earlier 4-week experiment, the aortic MR signal enhancement of hyperlipidemic rabbits given ␣v 3 -targeted paramagnetic nanoparticles averaged between
P1: OTA/XYZ P2: ABC c30 JWBS037-Chen January 25, 2011
13:43
Printer Name: Yet to Come
NANOMEDICINE APPLICATIONS IN CARDIOVASCULAR DISEASE
739
30
Signal Enhancement (%)
Statin Only
Control
25 20 15 10
Fumagillin 0 & 4 wks + No Statin
5 0 0
1
2
3 4 5 Week of Study
6
7
8
30 Statin Only Signal Enhancement (%)
25 20 15
Fumagillin 0 wk + Statin
10 5 Fumagillin 0 & 4 wks + Statin 0 0
1
2
3 4 5 Week of Study
6
7
8
FIGURE 30.7 MR signal enhancement up to 8 weeks post-treatment with targeted fumagillin nanoparticles and/or oral atorvastatin. (Top) Untreated and statin treated animals showed a constant level of angiogenesis in the aortic wall. Animals treated with targeted fumagillin nanoparticles at weeks 0 and 4 showed decreased angiogenesis (*p < 0.05) following each dose, which returned to baseline levels within 4 weeks. (Bottom) The combination of two fumagillin doses and statin produced a sustained decrease in angiogenesis (*p < 0.05). (Reproduced with permission from Winter et al. [86].)
20% and 25% at baseline (Fig. 30.7). The neovascular contrast enhancement among hyperlipidemic control rabbits was constant over the 8-week study. Atorvastatin alone did not significantly affect the MR signal enhancement from the aortic neovasculature over the trial. The ␣v 3 -targeted fumagillin nanoparticles elicited the same acute, antiangiogenic response as observed and discussed previously. The first dosage of ␣v 3 -targeted fumagillin nanoparticles decreased the angiogenic signal but neovascular recrudescence to the baseline level was apparent at 4 weeks. The second administration of ␣v 3 -targeted fumagillin
P1: OTA/XYZ P2: ABC c30 JWBS037-Chen January 25, 2011
740
13:43
Printer Name: Yet to Come
PERFLUOROCARBON NANOPARTICLES
nanoparticles mirrored the antiangiogenic effects observed after the first dosage without any apparent carry-over benefit. Continued treatment of dietary atorvastatin following a single dose of ␣v 3 -targeted fumagillin nanoparticles at baseline also produced no antiangiogenic benefit in the following 4-week interval. Serial treatment with ␣v 3 -targeted fumagillin nanoparticles at baseline and week 4 in conjunction with dietary atorvastatin resulted in the pharmacodynamic 4-week cyclic pattern during the first 4-week dosage period but the combination of the statin and the second fumagillin treatment resulted in sustained reduction on neovascular signal in weeks 6 and 8 rather than the return to baseline levels. This result suggests that the effect of statins requires some period of treatment, one month in this experiment, before its impact on angiogenesis is manifested. Alone, atorvastatin exerted no significant effect on plaque angiogenesis in this short study, but following treatment with ␣v 3 -targeted fumagillin, dietary statin therapy provided a chronic synergistic benefit that maintained the acute antineovascular drug effect. The potential role of antiangiogenic therapy in the treatment of atherosclerosis has garnered increasing scientific discussion both pro [136, 140] and con [122, 141, 142]. This work illustrated how a theranostic nanomedicine approach may be interwoven with standard clinical practice to provide a sustained antiangiogenic treatment regimen. Moreover, noninvasive MR longitudinal monitoring of atherosclerotic disease with ␣v 3 -targeted paramagnetic particles alone or enhanced with MR intraplaque hemorrhage assessment [143] offers an attractive, quantitative approach to continued medical management, particularly in asymptomatic patients.
30.5.3 Nanomedicine Approaches to Restenosis Following Angioplasty Vascular restenosis negates the clinical benefit of percutaneous angioplasty and can elicit recurrent anginal discomfort and myocardial damage or impede peripheral circulation. Drug-eluting stents (DES) have convincingly demonstrated that local release of antiproliferative agents ameliorate the restenosis response to coronary angioplasty [144–150]. However, placement of these devices in small target vessels, very long lesions, difficult anatomical locations (e.g., peripheral bifurcations or in kinked vessels), or in chronic total vascular occlusions may not always be feasible. Conversely, in large diameter arteries, such as iliac or femoral arteries, placement of stents with large inter-strut distances often requires higher drug loading, which may increase the potential for side effects or risk inadequate drug delivery [151]. Recently, late in-stent thrombosis has been associated with DES, reflecting the challenge of reendothelialization when delivering antiproliferative agents into the vessel wall from the stent–intima interface [152, 153]. Aggressive dual (and occasionally triple) antiplatelet therapy is employed for 6 months to a year to avoid acute thrombosis, but we now recognize that some patients are nonresponders to one or more of these drugs [154, 155]. Discontinuation of antiplatelet therapy secondary to noncompliance, bleeding complications, or the need for emergent surgery increases the likelihood for thrombotic complication. Moreover, the risk of late in-stent thrombosis may persist up to 30 months after DES implantation [156, 157]. More recently, DES have been shown to induce the accelerated formation of a fat-laden (yellow) atherosclerotic plaque on the healed stents, which commonly is associated with surface fibrin and plaque instability. Collectively, these data suggest that DES are effective but only palliative, and a final answer to the atherosclerotic restenosis issue is still awaited [158, 159].
P1: OTA/XYZ P2: ABC c30 JWBS037-Chen January 25, 2011
13:43
Printer Name: Yet to Come
NANOMEDICINE APPLICATIONS IN CARDIOVASCULAR DISEASE
741
Perfluorocarbon nanoparticles have been modified to target a variety of cellular and matrix biomarkers in vivo for imaging with MRI or ultrasound to delineate the morphology of the balloon-injured wall. [21, 22, 160] Recently, the effectiveness of ␣v 3 -targeted nanoparticles with rapamycin delivered locally through microfractures created by balloon overstretch injury was demonstrated to inhibit acute stenosis without delaying the rate of endothelial healing [84]. Femoral arteries of 6-month-old NZW rabbits, which had been fed a high-cholesterol diet for 4 months to induce atherogenesis, were randomized to local treatment using ␣v 3 targeted nanoparticles with 0.4 mol% rapamycin, with ␣v 3 -targeted nanoparticles without drug, with nontargeted nanoparticles with 0.4 mol% rapamycin, or saline for 5 min after balloon overstretch injury. Noninvasive MR imaging at 1.5 T demonstrated the presence of the integrin-targeted paramagnetic nanoparticles within the injured arterial wall segments 30–40 m after delivery and reestablishment of blood flow. T1 -weighted black blood MR images of vascular segments exposed to ␣v 3 -targeted paramagnetic nanoparticles, with or without drug, showed MR signal enhancement, whereas virtually no signal enhancement was detected following treatment with nontargeted paramagnetic nanoparticles or saline. Baseline MR time-of-flight angiography within 60 min. of angioplasty and treatment demonstrated patent arteries without luminal flow obstruction or vascular wall dissection in all treatment groups. Two weeks later, prominent vascular luminal irregularities were observed in vessels that had been exposed to ␣v 3 -targeted nanoparticles without drug, to nontargeted nanoparticles with rapamycin, or to saline. In contradistinction, segments treated with ␣v 3 -targeted rapamycin nanoparticles had minimal or no lumen irregularities detectable by MR angiograms (Fig. 30.8).
FIGURE 30.8 (a) MR time-of-flight angiogram 30 min after balloon stretch-injury depicting patent femoral arteries treated with ␣v 3 -integrin-targeted paramagnetic nanoparticles with rapamycin (left artery) and saline in the right artery. (b,c) MR angiograms 2 weeks after injury and treatment. (b) The ␣v 3 -integrin-targeted nanoparticles without drug (right) with arterial plaque versus the widely patent contralateral artery treated with ␣v 3 -integrin-targeted nanoparticles with rapamycin (left). (c) The ␣v 3 -integrin-targeted nanoparticles with rapamycin in the widely patent right femoral artery versus the partially occluded left artery treated with nontargeted nanoparticles with rapamycin (left). Arrows identify regions of intraluminal plaque due to balloon overstretch injury. (Reproduced with permission from Cyrus et al. [84].)
P1: OTA/XYZ P2: ABC c30 JWBS037-Chen January 25, 2011
742
13:43
Printer Name: Yet to Come
PERFLUOROCARBON NANOPARTICLES Average vascular stenosis
Average vascular stenosis
(a)
30 25 20 15 10 5 0
P = 0.034
P = 0.022
% stenosis
% stenosis
P = 0.004 P = 0.045
targeted
nontargeted
rapamycin
targeted
saline
no drug
(b)
30 25 20 15 10 5 0
P = 0.036 P = 0.048
targeted
nontargeted
rapamycin
targeted
saline
no drug
FIGURE 30.9 (a) Total average percentage lesion areas of the femoral arteries of NZW rabbits 2 weeks after balloon injury. Arterial segments were flash-frozen in OCT and alternate 7-m sections used for morphological analysis (H&E stains). (b) Maximum average stenosis within the injured and treated vascular segment. (Reproduced with permission from Cyrus et al. [84].)
Microscopic morphometric analysis of artery segments treated with ␣v 3 -targeted rapamycin nanoparticles revealed decreased (p < 0.05) neointimal proliferation throughout the balloon-injured segment by approximately 40% compared to segments exposed to nontargeted nanoparticles with rapamycin, ␣v 3 -integrin-targeted nanoparticles without drug, and the saline controls, respectively. Treatment with nontargeted nanoparticles with rapamycin or ␣v 3 -integrin-targeted nanoparticles without drug produced no significant improvement compared to saline controls (Fig. 30.9). Detailed analysis of intimal endothelial healing was performed on femoral arteries on postinterventional days 1, 7, 14, and 28 using longitudinally opened en face preparations of the balloon injured vascular segment. Relative to the initial area at risk, 80% of the endothelium was damaged 1 h after balloon injury. Gradual healing was observed over the subsequent 4-week time with no differences in the rate of endothelial healing observed among arteries treated with ␣v 3 -targeted rapamycin nanoparticles and any of the control groups (Fig. 30.10). These findings illustrate the potential benefit of targeted nanoparticulate agents to deliver and to achieve the antirestenotic benefits of rapamycin in lesions that are not amenable to stent placement or in cases where implantation of nondrug eluting stents may be preferred. Although not required for clinical application, the high paramagnetic contrast potential of these particles offered a unique tracking marker in the present study and provided important information in the context of MRI interventional procedures. The potential of this intramural approach to local postangioplasty nanoprevention of restenosis was recently corroborated using an alternative matrix-targeted nanoparticle incorporating prednisolone [161].
30.6 CONCLUSION Nanomedicine clearly offers unique tools to address intractable medical problems in cancer and cardiovascular disease from entirely new perspectives. Among the theranostic options emerging in this new wave of biotechnology development, the perfluorocarbon nanoparticles have shown robust potential in vivo for diagnosing, characterizing, treating, and following proliferating cancers, progressive atherosclerosis, and acute postangioplasty restenosis. Their unique chemical nature provides a stable platform for effective treatment of cancers with cytolytic peptides, particularly mellitin. In each application, imaging can
P1: OTA/XYZ P2: ABC c30 JWBS037-Chen January 25, 2011
13:43
Printer Name: Yet to Come
REFERENCES
(a)
743
(b)
Endothelial injury
% 100 80
Rapamycin
60
Saline
40 20 0 1 day
1 week
2 weeks
4 weeks
FIGURE 30.10 Area at risk of injured endothelium quantified on vascular en face preparations stained with Carstair’s stain. (a) Noninjured endothelium (yellow). (b) Injured endothelium with fibrin deposition (red). (c) Quantitation of injured endothelium in area at risk (100% = 1 cm excised vessel segment). Digitized images were analyzed on areas that had undergone balloon overstretch injury and were treated with ␣v 3 -integrin-targeted nanoparticles with 0.4 mol% rapamycin (n = 12), or saline control (n = 12). Vessels were excised on postinterventional days 1, 7, 14, and 28 (n = 3 per group and time point). (Reproduced with permission from Cyrus et al. [84].)
be used to confirm delivery, provide dosimetry, and noninvasively follow early response. Image-guided drug delivery based on nanotechnology is emerging as a powerful clinical opportunity.
REFERENCES 1. Krafft, M. Fluorocarbons and fluorinated amphiphiles in drug delivery and biomedical research. Adv. Drug. Deliv. Rev. 2001, 47, 209–228. 2. Geyer, R. P. “Bloodless” rats through the use of artificial blood substitutes. Fed. Proc. 1975, 34, 1499–1505. 3. Sloviter, H. A. Perfluoro compounds as artificial erythrocytes. Fed. Proc. 1975, 34, 1484–1487. 4. Mitsuno, T.; Ohyanagi, H.; Naito, R. Clinical studies of a perfluorochemical whole blood substitute (Fluosol-DA). Summary of 186 cases. Ann. Surg. 1982, 195, 60–69. 5. Kaufman, R. Clinical development of perfluorocarbon-based emulsions as red cell substitutes. In Emulsions and Emulsion Stability. Marcel Dekker: New York, 1996, pp. 343–367. 6. Modell, J. H.; Calderwood, H. W.; Ruiz, B. C.; Tham, M. K.; Hood, C. I. Liquid ventilation of primates. Chest 1976, 69, 79–81.
P1: OTA/XYZ P2: ABC c30 JWBS037-Chen January 25, 2011
744
13:43
Printer Name: Yet to Come
PERFLUOROCARBON NANOPARTICLES
7. Chang, S.; Zimmerman, N. J.; Iwamoto, T.; Ortiz, R.; Faris, D. Experimental vitreous replacement with perfluorotributylamine. Am. J. Ophthalmol. 1987, 103, 29–37. 8. Mattrey, R. F.; Long, D. M.; Multer, F.; Mitten, R.; Higgins, C. B. Perfluoroctylbromide: a reticuloendothelial-specific and tumor-imaging agent for computed tomography. Radiology 1982, 145, 755–758. 9. Mattrey, R. F.; Leopold, G. R.; vanSonnenberg, E.; Gosink, B. B.; Scheible, F. W.; Long, D. M. Perfluorochemicals as liver- and spleen-seeking ultrasound contrast agents. J. Ultrasound. Med. 1983, 2, 173–176. 10. Mattrey, R. F.; Hajek, P. C.; Gylys-Morin, V. M.; Baker, L. L.; Martin, J.; Long, D. C.; Long, D. M. Perfluorochemicals as gastrointestinal contrast agents for MR imaging: preliminary studies in rats and humans. AJR. Am. J. Roentgenol. 1987, 148, 1259–1263. 11. Mattrey, R. F.; Long, D. C. Potential role of PFOB in diagnostic imaging. Invest. Radiol. 1988, 23 (Suppl 1), S298–S301. 12. Lanza, G.; Wallace, K.; Scott, M.; Cacheris, W.; Abendschein, D.; Christy, D.; Sharkey, A.; Miller, J.; Gaffney, P.; Wickline, S. A novel site-targeted ultrasonic contrast agent with broad biomedical application. Circulation 1996, 94, 3334–3340. 13. Hughes, M. S.; Marsh, J. N.; Hall, C. S.; Fuhrhop, R. W.; Lacy, E. K.; Lanza, G. M.; Wickline, S. A. Acoustic characterization in whole blood and plasma of site-targeted nanoparticle ultrasound contrast agent for molecular imaging. J. Acoust. Soc. Am. 2005, 117, 964–972. 14. Hall, C. S.; Marsh, J. N.; Scott, M. J.; Gaffney, P. J.; Wickline, S. A.; Lanza, G. M. Time evolution of enhanced ultrasonic reflection using a fibrin-targeted nanoparticulate contrast agent. J. Acoust. Soc. Am. 2000, 108, 3049–3057. 15. Hall, C. S.; Lanza, G. M.; Rose, J. H.; Kaufmann, R. J.; Fuhrhop, R. W.; Handley, S. H.; Waters, K. R.; Miller, J. G.; Wickline, S. A. Experimental determination of phase velocity of perfluorocarbons: applications to targeted contrast agents. IEEE. Trans. Ultrason. Ferroelectr. Freq. Control 2000, 47, 75–84. 16. Marsh, J. N.; Hall, C. S.; Scott, M. J.; Fuhrhop, R. W.; Gaffney, P. J.; Wickline, S. A.; Lanza, G. M. Improvements in the ultrasonic contrast of targeted perfluorocarbon nanoparticles using an acoustic transmission line model. IEEE. Trans. Ultrason. Ferroelectr. Freq. Control 2002, 49, 29–38. 17. Marsh, J. N.; Hall, C. S.; Wickline, S. A.; Lanza, G. M. Temperature dependence of acoustic impedance for specific fluorocarbon liquids. J. Acoust. Soc. Am. 2002, 112, 2858–2862. 18. Lanza, G.; Trousil, R.; Wallace, K.; Rose, J.; Hall, C.; Scott, M.; Miller, J.; Eisenberg, P.; Gaffney, P.; Wickline, S. In vitro characterization of a novel, tissue-targeted ultrasonic contrast system with acoustic microscopy. J. Acoust. Soc. Am. 1998, 104, 3665–3672. 19. Marsh, J.; Hall, C.; Scott, M.; Fuhrhop, R.; Gaffney, P.; Wickline, S.; Lanza, G. Improvements in the ultrasonic contrast of targeted perfluorocarbon nanoparticles using an acoustic transmission line model. IEEE. Trans. Ultrason. Ferroelectr. Freq. Control 2002, 49, 29–38. 20. Lanza, G. M.; Wallace, K. D.; Fischer, S. E.; Christy, D. H.; Scott, M. J.; Trousil, R. L.; Cacheris, W. P.; Miller, J. G.; Gaffney, P. J.; Wickline, S. A. High-frequency ultrasonic detection of thrombi with a targeted contrast system. Ultrasound. Med. Biol. 1997, 23, 863–870. 21. Lanza, G. M.; Abendschein, D. R.; Hall, C. S.; Marsh, J. N.; Scott, M. J.; Scherrer, D. E.; Wickline, S. A. Molecular imaging of stretch-induced tissue factor expression in carotid arteries with intravascular ultrasound. Invest. Radiol. 2000, 35, 227–234. 22. Lanza, G. M.; Abendschein, D. R.; Hall, C. S.; Scott, M. J.; Scherrer, D. E.; Houseman, A.; Miller, J. G.; Wickline, S. A. In vivo molecular imaging of stretch-induced tissue factor in carotid arteries with ligand-targeted nanoparticles. J. Am. Soc. Echocardiogr. 2000, 13, 608–614. 23. Hall, C.; Marsh, J.; Scott, M.; Gaffney, P.; Wickline, S.; Lanza, G. Temperature dependence of ultrasonic enhancement with a site-targeted contrast agent. J. Acous. Soc. Am. 2001, 110, 1677–1684.
P1: OTA/XYZ P2: ABC c30 JWBS037-Chen January 25, 2011
13:43
Printer Name: Yet to Come
REFERENCES
745
24. Hughes, M. S.; Marsh, J. N.; Zhang, H.; Woodson, A. K.; Allen, J. S.; Lacy, E. K.; Carradine, C.; Lanza, G. M.; Wickline, S. A. Characterization of digital waveforms using thermodynamic analogs: detection of contrast-targeted tissue in vivo. IEEE. Trans. Ultrason. Ferroelectr. Freq. Control 2006, 53, 1609–1616. 25. Hughes, M. S.; McCarthy, J. E.; Marsh, J. N.; Arbeit, J. M.; Neumann, R. G.; Fuhrhop, R. W.; Wallace, K. D.; Znidersic, D. R.; Maurizi, B. N.; Baldwin, S. L.; Lanza, G. M.; Wickline, S. A. Properties of an entropy-based signal receiver with an application to ultrasonic molecular imaging. J. Acoust. Soc. Am. 2007, 121, 3542–3557. 26. Stark, D. D.; Weissleder, R.; Elizondo, G.; Hahn, P. F.; Saini, S.; Todd, L. E.; Wittenberg, J.; Ferrucci, J. T. Superparamagnetic iron oxide: clinical application as a contrast agent for MR imaging of the liver. Radiology 1988, 168, 297–301. 27. Weissleder, R.; Hahn, P. F.; Stark, D. D.; Elizondo, G.; Saini, S.; Todd, L. E.; Wittenberg, J.; Ferrucci, J. T. Superparamagnetic iron oxide: enhanced detection of focal splenic tumors with MR imaging. Radiology 1988, 169, 399–403. 28. Frank, H.; Weissleder, R.; Brady, T. J. Enhancement of MR angiography with iron oxide: preliminary studies in whole-blood phantom and in animals. AJR Am. J. Roentgenol. 1994, 162, 209–213. 29. Kresse, M.; Wagner, S.; Pfefferer, D.; Lawaczeck, R.; Elste, V.; Semmler, W. Targeting of ultrasmall superparamagnetic iron oxide (USPIO) particles tumor cells in vivo by using transferrin receptor pathways. Magn. Reson. Med. 1998, 40, 236–242. 30. Cunningham, C.; Arai, T.; Yang, P.; McConnell, M.; Pauly, J.; Conolly, S. Positive contrast magnetic resonance imaging of cells labeled with magnetic nanoparticles. Mag. Reson. Med. 2005, 53, 999–1005. 31. Dharmakumar, R.; Koktzoglou, I.; Li, D. Generating positive contrast from off-resonant spins with steady-state free precession magnetic resonance imaging: theory and proof-of-principle experiments. Phys. Med. Biol. 2006, 51, 4201–4215. 32. Mani, V.; Briley-Saebo, K. C.; Itskovich, V. V.; Samber, D. D.; Fayad, Z. A. Gradient echo acquisition for superparamagnetic particles with positive contrast (GRASP): sequence characterization in membrane and glass superparamagnetic iron oxide phantoms at 1.5 T and 3 T. Magn. Reson. Med. 2006, 55, 126–135. 33. Zurkiya, O.; Hu, X. Off-resonance saturation as a means of generating contrast with superparamagnetic nanoparticles. Magn. Reson. Med. 2006, 56, 726–732. 34. Stuber, M.; Gilson, W. D.; Sch¨ar, M.; Kedziorek, D. A.; Hofmann, L. V.; Shah, S.; Vonken, E. J.; Bulte, J. W. M.; Kraitchman, D. L. Positive contrast visualization of iron oxide-labeled stem cells using inversion-recovery with ON-resonant water suppression (IRON). Magn. Reson. Med. 2007, 58, 1072–1077. 35. Sipkins, D. A.; Cheresh, D. A.; Kazemi, M. R.; Nevin, L. M.; Bednarski, M. D.; Li, K. C. Detection of tumor angiogenesis in vivo by alphaV beta3-targeted magnetic resonance imaging. Nat. Med. 1998, 4, 623–626. 36. Lanza, G.; Lorenz, C.; Fischer, S.; Scott, M.; Cacheris, W.; Kaufman, R.; Gaffney, P.; Wickline, S. Enhanced detection of thrombi with a novel fibrin-targeted magnetic resonance imaging agent. Acad. Radiol. 1998, 5(Suppl 1), s173–s176. 37. Anderson, S. A.; Rader, R. K.; Westlin, W. F.; Null, C.; Jackson, D.; Lanza, G. M.; Wickline, S. A.; Kotyk, J. J. Magnetic resonance contrast enhancement of neovasculature with alpha(v)beta(3)-targeted nanoparticles. Magn. Reson. Med. 2000, 44, 433–439. 38. Yu, X.; Song, S-K.; Chen, J.; Scott, M.; Fuhrhop, R.; Hall, C.; Gaffney, P.; Wickline, S.; Lanza, G. High-resolution MRI characterization of human thrombus using a novel fibrintargeted paramagnetic nanoparticle contrast agent. Magn. Reson. Med. 2000, 44, 867–872. 39. Flacke, S.; Fischer, S.; Scott, M.; Fuhrhop, R.; Allen, J.; McLean, M.; Winter, P.; Sicard, G.; Gaffney, P.; Wickline, S.; Lanza, G. A novel MRI contrast agent for molecular imaging
P1: OTA/XYZ P2: ABC c30 JWBS037-Chen January 25, 2011
746
40.
41.
42.
43.
44.
45.
46.
47.
48. 49. 50.
51. 52.
53.
54.
13:43
Printer Name: Yet to Come
PERFLUOROCARBON NANOPARTICLES
of fibrin: implications for detecting vulnerable plaques. Circulation 2001, 104, 1280– 1285. Winter, P. M.; Caruthers, S. D.; Kassner, A.; Harris, T. D.; Chinen, L. K.; Allen, J. S.; Lacy, E. K.; Zhang, H.; Robertson, J. D.; Wickline, S. A.; Lanza, G. M. Molecular imaging of angiogenesis in nascent Vx2 rabbit tumors using a novel alpha(nu)beta3-targeted nanoparticle and 1.5 tesla magnetic resonance imaging. Cancer. Res. 2003, 63, 5838–5843. Winter, P. M.; Morawski, A. M.; Caruthers, S. D.; Fuhrhop, R. W.; Zhang, H.; Williams, T. A.; Allen, J. S.; Lacy, E. K.; Robertson, J. D.; Lanza, G. M.; Wickline, S. A. Molecular imaging of angiogenesis in early-stage atherosclerosis with alpha(v)beta3-integrin-targeted nanoparticles. Circulation 2003, 108, 2270–2274. Schmieder, A.; Winter, P.; Caruthers, S.; Harris, T.; Williams, T.; Allen, J.; Lacy, E.; Zhang, H.; Scott, M. J.; Wickline, S.; Lanza, G. MR molecular imaging of melanoma angiogenesis with avb3-Targeted paramagnetic nanoparticles. Magn. Reson. Med. 2005, 53, 621–627. Winter, P. M.; Caruthers, S. D.; Yu, X.; Song, S. K.; Chen, J.; Miller, B.; Bulte, J. W.; Robertson, J. D.; Gaffney, P. J.; Wickline, S. A.; Lanza, G. M. Improved molecular imaging contrast agent for detection of human thrombus. Magn. Reson. Med. 2003, 50, 411–416. Morawski, A. M.; Winter, P. M.; Crowder, K. C.; Caruthers, S. D.; Fuhrhop, R. W.; Scott, M. J.; Robertson, J. D.; Abendschein, D. R.; Lanza, G. M.; Wickline, S. A. Targeted nanoparticles for quantitative imaging of sparse molecular epitopes with MRI. Magn. Reson. Med. 2004, 51, 480–486. Morawski, A. M.; Winter, P. M.; Yu, X.; Fuhrhop, R. W.; Scott, M. J.; Hockett, F.; Robertson, J. D.; Gaffney, P. J.; Lanza, G. M.; Wickline, S. A. Quantitative “magnetic resonance immunohistochemistry” with ligand-targeted 19 F nanoparticles. Magn. Reson. Med. 2004, 52, 1255–1262. Caruthers, S. D.; Neubauer, A. M.; Hockett, F. D.; Lamerichs, R.; Winter, P. M.; Scott, M. J.; Gaffney, P. J.; Wickline, S. A.; Lanza, G. M. In vitro demonstration using 19 F magnetic resonance to augment molecular imaging with paramagnetic perfluorocarbon nanoparticles at 1.5 tesla. Invest. Radiol. 2006, 41, 305–312. Shimizu, M.; Kobayashi, T.; Morimoto, H.; Matsuura, N.; Shimano, T.; Nomura, N.; Itoh, S.; Yamazaki, M.; Iriguchi, N.; Yamamoto, T.; et al. Tumor imaging with anti-CEA antibody labeled 19 F emulsion. Magn. Reson. Med. 1987, 5, 290–295. Busse, L. J.; Pratt, R. G.; Thomas, S. R. Deconvolution of chemical shift spectra in two- or three-dimensional 19 F MR imaging. J. Comput. Assist. Tomogr. 1988, 12, 824–835. Bottomley, P. A. Human in vivo NMR spectroscopy in diagnostic medicine: clinical tool or research probe? Radiology 1989, 170, 1–15. Mason, R. P.; Antich, P. P.; Babcock, E. E.; Gerberich, J. L.; Nunnally, R. L. Perfluorocarbon imaging in vivo: a 19 F MRI study in tumor-bearing mice. Magn. Reson. Imaging. 1989, 7, 475–485. Lee, H. K.; Nalcioglu, O.; Buxton, R. B. Correction of chemical-shift artifacts in 19 F imaging of PFOB: a robust signed magnitude method. Magn. Reson. Med. 1992, 23, 254–263. Meyer, K. L.; Carvlin, M. J.; Mukherji, B.; Sloviter, H. A.; Joseph, P. M. Fluorinated blood substitute retention in the rat measured by fluorine-19 magnetic resonance imaging. Invest. Radiol. 1992, 27, 620–627. Mattrey, R. F.; Trambert, M. A.; Brown, J. J.; Young, S. W.; Bruneton, J. N.; Wesbey, G. E.; Balsara, Z. N. Perflubron as an oral contrast agent for MR imaging: results of a phase III clinical trial. Radiology 1994, 191, 841–848. Flogel, U.; Ding, Z.; Hardung, H.; Jander, S.; Reichmann, G.; Jacoby, C.; Schubert, R.; Schrader, J. In vivo monitoring of inflammation after cardiac and cerebral ischemia by fluorine magnetic resonance imaging. Circulation 2008, 118, 140–148.
P1: OTA/XYZ P2: ABC c30 JWBS037-Chen January 25, 2011
13:43
Printer Name: Yet to Come
REFERENCES
747
55. Bulte, J. W. Hot spot MRI emerges from the background. Nat. Biotechnol. 2005, 23, 945–946. 56. Ahrens, E. T.; Flores, R.; Xu, H.; Morel, P. A. In vivo imaging platform for tracking immunotherapeutic cells. Nat. Biotechnol. 2005, 23, 983–987. 57. Partlow, K. C.; Chen, J.; Brant, J. A.; Neubauer, A. M.; Meyerrose, T. E.; Creer, M. H.; Nolta, J. A.; Caruthers, S. D.; Lanza, G. M.; Wickline, S. A. 19 F magnetic resonance imaging for stem/progenitor cell tracking with multiple unique perfluorocarbon nanobeacons. FASEB. J. 2007, 21, 1647–1654. 58. Ruiz-Cabello, J.; Walczak, P.; Kedziorek, D.; Chacko, V.; Schmieder, A.; Wickline, S.; Lanza, G.; Bulte, J. In vivo “Hot Spot” MR imaging of neural stem cells using fluorinated nanoparticles. Magn. Reson. Med. 2008, 60, 1506–1511. 59. Keupp, J.; Mazurkewitz, P. Simultaneous 19 F/1 H imaging for quantification: calibration and sensitivity assessment. Paper presented at Proceedings of the International Society of Magnetic Resonance Medicine 2007. 60. Keupp, J.; Mazurkewitz, P.; Frasslin, X.; Schaeffter, T. Simultaneous 19 F and 1 H imaging on a clinical 3 T MR scanner. Paper presented at Proceedings of the International Society of Magnetic Resonance Medicine 2006. 61. Keupp, J.; Caruthers, S.; Rahmer, J.; Williams, T.; Wickline, S.; Lanza, G. Fluorine-19 MR molecular imaging of angiogenesis on Vx2 tumors in rabbits using ␣3-targeted nanoparticles. Paper presented at Proceedings of the International Society of Magnetic Resonance Medicine 2009. 62. Rahmer, J.; Keupp, J.; Caruthers, S.; Lips, O.; Williams, T.; Wickline, S.; Lanza, G. Dual resolution simultaneous 19 F/1 H in vivo imaging of targeted nanoparticles. Paper presented at Proceedings of the International Society of Magnetic Resonance Medicine 2009. 63. Rahmer, J.; Keupp, J.; Caruthers, S.; Lips, O.; Williams, T.; Wickline, S.; Lanza, G. 19 F/1 H simultaneous 3D radial imaging of atherosclerotic rabbits using self-navigated respiratory motion compensation. Paper presented at Proceedings of the International Society of Magnetic Resonance Medicine 2009. 64. Cheresh, D. A. Integrins in thrombosis, wound healing and cancer. Biochem. Soc. Trans. 1991, 19, 835–838. 65. Friedlander, M.; Theesfeld, C. L.; Sugita, M.; Fruttiger, M.; Thomas, M. A.; Chang, S.; Cheresh, D. A. Involvement of integrins alpha v beta 3 and alpha v beta 5 in ocular neovascular diseases. Proc. Natl. Acad. Sci. U.S.A. 1996, 93, 9764–9769. 66. De Nichilo, M.; Burns, G. Granulocyte-macrophage and macrophage colony-stimulating factors differentially regulate alpha v integrin expression on cultured human macrophages Proc. Natl. Acad. Sci. U.S.A. 1993, 90, 2517–2521. 67. Helluin, O.; Chan, C.; Vilaire, G.; Mousa, S.; DeGrado, W. F.; Bennett, J. S. The activation state of alpha v beta 3 regulates platelet and lymphocyte adhesion to intact and thrombin-cleaved osteopontin. J. Biol. Chem. 2000, 275, 18337–18343. 68. Itoh, H.; Nelson, P.; Mureebe, L.; Horowitz, A.; Kent, K. The role of integrins in saphenous vein vascular smooth muscle cell migration. J. Vasc. Surg. 1997, 25, 1061–1069. 69. Carreiras, F.; Denoux, Y.; Staedel, C.; Lehmann, M.; Sichel, F.; Gauduchon, P. Expression and localization of alpha v integrins and their ligand vitronectin in normal ovarian epithelium and in ovarian carcinoma. Gynecol. Oncol. 1996, 62, 260–267. 70. Kageshita, T.; Hamby, C. V.; Hirai, S.; Kimura, T.; Ono, T.; Ferrone, S. Differential clinical significance of alpha(v)beta(3) expression in primary lesions of acral lentiginous melanoma and of other melanoma histotypes. Int. J. Cancer 2000, 89, 153–159. 71. Hu, G.; Lijowski, M.; Zhang, H.; Partlow, K. C.; Caruthers, S. D.; Kiefer, G.; Gulyas, G.; Athey, P.; Scott, M. J.; Wickline, S. A.; Lanza, G. M. Imaging of Vx2 rabbit tumors with alpha(nu)beta3-integrin-targeted 111 In nanoparticles. Int. J. Cancer 2007, 120, 1951–1957.
P1: OTA/XYZ P2: ABC c30 JWBS037-Chen January 25, 2011
748
13:43
Printer Name: Yet to Come
PERFLUOROCARBON NANOPARTICLES
72. Weissleder, R.; Bogdanov, A., Jr.; Tung, C. H.; Weinmann, H. J. Size optimization of synthetic graft copolymers for in vivo angiogenesis imaging. Bioconjug. Chem. 2001, 12, 213–219. 73. Schmieder, A. H.; Winter, P. M.; Caruthers, S. D.; Harris, T. D.; Williams, T. A.; Allen, J. S.; Lacy, E. K.; Zhang, H.; Scott, M. J.; Hu, G.; Robertson, J. D.; Wickline, S. A.; Lanza, G. M. Molecular MR imaging of melanoma angiogenesis with alpha (v) beta (3)-targeted paramagnetic nanoparticles. Magn. Reson. Med. 2005, 53, 621–627. 74. Zetter, B. R. Angiogenesis and tumor metastasis. Annu. Rev. Med. 1998, 49, 407–424. 75. Folkman, J. Tumor angiogenesis. Adv. Cancer. Res. 1985, 43, 175–203. 76. Carmeliet, P. Mechanisms of angiogenesis and arteriogenesis. Nat. Med. 2000, 6, 389–395. 77. Izzedine, H.; Rixe, O.; Billemont, B.; Baumelou, A.; Deray, G. Angiogenesis inhibitor therapies: focus on kidney toxicity and hypertension. Am. J. Kidney. Dis. 2007, 50, 203–218. 78. Giles, F. J.; Bellamy, W. T.; Estrov, Z.; O’Brien, S. M.; Verstovsek, S.; Ravandi, F.; Beran, M.; Bycott, P.; Pithavala, Y.; Steinfeldt, H.; Reich, S. D.; List, A. F.; Yee, K. W. The antiangiogenesis agent, AG-013736, has minimal activity in elderly patients with poor prognosis acute myeloid leukemia (AML) or myelodysplastic syndrome (MDS). Leuk. Res. 2006, 30, 801–811. 79. Zakarija, A.; Soff, G. Update on angiogenesis inhibitors. Curr. Opin. Oncol. 2005, 17, 578– 583. 80. Eremina, V.; Jefferson, J. A.; Kowalewska, J.; Hochster, H.; Haas, M.; Weisstuch, J.; Richardson, C.; Kopp, J. B.; Kabir, M. G.; Backx, P. H.; Gerber, H. P.; Ferrara, N.; Barisoni, L.; Alpers, C. E.; Quaggin, S. E. VEGF inhibition and renal thrombotic microangiopathy. N. Engl. J. Med. 2008, 358, 1129–1136. 81. Lanza, G. M.; Yu, X.; Winter, P. M.; Abendschein, D. R.; Karukstis, K. K.; Scott, M. J.; Chinen, L. K.; Fuhrhop, R. W.; Scherrer, D. E.; Wickline, S. A. Targeted antiproliferative drug delivery to vascular smooth muscle cells with a magnetic resonance imaging nanoparticle contrast agent: implications for rational therapy of restenosis. Circulation 2002, 106, 2842–2847. 82. Winter, P.; Neubauer, A.; Caruthers, S.; Harris, T.; Robertson, J.; Williams, T.; Schmieder, A.; Hu, G.; Allen, J.; Lacy, E.; Wickline, S.; Lanza, G. Endothelial alpha(nu)beta(3)-integrin targeted fumagillin nanoparticles inhibit angiogenesis in atherosclerosis. Arterioscler. Thromb. Vasc. Biol. 2006, 26, 2103–2109. 83. Marsh, J.; Senpan, A.; Hu, G.; Scott, M.; Gaffney, P.; Wickline, S.; Lanza, G. Fibrin-targeted perfluorocarbon nanoparticles for targeted thrombolysis. Nanomedicine 2007, 2, 533–543. 84. Cyrus, T.; Zhang, H.; Allen, J. S.; Williams, T. A.; Hu, G.; Caruthers, S. D.; Wickline, S. A.; Lanza, G. M. Intramural delivery of rapamycin with alpha v beta3-targeted paramagnetic nanoparticles inhibits stenosis after balloon injury. Arterioscler. Thromb. Vasc. Biol. 2008, 28, 820–826. 85. Schmieder, A. H.; Caruthers, S. D.; Zhang, H.; Williams, T. A.; Robertson, J. D.; Wickline, S. A.; Lanza, G. M. Three-dimensional MR mapping of angiogenesis with {alpha}5{beta}1({alpha}{nu}{beta}3)-targeted theranostic nanoparticles in the MDA-MB435 xenograft mouse model. FASEB. J. 2008, 22, 4179–4189. 86. Winter, P.; Caruthers, S.; Zhang, H.; Williams, T.; Wickline, S.; Lanza, G. Antiangiogenic synergism of integrin-targeted fumagillin nanoparticles and atorvastatin in atherosclerosis. J. Am. Coll. Cardiol. Img. 2008, 1, 624–634. 87. Winter, P. M.; Schmieder, A. H.; Caruthers, S. D.; Keene, J. L.; Zhang, H.; Wickline, S. A.; Lanza, G. M. Minute dosages of alpha(nu)beta3-targeted fumagillin nanoparticles impair Vx2 tumor angiogenesis and development in rabbits. FASEB. J. 2008, 22, 2758–2767. 88. Soman, N.; Lanza, G.; Heuser, J.; Schlesinger, P.; Wickline, S. Synthesis and characterization of stable fluorocarbon nanostructures as drug delivery vehicles for cytolytic peptides. Nano. Lett. 2008, 8, 1131–1136.
P1: OTA/XYZ P2: ABC c30 JWBS037-Chen January 25, 2011
13:43
Printer Name: Yet to Come
REFERENCES
749
89. Partlow, K.; Lanza, G.; Wickline, S. Exploiting lipid raft transport with membrane targeted nanoparticles: a strategy for cytosolic drug delivery. Biomaterials 2008, 29, 3367–3375. 90. Liu, S.; Widom, J.; Kemp, C. W.; Crews, C. M.; Clardy, J. Structure of human methionine aminopeptidase-2 complexed with fumagillin. Science 1998, 282, 1324–1327. 91. Sin, N.; Meng, L.; Wang, M. Q.; Wen, J. J.; Bornmann, W. G.; Crews, C. M. The anti-angiogenic agent fumagillin covalently binds and inhibits the methionine aminopeptidase, MetAP-2. Proc. Natl. Acad. Sci. U.S.A. 1997, 94, 6099–6103. 92. Arfin, S. M.; Kendall, R. L.; Hall, L.; Weaver, L. H.; Stewart, A. E.; Matthews, B. W.; Bradshaw, R. A. Eukaryotic methionyl aminopeptidases: two classes of cobalt-dependent enzymes. Proc. Natl. Acad. Sci. U.S.A. 1995, 92, 7714–7718. 93. Wang, J.; Lou, P.; Henkin, J. Selective inhibition of endothelial cell proliferation by fumagillin is not due to differential expression of methionine aminopeptidases. J. Cell. Biochem. 2000, 77, 465–473. 94. Lowther, W. T.; Matthews, B. W. Structure and function of the methionine aminopeptidases. Biochim. Biophys. Acta 2000, 1477, 157–167. 95. Turk, B. E.; Griffith, E. C.; Wolf, S.; Biemann, K.; Chang, Y. H.; Liu, J. O. Selective inhibition of amino-terminal methionine processing by TNP-470 and ovalicin in endothelial cells. Chem. Biol. 1999, 6, 823–833. 96. Griffith, E. C.; Su, Z.; Niwayama, S.; Ramsay, C. A.; Chang, Y. H.; Liu, J. O. Molecular recognition of angiogenesis inhibitors fumagillin and ovalicin by methionine aminopeptidase 2. Proc. Natl. Acad. Sci. U.S.A. 1998, 95, 15183–15188. 97. Bergers, G.; Javaherian, K.; Lo, K. M.; Folkman, J.; Hanahan, D. Effects of angiogenesis inhibitors on multistage carcinogenesis in mice. Science 1999, 284, 808–812. 98. Castronovo, V.; Belotti, D. TNP-470 (AGM-1470): mechanisms of action and early clinical development. Eur. J. Cancer. 1996, 32A, 2520–2527. 99. Konno, H.; Tanaka, T.; Kanai, T.; Maruyama, K.; Nakamura, S.; Baba, S. Efficacy of an angiogenesis inhibitor, TNP-470, in xenotransplanted human colorectal cancer with high metastatic potential. Cancer 1996, 77, 1736–1740. 100. Shusterman, S.; Grupp, S. A.; Barr, R.; Carpentieri, D.; Zhao, H.; Maris, J. M. The angiogenesis inhibitor TNP-470 effectively inhibits human neuroblastoma xenograft growth, especially in the setting of subclinical disease. Clin. Cancer. Res. 2001, 7, 977–984. 101. Bhargava, P.; Marshall, J. L.; Rizvi, N.; Dahut, W.; Yoe, J.; Figuera, M.; Phipps, K.; Ong, V. S.; Kato, A.; Hawkins, M. J. A Phase I and pharmacokinetic study of TNP-470 administered weekly to patients with advanced cancer. Clin. Cancer. Res. 1999, 5, 1989–1995. 102. Kudelka, A. P.; Verschraegen, C. F.; Loyer, E. Complete remission of metastatic cervical cancer with the angiogenesis inhibitor TNP-470. N. Engl. J. Med. 1998, 338, 991–992. 103. Kudelka, A. P.; Levy, T.; Verschraegen, C. F.; Edwards, C. L.; Piamsomboon, S.; Termrungruanglert, W.; Freedman, R. S.; Kaplan, A. L.; Kieback, D. G.; Meyers, C. A.; Jaeckle, K. A.; Loyer, E.; Steger, M.; Mante, R.; Mavligit, G.; Killian, A.; Tang, R. A.; Gutterman, J. U.; Kavanagh, J. J. A Phase I study of TNP-470 administered to patients with advanced squamous cell cancer of the cervix. Clin. Cancer. Res. 1997, 3, 1501–1505. 104. Logothetis, C. J.; Wu, K. K.; Finn, L. D.; Daliani, D.; Figg, W.; Ghaddar, H.; Gutterman, J. U. Phase I trial of the angiogenesis inhibitor TNP-470 for progressive androgen-independent prostate cancer. Clin. Cancer. Res. 2001, 7, 1198–1203. 105. Offodile, R.; Walton, T.; Lee, M.; Stiles, A.; Nguyen, M. Regression of metastatic breast cancer in a patient treated with the anti-angiogenic drug TNP-470. Tumori 1999, 85, 51–53. 106. Kim, S.; Bell, K.; Mousa, S.; Varner, J. Regulation of angiogenesis in vivo by ligation of integrin alpha5beta1 with the central cell-binding domain of fibronectin. Am. J. Pathol. 2000, 156, 1345–1362.
P1: OTA/XYZ P2: ABC c30 JWBS037-Chen January 25, 2011
750
13:43
Printer Name: Yet to Come
PERFLUOROCARBON NANOPARTICLES
107. Boudreau, N.; Varner, J. The homeobox transcription factor Hox D3 promotes integrin alpha5beta1 expression and function during angiogenesis. J. Biol. Chem. 2004, 279, 4862– 4868. 108. Bayless, K.; Salazar, R.; Davis, G. RGD-dependent vacuolation and lumen formation observed during endothelial cell morphogenesis in three-dimensional fibrin matrices involves the alpha(v)beta(3) and alpha(5)beta(1) integrins. Am. J. Pathol. 2000, 156, 1673–1683. 109. Giuliani, A.; Pirri, G.; Nicoletto, S. Antimicrobial peptides: an overview of a promising class of therapeutics. Central. Eur. J. Biol. 2007, 2, 1–33. 110. Papo, N.; Shai, Y. Host defense peptides as new weapons in cancer treatment. Cell. Mol.. Life. Sci. 2005, 62, 784–790. 111. Hancock, R. E.; Sahl, H. G. Antimicrobial and host-defense peptides as new anti-infective therapeutic strategies. Nat. Biotechnol. 2006, 24, 1551–1557. 112. Hoskin, D. W.; Ramamoorthy, A. Studies on anticancer activities of antimicrobial peptides. Biochim. Biophys. Acta 2008, 1778, 357–375. 113. Zhang, L.; Falla, T. J. Antimicrobial peptides: therapeutic potential. Expert. Opin. Pharmacother. 2006, 7, 653–663. 114. Tosteson, M. T.; Tosteson, D. C. The sting. Melittin forms channels in lipid bilayers. Biophys. J. 1981, 36, 109–116. 115. Leuschner, C.; Hansel, W. Membrane disrupting lytic peptides for cancer treatments. Curr. Pharm. Des. 2004, 10, 2299–2310. 116. Hansel, W.; Leuschner, C.; Enright, F. Conjugates of lytic peptides and LHRH or betaCG target and cause necrosis of prostate cancers and metastases. Mol. Cell. Endocrinol. 2007, 269, 26–33. 117. Hansel, W.; Enright, F.; Leuschner, C. Destruction of breast cancers and their metastases by lytic peptide conjugates in vitro and in vivo. Mol. Cell. Endocrinol. 2007, 260–262, 183–189. 118. Kumar, C. S.; Leuschner, C.; Doomes, E. E.; Henry, L.; Juban, M.; Hormes, J. Efficacy of lytic peptide-bound magnetite nanoparticles in destroying breast cancer cells. J. Nanosci. Nanotechnol. 2004, 4, 245–249. 119. Soman, N.; Lanza, G.; Heuser, J.; Schlesinger, P.; Wickline, S. Synthesis and characterization of stable fluorocarbon nanostructures as drug delivery vehicles for cytolytic peptides. Nano. Lett. 2008, 8, 1131–1136. 120. Soman, N.; Baldwin, S.; Hu, G.; Marsh, J.; Lanza, G.; Heuser, J.; Arbeit, J.; Wickline, S.; Schlesinger, P. Molecularly targeted nanocarriers deliver cytolytic peptide mellitin specifically to tumor cells in mice, reducing tumor growth. J. Clin. Invest. 2009, 119, 2830–2842. 121. Moreno, P. R.; Purushothaman, K. R.; Fuster, V.; Echeverri, D.; Truszczynska, H.; Sharma, S. K.; Badimon, J. J.; O’Connor, W. N. Plaque neovascularization is increased in ruptured atherosclerotic lesions of human aorta: implications for plaque vulnerability. Circulation 2004, 110, 2032–2038. 122. Moreno, P. R.; Purushothaman, K. R.; Sirol, M.; Levy, A. P.; Fuster, V. Neovascularization in human atherosclerosis. Circulation 2006, 113, 2245–2252. 123. Virmani, R.; Kolodgie, F. D.; Burke, A. P.; Finn, A. V.; Gold, H. K.; Tulenko, T. N.; Wrenn, S. P.; Narula, J. Atherosclerotic plaque progression and vulnerability to rupture: angiogenesis as a source of intraplaque hemorrhage. Arterioscler. Thromb. Vasc. Biol. 2005, 25, 2054–2061. 124. Bjornheden, T.; Levin, M.; Evaldsson, M.; Wiklund, O. Evidence of hypoxic areas within the arterial wall in vivo. Arterioscler. Thromb. Vasc. Biol. 1999, 19, 870–876. 125. Boyle, J. J.; Wilson, B.; Bicknell, R.; Harrower, S.; Weissberg, P. L.; Fan, T. P. Expression of angiogenic factor thymidine phosphorylase and angiogenesis in human atherosclerosis. J. Pathol. 2000, 192, 234–242.
P1: OTA/XYZ P2: ABC c30 JWBS037-Chen January 25, 2011
13:43
Printer Name: Yet to Come
REFERENCES
751
126. Khatri, J. J.; Johnson, C.; Magid, R.; Lessner, S. M.; Laude, K. M.; Dikalov, S. I.; Harrison, D. G.; Sung, H. J.; Rong, Y.; Galis, Z. S. Vascular oxidant stress enhances progression and angiogenesis of experimental atheroma. Circulation 2004, 109, 520–525. 127. de Boer, O. J.; van der Wal, A. C.; Teeling, P.; Becker, A. E. Leucocyte recruitment in rupture prone regions of liid-rich plaques: a prominent role for neovascularization? Cardiovasc. Res. 1999, 41, 443–449. 128. Finn, A. V.; Kramer, M. C.; Vorpahl, M.; Kolodgie, F. D.; Virmani, R. Pharmacotherapy of coronary atherosclerosis. Expert. Opin. Pharmacother. 2009, 10, 1587–1603. 129. Pasternak, R. C.; Smith, S. C., Jr.; Bairey-Merz, C. N.; Grundy, S. M.; Cleeman, J. I.; Lenfant, C. ACC/AHA/NHLBI Clinical Advisory on the Use and Safety of Statins. Circulation 2002, 106, 1024–1028. 130. Girona, J.; La Ville, A. E.; Sola, R.; Plana, N.; Masana, L. Simvastatin decreases aldehyde production derived from lipoprotein oxidation. Am. J. Cardiol. 1999, 83, 846–851. 131. Kimura, M.; Kurose, I.; Russell, J.; Granger, D. N. Effects of fluvastatin on leukocyteendothelial cell adhesion in hypercholesterolemic rats. Arterioscler. Thromb. Vasc. Biol. 1997, 17, 1521–1526. 132. Verhoeven, B. A.; Moll, F. L.; Koekkoek, J. A.; van der Wal, A. C.; de Kleijn, D. P.; de Vries, J. P.; Verheijen, J. H.; Velema, E.; Busser, E.; Schoneveld, A.; Virmani, R.; Pasterkamp, G. Statin treatment is not associated with consistent alterations in inflammatory status of carotid atherosclerotic plaques: a retrospective study in 378 patients undergoing carotid endarterectomy. Stroke 2006, 37, 2054–2060. 133. Laufs, U.; La Fata, V.; Plutzky, J.; Liao, J. K. Upregulation of endothelial nitric oxide synthase by HMG CoA reductase inhibitors. Circulation 1998, 97, 1129–1135. 134. Sukhova, G. K.; Williams, J. K.; Libby, P. Statins reduce inflammation in atheroma of nonhuman primates independent of effects on serum cholesterol. Arterioscler. Thromb. Vasc. Biol. 2002, 22, 1452–1458. 135. Koutouzis, M.; Nomikos, A.; Nikolidakis, S.; Tzavara, V.; Andrikopoulos, V.; Nikolaou, N.; Barbatis, C.; Kyriakides, Z. S. Statin treated patients have reduced intraplaque angiogenesis in carotid endarterectomy specimens. Atherosclerosis 2007, 192, 457–463. 136. Kolodgie, F. D.; Narula, J.; Yuan, C.; Burke, A. P.; Finn, A. V.; Virmani, R. Elimination of neoangiogenesis for plaque stabilization: Is there a role for local drug therapy? J. Am. Coll. Cardiol. 2007, 49, 2093–2101. 137. Jain, R. K.; Finn, A. V.; Kolodgie, F. D.; Gold, H. K.; Virmani, R. Antiangiogenic therapy for normalization of atherosclerotic plaque vasculature: a potential strategy for plaque stabilization. Nat. Clin. Pract. Cardiovasc. Med. 2007, 4, 491–502. 138. Jain, R. Normalizing tumor vasculature with anti-angiogenic therapy: a new paradigm for combination therapy. Nat. Med. 2001, 7, 987–989. 139. Duda, D. G.; Jain, R. K.; Willett, C. G. Antiangiogenics: the potential role of integrating this novel treatment modality with chemoradiation for solid cancers. J. Clin. Oncol. 2007, 25, 4033–4042. 140. Doyle, B.; Caplice, N. Plaque neovascularization and antiangiogenic therapy for atherosclerosis. J. Am. Coll. Cardiol. 2007, 49, 2073–2080. 141. Khurana, R.; Simons, M.; Martin, J. F.; Zachary, I. C. Role of angiogenesis in cardiovascular disease: a critical appraisal. Circulation 2005, 112, 1813–1824. 142. Moreno, P. R.; Sanz, J.; Fuster, V. Promoting mechanisms of vascular health: circulating progenitor cells, angiogenesis, and reverse cholesterol transport. J. Am. Coll. Cardiol. 2009, 53, 2315–2323. 143. Takaya, N.; Yuan, C.; Chu, B.; Saam, T.; Polissar, N. L.; Jarvik, G. P.; Isaac, C.; McDonough, J.; Natiello, C.; Small, R.; Ferguson, M. S.; Hatsukami, T. S. Presence of intraplaque hemorrhage
P1: OTA/XYZ P2: ABC c30 JWBS037-Chen January 25, 2011
752
144.
145.
146.
147.
148.
149.
150.
151.
152.
153.
154.
155.
156.
13:43
Printer Name: Yet to Come
PERFLUOROCARBON NANOPARTICLES
stimulates progression of carotid atherosclerotic plaques: a high-resolution magnetic resonance imaging study. Circulation 2005, 111, 2768–2775. Moses, J. W.; Leon, M. B.; Popma, J. J.; Fitzgerald, P. J.; Holmes, D. R.; O’Shaughnessy, C.; Caputo, R. P.; Kereiakes, D. J.; Williams, D. O.; Teirstein, P. S.; Jaeger, J. L.; Kuntz, R. E. Sirolimus-eluting stents versus standard stents in patients with stenosis in a native coronary artery. N. Engl. J. Med. 2003, 349, 1315–1323. Stone, G. W.; Ellis, S. G.; Cox, D. A.; Hermiller, J.; O’Shaughnessy, C.; Mann, J. T.; Turco, M.; Caputo, R.; Bergin, P.; Greenberg, J.; Popma, J. J.; Russell, M. E. A polymer-based, paclitaxel-eluting stent in patients with coronary artery disease. N. Engl. J. Med. 2004, 350, 221–231. Goy, J. J.; Stauffer, J. C.; Siegenthaler, M.; Benoit, A.; Seydoux, C. A prospective randomized comparison between paclitaxel and sirolimus stents in the real world of interventional cardiology: the TAXi trial. J. Am. Coll. Cardiol. 2005, 45, 308–311. Holmes, D. R., Jr.; Leon, M. B.; Moses, J. W.; Popma, J. J.; Cutlip, D.; Fitzgerald, P. J.; Brown, C.; Fischell, T.; Wong, S. C.; Midei, M.; Snead, D.; Kuntz, R. E. Analysis of 1-year clinical outcomes in the SIRIUS trial: a randomized trial of a sirolimus-eluting stent versus a standard stent in patients at high risk for coronary restenosis. Circulation 2004, 109, 634–640. Sousa, J. E.; Costa, M. A.; Sousa, A. G.; Abizaid, A. C.; Seixas, A. C.; Abizaid, A. S.; Feres, F.; Mattos, L. A.; Falotico, R.; Jaeger, J.; Popma, J. J.; Serruys, P. W. Two-year angiographic and intravascular ultrasound follow-up after implantation of sirolimus-eluting stents in human coronary arteries. Circulation 2003, 107, 381–383. Kelbaek, H.; Thuesen, L.; Helqvist, S.; Klovgaard, L.; Jorgensen, E.; Aljabbari, S.; Saunamaki, K.; Krusell, L. R.; Jensen, G. V.; Botker, H. E.; Lassen, J. F.; Andersen, H. R.; Thayssen, P.; Galloe, A.; van Weert, A. The stenting coronary arteries in non-stress/benestent disease (SCANDSTENT) trial. J. Am. Coll. Cardiol. 2006, 47, 449–455. Windecker, S.; Remondino, A.; Eberli, F. R.; Juni, P.; Raber, L.; Wenaweser, P.; Togni, M.; Billinger, M.; Tuller, D.; Seiler, C.; Roffi, M.; Corti, R.; Sutsch, G.; Maier, W.; Luscher, T.; Hess, O. M.; Egger, M.; Meier, B. Sirolimus-eluting and paclitaxel-eluting stents for coronary revascularization. N. Engl. J. Med. 2005, 353, 653–662. Escolar, E.; Mintz, G. S.; Hong, M. K.; Lee, C. W.; Kim, J. J.; Fearnot, N. E.; Park, S. W.; Park, S. J.; Weissman, N. J. Relation of intimal hyperplasia thickness to stent size in paclitaxel-coated stents. Am. J. Cardiol. 2004, 94, 196–198. Kuchulakanti, P. K.; Chu, W. W.; Torguson, R.; Ohlmann, P.; Rha, S. W.; Clavijo, L. C.; Kim, S. W.; Bui, A.; Gevorkian, N.; Xue, Z.; Smith, K.; Fournadjieva, J.; Suddath, W. O.; Satler, L. F.; Pichard, A. D.; Kent, K. M.; Waksman, R. Correlates and long-term outcomes of angiographically proven stent thrombosis with sirolimus- and paclitaxel-eluting stents. Circulation 2006, 113, 1108–1113. Iakovou, I.; Schmidt, T.; Bonizzoni, E.; Ge, L.; Sangiorgi, G. M.; Stankovic, G.; Airoldi, F.; Chieffo, A.; Montorfano, M.; Carlino, M.; Michev, I.; Corvaja, N.; Briguori, C.; Gerckens, U.; Grube, E.; Colombo, A. Incidence, predictors, and outcome of thrombosis after successful implantation of drug-eluting stents. JAMA 2005, 293, 2126–2130. Wenaweser, P.; Dorffler-Melly, J.; Imboden, K.; Windecker, S.; Togni, M.; Meier, B.; Haeberli, A.; Hess, O. M. Stent thrombosis is associated with an impaired response to antiplatelet therapy. J. Am. Coll. Cardiol. 2005, 45, 1748–1752. Lev, E. I.; Patel, R. T.; Maresh, K. J.; Guthikonda, S.; Granada, J.; DeLao, T.; Bray, P. F.; Kleiman, N. S. Aspirin and clopidogrel drug response in patients undergoing percutaneous coronary intervention: the role of dual drug resistance. J. Am. Coll. Cardiol. 2006, 47, 27–33. McFadden, E. P.; Stabile, E.; Regar, E.; Cheneau, E.; Ong, A. T.; Kinnaird, T.; Suddath, W. O.; Weissman, N. J.; Torguson, R.; Kent, K. M.; Pichard, A. D.; Satler, L. F.; Waksman, R.; Serruys,
P1: OTA/XYZ P2: ABC c30 JWBS037-Chen January 25, 2011
13:43
Printer Name: Yet to Come
REFERENCES
157.
158.
159.
160.
161.
753
P. W. Late thrombosis in drug-eluting coronary stents after discontinuation of antiplatelet therapy. Lancet 2004, 364, 1519–1521. Rodriguez, A.; Mieres, J.; Fernandez-Pereira, C.; Vigo, C.; Rodriguez-Alemparte, M.; Berrocal, D.; Grinfeld, L.; Palacios, I. Coronary stent thrombosis in the current drug-eluting stent era: insights from the ERACI III trial. J. Am. Coll. Cardiol. 2006, 47, 205–207. Nakazawa, G.; Vorpahl, M.; Finn, A. V.; Narula, J.; Virmani, R. One step forward and two steps back with drug-eluting-stents: from preventing restenosis to causing late thrombosis and nouveau atherosclerosis. JACC. Cardiovasc. Imaging 2009, 2, 625–628. Higo, T.; Ueda, Y.; Oyabu, J.; Okada, K.; Nishio, M.; Hirata, A.; Kashiwase, K.; Ogasawara, N.; Hirotani, S.; Kodama, K. Atherosclerotic and thrombogenic neointima formed over sirolimus drug-eluting stent: an angioscopic study. JACC Cardiovasc. Imaging 2009, 2, 616–624. Cyrus, T.; Abendschein, D. R.; Caruthers, S. D.; Harris, T. D.; Glattauer, V.; Werkmeister, J. A.; Ramshaw, J. A.; Wickline, S. A.; Lanza, G. M. MR three-dimensional molecular imaging of intramural biomarkers with targeted nanoparticles. J. Cardiovasc. Magn. Reson. 2006, 8, 535–541. Joner, M.; Morimoto, K.; Kasukawa, H.; Steigerwald, K.; Merl, S.; Nakazawa, G.; John, M. C.; Finn, A. V.; Acampado, E.; Kolodgie, F. D.; Gold, H. K.; Virmani, R. Site-specific targeting of nanoparticle prednisolone reduces in-stent restenosis in a rabbit model of established atheroma. Arterioscler. Thromb. Vasc. Biol. 2008, 28, 1960–1966.
P1: OTA/XYZ P2: ABC c31 JWBS037-Chen February 2, 2011
21:53
Printer Name: Yet to Come
CHAPTER 31
Radioimmunonanoparticles for Cancer Imaging and Therapy ARUTSELVAN NATARAJAN Department of Radiology and Molecular Imaging Program at Stanford, Stanford University School of Medicine, Stanford, California, USA
Nanoparticles (NPs) designed with less than 200 nm in size are increasingly used in biomedical applications and specifically for cancer diagnosis and therapy. Many reports have indicated various strategies that have been adopted to construct nanoparticles with multifunctional properties to diagnose and treat cancer using a range of targeting agents and imaging modalities. Very recently, another new class of multifunctional NPs has been reported by utilizing radiopharmaceuticals or radionuclides for scintigraphic imaging of cancer inflammation/infection and sentinel lymph node detection. This novel class of molecules has been created by utilizing cancer targeting agents (antibodies, antibody fragments, aptamers, peptides, and small molecules) linked to nanoparticles that are tagged with nuclear imaging agents. Thus these novel NPs could provide the following opportunities: (1) targeting specific cancer cells to deliver drugs, (2) allowing simultaneous detection, imaging, and monitoring of drug delivery, and (3) possible use for targeted cancer therapy. Many reports have shown that radionanoparticles (RNPs) or radioimmunonanoparticles (RINPs) were already utilized in cancer diagnosis, imaging, and noninvasive hyperthermia using cancer targeting mAb and mAb fragments. Although considerable progress has been made to construct multifunctional NPs, several issues need to be addressed (e.g., toxicity, in vivo particle clearance) so as to have nanoparticle-based products for safe administration to human patients. Thus the use of RINPs could be the best strategy to monitor the fate of these NPs through bench to bed translational research without any modification at each stage from in vitro to clinical trial. In order to provide an update we have reviewed more than 25 articles from the recent literature related to passive NPs, active NPs (targeted), immunonanoparticles, RNPs, and RINPs used in cancer for diagnosis, imaging, and therapy. Special emphasis has been given to preclinical and clinical studies. Thus this chapter is organized into various sections based on the various components and applications of nanoplatforms, that is, nanoparticle core, targeting moiety, and imaging modalities.
Nanoplatform-Based Molecular Imaging Edited by Xiaoyuan Chen C 2011 John Wiley & Sons, Inc. Copyright
755
P1: OTA/XYZ P2: ABC c31 JWBS037-Chen February 2, 2011
756
21:53
Printer Name: Yet to Come
RADIOIMMUNONANOPARTICLES FOR CANCER IMAGING AND THERAPY
31.1 INTRODUCTION 31.1.1 Nanomaterials for Cancer Imaging and Therapy The use of nanomaterials has been exploited for several decades in biomedical applications, for example, to deliver and target chemotherapeutic agents and cancer targeting ligands [1–4]. Clinical studies of nanosized particles (e.g., superparamagnetic iron oxide (SPIO), liposome) linked to optical imaging and nuclear imaging agents are utilized in a wide variety of tumors and show detection of tumors with good sensitivity and specificity [5–7]. These particles were clearly demonstrated for detection of various soft tissues and bone inflammatory/infectious lesions, and performed equal to or better than imaging agents that are approved at present [1, 8]. Currently, significant research efforts are being made with various classes of nanoparticles (e.g., SPIO, gold, liposomes) to design and develop multifunctional targeting nanoplatforms for cancer diagnostic, imaging, and therapeutic agents. Although a wide variety of particles were utilized to target cancer (liposomes, polymer and protein nanoparticles, dendrimers, carbon-based nanoparticles such as fullerene), in vivo behavior still needs to be tuned for human patient studies. In this chapter attention is given to nontoxic nanomaterials such as radionanoparticles (RNPs) or radioimmunonanoparticles (RINPs) for detection and treatment of tumor, metastatic lymph nodes, and infectious/inflammatory diseases. 31.1.2 Key Parameters for the Development of Nanoparticle Platform for Imaging and Therapy Applications Nanoplatforms used for cancer diagnosis, imaging, and therapy consist of four basic components: (1) nanosized nontoxic core particle, (2) suitable surface modification with biocompatible materials, (3) cancer targeting moiety, and (4) signal molecules (Fig. 31.1). However, the choice of components may differ based on the specific applications.
FIGURE 31.1 Schematic diagram of radionanoparticles (RNPs) or radioimmunonanoparticles (RINPs) are prepared by linking radiometals via linkers to direct nanoparticles or through specific targeting moieties (e.g., antibody, antibody fragments, peptides, or small molecules). Nanoparticle cores were first coated with biocompatible polymers (dextran or PEG) to evade the reticuloendothelial system (RES) system, followed by conjugation of targeting moieties, and signal molecules to track particles, diagnose and image cancer tissues, and deliver therapeutic agents.
P1: OTA/XYZ P2: ABC c31 JWBS037-Chen February 2, 2011
21:53
Printer Name: Yet to Come
INTRODUCTION
757
Engineering of nanoparticles and surface modifications are important for prolonged circulation time and selective extravasations at pathological sites. For example, experimental and clinical tumor imaging may occur on the basis of neovascularization to discriminate tumor tissue from the surrounding normal tissue [9, 10]. Therefore pharmacokinetics (PK) and pharmacodynamics (PD) of the targeted nanoparticles are very important. This will depend on critical parameters such as particle size, chemistry, surface charge, and coating materials. For example, MRI contrast agents have been made “invisible” to macrophages by surface coating with PEG [11, 12]. In addition, identification of high-affinity ligands to clinically relevant targets, is important to increase targeting efficiency of the nanoparticles. Specific ligands to cell surface receptors, antibodies and antibody fragments, and aptamers can all involve such transport mechanisms to carry nanoparticles to their targets. For example, immunotargeting by use of monoclonal antibodies, chimeric antibodies, and humanized antibodies has now reached the stage of clinical applications. Antibody engineering enables generation of high-quality cancer targeting antibody fragments [13, 14] (scFv, di-scFv, and minibody) and facilitates the search for clinically relevant biomarkers; conjugation of these immune specific ligands and aptamers enables nanomaterials for specific targeting with improved clinical efficacy. Many imaging modalities are available for the diagnosis of cancer using nanoparticle platforms either alone or in combination with additional signal molecules (Table 31.1). For example, nanoparticles modified with gadolinium contrast agent provided contrast enhanced magnetic resonance imaging (MRI) for the assessment of tumor treatment response. MRI provides a wealth of information on local biology and pathology based on nuclear magnetic resonance signals, received from hydrogen nuclei present in the organism under different pathophysiological conditions [15–18]. Similarly, a new class of radioimmunonanoparticles has been developed by decorating nanoparticles with cancer targeting ligands and radionuclides for targeted imaging and therapy [19, 20]. Radionuclides have been linked to nanoparticles with and without ligands; thus these particles can be used first in small animals and primates prior to use in humans. Furthermore, radioactive signal would allow testing these particles rapidly in mice implanted with human cell targets to characterize and optimize the PK behavior. Radioactive signal tagged to NPs is easy to track and validate the assay in cell culture and small animal models. Later, the same molecular system can be translated into clinical studies. Hence radioactive signal provides a powerful feature to the nanoparticle-based drug, from cell culture to preclinical and to clinical translation studies. The development of RINPs or immunonanoparticles enables researchers to improve the targeted delivery and thermotherapy [19–21]. The same technique could potentially be applied for the development of novel and more effective diagnostic agents and screening of potential cancer biomarkers and drug candidates to extend the limits of molecular diagnostics. Thus RINPs could provide significant improvement over current clinical management of cancer patients. Many studies have been carried out using radiolabeled nanoparticles both in vitro and in vivo for various diseases. In this chapter our major emphasis is to apply radiolabeled nanoparticles in the following areas: to enhance the delivery and selectivity, to target specific biomarker and image various disease processes, and for scintigraphic imaging of cancer inflammation/infection and sentinel lymph node detection. This chapter is organized as follows. The first three sections provide various types of nanoparticle cores, radionuclides, and targeting moieties used for biomedical applications. The next three sections deal with the reported study of immunonanoparticles and passive and active targeting of RNPs and RINPs. The last sections discuss RINPs used for cancer
P1: OTA/XYZ P2: ABC c31 JWBS037-Chen February 2, 2011
758
21:53
Printer Name: Yet to Come
RADIOIMMUNONANOPARTICLES FOR CANCER IMAGING AND THERAPY
TABLE 31.1 Summary of the Differences Between Imaging Modalities and Their Possibilities for Nanoparticle Applicationsa Modality
Spatial Resolution
Depth
Temporal Sensitivity Resolution (mol/L)
Molecular Probe
Nanoparticle Design
PET
1–2 mm
No limit
10 s− min
10−11 –10−12
ng
SPECT
0.5–1 mm
No limit
min
10−10 –10−11
ng
Bioluminescence 3–5 mm
1–2 mm
sec–min
10−15 –10−17
g–mg
Fluorescence
2–3 mm
<1 mm
sec–min
10 −9 –10−12
g–mg
MRI
25–100 mm
No limit
min–hours 10−3 –10−5
CT
50–200 mm
No limit
min
10−1 –10−4
N/A
Ultrasound
50–500 mm
mm–cm
sec–min
10−1 –10−4
g–mg
Label outside, in the membrane, or inside (radionuclide) Label outside, in the membrane, or inside (radionuclide) Label inside (or outside), luminescent compound Label outside or inside, fluorescent compound Label outside, in the membrane, or inside, paramagnetic atom, particles Label inside (or outside), contrast media Label inside, gas-filled particles
g–mg
a Spatial
resolution, depth resolution, temporal resolution, and sensitivity) Source: From Kairemo et al. [18].
imaging, radiomagnetonanoparticles for hyperthermia, and finally the merits and demerits of these novel agents including cancer therapy, hyperthermia, and thermal ablation, and MNP-directed toxicity studies.
31.2 TYPES OF NANOMATERIALS USED IN CANCER IMAGING AND THERAPY 31.2.1 Magnetic Nanoparticles Magnetic nanoparticles (MNPs) provide many new opportunities in cancer imaging and therapy by improving the quality of MRI, site-specific drug delivery, manipulation of cell membranes, and hyperthermic treatment for malignant cells [22]. For example, SPIO coated with polyethylene glycol or dextran has been widely used to generate MRI contrast enhancement [23]. These NPs possess large magnetic moments and create magnetic field homogeneity. The contrast enhancement is due to this effect, which results in shortening T2, which leads to loss of signal, also known as “negative contrast.” MNPs are categorized
P1: OTA/XYZ P2: ABC c31 JWBS037-Chen February 2, 2011
21:53
Printer Name: Yet to Come
TYPES OF NANOMATERIALS USED IN CANCER IMAGING AND THERAPY
759
based on their size: SPIO-NPs of average size greater than 50 nm are usually subjected to uptake by Kupffer cells (reticuloendothelial system) and lower than 50 nm ultrasmall paramagnetic particles (USPIO) allow longer circulation times and final removal by the lymphatic system. Several SPIO-based MRI contrast agents, including Lumirem® and Endorem® , are in clinical use for imaging of the gastrointestinal tract and for liver and spleen tissue. In the latter application, the increased contrast between healthy and diseased tissue facilitates the detection of primary or secondary liver tumors [24]. Sinerem® is available for blood pool visualization with MRI. USPIOs can also be used for the detection of lymph node diseases [25–28]. 31.2.2 Liposomes Liposomes are small lipid vesicles in the size range of 50–1000 nm [29–31]. Their unique advantages lie in their ability to protect the drugs from degradation, targeting the drugs to the site of action, and reducing the toxicity and side effects of such drugs. Liposomes have been studied extensively as drug carriers, particularly for cancer therapy. Depending on their size and number of bilayers, liposomes can be classified into three categories: multilamellar vesicles (MLVs), large unilamellar vesicles (LUVs); and small unilamellar vesicles (SUVs). Liposomes are classified in terms of composition and mechanism of intracellular delivery into five types: conventional liposomes (CLs), pH-sensitive liposomes, cationic liposomes, immunoliposomes, and long-circulating liposomes (LCLs). The major problems associated with liposomes are their stability, batch-to-batch variation in yield and quality, difficulty in sterilization, and low drug-loading capacity. Although liposomes have been extensively studied for the last few decades, the only efficient nanoformulation available in the market is Doxil® (ALZA, Mountain View, CA) [32]. 31.2.3 Dendrimers Dendrimers are a class of repeatedly branched polymeric macromolecules with numerous arms extending from a center, resulting in a nearly perfect three-dimensional geometric pattern. Dendrimers can be synthesized via two major strategies: (1) divergent methods and (2) convergent methods, which differ in their direction of synthesis; either outward from the core or inward toward the core, respectively. Divergent methods were first introduced by Tomalia et al. [33] in the 1980s, when his group synthesized three-dimensional polyamindoamine (PAMAM) dendrimers by the growth of branches extending radially from a core site to the periphery [34]. Furthermore, PAMAM dendrimers contain tertiary amines and amide linkages, which allow for the binding of numerous targeting and guest molecules. The advantage of dendrimers is that they can be synthesized and can vary in size designed for specific applications. Many of the properties of dendrimers [35] suit the drug delivery systems due to their feasible topology, functionality, and dimensions; also, their size is very close to various important biological polymers and assemblies such as DNA and proteins, which are physiologically ideal [36]. One of the major applications of dendrimers is as a delivery vehicle for various anticancer drugs. Many investigators have explored the possibilities of delivering various drugs, such as doxorubicin, cisplatin, and 1-bromoacetyl5-fluorouracil (5FU), and PEGylation for the prolonged delivery of anticancer drugs [32]. Thus dendrimer-based NP system provide a controlled architecture that allows creation of a multivalent and multifunctional template with improved pharmacokinetic behaviors (e.g., cancer targeting moieties, radionuclide chelates, and signal molecules) [30].
P1: OTA/XYZ P2: ABC c31 JWBS037-Chen February 2, 2011
760
21:53
Printer Name: Yet to Come
RADIOIMMUNONANOPARTICLES FOR CANCER IMAGING AND THERAPY
31.2.4 Single-Walled Carbon Nanotubes (SWNTs) SWNTs have been utilized to deliver various biological molecules across the cellular membrane without cytotoxicity [37–39]. Further research is exploiting the intrinsic physical properties of SWNTs to impart unique features to enhance the usability of nanotube biocarriers for cancer imaging and therapy [37]. The unique feature of using SWNTs in the biological system is the strong absorbance of light in the 700–1100-nm near-infrared (NIR) [40, 41] window, to exploit biological systems that are transparent for optical stimulation of nanotubes inside the living cells to afford various useful functions. These intrinsic physical properties of SWNTs can be exploited to afford new types of biological transporters with useful functionalities. For example, selective cell destruction can be achieved by functionalization of SWNTs with a folate moiety, selective internalization of SWNTs inside cells labeled with folate receptor tumor markers [42], and NIR-triggered cell death, without harming receptor-free normal cells. 31.2.5 Gold Nanoparticles (GNPs) GNPs are attracting wide interest in biomedical applications because of their unique properties [43] and their potential use in photothermal therapy [44–46], biosensing [47–49], molecular imaging [50–52], and gene delivery [53, 54], for cancer treatment. There are many reviews that provide in-depth coverage of developments in the synthesis, surface modification, molecular assembly, and biological effects of gold nanomaterials. The application of GNPs in biosensing and cancer (i.e., photothermal therapy), molecular imaging, and gene delivery are well known [55]. These gold nanomaterials have been successfully synthesized in different shapes and sizes [56] such as gold nanospheres [57], nanobelts [58], nanocages [59, 60], nanoprisms [61], nanostars [62, 63], and nanorods [64]. These particles were further evaluated for biological applications. 31.2.6 Quantum Dots The semiconductor nanocrystals or quantum dots (QDs) made of cadmium selenide cadmium sulfide, or cadmium telluride, surrounded by an inert polymer coating, are used as fluorescent labels of live cells, receptors, and monogenic markers [65]. These nanocrystals range from ultraviolet to far-infrared, including the visible spectrum. These particles can be functionalized to recognize molecular targeting systems like ligands and monoclonal antibodies attached to the outer coating against a specific target or via receptor interactions with specialized ligands. However, limitations of QDs are low penetration depth of light through tissues, which hampers external in vivo imaging. Near-infrared fluorescent QDs have recently been developed which enable the visualization of lymph nodes and xenografted tumors [66, 67] in mice.
31.3 RADIOISOTOPES FOR CANCER IMAGING AND THERAPY Radionuclides are used in nuclear imaging and therapy, based on particle emissions (alpha, beta, and Auger electrons; Table 31.2 shows the properties of these particles). For cancer therapy these particles destroy the nuclear DNA strands by radiation-induced ionizations, excitations, chemical transmutations, and local charge effects. However, our focus here is
P1: OTA/XYZ P2: ABC c31 JWBS037-Chen February 2, 2011
21:53
Printer Name: Yet to Come
LIGANDS USED FOR TARGETED CANCER IMAGING, DIAGNOSIS, AND THERAPY
761
TABLE 31.2 Physical Characteristics of Therapeutic Radionuclides
 ␣ EC/ICa a EC,
Particles
Energy Emin –Emax
Range
Electrons Helium nuclei Auger electrons
Medium to high (0.05–2.3 MeV) High 5–9 MeV Very low (eV–keV)
0.2–12 mm 40–100 m Several nm
electron capture; IC, internal conversion.
on radionuclides used in nanoparticles for various cancer imaging, particle tracking, and therapy [68–70]. 1. ␣-Particle emitters (e.g., 211 At, 212 Bi, 213 Bi, 225 Ac) eject short-range (40–80 m), high-energy (4–8 MeV), highly cytotoxic helium nuclei. 2. -Particle emitters (e.g., 67 Cu, 90 Y, 131 I, 177 Lu, 186 Re, 188 Re) release electrons that have substantially lower cytotoxicity than ␣-particles but longer range (0.4–10 mm). 3. ␥ -Emitters (99m Tc and 111 In). 4. + (positron) Emitters (18 F, 64 Cu, 86 Y, 124 I). 5. Auger electron emitters (111 In, 123 I, 125 I) have a short range (<1 m) and are cytotoxic only when in close proximity to the nucleus.
31.4 LIGANDS USED FOR TARGETED CANCER IMAGING, DIAGNOSIS, AND THERAPY Typically, nanoparticles have a large surface area to conjugate functional groups of multiple diagnostic (e.g., optical, radioisotopic) and therapeutic (e.g., anticancer chemotherapeutics) agents. This led to bioaffinity nanoparticle probes for molecular and cellular imaging, targeted nanoparticle drugs for cancer therapy, and integrated nanodevices for early cancer detection and screening. Many approaches have been investigated to develop tumor-targeted NPs. One approach is conjugation of active tumor-targeting ligands or components to nanoparticles to provide preferential accumulation of nanoparticles in the tumor-bearing organ, in the tumor itself, in individual cancer cells, or in intracellular organelles inside cancer cells. This approach is based on specific interactions, such as lectin carbohydrate, ligand–receptor, and antibody–antigen. Lectin–carbohydrate is one of the classic examples of targeted drug delivery. Lectins are proteins of nonimmunological origin, capable of recognizing and binding to glycoproteins expressed on the cell surface [71]. Antibodies, antibody fragments, peptides, and small molecules have been used to target a variety of tumor-associated antigens. In particular, antibodies show their inherent specificity, stability, and versatility and have also been conjugated to various agents to serve either as a therapeutic agent or as an imaging modality or both. The U.S. FDA approved many antibody-based therapeutics to target a variety of cancers and cancer-associated antigens or to modulate the immune response (Table 31.3). In addition, antibodies, peptides, aptamers, and ligands that target vascular antigens, such as vascular endothelial (VE)-cadherin, VEGFR E-cadherin, or integrin, have themselves demonstrated therapeutic promise or have been conjugated to cytotoxic agents, such as radioisotopes, for cancer therapy.
P1: OTA/XYZ P2: ABC c31 JWBS037-Chen February 2, 2011
762
21:53
Printer Name: Yet to Come
RADIOIMMUNONANOPARTICLES FOR CANCER IMAGING AND THERAPY
TABLE 31.3 FDA Approved Therapeutics Ligand
FDA Approved ®
Mab
Rituxan
90
Zevalin® (ibritumomab tiuxetan) Bexxar® (tositumomab)
CD20
CD33
Mab
Mylotarg® (gemtuzumab ogozagomicin) Campath® (alemtuzumab)
Mab
Erbitux® (cetuximab)
EGFR
Mab
Vectibix® (panitumomab)
EGFR
Mab
Herceptin® (trastuzumab)
Her2/neu
Mab
Avastin® (bevacizumab)
VEGF
Mab
Zenapax® (daclizumab)
IL-2-diphtheria toxin
Ontak® (denileukin diftitox
IL-2 receptor (CD25) IL-2 receptor
Y-Mab
131
I-Mab
Drug-Mab
(rituximab)
Specific Target CD20
CD20
CD52
Cancer Target
Year
CD20-positive, B-cell non-Hodgkin lymphoma (NHL) B-cell non-Hodgkin lymphoma B-cell non-Hodgkin lymphoma CD33-positive acute myeloid leukemia B-cell chronic lymphocytic leukemia Metastatic colorectal cancer; squamous cell; cancer of the head and neck Refractory (EGFR-positive) metastatic colorectal cancer HER-2/neu-overexpressing metastatic breast cancer Metastatic colorectal cancer; nonsquamous, nonsmall cell lung cancer T-cell cancers; immununosuppression CD25-positive persistent or recurrent cutaneous T-cell lymphoma
1997
2002 2003 1997 1999 2000
2001
2004 2006
1998 2004
31.5 RADIONANOPARTICLES 31.5.1 Radionuclide Linked to Passive Targeting Nanoparticles Passive targeting NPs exploit the features of tumor biology that allow nanocarriers to accumulate in the tumor by the enhanced permeability and retention (EPR) effect. Passive targeting of nanoparticles was accomplished by two mechanisms: (1) particles should avoid uptake by the reticuloendothelial system (RES) and (2) particle size allows entrance into the tumor vasculature. Controlling the size and surface properties is possible by suitable modification of nanoparticles. For example, rapid vascularization in fast-growing cancerous tissues resulted in leaky, defective architecture and impaired lymphatic drainage [72–74]. This structure allows an EPR effect [75–79] resulting in the accumulation of nanoparticles at the tumor site. To maximize circulation times and targeting ability, the optimal size should be less than 100 nm in diameter and the surface should be hydrophilic to circumvent clearance by macrophages. A hydrophilic surface of the nanoparticles safeguards against plasma protein adsorption and can be achieved through hydrophilic polymer coatings such
P1: OTA/XYZ P2: ABC c31 JWBS037-Chen February 2, 2011
21:53
Printer Name: Yet to Come
PASSIVE TARGETING RADIOLABELED NANOPARTICLES
763
as PEG, poloxamines, and poloxamers polysaccharides, or through the use of branched or block amphiphilic copolymers [80–82]. The covalent linkage of amphiphilic copolymers (polylactic acid, polycaprolactone, polycyanonacrylate chemically coupled to PEG) is generally preferred, as it avoids aggregation and ligand desorption when in contact with blood components. Although many passively targeting NPs are in clinical trial [83], the first clinical trials reached in the mid-1980s, and the first products, based on liposomes and polymer–protein conjugates, were marketed in the mid-1990s [84]. An alternative passive targeting strategy is to utilize the unique tumor environment in a scheme called tumor-activated prodrug therapy. The drug is conjugated to a tumor-specific molecule and remains inactive until it reaches the target [85]. Overexpression of the matrix metalloproteinase (MMP) MMP-2 in melanoma has been shown in a number of preclinical as well as clinical investigations. Mansur et al. [86] reported a water-soluble maleimide derivative of doxorubicin (DOX) incorporating an MMP-2-specific peptide sequence (GlyPro-Leu-Gly-Ile-Ala-Gly-Gln) that rapidly and selectively binds to the cysteine-34 position of circulating albumin. The albumin–DOX conjugate is efficiently and specifically cleaved by MMP-2, releasing a DOX tetrapeptide (Ile-Ala-Gly-Gln-DOX) and subsequently DOX. The pH and redox potential have also been explored as drug release triggers at the tumor site [87]. Another passive targeting method is the direct local delivery of anticancer agents to tumors. This approach has the obvious advantage of excluding the drug from the systemic circulation. However, the administration procedure can be highly invasive, as it involves injections or surgical procedures. For some tumors, such as lung cancers, that are difficult to access, the technique is nearly impossible to use. Recently, Jankovic et al. [88] have developed 90 Y-labeled colloidal nanoparticles to prepare therapeutic radiopharmaceuticals. They compared various sizing techniques, including TEM analysis for the accurate measurement of size, shape, and chemical nature of these radiocolloids.
31.6 PASSIVE TARGETING RADIOLABELED NANOPARTICLES Chao-Ming et al. [89] reported an approach of labeling with the radioisotope 99 Tc directly on ferrite nanoparticles for diagnostic applications. In this approach, the reduced 99m Tc ions were mixed directly into the solution in which ferrite particles are formed, without complicated preprocess of chemical modification. The labeling efficiency of the radioNP was >90% and demonstrated the stability of the particles over a period of time. The radiolabeled ferrite nanoparticles were intravenously injected into the tail vein of male Wister rats (200–250 g). The scintigraphic images were monitored by planar imaging, using a gamma camera (GE, DST, XLi) with parallel collimator. The biodistribution of 99m Tc labeled NPs showed high uptake by the liver; a second large uptake is by lung, kidneys, and spleen. Little uptake was found in the heart and brain. The radioactive-NP accumulation in each organ is shown in Figure 31.2, at 5, 30, 60, and 180 min, respectively. This study demonstrated in rats that 80–90% of the injected dose was taken up by the liver within 5 min after injection. A further increase of radioactive particle concentration was accomplished at a specific site by applying an external magnet (Fig. 31.3). Thus this study provided the information that the phagocytic uptake of radiolabeled NPs was strongly related to the surface charge of the particles. Reducing the surface charge of ferrite nanoparticles may improve phagocytic uptake in magnetic targeting diagnosis. Surface modification of the radiolabeled ferrite may be obtained by adjusting the pH value in the aqueous reaction solution or by adding reducing reagents. As in previous reports [90–93], this study also
P1: OTA/XYZ P2: ABC c31 JWBS037-Chen February 2, 2011
764
21:53
Printer Name: Yet to Come
RADIOIMMUNONANOPARTICLES FOR CANCER IMAGING AND THERAPY
10000000 x10
Radioactivity (count)
9000000
5 min 30 min 60 min 180 min
8000000 7000000 6000000 5000000 4000000 3000000 2000000 1000000 0 He
ar
t
Li
ve
r
Lu
ng
h
n
ai
Br
ac
S
m to
Ki
r s e e en de ea tin stin le cr p ad es e l t t n S B In In Pa L. S.
ne
y
FIGURE 31.2 Biodistribution of radioactive nanospheres in organs of the rats after injection, without applied external magnet. (Reproduced with permission from Chao-Ming et al. [89].)
demonstrated the immune response toward the intravenously injected particles, which were recognized as foreign and were removed from blood circulation by the mononuclear phagocytic system (MPS). To overcome the mononuclear phagocyte system, in another experiment Chao-Ming et al. [94] utilized polyethylene glycol (PEG) to prepare biocompatible 99m Tc-ferrite nanoparticles. These new particles were used for in vivo biodistribution studies upon intravenous injection in rats. The scintigraphic image of the PEG conjugated ferrite NPs (Fig. 31.4 right) demonstrated more radio-NPs being dispersed out of liver and lung, as compared to that without PEG (Fig. 31.4 left). PEG has been known to link to solid surfaces via
FIGURE 31.3 The scintigraphic image obtained after intravenously injection of the 99m Tc labeled ferrite nanoparticles into a mice: (right) with application of an external magnet near the right thigh before injection and (left) placing the magnet on the left thigh. (Reproduced with permission from Chao-Ming et al. [94].)
P1: OTA/XYZ P2: ABC c31 JWBS037-Chen February 2, 2011
21:53
Printer Name: Yet to Come
PASSIVE TARGETING RADIOLABELED NANOPARTICLES
765
FIGURE 31.4 Scintigraphic images of rat upon intravenous injection of the ferrite-radiobeads (left), and ferrite-radiobeads conjugated with PEG (right). (Reproduced with permission from ChaoMing et al. [94].)
functional end groups, and thus it was used to modify the surface charge of the NPs [95]. The PEG modified radio-NPs have demonstrated evidence of reducing uptake by liver and lungs. The observed result may be attributed to the surface charge modification of ferrite radio-NPs, thus decreasing uptake by the liver so beads can be retained in the blood pool. Recently, Devaraj and colleagues [96] synthesized an 18 F-modified trimodal nanoparticle (18 F-CLIO) and tested it for in vivo characterization that shows promise for PET, fluorescence molecular tomography, and MR imaging. These particles were crosslinked to dextran held together in core–shell formation by an SPIO core and functionalized with the radionuclide 18 F in high yield via “click” chemistry. Similarly, Jarrett et al. [97] have developed dual-mode PET/MRI active nanoparticle probes that target vascular inflammation for combined magnetic resonance (MR) and positron emission tomography (PET) imaging, using iron oxide nanoparticles coated with dextran sulfate that linked to positron-emitting copper-64 isotopes.
31.6.1 RNP – Liposome Core A recent review details the development of liposome-based nanoparticles with gammaemitting radionuclides for scintigraphic imaging of cancer, inflammation/infection, and sentinel lymph node detection [98]. In this review, special emphasis was given to the clinical studies performed with liposome-based radionanoparticles for detection of tumors, infectious/inflammatory sites or metastatic lymph nodes. Many articles were cited; the clinical studies with liposome radiopharmaceuticals demonstrated that a wide variety of tumors could be detected with good sensitivity and specificity. These particles also clearly detect various soft tissue and bone inflammatory/infectious lesions, and performed equal to or better than infection imaging agents that are approved at present.
P1: OTA/XYZ P2: ABC c31 JWBS037-Chen February 2, 2011
766
21:53
Printer Name: Yet to Come
RADIOIMMUNONANOPARTICLES FOR CANCER IMAGING AND THERAPY
31.6.2 Nuclear Medicine Imaging and Therapy In a short review Lucignani [99] summarized a number of relatively new technological approaches for the synthesis of radiotracers. In the field of therapeutic nuclear medicine, an interesting development comes from the synthesis of lanthanide nanoparticles in the form of clusters of 153 Smatoms, meant for peptide mediated tumor treatment [100]. Liposome NPs were utilized to carry radioactive compounds as radiotracers linked to multiple locations in liposomes. One option is the hydrated compartment inside the liposome, another the lipid core into which especially hydrophobic conjugates can be attached, and the third option is the outer lipid leaflet where molecules can be bound by covalent linkage. Nonlabeled peptide-targeted liposomes have been used in experimental cancer therapy, and augmented killing (approximately fourfold) of U937 leukemia and HT1080 sarcoma cells was obtained by the peptide-targeted delivery of doxorubicin-containing liposomes. 31.6.3 Disadvantages In nearly all work with particulate carriers, it is clear that interaction of the particulates with phagocytic cells results in significant accumulation in liver and spleen. Furthermore, the reticuloendothelial system (RES) is the major source of particulate uptake and clearance in the body [101–104]. Diverse and specialized phagocytic cells comprising this system are present in different organs (Kupffer cells in liver, Langerhans cells in skin, etc.). Many methods have been devised to temporarily inhibit the RES, mostly relying on injection of a particulate material, such as carrageen, iron, or silica particles, to competitively incapacitate the RE cell function [105, 106]. Clodronate liposomal preparations, which are naturally targeted NPs engulfed by active phagocytic cells, have recently proved to be an excellent method for temporary phagocytic inhibition. These preparations are effective, have few side effects, and are very stable [107, 108].
31.7 ACTIVE TARGETING RADIOACTIVE NANOPARTICLES The principle of active targeting is well known in diagnostic and therapeutic nuclear medicine [109], in which radiolabeled biomolecules with unique biological properties have found widespread applications. However, combining nuclear medicine, tumor targeting moiety, and nanoparticles for cancer diagnosis, imaging, and therapy is gaining more attention recently. As discussed in the previous section, nanomaterials alone without targeting agents were accumulated in the RES [110, 111]. To direct the particles to a desired site, NPs must be coupled to a specific targeting agent (e.g., peptides, MAb, and small molecules). Thus a ligand-linked NP system provides unique opportunities to study biodistribution, vascular targeting, and stability of the NPs. For example, the targeting ability of G-250MAb conjugated to MRI contrast agents wrapped in an neutral liposome was tested in renal cell carcinoma [112]. A large fraction of the injected antibody accumulates at the target within minutes of injection, with little uptake by other organs. Most of the commercially available nanomaterials and contrast agents do not have active targeting characteristics. For active targeting, newer versions of nanoparticles have been generated with suitable surface and functional modifications that can be used to decorate ligands. Thus particles can readily be conjugated to monoclonal antibodies or receptortargeting agents. A widely investigated modality for the recognition of cancer cells is the
P1: OTA/XYZ P2: ABC c31 JWBS037-Chen February 2, 2011
21:53
Printer Name: Yet to Come
ACTIVE TARGETING RADIOACTIVE NANOPARTICLES
767
use of monoclonal antibodies. The recognition of cell surface antigens has been successful in various instances, especially in nonsolid tumors such as non-Hodgkin lymphoma [113]. For example, Herceptin® conjugated MRI probes allow successful monitoring of in vivo behavior of human cancer cells in a mouse model [114]. Of particular interest would be the active targeting with MRI contrast agents covalently linked to monoclonal antibodies recognizing epitopes with internalizing properties, thus allowing the magnetic labeling of specific cancer cells [8]. In another approach, instead of antibodies, nonimmunogenic folate was utilized as a targeting agent [115]. Folate receptor is expressed in various tumors, including ovarian, colorectal, and lung malignancies, since DNA methylation is dependent on folate. Moreover, the folate receptor is absent in most human normal tissues. Thus nanoparticles decorated with folic acid demonstrated internalization into the cells through receptor mediation [115–117]. In another study Kennel et al. [118] developed a model system of radiolabeled NPs, based on quantum dots, to investigate the competition between efficient vascular targeting and interaction of the NP with the RES. They prepared 125m Te-labeled CdTe NPs (125m Te half-life similar to 125 I; t 12 = 59 days) capped with ZnS. These particles were synthesized and derivatized with mercapto acetic acid and then conjugated with 201B antibody to target murine thrombo modulin expressed in the lumen of lung blood vessels. The MAb-targeted NPs were tested for targeting performance in vivo using single-photon emission computed tomography (SPECT)/computed tomography (CT) imaging, tissue autoradiography, and biodistribution. Biodistribution was also determined in mice that had been depleted for phagocytic cells by use of clodronate-loaded liposomes. The study results showed that MAb 201B conjugated to Cd125m Te/ZnS NPs was accumulated in lung (>400% injected dose [ID]/g) within 1 h of intravenous injection compared with control antibody-coupled NPs that did not accumulate in lung (<10% ID/g) but accumulated in liver and spleen. Images from micro-SPECT/CT (Table 31.4, Figs. 31.5 and 31.6) and autoradiography studies demonstrated specific uptake in lung and uniform distribution by the targeted NPs and with minor accumulation in liver and spleen. The biodistribution of targeted NPs in normal mice versus phagocytic cells depleted mice also TABLE 31.4 Biodistribution of 125 I MAb 201B-CdTe NP in Normal and Clodronate-Treated Mice at 1 Houra Percent ID per gram (×± , n = 5) Normal Mice Tissue
125
Liver Spleen Lung
3.4 ± 0.4 6.4 ± 0.9 268 ± 33
I MAb
Clodronate Mice
125
I MAb NP
57 ± 5 87 ± 6 48 ± 12
125
I MAb NP
5.9 ± 0.5 268 ± 44 169 ± 29
a MAb 201B was radioiodinated with 125 I using chloramine T. The product was dialyzed free of small molecules and coupled with cold CdTe NP. Mice receiving MAb only were injected intravenously with 200 L of 125 I MAb 201B (8 g, 22 Ci). Mice receiving MAb coupled with NP were injected with 200 L PBS containing 8 g, 22 Ci MAb coupled to 22 g cold CdTE NP. Clodronate treatment of mice was with 100 L of liposomal suspension injected intravenously 24 h prior to radioisotope administration. Source: Kennel et al. [118].
P1: OTA/XYZ P2: ABC c31 JWBS037-Chen February 2, 2011
768
21:53
Printer Name: Yet to Come
RADIOIMMUNONANOPARTICLES FOR CANCER IMAGING AND THERAPY
(a) MAb 201B targeted NP
(b) Control MAb targeted NP
FIGURE 31.5 Three-dimensional tomographic images of SPECT/CT scans from mice injected with Cd 125m Te NP as described in the footnote for Table 31.4 1 day after injection. SPECT images of (a) MAb 201B targeted NP 201B or (b) control MAb 33 NP. Three views with different orientations are shown with color representing the SPECT data. The fourth image in each panel is superimposed SPECT and CT data ventral view. (Reproduced with permission from Kennel et al. [118].)
provided evidence that in mice treated with clodronate liposomes, accumulation of NPs in liver was reduced by fivefold, while accumulation in lung at 1 h was enhanced by ∼50%. By 24 h, loss of the targeted NPs from lung was inhibited severalfold, while accumulation in liver and spleen remained constant. Thus the treated mice had a larger accumulation and retention of the NPs at the target site and a decrease in dose to other organs except spleen. This study clearly indicates that MAb directed NP preparations exhibit a difference between the effectiveness of the targeting agent and the natural tendency for RES uptake of the particles. Temporary inhibition of the RES may enhance the usefulness of NPs for drug and radioisotope delivery. In another study, investigators targeted tumor antigens or relied on the leaky tumor vasculature and the phagocytic activity of tumor cells, but accumulation of NPs to sites outside the vascular space is limited [102, 119–121]. For this reason, many recent studies
P1: OTA/XYZ P2: ABC c31 JWBS037-Chen February 2, 2011
21:53
Printer Name: Yet to Come
ACTIVE TARGETING RADIOACTIVE NANOPARTICLES
769
FIGURE 31.6 Mice injected with 123 I-labeled peptide-targeted PEGylated liposomes. On the left the mouse was preinjected with a large amount of cold control peptide (displacement study), the tumor uptake is weak; whereas sufficient tumor uptakes are seen in mice with no peptide blockade (middle and right). (Reproduced with permission from Kairemo et al. [18].)
have taken advantage of targets in the vascular space such as markers of angiogenesis [122–125]. Recently, antibody-conjugated carbon nanotubes ranging up to 200 m in length were studied with radioisotope linked via a chelator [126] at other sites in the nanotubes. Several targeting constructs were tested and found to be useful for targeting the nanotubes to human lymphoma xenografts; however, in this study liver uptake was quite high. Another work shows 64 Cu attached to commercial CdTe NPs and the biodistribution being monitored using dynamic positron emission tomography (PET) imaging [123, 127]. In these studies, the circulation half-life of the NPs was about 2 min, and coating of the NP surface with polyethylene glycol (PEG) extended this serum half-life time to only about 6 min, with the NPs still accumulating largely in spleen and liver [127]. Kobayashi et al. [128] prepared a radioconjugate of polyamidoamine (PAMAM) dendrimer-1B4 M-anti Tac IgG MAb labeled with 111 In and 88 Y; they demonstrated >99% radiochemical purity, good immunoreactivity, and significant tumor accumulation in female athymic nude mice. In another study, the same group of researchers prepared murine monoclonal IgG1 labeled with 111 In linked to PAMAM, which demonstrated minimal loss of immunoreactivity and high specific activity [129]. Biodistribution studies of these preparations indicated fast blood clearance of the radiolabeled dendrimer conjugate and specific tumor accumulation. Both studies showed high accumulations of PAMAM-1B4M-antibody conjugates in the liver, kidney, and spleen. However, this nonspecific localization could be
P1: OTA/XYZ P2: ABC c31 JWBS037-Chen February 2, 2011
770
21:53
Printer Name: Yet to Come
RADIOIMMUNONANOPARTICLES FOR CANCER IMAGING AND THERAPY
significantly reduced when 1B4M chelates were saturated with stable indium or yttrium [128, 129]. In addition, PAMAM dendrimers incorporating 111 In-labeled antisense oligonucleotides have been studied as a tracer for solid tumor imaging [130–132]. Scintigraphic imaging of the radiolabeled dendrimer conjugate in tumor–bearing nude mice showed clear tumor localization with sufficient contrast to background organs. The biodistribution data revealed significantly enhanced tumor delivery of the dendrimer conjugate as compared to the free oligonucleotides [131]. Similarly, in another set of studies, PAMAM dendrimers were conjugated to isocyanato polyhedral borane and the boronated dendrimers were then conjugated to a monoclonal antibody against murine B16 melanoma [133]. However, in vivo, the conjugate largely accumulated in liver and spleen. Subsequently, boronated PAMAM dendrimers were conjugated to EGF to produce a stable conjugate containing approximately 960 atoms of boron per conjugate [133]. In vitro studies in human malignant glioma cells indicated high affinity binding of the conjugate to EGF receptor and subsequent internalization by endocytosis. In vivo evaluation in EGF receptor positive glioma in Fischer rats showed that intravenous (IV) injection of the conjugate resulted in large accumulation in the liver and spleen with low uptake in the tumor [134]. In contrast, intratumoral injection of the conjugate resulted in high tumor uptake with insignificant localization in liver and spleen. Targeting the EGF receptor was achieved using cetuximab-MAb against EGF receptor [134]. A heavily boronated generation 5 PAMAM dendrimer conjugated to cetuximab was synthesized containing ∼1100 boron atoms per conjugate. Biodistribution studies were carried out after intratumoral injection of the conjugate in brain tumor-bearing rats. At 24 h the conjugate uptake in tumor was sevenfold greater than in normal brain tissue [134]. Based on the favorable tumor uptake results, therapy studies were initiated in rats with intracerebral glioma. Following treatment, the animals showed dramatic increase in mean survival times (45 ± 3 days) as compared to the untreated controls (25 ± 3 days). Liang et al. [135] developed rhenium-188 radionuclide conjugated magnetic nanoparticles for radiotherapy. Its half-life of 16.9 h and -emissions of 2.12 MeV (71.6%) and 1.96 MeV (25.1%) are suitable for therapy with an average penetration range of 2.6 mm, and the ␥ -emission of 155 keV (15%) allows imaging with a ␥ -camera. In this study, they synthesized SPIO particles and modified them with amino-silane, and Hepama-1 MAb against liver cancer was conjugated to functional SPION by the crosslinker of glutaraldehyde to prepare immuno-SPION. The modified particles were radiolabeled with rhenium-188 for biomagnetically targeted therapy. Previously, H¨afeli et al. [90, 91, 136] demonstrated this technique in both animal and cell culture studies using yttrium-90 and rhenium-188. Hepama-1 MAb was conjugated to 10-nm SPIO magnetite nanoparticles and radiolabeled with 188 Re, 188 ReO4 − on the surface of the modified MAb. The specific activity was 0.1 mCi 188 ReO4 − per 5 mg of NPs; the best labeling condition was optimized in a volume of 100 L for 50 min at 37 ◦ C. The labeling efficiency was about 90% and the stability of the product in serum albumin was 80% after 10 h. In vitro results showed that 188 Re-labeled particles killed SMMC-7721 cells effectively [136]. In another approach, biocompatible radioactive holmium-loaded nanoparticles were developed for a targeted multitherapy technique. For this, holmium acetylacetonate has been encapsulated in poly-l-lactide (PLLA)-based nanoparticles by the oil-in-water emulsion–solvent evaporation method. The NPs were irradiated in a nuclear reactor with a neutron flux of 1.1 × 10 [13] n/cm2 /s for 1 h, which yielded a specific activity of about 27.4 GBq/g of NPs being sufficient for our desired application. These RINPs may
P1: OTA/XYZ P2: ABC c31 JWBS037-Chen February 2, 2011
21:53
Printer Name: Yet to Come
ACTIVE TARGETING RADIOACTIVE NANOPARTICLES
771
represent a novel nanoparticle-based radiopharmaceutical for intratumoral radiotherapy [137, 138]. Jestin et al. [139] developed novel bispecific antitumor antihapten antibody coated nanocapsules labeled with a variety of radionuclides [139]. Specific constructs (DSPEPEG-haptens, DSPE-PEG-BCA, and DSPE-PEG-BH) were synthesized to develop new radiolabeled vector formulations. In this construct PEG was used to increase the residence time in blood. Serum stability studies showed that this 99m Tc-labeling method was efficient for at least 20 h. Several methods were used for 111 In-labeling by using DSPE-PEGDTPA and for 125 I labeling with DSPE-PEG-BH. This study provides nanocapsule labeling feasibility with a variety of radionuclides and their stability in serum was tested. Kim et al. [140] developed novel radioactive magnetic fluids for the treatment and diagnosis of cancer. These magnetic nanoparticles of Cux Fe1-x O were first coated with surfactant utilizing decanoic acid and nonanoic acid. These tightly bonded decanoic acid and nonanoic acid layers increased particle–particle repulsion and allowed homogeneous and stable dispersion of these particles in water. Also, the exposed carboxylic acid of the surfactant prevented the magnetic nanoparticles from being oxidized by air. The radioactive Cu2+ component of the magnetic fluids emits beta rays and gamma rays. The beta radiation effectively serves to kill tumor cells, while the gamma radiation is easily imaged with a gamma camera. Kim et al. [140] determined that the radioactive magnetic fluids could be used as a therapeutic drug or diagnostic reagent for cancer treatment and imaging. Esenaliev [141] developed targeted nanoparticles by the attachment of antibodies directed against antigens in tumor vasculature for selective delivery to tumor blood vessel walls [142]. However, this system utilized the interaction of electromagnetic pulses or ultrasonic radiation with nanoparticles for the enhancement of drug delivery in solid tumors. Upon delivery, perforation can be induced by ultrasonic waves or local heating of the particles by pulsed electromagnetic radiation, resulting in perforation of tumor blood vessels, microconvection in the interstitium, and perforation of cancer cell membrane, therefore providing enhanced delivery of macromolecular therapeutic agents from blood into cancer cells with minimal thermal and mechanical damage to normal tissues. Ex vivo animal tissue studies were performed to determine particle penetration due to laser radiation and sonication utilizing activated carbon particles 1 m in diameter placed within various animal tissues. Fluorescein isothiocyanate-dextran was utilized to study penetration of macromolecules in the tissues. Li et al. [143] developed novel nanoparticle-based therapeutic agents of anti-Flk-1 MAbNP-90 Y and IA-NP-90 Y that demonstrated potential therapeutic efficacy for treatment of a variety of tumor types. The study results showed that targeted radiotherapy works using different targeting agents on a nanoparticle, to target both the integrin ␣v 3 and the vascular endothelial growth factor receptor. The study investigated the potential therapeutic efficacy of three-component treatment regimens using two murine tumor models. Integrin ␣v 3 and vascular endothelial growth factor receptor 2 (Flk-1) have been shown to be involved in tumor-induced angiogenesis. The anti-Flk-1 MAb conjugated NP radiolabeled with 90 Y to target melanoma (murine tumor models K1735-M2) and similarly the small molecule integrin antagonist (IA) linked to NP-radiolabeled with 90 Y to target colon adenocarcinoma (4-[2-(3,4,5,6-tetrahydropyrimidin-2-ylamino)ethoxy]-benzoyl-2(5)-aminoethylsulfonylamino-beta-alanine, which binds to the integrin ␣v 3 , CT-26) were utilized to evaluate treatment efficacy. Lee et al. [144] developed a bifunctional iron oxide nanoparticle probe (45 ± 10 nm) using polyaspartic to link PET and MRI probes linked to cyclic arginine–glycine–aspartic
P1: OTA/XYZ P2: ABC c31 JWBS037-Chen February 2, 2011
772
21:53
Printer Name: Yet to Come
RADIOIMMUNONANOPARTICLES FOR CANCER IMAGING AND THERAPY
(RGD) peptides for integrin ␣v 3 targeting in a mouse model. The 64 Cu-labeled DOTA linked to NPs was used as the PET tracer. These RGD-coated radiolabled NPs were tested in vitro and in vivo to determine receptor targeting efficacy and feasibility for dual PET/MRI imaging. Both small-animal PET and T2-weighted MRI demonstrated integrin-specific delivery of RGD-aspartic nanoparticles and prominent reticuloendothelial system uptake; and also provided simultaneous dual imaging (PET and MRI) of tumor. These bifunctional radiopeptide NPs could allow for earlier tumor detection with a high degree of accuracy and may provide further insight into the molecular mechanisms of cancer [144].
31.8 RADIOMAGNETONANOPARTICLES FOR FOCUSED CANCER THERAPY Radionanoparticles were first developed by immobilizing radiometals on magnetic microspheres to kill cancer cells. Focused radiotherapy was achieved by dragging microspheres to the vicinity of tumor tissue using an external magnetic field. Blanchard et al. [145] first used this system for cancer therapy using resin microspheres loaded with 90 Y by IV injection for patients who had astrocytomas or liver, prostate, lung, or tongue cancers. However, these attempts were not successful due to the severity of the patients’ condition, along with undefined particle sizes. Later, Houle et al. [146] and Mantravadi et al. [147] developed 90 Y linked glass microspheres for the treatment of hepatomas, with partial success. In this system the glass microsphere causes anatomical distress to the patient. To overcome this problem Hafelie et al. [91] developed biodegradable magnetic microspheres made from poly-l-lactic acid using 90 Y radioisotopes. In this approach directing the particles from outside the body to lesions near the body’s surface provided very efficient targeting. These particles could be a new class of radioactive magnetic nanoparticles for cancer therapy. This system was further developed for hyperthermia treatment; thermal therapy using an external alternating magnetic field enhanced the temperature at the tumor site.
31.9 RADIOIMMUNONANOPARTICLES FOR HYPERTHERMIA DeNardo et al. [19, 21] utilized 111 In-MAb-linked iron oxide nanoparticles to target breast cancer to deliver tumor-specific thermal therapy (Rx) for metastatic cancer. Heat was induced by an externally applied alternating magnetic field (AMF). The Rx potential of this thermal therapy was evaluated for in vivo tumor targeting and efficacy, and predictive radionuclide-based heat dosimetry was studied using 111 In-ChL6-NPs (ChL6 is chimeric L6) in a human breast cancer xenograft model. These radiolabled NPs were purified and well characterized (20–30 Ci/2.2 mg of NPs) and injected intravenously into mice bearing HBT3477 human breast cancer xenografts. Pharmacokinetic data were obtained at 1, 2, 3, and 5 days after treatment. AMF was applied 72 h after bioprobe injection at amplitudes of 1410 (113 kA/m), 1300 (104 kA/m), and 700 (56 kA/m) oersteds (Oe) at 30%, 60%, and 90% “on” time (duty), respectively, and at 1050 Oe (84 kA/m) at 50% and 70% duty over the 20-min treatment. Treated and control mice were monitored for 90 days. Tumor total heat dose (THD) delivered by the NPs, that is, induced by AMF, was calculated for each Rx group using the 111 In-MAb-NPs tumor concentration and premeasured particle heat response to AMF amplitudes. Tumor growth delay was analyzed by Wilcoxon
P1: OTA/XYZ P2: ABC c31 JWBS037-Chen February 2, 2011
21:53
Printer Name: Yet to Come
RADIOIMMUNONANOPARTICLES FOR HYPERTHERMIA
773
rank sum comparison of time to double, triple, and quintuple tumor volume in each group, and all groups were compared with the controls. Mean tumor concentration of 111 In-MAbNPs at 48 h was 14 ± 2% injected dose per gram; this concentration 24 h before AMF treatment was used to calculate THD. No particle-related toxicity was observed. Toxicity was observed at the highest AMF amplitude-duty combination of 1300 Oe and 60% over 20 min; 6 of 10 mice died acutely. Tumor growth delay occurred in all of the other groups, correlated with heat dose and, except for the lowest heat dose group, was statistically significant when compared with the untreated group. Electron microscopy showed 111 InMAb-NPs on tumor cells and cell death by necrosis at 24 and 48 h after AMF. Radioactive mAb-guided NPs effectively targeted human breast cancer xenografts in mice. The thermal dose delivered by 111 In-MAb-NPs in tumor by external AMF correlated with tumor growth delay. Liang et al. [148] successfully developed magnetic nanoparticles with glutaraldehyde crosslinker to link humanized MAb (Hepama-1) against liver cancer to prepare RINPs. These 20-nm magnetite NPs were coated with a fine layer of silane on the surface of the magnetite core and labeled with 188 Re. The labeling efficiency was about 90% and the stability in serum albumin was 85% until 10 h. These particles were designed for biomagnetic radiotherapy [148]. Fu-Wen and co-workers [149] developed radionanoparticles with a novel approach by directly mixing with radioisotope 99m Tc into the reaction solution of ferrite nanoparticles to achieve high labeling efficiency. These radiolabeled nanoparticles preserved their small size and exhibited enough surface space for further biochemical conjugation for therapeutic applications, for example polyethyelene glycol and therapeutic herbs (Gentiana comnination and Trionyx sinensis) for chronic hepatitis. Therapeutic herbal modified radionanoparticles were tested in an in vivo biodistribution study in rats and monitored by a gamma camera. Natarajan et al. [20] developed NPs of various size and type (20-nm SPIO; 30- and 100nm Nano Ferrite) to deliver more heat induced by an external alternating magnetic field. These particles were constructed with dextran and PEG-coated iron oxide that covalently attached to radiolabled MAb (i.e., radioimmunonanoparticles, RINPs) to target breast cancer cells selectively. These particles were well characterized to study pharmacokinetics in athymic mice bearing human breast cancer (HBT 3477) xenografts. RINPs (2.2 mg) were injected intravenously and whole body; blood and tissue data were collected at 4, 24, and 48 h. Mean tumor uptakes (% ID/g ± SD) for each particle—20 nm, 30 nm, and 100 nm RINPs—at 48 h were 9.00 ± 0.8 (20 nm), 3.0 ± 0.3 (30 nm), and 4.5 ± 0.8 (100 nm), respectively; the range of tissue uptakes were liver (16–32 ± 1–8), kidney (7.0–15 ± 1), spleen (8–17 ± 3–8), lymph nodes (5–6 ± 1–2), and lung (2.0–4 ± 0.1–2). This study intended to select optimal particle size for focused hyperthermia. The study demonstrated that, although 100-nm RINPs targeted tumor less than 20-nm particles, their heating capacity is typically six times greater, suggesting the 100-nm RINPs could still deliver better therapy with AMF. This study also showed that conjugation of NPs with suitable nuclear imaging and molecular targeting agents provides opportunities for imaging, targeting, and treating cancer cells selectively by thermotherapy. In another study, Natarajan et al. [150] developed a 20-nm NP-based radioconjugate 111 -In-DOTA-di-scFv-NP), using recombinantly generated antibody fragments, di-scFv-c, for imaging and therapy of anti-MUC-1-expressing cancers. Since aberrant MUC-1 is abundantly expressed on the majority of human epithelial cancers, anti-MUC-1 discFv-c (50 kDa) was engineered, generated, and selected to link maleimide functionalized nanoparticles (NP-M). The characterization of the RINP was carried out by polyacrylamide
P1: OTA/XYZ P2: ABC c31 JWBS037-Chen February 2, 2011
774
21:53
Printer Name: Yet to Come
RADIOIMMUNONANOPARTICLES FOR CANCER IMAGING AND THERAPY
gel electrophoresis, cellulose acetate electrophoresis, size-exclusion chromatography, and tumor cell binding. Pharmacokinetics and whole-body autoradiography studies demonstrated that a 5% injected dose was targeted in tumor after 24 h.
31.10 RADIONANOPARTICLES IN CANCER IMAGING AND DIAGNOSIS Rossin et al. [151] developed a RINP of 64 Cu anti-ICAM MAb-NPs (latex beads) for PET imaging targeted to ICAM-1 for lung imaging. NPs targeting intercellular adhesion molecule (ICAM-1) hold promise as a means of delivering therapeutics to the pulmonary endothelium in patients with acute and chronic respiratory diseases. The study results showed that lungs of mice administered 64 Cu-anti-ICAM MAb-NPs were clearly imaged by small-animal PET at 1, 4, and 24 h after administration. Both biodistribution and smallanimal imaging showed a three- to fourfold higher uptake in the lungs of mice injected with ICAM-targeted NPs relative to that of the control group. Furthermore, increased contrast in the lungs was achieved when upregulating ICAM-1 with lipopolysaccharides administration. Although both in vivo metabolism of 64 Cu-DOTA conjugates and release of prototype nanocarriers from endothelial cells led to a signal decrease with time, the lungs of mice injected with radiolabeled anti-ICAM NPs were clearly imaged by microPET up to 24 h after injection. This imaging approach holds promise as a new drug delivery agent for respiratory diseases. 31.10.1 Disadvantages Antibodies and targeting proteins linked to NPs sometimes undergo aggregation, causing the NP size to increment. NPs used as carriers with very large size after modification may have disadvantages as shown, for example, by one study using liposome NPs [103, 152]. Additionally, a few other problems were also reported, such as unexpected infusionrelated adverse reaction, high accumulation in normal tissues, and complex manufacturing procedures [153]. To address these problems, continuing research needs to be focused on the following areas: nanomaterial design, nanomaterial biology, and improved labeling methodologies. 31.10.2 Toxicity Zhu and colleagues [154] studied the fate of iron oxide nanoparticles having a size of 22 nm; radioactive isotope 59 Fe-labeled ferric oxide nanoparticles were intratracheally instilled into male Sprague–Dawley rats at a dose of 4 mg/rat. Extrapulmonary distribution of 59 Fe2 O3 in organs and its metabolism in lung, blood, urine, and feces were measured after 50 days exposure. Phagocytosis and clearance of agglomerated nano-59 Fe2 O3 by monocytes/ macrophages were observed by histopathology and inductively coupled plasma-mass spectrometry examination. Study results indicated intratracheal-instilled nano-59 Fe2 O3 could pass through the alveolar–capillary barrier into systemic circulation within 10 min, consistent with the one-compartment kinetic model. The nano-59 Fe2 O3 in the lung was distributed to organs rich in mononuclear phagocytes, including liver, spleen, kidney, and testicle. The plasma elimination half-life of nano-59 Fe2 O3 was 22.8 days and the lung clearance rate was 3.06 g/day, indicating that systemic accumulation and lung retention had occurred. The deposited nano-59 Fe2 O3 in interstitial lung was probably contributed by the particles
P1: OTA/XYZ P2: ABC c31 JWBS037-Chen February 2, 2011
21:53
Printer Name: Yet to Come
REFERENCES
775
escaping alveolar macrophage phagocytosis and macrophage clearance function overloading. Furthermore, results suggested that the effect of Fe2 O3 nanoparticle exposure, even at low concentration, should be assessed because of the potential lung and systemic cumulative toxicity of the nanoparticles.
31.11 CONCLUSION Nanotechnology in the biomedical field provides versatile opportunities to develop safer and more effective diagnostic and therapeutic modalities for cancer therapy. The medical use of nanoplatforms for thermotherapy, drug delivery, noninvasive cancer diagnosis, and imaging has already been demonstrated by various researchers. Multifunctional nanoparticles have been increasingly developed by conjugating with various cancer-specific ligands and nuclear or optical imaging agents to deliver multiple anticancer drugs to a tumor region and also simultaneously track and image critical biological pathways that are involved in cancer proliferation. A combination of immunotherapy could be used with two or multiple imaging modalities and nanotechnology could provide powerful diagnostic applications in cancer therapy and response management. For example, detection of tumor biomarkers at very low concentration with very high sensitivity has been achieved using antibodycoated magnetonanoparticles; the concentration levels are far smaller than those detectable using conventional imaging and diagnostic technologies. Radioimmunonanoparticles are expected to increase the efficacy of therapeutic agents while imaging and tracking the particles in vivo. Thus multifunctional nanoplatforms play an increasingly important role in the fields of cancer diagnosis, imaging, and therapy, and we expect that this technology will continue to grow over the next few decades. Although many nanoplatforms are still at the initial stage of development, these products need to follow rigorous safety regulations and to improve process and scale-up manufacturing to generate safe products for human clinical trials. Therefore development of suitable RINPs or multifunctional NPs with optimal characteristics remains obscure. For example, use of NPs for therapeutic applications requires suitable design and methods to overcome the nonspecific uptake by phagocytic cells and by nontargeted cells. Similarly, targeting cell surface markers could be another major challenge, with regard to solid tumors, and tumor vasculature. In spite of these challenges, rapid progress has been made in nanotechnology, especially in multifunctional NP-based approaches for multimodality diagnostics, imaging, and therapy for cancer. The application of these nanoplatforms is progressing rapidly but has a long way to go.
REFERENCES 1. Liong, M.; Lu, J.; Kovochich, M.; Xia, T.; Ruehm, S. G.; Nel, A. E.; Tamanoi, F.; Zink, J. I. ACS Nano 2008, 2, 889–896. 2. Kim, J.; Piao, Y.; Hyeon, T. Chem. Soc. Rev. 2009, 38, 372–390. 3. Goldberg, M.; Langer, R.; Jia, X. J. Biomater. Sci. Polym. Ed. 2007, 18, 241–268. 4. Kim, K. Y. Nanomedicine 2007, 3, 103–110. 5. Kattan, J.; Droz, J.-P.; Couvreur, P.; Marino, J.-P.; Boutan-Laroze, A.; Rougier, P.; Brault, P.; Vranckx, H.; Grognet, J.-M.; Morge, X.; Sancho-Garnier, H. Investigational New Drugs 1992, 10, 191–199.
P1: OTA/XYZ P2: ABC c31 JWBS037-Chen February 2, 2011
776
21:53
Printer Name: Yet to Come
RADIOIMMUNONANOPARTICLES FOR CANCER IMAGING AND THERAPY
6. Alexiou, C.; Arnold, W.; Klein, R. J.; Parak, F. G.; Hulin, P.; Bergemann, C.; Erhardt, W.; Wagenpfeil, S.; Lubbe, A. S. Cancer Res. 2000, 60, 6641–6648. 7. Jordan, A.; Scholz, R.; Wust, P.; F¨ahling, H.; Roland, F. J. Magn. Magn. Mater. 1999, 201, 413–419. 8. Huh, Y. M.; Jun, Y. W.; Song, H. T.; Kim, S.; Choi, J. S.; Lee, J. H.; Yoon, S.; Kim, K. S.; Shin, J. S.; Suh, J. S.; Cheon, J. J. Am. Chem. Soc. 2005, 127, 12387–12391. 9. Enochs, W. S.; Harsh, G.; Hochberg, F.; Weissleder, R. J. Magn. Reson. Imaging 1999, 9, 228–232. 10. Keller, T. M.; Michel, S. C.; Frohlich, J.; Fink, D.; Caduff, R.; Marincek, B.; Kubik-Huch, R. A. Eur. Radiol. 2004, 14, 937–944. 11. Harris, J. M.; Chess, R. B. Nat. Rev. Drug. Discov. 2003, 2, 214–221. 12. Kommareddy, S.; Tiwari, S. B.; Amiji, M. M. Technol. Cancer Res. Treat. 2005, 4, 615–625. 13. Goldenberg, M. M. Clin. Ther. 1999, 21, 309–318. 14. Leget, G. A.; Czuczman, M. S. Curr. Opin. Oncol. 1998, 10, 548–551. 15. Arbab, A. S.; Wilson, L. B.; Ashari, P.; Jordan, E. K.; Lewis, B. K.; Frank, J. A. NMR. Biomed. 2005, 18, 383–389. 16. Hoehn, M.; Wiedermann, D.; Justicia, C.; Ramos-Cabrer, P.; Kruttwig, K.; Farr, T.; Himmelreich, U. J. Physiol. 2007, 584, 25–30. 17. Petry, K. G.; Boiziau, C.; Dousset, V.; Brochet, B. Neurotherapeutics 2007, 4, 434–442. 18. Kairemo, K.; Erba, P.; Bergstrom, K.; Pauwels, E. K. Curr. Radiopharm. 2008, 1, 30–36. 19. DeNardo, S. J.; DeNardo, G. L.; Natarajan, A.; Miers, L. A.; Foreman, A. R.; Gruettner, C.; Adamson, G. N.; Ivkov, R. J. Nucl. Med. 2007, 48, 437–444. 20. Natarajan, A.; Gruettner, C.; Ivkov, R.; DeNardo, G. L.; Mirick, G.; Yuan, A.; Foreman, A.; DeNardo, S. J. Bioconjug. Chem. 2008, 19, 1211–1218. 21. DeNardo, S. J.; DeNardo, G. L.; Miers, L. A.; Natarajan, A.; Foreman, A. R.; Gruettner, C.; Adamson, G. N.; Ivkov, R. Clin. Cancer Res. 2005, 11, 7087s–7092s. 22. Berry, C. C.; Curtis, A. S. G. J. Phys. D: Appl. Phys. 2003, 36, R198–R206. 23. Lok, C. Nature 2001, 412, 372–374. 24. Kanematsu, M.; Kondo, H.; Goshima, S.; Kato, H.; Tsuge, U.; Hirose, Y.; Kim, M. J.; Moriyama, N. Eur. J. Radiol. 2006, 58, 217–228. 25. Stadnik, T.; Everaert, H.; Makkat, S.; Sacr´e, R.; Lamote, J.; Bourgain, C. Eur. Radiol. 2006, 16, 2153–2160. 26. Harisinghani, M. G.; Saini, S.; Slater, G. J.; Schnall, M. D.; Rifkin, M. D. J. Magn. Reson. Imaging 1997, 7, 161–163. 27. Anzai, Y.; Prince, M. R. J. Magn. Reson. Imaging 1997, 7, 75–81. 28. Anzai, Y.; McLachlan, S.; Morris, M.; Saxton, R.; Lufkin, R. B. AJNR. Am. J. Neuroradiol. 1994, 15, 87–94. 29. Fahmy, T. M.; Fong, P. M.; Park, J.; Constable, T.; Saltzman, W. M. AAPS J. 2007, 9, E171–F180. 30. Torchilin, V. P. AAPS J. 2007, 9, E128.-E147. 31. Torchilin, V.; Weissing, V., Eds. Liposomes: A Practical Approach 2nd ed. Oxford University Press: New York, 2003. 32. Bharali, D. J.; Khalil, M.; Gurbuz, M.; Simone, T. M.; Mousa, S. A. Int. J. Nanomed. 2009, 4, 1–7. 33. Tomalia, D. A.; Dewald, J.; 4507466, U. S. p. U., Ed.; The Dow Chemical Corporation (Midland, MI): USA, 1985 March 26. 34. Tomalia, D. A.; Baker, H.; Dewald, J.; Hall, M.; Kallos, G.; Martin, S.; Roeck, J.; Ryder, J.; Smith, P. Polym. J. 1985, 17, 117–132.
P1: OTA/XYZ P2: ABC c31 JWBS037-Chen February 2, 2011
21:53
Printer Name: Yet to Come
REFERENCES
777
35. Tomalia, D. A.; Reyna, L. A.; Svenson, S. Biochem. Soc. Trans. 2007, 035, 61–67. 36. Gillies, E. R.; Frechet, J. M. Drug Discov. Today 2005, 10, 35–43. 37. Kam, N. W.; O’Connell, M.; Wisdom, J. A.; Dai, H. Proc. Natl. Acad. Sci. U.S.A. 2005, 102, 11600–11605. 38. Bianco, A.; Kostarelos, K.; Partidos, C. D.; Prato, M. Chem. Commun. 2005, 571–577. 39. Cherukuri, P.; Bachilo, S. M.; Litovsky, S. H.; Weisman, R. B. J. Am. Chem. Soc. 2004, 126, 15638–15639. 40. K¨onig, K. J. Microsc. 2000, 200, 83–104. 41. O’Connell, M. J.; Bachilo, S. M.; Huffman, C. B.; Moore, V. C.; Strano, M. S.; Haroz, E. H.; Rialon, K. L.; Boul, P. J.; Noon, W. H.; Kittrell, C.; Ma, J.; Hauge, R. H.; Weisman, R. B.; Smalley, R. E. Science 2002, 297, 593–596. 42. Lu, Y.; Sega, E.; Leamon, C. P.; Low, P. S. Adv. Drug Deliv. Rev. 2004, 56, 1161–1176. 43. Da-Peng Yang, D.-X. C. Chemistry–An Asian Journal 2008, 3, 2010–2022. 44. Huang, X.; Jain, P. K.; El-Sayed, I. H.; El-Sayed, M. A. Nanomedicine 2007, 2, 681–693. 45. Huff, T. B.; Tong, L.; Zhao, Y.; Hansen, M. N.; Cheng, J. X.; Wei, A. Nanomedicine 2007, 2, 125–132. 46. Leyu Wang, J. B., Yujing Li, Yu. Huang, Angew. Chem. 2008, 120, 2473–2476. 47. Liu, X.; Dai, Q.; Austin, L.; Coutts, J.; Knowles, G.; Zou, J.; Chen, H.; Huo, Q. J. Am. Chem. Soc. 2008, 130, 2780–2782. 48. Yu, C.; Irudayaraj, J. Anal. Chem. 2007, 79, 572–579. 49. Chen, C. D.; Cheng, S. F.; Chau, L. K.; Wang, C. R. Biosens. Bioelectron. 2007, 22, 926–932. 50. Durr, N. J.; Larson, T.; Smith, D. K.; Korgel, B. A.; Sokolov, K.; Ben-Yakar, A. Nano Lett. 2007, 7, 941–945. 51. Imura, K.; Nagahara, T.; Okamoto, H. J. Phys. Chem. B. 2005, 109, 13214–13220. 52. Imura, K.; Nagahara, T.; Okamoto, H. J. Chem. Phys. 2005, 122, 154701. 53. Chen, C. C.; Lin, Y. P.; Wang, C. W.; Tzeng, H. C.; Wu, C. H.; Chen, Y. C.; Chen, C. P.; Chen, L. C.; Wu, Y. C. J. Am. Chem. Soc. 2006, 128, 3709–3715. 54. Salem, A. K.; Searson, P. C.; Leong, K. W. Nat. Mater. 2003, 2, 668–671. 55. Ghosh, P.; Han, G.; De, M.; Kim, C. K.; Rotello, V. M. Adv. Drug Deliv. Rev. 2008, 60, 1307–1315. 56. Andrea, R.; Tao, S. H.; Peidong, Y. Small 2008, 4, 310–325. 57. Zhao, N.; Wei, Y.; Sun, N.; Chen, Q.; Bai, J.; Zhou, L.; Qin, Y.; Li, M.; Qi, L. Langmuir 2008, 24, 991–998. 58. Smetana, A. B.; Wang, J. S.; Boeckl, J.; Brown, G. J.; Wai, C. M. Langmuir 2007, 23, 10429–10432. 59. J. Chen, B. W; Z.-Y. Li; Campbell, D.; Saeki, F.; Cang, H.; Au, L.; Lee, J.; Li, X.; Xia, Y. Adv. Mater. 2005, 17, 2255–2261. 60. Skrabalak, S. E.; Au, L.; Li, X.; Xia, Y. Nat. Protoc. 2007, 2, 2182–2190. 61. Millstone, J. E.; Mirkin, C. A. Adv. Funct. Mater. 2006, 16, 1209–1214. 62. Hao, F.; Nehl, C. L.; Hafner, J. H.; Nordlander, P. Nano Lett. 2007, 7, 729–732. 63. Kumar, P. S.; Pastoriza-Santos, I.; Rodriguez-Gonzalez, B.; Abajo, F. J. G. d.; Liz-Marzan, L. M. Nanotechnology 2008, 19, 015606. 64. Gole, A.; Murphy, C. J. Chem. Mater. 2004, 16, 3633–3640. 65. Jaiswal, J. K.; Mattoussi, H.; Mauro, J. M.; Simon, S. M. Nat. Biotechnol. 2003, 21, 47–51. 66. Cai, W.; Shin, D. W.; Chen, K.; Gheysens, O.; Cao, Q.; Wang, S. X.; Gambhir, S. S.; Chen, X. Nano Lett. 2006, 6, 669–676.
P1: OTA/XYZ P2: ABC c31 JWBS037-Chen February 2, 2011
778
21:53
Printer Name: Yet to Come
RADIOIMMUNONANOPARTICLES FOR CANCER IMAGING AND THERAPY
67. Kim, S.; Lim, Y. T.; Soltesz, E. G.; De Grand, A. M.; Lee, J.; Nakayama, A.; Parker, J. A.; Mihaljevic, T.; Laurence, R. G.; Dor, D. M.; Cohn, L. H.; Bawendi, M. G.; Frangioni, J. V. Nat. Biotechnol. 2004, 22, 93–97. 68. Sofou, S. Int. J. Nanomed. 2008, 3, 181–199. 69. Dearling, J. L.; Pedley, R. B. Clin. Oncol. (R. Coll. Radiol.) 2007, 19, 457–469. 70. Kassis, A. I.; Adelstein, S. J. J. Nucl. Med. 2005, 46 (Suppl 1), 4S–12S. 71. Kim, G. J.; Nie, S. Mater. Today 2005, 8, 28–33. 72. Mortimer, P. S. In Rook’s Textbook of Dermatology, 7th ed. Burns, T., Ed. Wiley-Blackwell: Malden, MA, 2008, pp. 2495–2522. 73. Brigger, I.; Dubernet, C.; Couvreur, P. Adv. Drug Deliv. Rev. 2002, 54, 631–651. 74. Muthu, M. S.; Singh, S. Nanomedicine 2009, 4, 105. 75. Duncan, R. Nat. Rev. Drug. Discov. 2003, 2, 347–360. 76. Jain, R. K. Annu. Rev. Biomed. Eng. 1999, 1, 241–263. 77. Jain, R. K. J. Control. Release 2001, 74, 7–25. 78. Matsumura, Y.; Maeda, H. Cancer Res. 1986, 46, 6387–6392. 79. Jain, R. K. Clin. Cancer Res. 1999, 5, 1605–1606. 80. Park, E. K.; Lee, S. B.; Lee, Y. M. Biomaterials 2005, 26, 1053–1061. 81. Davis, F. F. Adv. Drug Deliv. Rev. 2002, 54, 457–458. 82. Moghimi, S. M.; Hunter, A. C. Trends Biotechnol. 2000, 18, 412–420. 83. Arruebo, M.; Fern´andez-Pacheco, R.; Ibarra, M. R.; Santamar´ıa, J. Nano Today 2007, 2, 22–32. 84. Peer, D.; Karp, J. M.; Hong, S.; Farokhzad, O. C.; Margalit, R.; Langer, R. Nat. Nano 2007, 2, 751–760. 85. Chari, R. V. Adv. Drug. Deliv. Rev. 1998, 31, 89–104. 86. Mansour, A. M.; Drevs, J.; Esser, N.; Hamada, F. M.; Badary, O. A.; Unger, C.; Fichtner, I.; Kratz, F. Cancer Res. 2003, 63, 4062–4066. 87. Guo, X.; Szoka, F. C., Jr. Acc. Chem. Res. 2003, 36, 335–341. 88. Jankovic, D.; Maksin, T.; Djokic, D.; Milonjic, S.; Nikolic, N.; Mirkovic, M.; Vranjes-Djuric, S. J. Microsc. 2008, 232, 601–604. 89. Chao-Ming, F.; Yuh-Feng, W.; Yu-Chiang, C.; Shih-Hung, H.; Ming-Da, Y. IEEE. Trans. Magnetics 2004, 40, 3003–3005. 90. H¨afeli, U. O.; Pauer, G. J. J. Magn. Magn. Mater. 1999, 194, 76–82. 91. H¨afeli, U. O.; Sweeney, S. M.; Beresford, B. A.; Humm, J. L.; Macklis, R. M. Nucl.. Med. Biol. 1995, 22, 147–155. 92. Haik, Y.; Pai, V.; Chen, C.-J. J. Magn. Magn. Mater. 1999, 194, 254–261. 93. Sieben, S.; Bergemann, C.; L¨ubbe, A.; Brockmann, B.; Rescheleit, D. J. Magn. Magn. Mater. 2001, 225, 175–179. 94. Chao-Ming, F.; Yuh-Feng, W.; Yu-Feng, G.; Tang-Yi, L.; Jainn-Shiun, C. IEEE Trans. Magnetics 2005, 41, 4120–4122. 95. Gref, R. Synthesis, Functionalization and Surface Treatment of Nanoparticles. Baraton, M.-I., Ed., American Scientific Publishers: Valencia, CA, 2003. 96. Devaraj, N. K.; Keliher, E. J.; Thurber, G. M.; Nahrendorf, M.; Weissleder, R. Bioconjug. Chem. 2009, 20, 397–401. 97. Jarrett, B. R.; Gustafsson, B.; Kukis, D. L.; Louie, A. Y. Bioconjug. Chem. 2008, 19, 1496–1504. 98. Goins, B. A. Expert Opin. Med. Diagn. 2008, 2, 853–873. 99. Lucignani, G. Eur. J. Nucl. Med. Mol. Imaging 2006, 33, 849–851. 100. Ascencio, J. A.; Rincon, A. C.; Canizal, G. J. Phys. Chem. B. 2005, 109, 8806–8812.
P1: OTA/XYZ P2: ABC c31 JWBS037-Chen February 2, 2011
21:53
Printer Name: Yet to Come
REFERENCES
779
101. Cheng, J.; Teply, B. A.; Sherifi, I.; Sung, J.; Luther, G.; Gu, F. X.; Levy-Nissenbaum, E.; Radovic-Moreno, A. F.; Langer, R.; Farokhzad, O. C. Biomaterials 2007, 28, 869–876. 102. Fang, C.; Shi, B.; Pei, Y. Y.; Hong, M. H.; Wu, J.; Chen, H. Z. Eur. J. Pharm. Sci. 2006, 27, 27–36. 103. Klibanov, A. L.; Maruyama, K.; Beckerleg, A. M.; Torchilin, V. P.; Huang, L. Biochim. Biophys. Acta. 1991, 1062, 142–148. 104. Owens, D. E., 3rd; Peppas, N. A. Int. J. Pharm. 2006, 307, 93–102. 105. Sanderson, C. J.; Taylor, G. A. Immunology 1976, 30, 117–121. 106. Wong, D. M.; Varesio, L. Methods Enzymol. 1984, 108, 307–313. 107. Lasbury, M. E.; Durant, P. J.; Ray, C. A.; Tschang, D.; Schwendener, R.; Lee, C. H. J. Immunol. 2006, 176, 6443–6453. 108. Van Rooijen, N. J. Immunol. Methods 1989, 124, 1–6. 109. Bischof Delaloye, A. Semin. Nucl. Med. 2000, 30, 186–194. 110. Chen, J.; Wu, H.; Han, D.; Xie, C. Cancer Lett. 2006, 231, 169–175. 111. Date, A. A.; Joshi, M. D.; Patravale, V. B. Adv. Drug. Deliv. Rev. 2007, 59, 505–521. 112. Press, O. W.; Rasey, J. Semin. Oncol. 2000, 27, 62–73. 113. Lin, M. Z.; Teitell, M. A.; Schiller, G. J. Clin. Cancer Res. 2005, 11, 129–138. 114. Shinkai, M.; Le, B.; Honda, H.; Yoshikawa, K.; Shimizu, K.; Saga, S.; Wakabayashi, T.; Yoshida, J.; Kobayashi, T. Jpn. J. Cancer Res. 2001, 92, 1138–1145. 115. Sun, C.; Sze, R.; Zhang, M. J. Biomed. Mater. Res. A 2006, 78, 550–557. 116. Moghimi, S. M.; Hunter, A. C.; Murray, J. C. Pharmacol. Rev. 2001, 53, 283–318. 117. Sudimack, J.; Lee, R. J. Adv. Drug. Deliv. Rev. 2000, 41, 147–162. 118. Kennel, S. J.; Woodward, J. D.; Rondinone, A. J.; Wall, J.; Huang, Y.; Mirzadeh, S. Nucl. Med. Biol. 2008, 35, 501–514. 119. Kirpotin, D. B.; Drummond, D. C.; Shao, Y.; Shalaby, M. R.; Hong, K.; Nielsen, U. B.; Marks, J. D.; Benz, C. C.; Park, J. W. Cancer Res. 2006, 66, 6732–6740. 120. Medarova, Z.; Pham, W.; Farrar, C.; Petkova, V.; Moore, A. Nat. Med. 2007, 13, 372–377. 121. Rossin, R.; Pan, D.; Qi, K.; Turner, J. L.; Sun, X.; Wooley, K. L.; Welch, M. J. J. Nucl. Med. 2005, 46, 1210–1218. 122. Bibby, D. C.; Talmadge, J. E.; Dalal, M. K.; Kurz, S. G.; Chytil, K. M.; Barry, S. E.; Shand, D. G.; Steiert, M. Int. J. Pharm. 2005, 293, 281–290. 123. Cai, W.; Chen, K.; Li, Z. B.; Gambhir, S. S.; Chen, X. J. Nucl. Med. 2007, 48, 1862–1870. 124. Hu, G.; Lijowski, M.; Zhang, H.; Partlow, K. C.; Caruthers, S. D.; Kiefer, G.; Gulyas, G.; Athey, P.; Scott, M. J.; Wickline, S. A.; Lanza, G. M. Int. J. Cancer 2007, 120, 1951–1957. 125. Smith, B. R.; Heverhagen, J.; Knopp, M.; Schmalbrock, P.; Shapiro, J.; Shiomi, M.; Moldovan, N. I.; Ferrari, M.; Lee, S. C. Biomed. Microdevices 2007, 9, 719–727. 126. McDevitt, M. R.; Chattopadhyay, D.; Kappel, B. J.; Jaggi, J. S.; Schiffman, S. R.; Antczak, C.; Njardarson, J. T.; Brentjens, R.; Scheinberg, D. A. J. Nucl. Med. 2007, 48, 1180–1189. 127. Schipper, M. L.; Cheng, Z.; Lee, S. W.; Bentolila, L. A.; Iyer, G.; Rao, J.; Chen, X.; Wu, A. M.; Weiss, S.; Gambhir, S. S. J. Nucl. Med. 2007, 48, 1511–1518. 128. Kobayashi, H.; Wu, C.; Kim, M. K.; Paik, C. H.; Carrasquillo, J. A.; Brechbiel, M. W. Bioconjug. Chem. 1999, 10, 103–111. 129. Kobayashi, H.; Sato, N.; Saga, T.; Nakamoto, Y.; Ishimori, T.; Toyama, S.; Togashi, K.; Konishi, J.; Brechbiel, M. W. Eur. J. Nucl. Med. Mol. Imaging 2000, 27, 1334–1339. 130. Wangler, C.; Moldenhauer, G.; Eisenhut, M.; Haberkorn, U.; Mier, W. Bioconjug. Chem. 2008, 19, 813–820.
P1: OTA/XYZ P2: ABC c31 JWBS037-Chen February 2, 2011
780
21:53
Printer Name: Yet to Come
RADIOIMMUNONANOPARTICLES FOR CANCER IMAGING AND THERAPY
131. Sato, N.; Kobayashi, H.; Saga, T.; Nakamoto, Y.; Ishimori, T.; Togashi, K.; Fujibayashi, Y.; Konishi, J.; Brechbiel, M. W. Clin. Cancer Res. 2001, 7, 3606–3612. 132. Mamede, M.; Saga, T.; Ishimori, T.; Higashi, T.; Sato, N.; Kobayashi, H.; Brechbiel, M. W.; Konishi, J. J. Control Release 2004, 95, 133–141. 133. Wu, G.; Barth, R. F.; Yang, W.; Lee, R.; Tjarks, W.; Backer, M. V.; Backer, J. M. Nanotechnol. Life Sci. 2007, 6, 33. 134. Yang, W.; Barth, R. F.; Wu, G.; Kawabata, S.; Sferra, T. J.; Bandyopadhyaya, A. K.; Tjarks, W.; Ferketich, A. K.; Moeschberger, M. L.; Binns, P. J.; Riley, K. J.; Coderre, J. A.; Ciesielski, M. J.; Fenstermaker, R. A.; Wikstrand, C. J. Clin. Cancer Res. 2006, 12, 3792–3802. 135. Liang, S.; Wang, Y.; Yu, J.; Zhang, C.; Xia, J.; Yin, D. J. Mater. Sci. Mater. Med. 2007, 18, 2297–2302. 136. H¨afeli, U. O.; Roberts, W. K.; Pauer, G. J.; Kraeft, S.-K.; Macklis, R. M. Appl. Radiat. Isotopes 2001, 54, 869–879. 137. Hamoudeh, M.; Fessi, H.; Salim, H.; Barbos, D. Drug. Dev. Ind. Pharm. 2008, 34, 796–806. 138. Hamoudeh, M.; Fessi, H.; Mehier, H.; Faraj, A. A.; Canet-Soulas, E. Int. J. Pharm. 2008, 348, 125–136. 139. Jestin, E.; Mougin-Degraef, M.; Faivre-Chauvet, A.; Remaud-Le Saec, P.; Hindre, F.; Benoit, J. P.; Chatal, J. F.; Barbet, J.; Gestin, J. F. Q. J. Nucl. Med. Mol. Imaging 2007, 51, 51–60. 140. Kim, C.-O.; Kim, J.-H.; Huang, Y.-Q.; Park, S.-I. US2005019257A1 2000. 141. Esenaliev, R. O. The Board of Regents of the University of Texas System USA, 2000; Vol. WO/2000/002590. 142. Praetorius, N. P.; Mandal, T. K. Recent. Pat. Drug. Deliv. Formul. 2007, 1, 37–51. 143. Li, L.; Wartchow, C. A.; Danthi, S. N.; Shen, Z.; Dechene, N.; Pease, J.; Choi, H. S.; Doede, T.; Chu, P.; Ning, S.; Lee, D. Y.; Bednarski, M. D.; Knox, S. J. Int. J. Radiat. Oncol. Biol. Phys. 2004, 58, 1215–1227. 144. Lee, H. Y.; Li, Z.; Chen, K.; Hsu, A. R.; Xu, C.; Xie, J.; Sun, S.; Chen, X. J. Nucl. Med. 2008, 49, 1371–1379. 145. Blanchard, R. J.; Lafave, J. W.; Kim, Y. S.; Frye, C. S.; Ritchie, W. P.; Perry, J. F., Jr. Cancer 1965, 18, 375–380. 146. Houle, S.; Yip, T. K.; Shepherd, F. A.; Rotstein, L. E.; Sniderman, K. W.; Theis, E.; Cawthorn, R. H.; Richmond-Cox, K. Radiology 1989, 172, 857–860. 147. Mantravadi, R. V.; Spigos, D. G.; Tan, W. S.; Felix, E. L. Radiology 1982, 142, 783–786. 148. Liang, S.; Wang, Y.; Zhang, C.; Liu, X. J. Radioanal. Nucl. Chem. 2006, 269, 3–7. 149. Chao-Ming, F.; Yuh-Feng, W.; Tang-Yi, L.; Yu-Feng, G.; Fu-Wen, L. IEEE. xplore. 2005, 1. 150. Natarajan, A.; Xiong, C. Y.; Gruettner, C.; DeNardo, G. L.; DeNardo, S. J. Cancer Biother. Radiopharm. 2008, 23, 82–91. 151. Rossin, R.; Muro, S.; Welch, M. J.; Muzykantov, V. R.; Schuster, D. P. J. Nucl. Med. 2008, 49, 103–111. 152. Hughes, B. J.; Kennel, S.; Lee, R.; Huang, L. Cancer Res. 1989, 49, 6214–6220. 153. Manninger, S. P.; Muldoon, L. L.; Nesbit, G.; Murillo, T.; Jacobs, P. M.; Neuwelt, E. A. AJNR. Am. J. Neuroradiol. 2005, 26, 2290–2300. 154. Zhu, M. T.; Feng, W. Y.; Wang, Y.; Wang, B.; Wang, M.; Ouyang, H.; Zhao, Y. L.; Chai, Z. F. Toxicol. Sci. 2009, 107, 342–351.
P1: OTA/XYZ P2: ABC c32 JWBS037-Chen January 25, 2011
13:56
Printer Name: Yet to Come
PART IV
TRANSLATIONAL NANOMEDICINE
P1: OTA/XYZ P2: ABC c32 JWBS037-Chen January 25, 2011
13:56
Printer Name: Yet to Come
CHAPTER 32
Current Status and Future Prospects for Nanoparticle-Based Technology in Human Medicine ´ ´ NURIA SANVICENS, FATIMA FERNANDEZ, J.-PABLO SALVADOR, and M.-PILAR MARCO Applied Molecular Receptors Group, CIBER de Bioingenier´ıa, Biomateriales y Nanotecnolog´ıa, IQAC-CSIC, Barcelona, Spain
32.1 INTRODUCTION The nanoscale world—which ranges from 1 to 100 nm—is an emerging trend that has fascinated the scientific community. The potential and opportunities of “nano” things seem endless in all fields. Thus it is not a surprise that nanoscience has raised the enthusiasm of biologists, chemists, engineers, physicists, technologists, and many others. Medicine is not an exception. The clinical use of numerous materials and devices of nanometric dimensions demonstrates that nanobiotechnology—the convergence of engineering and biology—has already had an impact on health care [1]. Nanoparticles are attractive constructs with distinctive chemical and physical features associated with their being of nanoscale size. Box 32.1 and Table 32.1 summarize the characteristics of some of the nanoparticles that are currently used in human medicine. Nanoparticles, which can be designed to have unique properties based on their small size, large surface area, chemical composition, solubility, and geometry, might be used to overcome the limitations of existing technologies. In this manner, the application of nanoparticles to medicine holds great promise for revolutionizing the medical landscape. Already, nanoparticle-based therapeutics and diagnostic agents have been commercialized for the treatment of numerous pathologies (see Table 32.2). Likewise, products that are being evaluated in clinical trials promise new ways to diagnose and to monitor disease and to specifically deliver medical agents. In this chapter, we highlight some of the nanoparticle-based novel technologies for molecular imaging, diagnosis, and drug delivery formulations. Other nanodevices such as nanomachines [2], nanowires [3], cantilevers [4], and nanofibers [5] with applications to medicine are beyond the scope of this chapter. Limitations and future challenges of nanoparticle-based systems are also discussed. Special emphasis is given to the concern
Nanoplatform-Based Molecular Imaging Edited by Xiaoyuan Chen C 2011 John Wiley & Sons, Inc. Copyright
783
P1: OTA/XYZ P2: ABC c32 JWBS037-Chen January 25, 2011
784
13:56
Printer Name: Yet to Come
CURRENT STATUS AND FUTURE PROSPECTS FOR NANOPARTICLE
BOX 32.1 EXAMPLES OF NANOPARTICLES CURRENTLY USED IN HUMAN MEDICINE
Liposomes Liposomes are amphiphilic phospholipid vesicles (50–100 nm) with a bilayer membrane structure similar to that of biological membranes and an internal aqueous phase. Liposomes can be classified according to size and number of layers into multi-, oligo-, or unilamellar. They can also be classified in terms of composition and mechanism of intracellular delivery into conventional lyposomes, pH-sensitive delivery liposomes, cationic liposomes, immunoliposomes, and long-circulating liposomes. Liposomes’ amphiphilic nature allows them to transport hydrophilic drugs entrapped within their aqueous interior and hydrophobic drugs dissolved into the membrane. The liposome surface can be modified with ligands and/or polymers to increase drug delivery specificity [169]. Moreover, a new generation of liposomes called “stealth liposomes” has the ability to evade interception by the immune system, providing prolonged blood circulation times [170]. Dendrimers Dendrimers are highly branched synthetic polymers (10–100 nm) with layered architectures consisting of a central core, an internal region, and numerous terminal groups that determine dendrimer characteristics. A dendrimer can be prepared using multiple types of chemistry, the nature of which defines the dendrimer solubility and biological activity. The high number of functional groups on the dendrimer’s surface allows the presentation of multiple ligands on a single dendrimer molecule, which makes dendrimers excellent drug and imaging agent carriers [171]. Polymeric Micelles These are spherical colloidal nanoparticles formed by the self-assembly of amphiphilic copolymers (e.g., polylactic acid, poly(cyano)acrylates, polyethylene imine, polycaprolactone) in aqueous media. In this manner, polymeric micelles have a hydrophobic core stabilized by a hydrophilic corona. Their nature allows entrapping drugs or contrasting agents within the micelle hydrophobic interior and linking polar molecules to their surface. Drugs with intermediate polarity can be distributed in intermediate positions. Similar to liposomes, the micelle’s surface can be modified with ligand molecules for targeted delivery to specific cells. On the other hand, polymeric micelle size (>100 nm) offers benefits over the liposome such as avoiding RES clearance, and restricting activation of the human complement system and allowing passive targeting to cancerous tissues through the enhanced permeation and retention effect [172]. Polymeric Nanoparticles These are prepared by physical entrapment or covalent link of biomolecules or drugs to polymeric chains. Polysaccharide chitosan nanoparticles are an example of polymeric
P1: OTA/XYZ P2: ABC c32 JWBS037-Chen January 25, 2011
13:56
Printer Name: Yet to Come
INTRODUCTION
785
nanoparticles. Depending on their chemical composition, they can carry a wide variety of compounds, which makes polymeric nanoparticles ideal drug delivery systems. In addition, they can incorporate a metallic core, providing the particle with optical, magnetic, or hyperthermic properties. The use of absorbable or degradable polymers, such as polyesters, gives polymeric nanoparticles a high degree of biocompatibility [173]. Carbon Nanotubes These are carbon-based nanoparticles composed of coaxial graphite sheets (<100 nm) rolled up into cylinders. They belong to the family of fullerenes. They can be obtained as single- (one graphite sheet) or multiwalled nanotubes (several concentric graphite sheets). They exhibit excellent strength and electrical properties and are efficient heat conductors. Due to their metallic or semiconductor nature, nanotubes are used as biosensors. Carbon nanotubes can be made water soluble by surface functionalization. They are used as molecular biosensors, as drug carriers, and as tissue repair scaffolds [174]. Fullerenes are another example of carbon-based nanomaterials. Magnetic Nanoparticles These are spherical nanocrystals (10–20 nm) usually composed of an Fe2+ O3 and Fe3+ O4 core surrounded with an organic shell of surfactants or polymers, or with an inorganic layer such as silica or carbon. These nanoparticles posses large magnetic moments when brought into a magnetic field, which make them excellent as biomolecule labeling agents in bioassays and as magnetic resonance imaging contrast agents. They are also amenable to surface functionalization for active targeting in vivo or for in vitro diagnosis [175]. Gold Nanoparticles These are an example of metallic nanoparticles; others are Au, Ni, Pt, and TiO2 nanoparticles. Gold nanoparticles (<50 nm) are prepared with different geometries: gold nanospheres, nanoshells, nanorods, or nanocages. They show localized surface plasmon resonant properties: under light irradiation, the conduction electrons are driven by the associated electric field to a collective oscillation at a resonant frequency, absorbing light and emitting photons with the same frequency in all directions. Gold nanoparticles are excellent labels for biosensors as they can be detected by numerous techniques, such as optic absorption, fluorescence, and electric conductivity [176]. Quantum Dots These are fluorescent semiconductor nanocrystals (2–10 nm) made of a central core of a combination of elements from groups II–VI (CdSe, CdTe, CdS, PbSe, ZnS, and ZnSe) or III–V (GaAs, GaN, InP, and InAs) on the periodic table, overcoated with a layer of ZnS. Quantum dots have unique optical properties including narrow emission range, sizeand composition-tunable emission spectra, high quantum yield, low photobleaching, and resistance to chemical degradation [32].
P1: OTA/XYZ P2: ABC c32 JWBS037-Chen January 25, 2011
786
13:56
Printer Name: Yet to Come
CURRENT STATUS AND FUTURE PROSPECTS FOR NANOPARTICLE
TABLE 32.1 Nanoparticles: Applications in Human Medicine, Advantages and Disadvantages Nanoparticle
Applications
Advantages
Limitations
a
Drug and gene delivery Imaging agent carriers
Dendrimers
Drug and gene delivery Imaging agent carriers Intrinsic drug properties Tissue repair scaffolds
Clearance by RES Opsonization Instability Poor control over release of the drug Potential toxicity Control over the ratio of components is challenging High cost
Polymeric micelles
Systemic delivery of water-insoluble drugs Drug and gene delivery
Excellent circulation, penetration, and diffusion properties Biocompatible Biodegradable Good stability Control over biodistribution and pharmacokinetics by controlling dendrimer size Long-circulating life Biocompatible Passive targeting Core-dependent properties: optical, magnetic, hiperthermic Biocompatible Excellent strength Good heat conductors Electrical properties Capability to cross cell membranes Magnetic properties Specific targeting Afford excellent sensitivity
Liposomes
Polymeric nanoparticles
Carbon-based nanoparticles
Diagnosis: biosensors Drug and gene delivery Tissue repair scaffolds
Magnetic nanoparticles
Contrast agents for MR imaging Hyperthermia Tissue repair Diagnosis: bioassays Drug and gene delivery Imaging agents Diagnosis: bioassays Drug and gene delivery Phototermal therapy Biomolecule labeling In vivo imaging Bioassays
Metallic nanoparticles
Quantum dots
High stability Excellent properties for optical detection Unique size-controlled optical properties
In vivo stability and drug retention are challenging Rapid clearance
Potential toxicity
Poor penetration depth Inability to overcome biological barriers Clearance by RES Potential toxicity Instability Potential toxicity
Poor penetration depth Potential toxicity
a Apart from liposomes, other nanoparticles derived from natural materials include chitosan, dextrane, gelatine, starch, and alginated nanoparticles.
about the safety of nanoparticles for human health, which has not yet been addressed in a satisfactory manner. 32.2 NANOPARTICLE-BASED MOLECULAR IMAGING SYSTEMS Until a few decades ago, imaging diagnosis was limited to the description of anatomic structures. The development of techniques such as magnetic resonance imaging (MRI), computed tomography (CT), positron emission spectroscopy (PET) and near-infrared fluorescence has
P1: OTA/XYZ P2: ABC c32 JWBS037-Chen January 25, 2011
13:56
Printer Name: Yet to Come
NANOPARTICLE-BASED MOLECULAR IMAGING SYSTEMS
787
TABLE 32.2 Examples of Commercialized Nanoparticle-Based Products Nanoparticle Component Liposomes
Product ®
Doxil
®
Inflexal Epaxal
®
AmBiosome
®
Dendrimers
Vivagel
Polymeric micelles
Abelcet
Indication
Doxorubicin
Sequus Pharmaceutical (Santa Monica, CA, USA)
Antigens influenza virus Antigen hepatitis A virus Amphotericin B
Berna Biotech AG (Basel, Switzeland)
Ovarian and breast cancer, Kaposi sarcoma in AIDS Influenza vaccine
Amphotericin B
®
®
Lumiren Gastro® mark ® Endorem ®
Combidex
®
Dynal magnetic Magnetic beads nanoparticles
17 estradiol
Ferumoxisil
Ferumoxide
Gilead Science (Foster City, CA, USA) Starpharma (Melbourne, Australia) The Liposome Company (Princeton, NJ, USA) Novavax, Inc. (Rockville, MD, USA) AMAG Pharmaceuticals Inc. (Cambrige, MA, USA)
Furumoxtran-10
—
Dynal/Invitrogen (Oslo, Norway)
MACS Technology
—
SensationTM Technology
—
Miltenyi Biotec (Bergisch Gladbach, Germany) Nanomix (Emeryville, CA, USA)
®
Carbon nanotubes
Company
SPL7013 Gel
®
Estrasorb
SPIO
®
Drug
Hepatitis A vaccine Fungal infections
Vaginal microbicide Fungal infections
Treatment of postmenopausal symptoms Bowel contrast agent Liver/spleen imaging Lymph node metastases detection Clinical cell separation for immunodiagnostics
Respiration function monitoring (Continued)
P1: OTA/XYZ P2: ABC c32 JWBS037-Chen January 25, 2011
788
13:56
Printer Name: Yet to Come
CURRENT STATUS AND FUTURE PROSPECTS FOR NANOPARTICLE
TABLE 32.2 (Continued) Nanoparticle Component
Product
Drug ®
Gold Verigene system nanoparticles —
—
—
AuroLaseTM Therapy
®
Nanogold
®
—
—
Singlepath ® Duopath
—
Immunocap ® Rapid
—
Company
Indication
Nanosphere Inc. (Northbrook, IL, USA) British Biocell (Cardiff, UK) Amersham/GE (Little Chanfont, UK) Nymox (Hasbrouk Heights, NJ, USA) Nanospectra Biosciences Inc. (Houston, TX, USA) Nanoprobes Inc. (Yaphank, NY, USA) Merk Kgba (Darmstadt, Germany) Phadia (Uppsala, Sweden)
Detection of nuclei acid and protein targets Detection of pregnancy, ovulation, HIV
Thermal destruction of solid tumours Imaging
Lateral flow test for foodborne food detection Lateral flow test for allergy detection
facilitated the observation of inner spaces of the body by noninvasive means. Despite these technological advances and their impact on clinical medicine, none of the above-mentioned techniques has all the requirements for the accurate imaging diagnosis of numerous pathologies. MRI and CT are noninvasive techniques that allow three-dimensional (3D) imaging, but they lack target sensitivity. On the other hand, PET does have high target sensitivity but it shows very poor spatial resolution. Similarly, near-infrared fluorescence shows good sensitivity but low tissue penetration and poor 3D spatial resolution. To overcome these limitations, imaging techniques have been combined in what is known as multimodal imaging, which shows great potential for clinical and preclinical diagnosis. One example is the coregistration of in vivo fluorescence imaging with anatomical imaging modalities such as MRI [6] or the alignment of the whole-body anatomical (CT) and functional (PET) images [7]. Unfortunately, multimodal imaging does not provide information of biological systems at the molecular level. In this context, the incorporation of nanoparticles has provoked a revolution in imaging diagnosis, since nanoparticles permit imaging at the cellular level, which implies more precise diagnosis with high-quality images. Magnetic nanoparticles are the most employed nanomaterials in imaging diagnosis. They are noninvasive and show good application in multidimensional tomography and provide high spatial resolution. Moreover, magnetic nanoparticles afford better sensitivity than other contrast agents based on nanoparticles. The use of magnetic nanoparticles in MRI is well established. Indeed, they have already been commercialized for direct application in
P1: OTA/XYZ P2: ABC c32 JWBS037-Chen January 25, 2011
13:56
Printer Name: Yet to Come
NANOPARTICLE-BASED MOLECULAR IMAGING SYSTEMS
789
MRI [8, 9]. Iron oxide nanoparticles used as MR contrast agents are normally coated with silica or dextran, which prevents nanoparticle aggregation in aqueous media, an essential requirement for biomedical applications. AMI-121 [10], AMI-25 [11], AMI-227 [12], and SHU 55A [13] are examples of commercially available iron oxide nanoparticles coated with silica or dextran. Several reviews have described the use of magnetic nanoparticles for biomedical applications in cancer, apoptosis, or cardiovascular imaging [8, 9, 14–17]. For example, Lee and co-workers prepared magnetic ferrite nanoparticles, in which they incorporated an organic dye for fluorescence detection and mouse mesenchymal stem cells to facilitate nanoparticle direction to the myocardial infarction region. Using MRI, the authors successfully tracked the magnetic nanoparticles in the hearts of mice at day 7 after inducing myocardial infarction [18]. Turvey and collaborators used iron oxide nanoparticles to visualize the inflammatory lesions produced by diabetes in the pancreatic islets [19]. This study described for the first time the capability of magnetic nanoparticle-MRI to identify and quantify the vascular volume and the permeability changes associated with inflammation of the pancreas during the development of autoimmune diabetes in the mouse (see Fig. 32.1). Magnetic nanoparticles coupled to cancer cell-targeting antibodies have been used to diagnose cancer in vitro [20] and in vivo [21, 22] (see Fig. 32.2). Nevertheless, the main goal in noninvasive detection of cancer is to diagnose the disease at the earliest stage. Conventional MRI is able to detect cancer with a sensitivity of 1 cm3 [8]. Yet magnetic nanoparticles could afford better sensitivities due to their intrinsic properties that selectively Day 6 after CPA
Control
T2 (ms)
MRI with pancreatic T2 pseudocoloring
45.0 39.3 33.5 27.8
Representative islet histology
22.0
FIGURE 32.1 Noninvasive response of magnetic nanoparticles labeled with autoreactive T cells by MRI to reveal the severity of insulitis. Young female BDC2.5/NOD mice were imaged on day 6 after treatment with cyclophosphamide, which destabilizes the immunoregulatory balance and induces autoimmune diabetes in 100% of animals, and 24 h after injection of the magnetic nanoparticles. A pseudocolor was assigned to the pancreas, reflecting the T2 value of the organ. Photomicrographs show representative islet histology from these animals. Magnification, 20×. (Reproduced with permission from Turvey et al. [19].)
P1: OTA/XYZ P2: ABC c32 JWBS037-Chen January 25, 2011
790
13:56
Printer Name: Yet to Come
CURRENT STATUS AND FUTURE PROSPECTS FOR NANOPARTICLE
ScFvEGFR-10 Pre
L
Post 30 hr
Post 5 hr
Pancreatic tumor
R Pre
L
Post 5 hr
R
L
R
FIGURE 32.2 Single chain epidermal growth factor receptor antibody conjugated iron oxide (ScFvEGFR-IO) nanoparticles as a molecular-targeted in vivo tumor imaging agent. The target specificity of ScFvEGFR-IO nanoparticles was examined by MRI using an orthotopic human pancreatic xenograft model. Figure shows the MRI of a tumor-bearing mouse. ScFvEGFR-IO nanoparticles (8 nmol kg−1 body weight) were injected into the mouse through the tail vein. Pre- and post-contrast MRI at 5 and 30 h were collected. Upper and lower panels showed the MRI from different sectional levels of the same mouse. The areas of the pancreatic tumor were marked as a dash-lined circle (red). The pancreatic tumor area showed a bright signal before receiving the nanoparticle. After injection of the targeted IO nanoparticles, a marked MRI contrast decrease was detected in the tumor (darker), which delineated the area of the tumor lesion. MRI contrast change is also found in the liver (green arrow) and spleen. These MRIs are representative results of five mice that received ScFvEGFR-IO nanoparticles. Lower right is the picture of tumor and spleen tissues, showing sizes and locations of two intra-pancreatic tumor lesions (arrows) that correspond with the tumor images of MRI. (Reproduced with permission from Yang et al. [180].)
enhance the MR signal of cancer cells. One example is the work by Lee and collaborators, who developed innovative magnetism-engineered iron oxide nanoprobes to enhance MRI sensitivity for the detection of cancer markers [23]. In this same context, the application of USPIO (ultrasmall particle of iron oxide) to diagnose lymph node metastasis in patients with endometrial and cervical cancer enhanced the sensitivity of MRI [24]. Deregulation of apoptosis—programmed self-destruction of cells—has been well documented in several human pathologies, including cancer, neurodegenerative diseases, and AIDS [25]. Therefore noninvasive detection of apoptosis with magnetic nanoparticles could be a useful tool for monitoring disease progression and for assessing response to therapy. A good example is the work by Zhao and collaborators who employed SPIO nanoparticles to detect cells undergoing apoptosis in vivo in a tumor treated with chemotherapeutic drugs [26]. In this same context, van Tilborg and co-workers developed a bimodal contrast agent
P1: OTA/XYZ P2: ABC c32 JWBS037-Chen January 25, 2011
13:56
Printer Name: Yet to Come
791
NANOPARTICLE-BASED MOLECULAR IMAGING SYSTEMS
(a)
(b)
1.80E+10 1.60E+10
Before
2 min
60 min
120 min
Photo counts/s
1.40E+10 1.20E+10 1.00E+10 8.00E+09 6.00E+09 4.00E+09 2.00E+09 0.00E+00 0.033
1
2
3
4
8
Time (h)
180 min
240 min
480 min
480 min
FIGURE 32.3 Distribution and retention of wheat germ agglutinin-conjugated quantum dotsloaded nanoparticles (WGA-QDs-NP) in the brain following intranasal administration. (a) Representative optical images taken before, 2 min, and 1, 2, 3, 4, and 8 h after dose. (b) Quantification of the luminescence signal from WGA-QDs-NP in the brain over time following intranasal administration (n = 6 for the time points of 2 min and 1, 2, 3 h; n = 3 for the time points of 4 and 8 h). 176 × 85 mm (300 × 300 DPI). (Reproduced with permission from Gao et al. [33].)
based on iron oxide nanoparticles coupled to fluorescent lipids that enable the detection of apoptotic cells with both MRI and optical techniques [27]. Magnetic nanoparticles have also been used to visualize noninvasively transgene expression in vivo by MRI, which is of great importance for gene therapy monitoring [28]. As highlighted, simple magnetic nanoparticles function as MRI contrast enhancement probes. Besides, magnetic nanoparticles can be coupled to other functional moieties, affording multimodal imaging probes, which nowadays constitute the state-of-the-art in imaging diagnosis [29]. For instance, fluorescent tags can be linked to magnetic nanoparticles for dual-mode MRI–optical imaging [30] or radionuclide labels for dual-mode MRI–PET imaging [31]. The application of quantum dots as contrast agents in cellular and in vivo imaging has significantly increased in the last years due to their unique optical properties [32]. For instance, Gao et al. [33] prepared quantum dots coated with polyethylene glycol (PEG) and polylactic acid for brain imaging. In this study, quantum dots were coupled to wheat germ agglutinin to facilitate brain delivery (see Fig. 32.3). This transport of imaging agents to the brain is highly important for the diagnosis and treatment of central nervous system diseases. Similarly, the potential of near-infrared-emitting quantum dots for in vivo fluorescence imaging of lymph nodes has been demonstrated [34]. In vivo targeting and imaging of human prostate cancer growing in nude mice has been achieved by encapsulation of quantum dots with an amphiphilic triblock copolymer and coupling to tumor-targeting ligands [35]. Carbon-based nanoparticles have recently been applied to molecular imaging as well. For example, single-walled carbon nanotubes (SWNTs) coupled to a cyclic tripeptide with high affinity to the tumor neovasculature, overexpressing ␣v 3 integrin, were used as contrast agents for photoacoustic imaging of tumors in living mice [36]. Nanocarbonbased nanoparticles have also been combined with 111 In for PET imaging of the radioactive nanotube biodistribution and clearance in rats [37]. Altogether, these results highlight the
P1: OTA/XYZ P2: ABC c32 JWBS037-Chen January 25, 2011
792
13:56
Printer Name: Yet to Come
CURRENT STATUS AND FUTURE PROSPECTS FOR NANOPARTICLE
potential of nanoparticles to improve the resolution and the precision with which single events can be monitored. Moreover, nanoparticles will help to monitor responses to drug therapies, to track cell migration, and to detect early-stage diseases.
32.3 NANOPARTICLE-BASED MOLECULAR DIAGNOSTIC SYSTEMS Current molecular diagnostic methods do not always afford the rapid, selective, and single cell or molecule detection that is critical in diagnosing genetic diseases, infections, neurological disorders, metabolic diseases, autoimmune diseases, inflammations, and cancer or in discovering biomarkers. In the last years, the need for more accurate, selective, and sensitive targeting toward molecules or cells indicative of diseases has promoted the incorporation of nanoparticles in medical diagnostic systems. Such is the expectation created by nanoparticle-based molecular systems that even traditional and robust detection methods such as mass spectrometry have been combined with nanoparticles [38, 39]. 32.3.1 Nanoparticle-Based Optical Methods of Diagnosis Optical methods of diagnosis based on nanoparticles are not new. In the late 1980s, gold nanoparticles were incorporated in lateral flow test assays for the in vitro diagnosis of HIV, pregnancy, or ovulation. Gold nanoparticles afforded high stability to these tests and in this manner, false positives were avoided. In 1996, Mirkin and co-workers developed a simple method for DNA detection based on metallic nanoparticles. Gold nanoparticles modified with different oligonucleotide probes provoked a change in the solution color from red to blue in the presence of the target DNA due to the light-scattering properties of aggregated nanoparticles [40]. The sensitivity of this method in the low nanomolar range was later improved down to 500 pM using gold nanoparticles combined with latex microspheres [41]. Similarly, Storhoff and collaborators, taking advantage of the light-scattering properties of gold nanoparticles, detected DNA at concentrations of 33 fM [42]. This change in the localized plasmon of metallic nanoparticles has continued being exploited for numerous applications. Gold nanoparticles modified with peptide nucleic acids have been used in a colorimetric DNA assay to detect the presence of single base mismatches [43]. Variations in the plasmon of silver nanotriangles have been employed for the determination of the pathogenic Alzheimer disease marker amyloid beta-derived diffusible ligands [44]. Recently, light scattering of gold nanoparticles has been applied for the detection of human immunoglobulin in serum samples [45]. Similarly, the tunable plasmon resonant response of gold nanoshells has been used to detect sub-ng/mL concentrations of immunoglobulins in saline, serum, and whole blood [46]. In this same context, Niemeyer and Ceyhan developed a sandwich immunoassay using gold nanoparticles for protein quantification [47]. The bio-barcode assay combines the separation and concentration capability of magnetic particles with metallic nanoparticle-based optical detection (see Fig. 32.4). This system relies on the concentration afforded by magnetic particles conjugated to an antibody that specifically recognizes the target of interest and the detection provided by metallic nanoparticles encoded with DNA—barcodes—and antibodies that can sandwich the target captured by the magnetic particles. Using this method the prostate specific antigen—present in the sera of prostate cancer patients—was detected at a 30-attomolar concentration. Similarly, amyloid-beta-derived diffusible ligands were detected at femtomolar levels in cerebral spinal fluid [48]. Additional modifications of the bio-barcode assay by incorporating two
P1: OTA/XYZ P2: ABC c32 JWBS037-Chen January 25, 2011
13:56
Printer Name: Yet to Come
NANOPARTICLE-BASED MOLECULAR DIAGNOSTIC SYSTEMS
1
2
793
3
target DNA strands
NP probes
MMP probe
probe separation and DNA release from NP surface with DTT
barcode DNA detection scanometric detection with a universal probe
barcode DNA
magnetic field
FIGURE 32.4 The multiplexed biobarcode assay for scanometric DNA detection. Magnetic separation of the target is followed by the dehybridation of the oligonucleotides immobilized onto the metallic surface. Then, identification of the oligonucleotide sequence release from the metallic nanoparticles allows the determination of the presence of the target DNA. Signal amplification is achieved due to the large number of oligonucleotides onto the metallic nanoparticle. (Reproduced with permission from Stoeva et al. [181]).
nanoparticle–oligonucleotide conjugates led to the detection of oligonucleotide sequences associated with the anthrax lethal factor at a 500-zeptomolar (zM = 10−21 ) concentration [49]. Until now, most of the magnetic particles that have been employed in the field of diagnosis were in the micrometric dimensions. Recently, magnetic particles in the nanoscale range have been incorporated into optical detection methods. For instance, El-Boubbou and co-workers reported a glycomagnetic nanoparticle-based system, which detected by fluorescence microscopy three different Escherichia coli strains on the basis of the response patterns of two distinctive carbohydrates attached onto magnetic nanoparticles [50]. Magnetic/luminescent core–shell nanoparticles (core: Fe3 O4 /luminescent Eu:Gd2 O3 ) have been used to quantify the antibiotic resistance gene tetQ [51]. A novel application of magnetic/luminescent core–shell nanoparticles (core: Co:Nd:Fe2 O3 /luminescent Eu:Gd2 O3 ) as internal standard has been described in a recent developed platform for multiprotein analysis. In this work, magnetic properties of the nanoparticles allow sample manipulation, while their optical properties enable the internal calibration of the detection system [52]. In 1999, Mitchell and co-workers reported for the first time the immobilization of DNA onto quantum dot surfaces. This work opened the door to novel quantum dot–biomolecule building block applications in human medicine [53]. Since then, quantum dot conjugates have been extensively applied to molecular diagnosis. For example, quantum dots functionalized with oligonucleotides have been used in DNA detection. A single pair mutation in the human p53 tumor suppressor gene, which has been found to be mutated in more than 50% of the known human cancers, was detected with functionalized quantum dots using a microarray platform. In this same work, the presence of hepatitis B and C viruses was
P1: OTA/XYZ P2: ABC c32 JWBS037-Chen January 25, 2011
794
13:56
Printer Name: Yet to Come
CURRENT STATUS AND FUTURE PROSPECTS FOR NANOPARTICLE
FRET
Acceptor emission (hv)
Quantun dot Excitation (hv)
Quantum dot emission (hv)
FIGURE 32.5 Schematic of single-quantum dot based nanosensor for detection of sequencespecific DNA by FRET. The presence of a specific DNA target leads to formation of a nanocomplex comprising a quantum dot, DNA probes, and targets, which causes FRET between a quantum dot (energy donor) and a energy acceptor (i.e. fluorophore) labeled on a DNA probe.
detected using multicolor quantum dots conjugated to two sequences specific to the alleles of these viruses [54]. Agrawal and collaborators developed a single-molecule-counting method for real-time detection of single genes, proteins, and intact viruses based on the two-color fluorescent coincident detection of green and red quantum dot bioconjugates [55]. Quantum dots modified with streptavidin have been employed for the specific detection of in vivo biotinylated bacteriophages with a sensitivity of 10 bacterial cells per milliliter [56]. Quantum dot–streptavidin conjugates have also been applied to reversephase protein microarrays. This is an emerging high-throughput technology that allows the quantitative measurement of protein levels and posttranslational modifications of signaling proteins in clinical specimens [57, 58]. The fluorescence resonance energy transfer (FRET) phenomenon has been employed as a strategy to develop biological quantum dot-based optical detection systems, with the quantum dots as the FRET donor (see Fig. 32.5). There are numerous examples of FRET-based quantum dot biosensors in the literature [59]. For example, these systems have been used to develop bioanalyses of proteins such as thrombin [60], of protease activity (i.e., collagenase) [61], and of nucleic acids such as the point mutation typical of some ovarian tumors in clinical samples [62]. Apart from metallic nanoparticles and quantum dots, other nanoparticles have been incorporated in optical detection systems. SWNTs have been exploited for the direct determination of specific haplotypes that code for genetic disorders in 10-kilobase-size DNA fragments by atomic force microscopy [63]. Bioconjugated dye-doped silica nanoparticles have been employed as labels for chip-based sandwich DNA assays [64] and for the quantification of bacteria down to a level of single-cell detection [65].
32.3.2 Nanoparticle-Based Magnetic Methods of Diagnosis The magnetic properties of some nanoparticles have also been exploited in the diagnostic field. Shen and co-workers employed arrays of magnetic tunnel junction sensors to detect target DNA labeled with Fe3 O4 nanoparticles with a detection limit below 100 nM [66]. A nanoparticle-based magnetic method has been applied to the detection of the C-reactive protein in whole blood in only 5 min using antibody-conjugated dextran Fe2 O3 nanoparticles
P1: OTA/XYZ P2: ABC c32 JWBS037-Chen January 25, 2011
13:56
Printer Name: Yet to Come
NANOPARTICLE-BASED MOLECULAR DIAGNOSTIC SYSTEMS
795
as superparamagnetic labels with a limit of detection of 3 mg/L. In this work, the C-reactive protein forms a sandwich complex between the Fe2 O3 nanoparticles and silica microparticles, which is sedimented under normal gravitation. A magnetic permeability increase in the sediment due to the presence of the complexed superparamagnetic nanoparticles was determined using an inductance-based transducer [67]. A recent publication reports the development of conjugated nanoparticle-based magnetic relaxation switch biosensors and their application to protein analysis. In presence of the analyte, Fe2 O3 nanoparticles form a cluster that induces a change in relaxation parameter [68]. 32.3.3 Nanoparticle-Based Electrical Methods of Diagnosis In recent years, electrochemical devices for genetic testing and disease-related protein detection have taken advantage of the unique electrical properties, robustness, and high surface area of metallic nanoparticles to improve their speed, selectivity, and detectability (see Fig. 32.6). A good example is the array-based electrical detection method developed
FIGURE 32.6 Approaches used for DNA detection by labelling with metallic nanoparticles. Strategies involve a probe DNA immobilization on a transducing platform following hybridization with target DNA and further with nanoparticle-modified DNA probes. (A) Conductivity assay in which gold is accumulated in the gap and later on a silver enhancement procedure in the presence of hydroquinone is performed. (B) Electrochemical stripping assays based on labeling with gold nanoparticles which were then dissolved with HBr/Br2 and detected by stripping techniques. (C) The same as (B) but the gold nanoparticles are first covered with silver by a deposition treatment and then detected by stripping techniques via a silver enhanced signal. (D) Multilabelling by the use of three different quantum dots and the simultaneous detection of the three DNA targets. (Reprinted with permission from Merkoc¸i [182].)
P1: OTA/XYZ P2: ABC c32 JWBS037-Chen January 25, 2011
796
13:56
Printer Name: Yet to Come
CURRENT STATUS AND FUTURE PROSPECTS FOR NANOPARTICLE
by Park and co-workers [69], which detected DNA at a 500-fM concentration with a point mutation selectivity factor of 100,000:1. Several methods based on the electrochemical stripping of colloidal gold have been developed for DNA detection. Wang and collaborators reported a system in which the target oligonucleotide was captured with magnetic particles conjugated to the oligonucleotide probe and then bound to streptavidin-coated gold nanoparticles. Detection was done by dissolution of the gold nanoparticles and the subsequent potentiometric stripping measurements of the dissolved metal tag at carbon electrodes [70]. Deposition of gold onto piezoelectric crystals allowed DNA detection down to a 1-fM concentration. DNA immobilized on the crystal was hybridized with the target strand. Next, a biotinylated strand hybridized the second half of the target strand. Finally, an avidin–gold nanoparticle conjugate was added to the system and gold was electrochemically deposited on the surface of the metal nanoparticle, providing signal amplification [71]. Metallic nanoparticles have also been employed for the electrochemical detection of proteins. For example, gold nanoparticles have been used in an aptamer-based sandwich assay for thrombin detection. The primary aptamer was immobilized on the surface of a screen-printed carbon electrode, while the secondary aptamer was linked to the gold nanoparticles. In this configuration, the electrochemical reduction current of the gold nanoparticles is proportional to the concentration of the target analyte [72]. Recently, an electrochemical immunosensor has been developed for the amperometric detection of the carcinoembryonic antigen using carbon paste electrodes. The properties of this material were improved by incorporation of a combination of gold and magnetic nanoparticles, which rendered a sensibility of 0.13 ng/mL [73]. Quantum dots have also been employed in electrical systems, which use the inorganic nanocrystal electrochemical stripping for detection. One example is the work by Liu and co-workers. These authors developed a rapid single-nucleotide polymorphism coding system using quantum dots of diverse composition (ZnS, CdS, PbS, CuS), which provided different potential voltammetric signatures [74]. Each quantum dot was functionalized with a different mononucleotide. Via base paring, the nanoparticle conjugates interacted with the mismatch in the target, yielding a distinct electronic fingerprint that reflected the identity of the mismatch. These same authors developed similar detection system using cadmium phosphate-loaded apoferritin nanoparticles instead of quantum dots [75]. Wang and collaborators have also exploited the electrodiversity provided by quantum dots of diverse composition for the simultaneous detection of multiple DNA targets [76]. Using a dual hybridization event, with probes linked to the tagged quantum dots and to magnetic beads, these authors have not only identified the corresponding DNA targets, but they have measured their levels up to femtomolar detection limits as well. Quantum dot–based electrochemical stripping transduction has also been used for multiplexed protein assays such as the quantum dot–aptamer-based method for the dual detection of thrombin and lysozyme at attomole limits [77] or the multianalyte electrical sandwich immunoassay for 2 -microglobulin, IgG, bovine serum albumin, and C-reactive protein [78]. Carbon nanotubes’ excellent mechanical, electrical, and electrochemical properties have encouraged the application of these nanoparticles as transducer components in electrochemical sensors. For instance, classical carbon electrodes have been modified with carbon nanotubes to enhance electron transfer. In this context, Wang and co-workers developed multiwalled carbon nanotube (MWNT) modified glassy-carbon electrodes for improved detection of DNA hybridization, which were used for measuring nucleic acid segments related to the breast cancer gene BRCA1 [79]. Similarly, MWNT were added to carbon paste electrodes to improve cholesterol detection in blood [80]. To achieve ultrasensitive DNA
P1: OTA/XYZ P2: ABC c32 JWBS037-Chen January 25, 2011
13:56
Printer Name: Yet to Come
NANOPARTICLE-BASED DRUG DELIVERY SYSTEMS
797
detection, oligonucleotides have been covalently linked to carbon nanotubes on an SiO2 substrate [81] or on a glass carbon electrode [82]. Likewise, carbon nanotubes have been combined with aptamers [83] and redox enzymes and have been employed as a platform for investigating surface–protein and protein–protein binding as well [84]. At the present time, the main applications of nanoparticle-based diagnostic methods are biomarker discovery, cancer diagnosis, and detection of infectious microorganisms. Nanoparticles have enabled an increase in the limits of detection of diagnostic techniques, establishment point-of- care diagnosis, and integration of diagnosis with therapeutics. It is expected that nanoparticles will enable the selective tagging of a wide range of medically important targets, including bacteria, biomarkers, and individual molecules such as proteins and DNA. Future trends in diagnosis will continue in miniaturization of biochips technology to the nanometric, dimensions, achieving the principal goal of reducing waiting time for test results.
32.4 NANOPARTICLE-BASED DRUG DELIVERY SYSTEMS Genomic and proteomic mediated identification of new therapeutic targets, altogether with combinatorial chemistry procedures, have substantially increased the number of lead molecules to be considered as drug candidates. However, out of 5000 compounds that enter clinical testing, only 5 lead compounds make it to human trials and only one enters the market for clinical use [85]. New molecular entities fail during development mainly because of low aqueous solubility and permeability, high lipophilicity, poor stability, poor pharmacokinetics and pharmacodynamics, and metabolic instability. In this same context, FDA approved drugs confront problems of low efficacy or nonspecific effects. The introduction of nanoparticles as drug delivery systems has helped to traverse these fundamental problems. Nanoparticles can solve issues associated with current therapeutics due to their ability to deliver water-insoluble compounds, to improve stability as well as to localize drug delivery [86] (see Table 32.3). Moreover, the high surface area to volume ratio of nanoparticles allows for a large number of therapeutic molecules to be attached to an
TABLE 32.3 Advantages Provided by Nanoparticles in Drug Delivery Delivery of water-insoluble drugs Increased efficacy Extended drug life
Improved safety
Enhanced permeability and retention effect
Increase rate of dissolution Mask drug lipophilicity Delivery of drugs in the optimum dosage Drug reservoirs for controlled and sustained release Prevention of drug loss through rapid clearance and metabolism Prevention of protein binding to drugs Improved bioavailability Immunocompatibility Reduced systemic toxicity Selective targeting Controlled biodistribution Ability to deliver drugs across biological barriers Improved cellular uptake Improved tissue biodistribution profile
P1: OTA/XYZ P2: ABC c32 JWBS037-Chen January 25, 2011
798
13:56
Printer Name: Yet to Come
CURRENT STATUS AND FUTURE PROSPECTS FOR NANOPARTICLE
individual construct, which contributes to enhance the effectiveness of the drug. Liposomes, dendrimers, polymeric micelles, ceramic nanoparticles, and polymeric nanoparticles are the most extensively investigated nanoparticles for drug delivery and possess the most appropiate characteristics for encapsulation of many drugs [87, 88]. Yet the choice of the most suitable nanoparticle to use as drug carrier depends on the type of compound to be transported, material safety, and route of administration.
32.4.1 Delivery of Water-Insoluble Compounds Solubility is one of the major factors that influence drug effectiveness [89]. Significantly, approximately 40% of newly developed drugs are rejected in early discovery because of low water solubility. Nanoparticles offer the opportunity to formulate poorly water-soluble compounds by drug entrapment within a hydrophobic environment. In this way, drug lipophilicity is masked and replaced by the hydrophilic characteristics of the nanoparticle corona. At the same time, the nanoparticles is small size improves the dissolution rate and saturation solubility, which altogether results in enhanced in vivo drug performance. Liposomes, dendrimers, and polymeric micelles have been extensively assessed as carrier systems for poorly soluble drugs [90–92]. Alternatively, another strategy to enhance drug solubility consists in formulating the therapeutic molecules as nanocrystalline particles ® [93]. One example is Triglide , which is a nanocrystalline fenofibrate formulation for the treatment of lipid disorders.
32.4.2 Extended Drug Life To effectively deliver a drug to the target site, nanoparticles must have the ability to remain in the bloodstream for a considerable time without being eliminated. It has been established that nanoparticle physicochemical characteristics such as size, charge, and surface chemistry dictate their fate in vivo [94]. Surface nonmodified nanoparticles can be adsorbed by proteins, which results in opsonization and entrapment by the mononuclear phagocyte system and the subsequent clearance from circulation. A hydrophilic surface facilitates escape from macrophage capture [95]. For example, nanoparticle surface modifications with PEG [96], modified acrylic acid polymer [97], and polypeptide coatings [98] have demonstrated minimization of unwanted recognition, reduced immunogenicity, and increased nanoparticle circulation half-life and in turn improve the pharmacokinetics and therapeutic index of the drug. It has been hypothesized that this behavior is due to the repulsive effect exerted by the dynamic molecular “cloud” formed by the polymeric coating over the nanoparticle surface [99]. Nevertheless, not only hydrophilicity but other characteristics of the coating polymer such as length, surface density, and conformation influence nanoparticle phagocytosis as well. Nanoparticle size also affects how particles travel through the body. Nanoparticles bigger than 200 nm have shown increased clearance compare to smaller nanoparticles regardless of the composition. For instance, Fang and co-workers demonstrated that serum protein adsorption, murine macrophage uptake, and blood clearance kinetics of pegylated nanoparticles with identical formulation correlated with their size, with smaller nanoparticles showing the best bioavailability [100]. A similar result was shown by Liu and collaborators, who examined the effect of liposome size on liposome circulation time in mice [101].
P1: OTA/XYZ P2: ABC c32 JWBS037-Chen January 25, 2011
13:56
Printer Name: Yet to Come
MULTIFUNCTIONAL NANOPARTICLES
799
Surface functionality also determines nanoparticle life span and fate during circulation. Positively charged nanoparticles tend to form aggregates in the presence of negatively charged serum proteins. Consequently, positively charged nanoparticles often show more rapid blood clearance than neutral and negatively charged ones. In this manner, anionic dendrimers display increased blood circulating time compared to their cationic counterparts [102]. On the other hand, a neutral and hydrophilic surface leads to a longer half-life than a positive or negative one. This effect was shown by Zahr and collaborators using PEG modified core–shell nanoparticles, which had the lowest percentage of uptake when compared to positively and negatively charged nanoshells [103]. 32.4.3 Selective Targeting Among many improvements, nanoparticles have the potential to achieve site-specific delivery of therapeutics. For instance, hydrophilic nanoparticles ranging from 10 to 100 nm in size can passively target and nonspecifically accumulate within tissues with pathophysiological structural abnormalities in the lymphatic drainage. This phenomenon that occurs in tumors and inflammatory and infectious sites is known as the enhanced permeability and retention (EPR) effect [104]. Tumors and inflammatory and infections sites are also characterized by abnormalities in vascular endothelium cells, which have led to the design of nanoparticles that recognize these biological and biophysical differences and adhere to the endothelial system under the flow [105]. In addition to these passive targeting mechanisms, ligands or antibodies directed against the target have been attached onto the nanoparticle surface to amplify the specificity of nanoparticles. The active targeting strategy is described in Section 32.5 in detail. Most of the newly discovered drugs are unable to cross the blood brain barrier. Nanoparticles may also have a significant impact in the selective transport of drugs across biological barriers. Although this ideal goal has not been achieved yet, a considerable amount of research is being currently done in this area. For instance, it has been demonstrated that doxorubicin bound to polysorbate-coated nanoparticles crosses the blood–brain barrier in vivo and that glioblastoma animal models treated with doxorubicin– nanoparticle conjugates show higher survival times with minimal toxicity [106]. Similarly, using different nanoparticle formulations, transport of peptides [107], oligonucleotides, genes [108], and contrast agents across the blood–brain barrier has been accomplished in vivo. Conventional pharmaceuticals have struggled with solubility, systemic toxicity, and adverse effects. Nanoparticles have the potential to deliver poorly water-soluble drugs and to achieve site-specific delivery of therapeutics, thus differentiating between normal and pathological cells. Moreover, they can reduce the cost and the time of drug discovery, design, and development. Unfortunately, up to now, none of the developed nanoparticles fulfill all the desired characteristics of an ideal drug delivery system and for that to happen nanoparticles will have to get smarter. In this context, multifunctional and multimodal nanoparticles are now being actively investigated and are on the horizon as the next generation of nanoparticles.
32.5 MULTIFUNCTIONAL NANOPARTICLES Over the past years, multifunctional nanoparticles have emerged as a promising approach that could afford endless opportunities for multimodal imaging and simultaneous diagnosis
P1: OTA/XYZ P2: ABC c32 JWBS037-Chen January 25, 2011
800
13:56
Printer Name: Yet to Come
CURRENT STATUS AND FUTURE PROSPECTS FOR NANOPARTICLE
FIGURE 32.7 Multifuncional nanoparticles for drug delivery. Multifunctional nanocarriers may combine a specific targeting agent, with nanoparticles for imaging, a cell-penetrating agent, a stimulussensitive element of drug release, a stabilizing polymer to ensure biocompatibility, and the therapeutic compound. These multiple functionalities have to be modulated and coordinated in such a manner that they do not interfere with one another. For this purpose, functionalities have to be physically separated onto the nanoparticle surface, which can be achieved using orthogonal chemistries. (Reproduced with permission from Sanvicens and Marco [143].)
and therapy [109]. Whereas monofunctional nanoparticles only provide a single function— that is, liposomes, dendrimers, and polymeric micelles can transport drugs and quantum dots—metallic and magnetic nanoparticles are excellent imaging tools; however, none of them possess the inherent property to distinguish between healthy and unhealthy cells or tissues. Multifunctional nanoparticles combine different properties in a single stable construct. Figure 32.7 illustrates the schematic structure of a multifunctional nanoparticle. For active targeting, monofunctional nanoparticles can be linked to a specific molecular probe that differentially recognizes unique and distinct surface molecular features of the target cells. In this manner, nanoparticles are able to distinguish between normal and pathological tissues. Active specific target recognition improves delivery efficiency and reduces collateral toxicity and adverse effects. Some of the strategies that have been exploited for specific targeting include the attachment onto the nanoparticle surface of antibodies [110, 111], peptides [112, 113], proteins [114], aptamers [115, 116], carbohydrates [117], and small molecules [118, 119]. In this same context, magnetic nanoparticles attracted by an external magnetic field have been used to increase site-specific delivery [120]. Targeteddelivery has been also achieved using low-frequency ultrasound [121]. Once nanoparticles reach the target site, they have to get inside the cell and release the drug to the cytoplasm. Due to the lipophilic nature of the biological membranes, nanoparticles—despite their small size—do not easily enter into cells. Some nanoparticles such as liposomes or nanoparticles coated with folate or transferrin exploit the mechanism of receptor-mediated endocytosis to move across the plasma membrane [122]. Yet endosomal uptake raises several problems. Accumulation in the endo-/lysosomal compartment may lead to drug inactivation. What is more, certain therapeutic agents exert their action in specific organelles, and therefore they need to escape from the lysosomes. To traverse
P1: OTA/XYZ P2: ABC c32 JWBS037-Chen January 25, 2011
13:56
Printer Name: Yet to Come
MULTIFUNCTIONAL NANOPARTICLES
801
TABLE 32.4 Strategies for the Intracellular Delivery of Nanoparticles Mechanism
Functiona
Target
References
Peptide mediated
TAT RGD motif MPAP
126 140 177
Polymer mediated Protein/enzyme mediated Lipid mediated
PEI-grafted-PEG Matrix metalloprotease 2 and 7 Liposomes
U-87 MG astrocytoma cells SLK cells 9L-GFP and 9L-RFP gliosarcoma cells Mice with 9L-GFP, 9L-RFP and LS174T induced tumors Hela cells HT-1080 cells A549 cells C57B16 mice
123 178 137
a Abbreviations: MPAP, myristoylated polyarginine peptides; PEI, polyethylenemine; PEG, polyethylene glycol; TAT, trans-activating transcriptional activator.
this shortcoming, strategies such as surface polymeric coatings [123], cationic liposomes [124], and fusogenic peptides [125] have been designed to facilitate nanoparticle endosome release. Alternatively, numerous approaches have been investigated to enhance direct nanoparticle delivery to the cytoplasm. Such functions not only modify nanoparticle pharmacokinetics and biodistribution, but increase drug efficacy as well. Table 32.4 summarizes the plasma membrane permeation enhancers most commonly used. Strategies to enable nanoparticles to access specific organelles have been reviewed by Breunig et al. [122]. Ideally, nanoparticle-based drug delivery systems should be stable during transport, keeping the drug encapsulated during systemic circulation, but degrade and release their cargo at the target site. This goal has been achieved incorporating a stimuli-responsive function of controlled drug release into nanoparticles. Some of the approaches that have been investigated exploit unique features of the pathological sites. For example, pH-, temperature- and redox-sensitive nanocarriers have been designed taking into consideration the distinctive decreased extracellular pH value and increased temperature and redox activity that characterize the inflammatory and tumor microenvironment [126–128]. Another possibility is to apply an external stimulus to release the drug at the desired site. This is the case of magnetic-, photo-, and photothermal-sensitive nanoparticles [129–131]. Recently, smart multifunctional nanoparticles have been engineered to sense drug release and to evaluate its therapeutic efficacy. For example, Bagalkot and collaborators have designed quantum dot-based multifunctional nanoparticles that change in fluorescence—based on the FRET mechanism—to report drug release into the cell [132]. Multifunctional nanoparticles show great potential to overcome the problems of gene delivery. The successful outcome for gene therapies based on nonviral systems still remains elusive. One of the main limitations of nonviral systems is their low transfection efficiency due to restricted passage through the barriers present in the body. In this area, multifunctional nanoparticles have demonstrated their potential to circumvent biological barriers [133]. Another important bottleneck is the specific delivery of DNA into the nucleus. As described in this section, the nanoparticle surface can be functionalized with different ligands to improve cellular uptake, nanoparticle internalization, and escape from the endosomes. However, DNA has to move from the cytoplasm to the nucleus for efficient transfection and successful gene therapy. Therefore crossing the nuclear plasma constitutes another
P1: OTA/XYZ P2: ABC c32 JWBS037-Chen January 25, 2011
802
13:56
Printer Name: Yet to Come
CURRENT STATUS AND FUTURE PROSPECTS FOR NANOPARTICLE
barrier to gene delivery. In this context, nanoparticles have been conjugated with nuclear localization signals, which are peptide sequences that specifically interact with cytoplasmic factors, which, in turn, target molecules to the nucleus [122].
32.5.1 Nanohybrids: A Novel Type of Multifunctional Nanoparticles Multifunctional nanoparticles built with an imaging core scaffold—such as quantum dots, fluorescent core–shell silica nanoparticles, or iron oxide nanocrystals—combined with other functionalities, allow simultaneous drug delivery, monitoring of the transport process, and reporting of nanoparticle accumulation in the target site, therefore providing real-time monitoring of nanoparticle biodistribution [111, 124, 134–136]. In this area, multimodal nanoparticles—engineered nanohybrids of two or more nanoparticles in a single construct—have emerged as a novel tool for combinatory imaging, diagnosis, and therapy. Al-Jamal and co-workers described the preparation of functionalized quantum dot–liposome hybrids, which combined quantum dot imaging properties with liposomemediated cellular binding and internalization into cells and tissues [137]. Likewise, iron oxide, silica, and polystyrene nanoparticles have been incorporated into liposomes to enhance their compatibility with biological media [138]. To improve the delivery of the poor water-soluble drug campthothecin, the therapeutic was incorporated into negatively charged micelles, which were in turn encapsulated in magnesium–aluminum layered double hydroxide nanoparticles [139]. Other examples of nanohybrids include multifunctional polymeric micelles composed of a superparamagnetic iron oxide cluster [140] and iron oxide nanocrystals covered by a mesoporous silica shell [141]. In recent work, Park and co-workers reported that micellar hybrid nanoparticles composed of magnetic nanoparticles and quantum dots had the in vivo capacity for simultaneous multimodal dual diagnosis and therapy [142]. Other applications of nanohybrids such as detection, imaging, and diagnosis have been described in Section 32.2 of this chapter. Despite these results, this field is still in a very preliminary stage. But undoubtedly, nanohybrids will constitute a versatile platform for multiple biomedical applications.
32.6 NANOPARTICLE TOXICITY Nanoparticles can be the next breakthrough in pharmacology and human medicine, with extraordinary opportunities for application in early-stage cancer diagnosis, pathogen detection, or gene and drug delivery. These applications of nanoparticle-based systems will imply human exposure to nanoparticles through the skin, the respiratory tract, the gastrointestinal tract, and the lymphatics. In spite of this, until recently it has not been taken into consideration that nanoparticles may constitute a potential hazard to human health. In this context, we still do not know the mechanisms of entry of nanoparticles into the organism, how they are distributed in the body, how they are taken up by the cell, how they move through tissues, and how they affect different organs. Whether nanoparticles are a potential hazard to human health is still a controversial subject. Published reports are few, inconclusive, and sometimes contradictory [143]. Nanoparticle toxicity seems to depend on too many factors such as chemical composition [144], size and surface area [145], shape [146], concentration [147], and surface coating [148]. On top of that, experimental conditions also affect the results. For instance, Patra and co-workers
P1: OTA/XYZ P2: ABC c32 JWBS037-Chen January 25, 2011
13:56
Printer Name: Yet to Come
NANOPARTICLE TOXICITY
803
demonstrated that whereas 33-nm citrate-capped gold nanospheres induced cell death in human carcinoma cells, they were not toxic to baby hamster kidney and human hepatocellular liver carcinoma cells [149]. Quantum dots also show this cell-selective response. Both IMR-32 human neuroblastoma cells and MCF-7 human breast cancer cells undergo apoptosis in the presence of quantum dots. However, each cell line shows a completely different cell death mechanism, capase-dependent [150] or caspase-independent [151]. It has been suggested that this inconsistency of results could be due to the influence that nanoparticle properties could have in cytotoxicity assays. Nanoparticles high adsorption capacity, catalytic activity, and optical or magnetic properties may interfere with assay components or detection systems, which may introduce artifacts into the studies [152]. Consequently, nanoparticle toxicity studies should be carried out in the broadest context and beyond simple cell lines, since both the physicochemical characteristics of the nanoparticles and the experimental conditions influence the results. Several studies have demonstrated no toxicity of nanoparticles in vivo [153–155]. Nevertheless, most of these experiments do not involve nanoparticle long-term safety analysis. Nanoparticle interaction and accumulation in tissues has been described in the liver, the spleen, the lymph node, and the bone marrow [112, 156, 157], which could potentially amplify nanoparticle toxicity due to intracellular changes. For example, quantum dotbased probes have shown no acute toxicity in some in vivo models [155, 158]. However, a recent work by Mancini and co-workers described how polymer-coated quantum dots can be degraded by reactive oxygen species, which could have important implications regarding their in vivo long-term safety [159]. Therefore in order to avoid collateral damage, once nanoparticles have exerted their function in the body, accumulation of residual nanomaterial should be avoided. In this context, chemical strategies have been proposed to reduce nanotoxicological issues and to facilitate nanoparticle renal clearance: for instance, modification of the nanoparticle surface with biocompatible coatings such as PEG [160], phophatidylcholine [161], acetylated poly(amidoamine) [162], and cysteine [163, 164]. Another approach involves the development of novel biocompatible materials with low toxicity, for example, poly(d,l-lactide-co-glycolide) [165], poly(-amino esters) [166], poly(ε-caprolactone) [167], and poly(glycerol-adipate) [168]. Today, many questions await resolution in the nanotoxicology field. The major challenges that lie ahead involve the full characterization of nanomaterials, the complete evaluation of their potential hazard, and the development of novel methodologies to assess nanomaterial exposure. In this context, nanoregulation is undergoing major changes to encompass environmental, health, and safety issues. The European Center for Ecotoxicology and Toxicology of Chemicals (OECD) is currently working on building guidelines for risk assessment and management of nanomaterials. The same goal is pursued in the strategic research agenda for nanomedicine of the European Nanomedicine Technology (ENT) Platform. The Seventh Research Framework Programme of the European Union includes among its objectives the study of the “impact of engineered nanoparticles on health and the environment” and the “validation, adaptation and/or development of risk assessment methodology for engineered nanoparticles” (FP7 Cooperation Work Programme 2008). The American National Nanotech Initiative’s (NNI) strategic plan from December 2007 prioritizes the understanding of the potential risks of nanotechnology and the development of the means to manage them. The U.S. Food and Drug Administration (FDA) has formed a NanoTechnology Interest Group (NTIG) to facilitate the regulation of nanotechnology products (Task Force July 2007) and will invest over US$96 million in environmental, health, and safety R&D to address the potential hazards linked to nanotechnology.
P1: OTA/XYZ P2: ABC c32 JWBS037-Chen January 25, 2011
804
13:56
Printer Name: Yet to Come
CURRENT STATUS AND FUTURE PROSPECTS FOR NANOPARTICLE
32.7 CONCLUSION Nanomedicine, the application of nanotechnology to health, holds great promise for the future, since it offers significant improvements in performance compared with existing technologies. In this context, nanoparticles have raised a high expectation in the development of new treatments and diagnosis. To date, nanoparticle applications in medicine have provided novel contrast agents for diagnosis, nanoparticle-based biosensors with improved sensitivity, and targeted drug delivery systems with enhanced drug-therapeutic efficiency and pharmacological characteristics. On these bases, in the a near future, the combination of nanoparticles with other nanotechnology-based materials has the potential to help us understand biological processes at the molecular level and to accurately target and quantify molecules indicative of cellular disorders at the single-molecule level. These advances will allow US to examine more molecular targets in vivo, identify disease at the earliest possible stage, design personalized delivery systems for highly specific applications, and develop programmable drug delivery systems that release the appropriate concentration of drug in response to biomarkers. On the other hand, at the current state, we cannot conclude that nanoparticles may provoke unseen adverse effects in humans. The advantage of nanoparticles over conventional medicine lies in their size. Unfortunately, it is the nanoparticles’ small dimensions that make them potentially dangerous for human health. Consequently, it is mandatory that nanoparticles are thoroughly examined for biocompatibility, biodistribution, and clearance to ensure their safe clinical use in humans. Therefore future research should strive to completely understand the factors influencing nanoparticle toxicity and to develop suitable cytotoxic tests for this purpose.
ACKNOWLEDGMENTS CIBER-BBN is an initiative funded by the VI National R&D&i Plan 2008–2011, Iniciativa Ingenio 2010, Consolider Program, CIBER Actions and financed by the Instituto de Salud Carlos III with assistance from the European Regional Development Fund.
ABBREVIATIONS CT, computed tomography; EPR, enhanced permeability and retention; FRET, fluorescence resonance energy transfer; MRI, magnetic resonance imaging; MWNT, multiwalled carbon nanotube; PEG, polyethylene glycol; PET, positron emission spectroscopy; SWNT, singlewalled carbon nanotube; USPIO, ultrasmall particle of iron oxide.
REFERENCES 1. Farokhzad, O. C.; Langer, R. Nanomedicine: developing smarter therapeutic and diagnostic modalities. Adv. Drug Deliv. Rev. 2006, 58(14), 1456–1459. 2. Park, H. H.; Jamison, A. C.; Lee, T. R. Rise of the nanomachine: the evolution of a revolution in medicine. Nanomedicine 2007, 2(4), 425–439.
P1: OTA/XYZ P2: ABC c32 JWBS037-Chen January 25, 2011
13:56
Printer Name: Yet to Come
REFERENCES
805
3. Patolsky, F.; Zheng, G.; Lieber, C. M. Nanowire sensors for medicine and the life sciences. Nanomedicine 2006, 1(1), 51–65. 4. Hwang, K. S.; et al. In-situ quantitative analysis of a prostate-specific antigen (PSA) using a nanomechanical PZT cantilever. Lab on a Chip 2004, 4(6), 547–552. 5. Kumbar, S. G.; et al. Polymeric nanofibers as novel carriers for the delivery of therapeutic molecules. J. Nanosci. Nanotechnol. 2006, 6(9-10), 2591–2607. 6. Ntziachristos, V.; et al. Current MRI and diffuse optical tomography of breast after indocyanine green enhancement. Proc. Natl. Acad. Sci. U.S.A 2000, 97(6), 2767–2772. 7. Beyer, T.; et al. A combined PET/CT scanner for clinical oncology. J. Nucl. Med. 2000, 41(8), 1369–1379. 8. Mirkin, C. A.; Niemeyer, C. M.; Nanobiotechnology II. More concepts and applications. Adv. Mater. 2007, 19, 3393–3394. 9. Thorek, D.; et al. Superparamagnetic iron oxide nanoparticle probes for molecular imaging. Ann. Biomed. Eng. 2006, 34(1), 23–38. 10. Bach-Gansmo, T. Ferrimagnetic susceptibility contrast agents. Acta Radiol. Suppl. 1993, 387: 1–30. 11. Weissleder, R.; et al. Superparamagnetic iron oxide: pharmacokinetics and toxicity. Am. J. Roentgenol. 1989, 152(1), 167–173. 12. Bartolozzi, C.; et al. Adbominal MR: liver and pancreas. Eur. Radiol. 1999, 9(8), 1496–1512. 13. Reimer, P.; et al. Clinical results with Resovist: a phase 2 clinical trial. Radiology 1995, 195(2), 489–496. 14. McCarthy, J. R.; Weissleder, R. Multifunctional magnetic nanoparticles for targeted imaging and therapy. Adv. Drug Deliv. Rev. 2008, 60(11), 1241–1251. 15. Sun, C.; Lee, J. S. H.; Zhang, M. Magnetic nanoparticles in MR imaging and drug delivery. Adv. Drug Deliv. Rev. 2008, 60(11), 1252–1265. 16. Ito, A.; et al. Medical application of functionalized magnetic nanoparticles. J. Biosci. Bioeng. 2005, 100(1), 1–11. 17. Sosnovik, D.; Nahrendorf, M.; Weissleder, R. Magnetic nanoparticles for MR imaging: agents, techniques and cardiovascular applications. Basic Res. Cardiol. 2008, 103(2), 122–130. 18. Lee, K.; et al. Magnetic resonance tracking of multifunctional nanoparticle-labeled mouse mesenchymal stem cells in a mouse model of myocardial infarction. Curr. Appl. Phys. 2009, 9(1), S12–S14. 19. Turvey, S. E.; et al. Noninvasive imaging of pancreatic inflammation and its reversal in type 1 diabetes. J. Clin. Invest. 2005, 115(9), 2454–2461. 20. Artemov, D.; et al. MR molecular imaging of the Her-2/neu receptor in breast cancer cells using targeted iron oxide nanoparticles. Magn. Reson. Med. 2003, 49(3), 403–408. 21. Huh, Y.-M.; et al. In vivo magnetic resonance detection of cancer by using multifunctional magnetic nanocrystals. J. Am. Chem. Soc. 2005, 127(35), 12387–12391. 22. Tiefenauer, L. X.; et al. In vivo evaluation of magnetite nanoparticles for use as a tumor contrast agent in MRI. Magn. Reson. Imaging. 1996, 14(4), 391–402. 23. Lee, J. H.; et al. Artificially engineered magnetic nanoparticles for ultra-sensitive molecular imaging. Nat. Med. 2007, 13(1), 95–99. 24. Rockall, A. G.; et al. Diagnostic performance of nanoparticle-enhanced magnetic resonance imaging in the diagnosis of lymph node metastases in patients with endometrial and cervical cancer. J. Clin. Oncol. 2005, 23(12), 2813–2821. 25. Sanvicens, N.; et al. Oxidative stress-induced apoptosis in retinal photoreceptor cells is mediated by calpains and caspases and blocked by the oxygen radical scavenger CR-6. J. Biol. Chem. 2004, 279(38), 39268–39278.
P1: OTA/XYZ P2: ABC c32 JWBS037-Chen January 25, 2011
806
13:56
Printer Name: Yet to Come
CURRENT STATUS AND FUTURE PROSPECTS FOR NANOPARTICLE
26. Zhao, M.; et al. Non-invasive detection of apoptosis using magnetic resonance imaging and a targeted contrast agent. Nat. Med. 2001, 7(11), 1241–1244. 27. van Tilborg, G. A.; et al. Annexin A5-functionalized bimodal lipid-based contrast agents for the detection of apoptosis. Bioconjug. Chem. 2006, 17(3), 741–749. 28. Weissleder, R.; et al. In vivo magnetic resonance imaging of transgene expression. Nat. Med. 2000, 6(3), 351–355. 29. Cheon, J.; Lee, J. H. Synergistically integrated nanoparticles as multimodal probes for nanobiotechnology. Acc. Chem. Res. 2008, 41(12), 1630–1640. 30. Mulder, W. J.; et al. Quantum dots with a paramagnetic coating as a bimodal molecular imaging probe. Nano Lett. 2006, 6(1), 1–6. 31. Choi, J. S.; et al. A hybrid nanoparticle probe for dual-modality positron emission tomography and magnetic resonance imaging. Angew. Chem. Int. Ed. Engl. 2008, 47(33), 6259–6262. 32. Medintz, I. L.; et al. Quantum dot bioconjugates for imaging, labelling and sensing. Nat. Mater. 2005, 4(6), 435–446. 33. Gao, X.; et al. Quantum dots bearing lectin-functionalized nanoparticles as a platform for in vivo brain imaging. Bioconjug. Chem. 2008, 19(11), 2189–2195. 34. Kim, S.; et al. Near-infrared fluorescent type II quantum dots for sentinel lymph node mapping. Nat. Biotechnol. 2004, 22(1), 93–97. 35. Gao, X.; et al. In vivo cancer targeting and imaging with semiconductor quantum dots. Nat. Biotechnol. 2004, 22(8), 969–976. 36. De La Zerda, A.; et al. Carbon nanotubes as photoacoustic molecular imaging agents in living mice. Nat. Nanotechnol. 2008, 3(9), 557–562. 37. Lacerda, L.; et al. Dynamic imaging of functionalized multi-walled carbon nanotube systemic circulation and urinary excretion. Adv. Mater. 2008, 20(2), 225–230. 38. Terracciano, R.; et al. Selective binding and enrichment for low-molecular weight biomarker molecules in human plasma after exposure to nanoporous silica particles. Proteomics 2006, 6(11), 3243–3250. 39. Gaspari, M.; et al. Nanoporous surfaces as harvesting agents for mass spectrometric analysis of peptides in human plasma. J. Proteome Res. 2006, 5(5), 1261–1266. 40. Mirkin, C. A.; et al. A DNA-based method for rationally assembling nanoparticles into macroscopic materials. Nature 1996, 382(6592), 607–609. 41. Reynolds, R. A.; Mirkin, C. A.; Letsinger, R. L. A gold nanoparticle/latex microspherebased colorimetric oligonucleotide detection method. Pure Appl. Chem. 2000, 72(1-2), 229– 235. 42. Storhoff, J. J.; et al. Homogeneous detection of unamplified genomic DNA sequences based on colorimetric scatter of gold nanoparticle probes. Nat. Biotechnol. 2004, 22(7), 883–887. 43. Chakrabarti, R.; Klibanov, A. M. Nanocrystals modified with peptide nucleic acids (PNAs) for selective self-assembly and DNA detection. J. Am. Chem. Soc. 2003, 125(41), 12531– 12540. 44. Haes, A. J.; et al. A localized surface plasmon resonance biosensor: First steps toward an assay for Alzheimer’s disease. Nano Lett. 2004, 4(6), 1029–1034. 45. Du, B. A.; Li, Z. P.; Cheng, Y. Q. Homogeneous immunoassay based on aggregation of antibodyfunctionalized gold nanoparticles coupled with light scattering detection. Talanta 2008, 75(4), 959–964. 46. Hirsch, L. R.; et al. A whole blood immunoassay using gold nanoshells. Anal. Chem. 2003, 75(10), 2377–2381. 47. Niemeyer, C. M.; Ceyhan, B. DNA-directed functionalization of colloidal gold with proteins. Angew. Chem. Int. Ed. Engl. 2001, 40(19), 3685–3688.
P1: OTA/XYZ P2: ABC c32 JWBS037-Chen January 25, 2011
13:56
Printer Name: Yet to Come
REFERENCES
807
48. Georganopoulou, D. G.; et al. Nanoparticle-based detection in cerebral spinal fluid of a soluble pathogenic biomarker for Alzheimer’s disease. Proc. Natl. Acad. Sci. U.S.A. 2005, 102(7), 2273–2276. 49. Nam, J. M.; Stoeva, S. I.; Mirkin, C. A. Bio-bar-code-based DNA detection with PCR-like sensitivity. J. Am. Chem. Soc. 2004, 126(19), 5932–5933. 50. El-Boubbou, K.; Gruden, C.; Huang, X. Magnetic glyco-nanoparticles: a unique tool for rapid pathogen detection, decontamination, and strain differentiation. J. Am. Chem. Soc. 2007, 129(44), 13392–3. 51. Son, A.; et al. Quantitative detection of antibiotic resistance genes using magnetic/luminescent core-shell nanoparticles - art. no. 68650P. Nanoscale Imaging, Sensing, and Actuation for Biomedical Applications V 2008, 6865: p. P8650-P8650. 52. Nichkova, M.; et al. Multiplexed immunoassays for proteins using magnetic luminescent nanoparticles for internal calibration. Anal. Biochem. 2007, 369(1), 34–40. 53. Mitchell, G. P.; Mirkin, C. A.; Letsinger, R. L.; Programmed assembly of DNA functionalized quantum dots. J. Am. Chem. Soc. 1999, 121(35), 8122–8123. 54. Gerion, D.; et al. Room-temperature single-nucleotide polymorphism and multiallele DNA detection using fluorescent nanocrystals and microarrays. Anal. Chem. 2003, 75(18), 4766–4772. 55. Agrawal, A.; et al. Counting single native biomolecules and intact viruses with color-coded nanoparticles. Anal. Chem. 2006, 78(4), 1061–1070. 56. Edgar, R.; et al. High-sensitivity bacterial detection using biotin-tagged phage and quantum-dot nanocomplexes. Proc. Natl. Acad. Sci. U.S.A. 2006, 103(13), 4841–4845. 57. Liotta, L. A.; et al. Protein microarrays: Meeting analytical challenges for clinical applications. Cancer Cell 2003, 3(4), 317–325. 58. Geho, D.; et al. Pegylated, steptavidin-conjugated quantum dots are effective detection elements for reverse-phase protein microarrays. Bioconjug. Chem. 2005, 16(3), 559–566. 59. Medintz, I. L.; et al. Self-assembled nanoscale biosensors based on quantum dot FRET donors. Nat. Mater. 2003, 2(9), 630–638. 60. Levy, M.; Cater, S. F.; Ellington, A. D. Quantum-dot aptamer beacons for the detection of proteins. Chembiochem. 2005, 6(12), 2163–2166. 61. Shi, L.; et al. Synthesis and application of quantum dots FRET-based protease sensors. J. Am. Chem. Soc. 2006, 128(32), 10378–10379. 62. Zhang, C. Y.; et al. Single-quantum-dot-based DNA nanosensor. Nat. Mater. 2005, 4(11), 826–831. 63. Woolley, A. T.; et al. Direct haplotyping of kilobase-size DNA using carbon nanotube probes. Nat. Biotechnol. 2000, 18(7), 760–763. 64. Zhao, X. J.; Tapec-Dytioco, R.; Tan, W. H. Ultrasensitive DNA detection using highly fluorescent bioconjugated nanoparticles. J. Am. Chem. Soc. 2003, 125(38), 11474–11475. 65. Zhao, X. J.; et al. A rapid bioassay for single bacterial cell quantitation using bioconjugated nanoparticles. Proc. Natl. Acad. Sci. U.S.A. 2004, 101(42), 15027–15032. 66. Shen, W.; et al. Quantitative detection of DNA labeled with magnetic nanoparticles using arrays of MgO-based magnetic tunnel junction sensors. Appl. Phys. Lett. 2008, 93(3), 033903. 67. Braimi, F.; et al. Rapid one-step whole blood C-reactive protein magnetic permeability immunoassay with monoclonal antibody conjugated nanoparticles as superparamagnetic labels and enhanced sedimentation. Anal. Bioanal. Chem. 2006, 384(3), 651–657. 68. Lowery, T. J.; et al. Single-coil, multisample, proton relaxation method for magnetic relaxation switch assays. Anal. Chem. 2008, 80(4), 1118–1123. 69. Park, S. J.; Taton, T. A.; Mirkin, C. A. Array-based electrical detection of DNA with nanoparticle probes. Science 2002, 295(5559), 1503–1506.
P1: OTA/XYZ P2: ABC c32 JWBS037-Chen January 25, 2011
808
13:56
Printer Name: Yet to Come
CURRENT STATUS AND FUTURE PROSPECTS FOR NANOPARTICLE
70. Wang, J.; et al. Metal nanoparticle-based electrochemical stripping potentiometric detection of DNA hybridization. Anal. Chem. 2001, 73(22), 5576–5581. 71. Weizmann, Y.; Patolsky, F.; Willner, I. Amplified detection of DNA and analysis of single-base mismatches by the catalyzed deposition of gold on Au-nanoparticles. Analyst 2001, 126(9), 1502–1504. 72. Kerman, K.; Tamiya, E. Aptamer-functionalized Au nanoparticles for the electrochemical detection of thrombin. J. Biomed. Nanotechn. 2008, 4(2), 159–164. 73. Tang, D. Y.; Xia, B. Y. Electrochemical immunosensor and biochemical analysis for carcinoembryonic antigen in clinical diagnosis. Microchimica Acta 2008, 163(1-2), 41–48. 74. Liu, G. D.; Lee, T. M. H.; Wang, J. S. Nanocrystal-based bioelectronic coding of single nucleotide polymorphisms. J. Am. Chem. Soc. 2005, 127(1), 38–39. 75. Liu, G. D.; Lin, Y. H. Electrochemical quantification of single-nucleotide polymorphisms using nanoparticle probes. J. Am. Chem. Soc. 2007, 129(34), 10394–10401. 76. Wang, J.; Liu, G. D.; Merkoci, A. Electrochemical coding technology for simultaneous detection of multiple DNA targets. J. Am. Chem. Soc. 2003, 125(11), 3214–3215. 77. Hansen, J. A.; et al. Quantum-dot/aptamer-based ultrasensitive multi-analyte electrochemical biosensor. J. Am. Chem. Soc. 2006, 128(7), 2228–2229. 78. Liu, G.; et al. Electrochemical coding for multiplexed immunoassays of proteins. Anal. Chem. 2004, 76(23), 7120–7130. 79. Wang, J.; Kawde, A. N.; Musameh, M. Carbon-nanotube-modified glassy carbon electrodes for amplified label-free electrochemical detection of DNA hybridization. Analyst 2003, 128(7), 912–916. 80. Li, G.; et al. Study of carbon nanotube modified biosensor for monitoring total cholesterol in blood. Biosens. Bioelectron. 2005, 20(10), 2140–2144. 81. Li, J.; et al. Carbon nanotube nanoelectrode array for ultrasensitive DNA detection. Nano Lett. 2003, 3(5), 597–602. 82. Cai, H.; et al. Carbon nanotube-enhanced electrochemical DNA biosensor for DNA hybridization detection. Anal. Bioanal. Chem. 2003, 375(2), 287–293. 83. Maehashi, K.; et al. Label-free protein biosensor based on aptamer-modified carbon nanotube field-effect transistors. Anal. Chem. 2007, 79(2), 782–787. 84. Chen, R. J.; et al. Noncovalent functionalization of carbon nanotubes for highly specific electronic biosensors. Proc. Natl. Acad. Sci. U.S.A. 2003, 100(9), 4984–4989. 85. Bawa, R. Nanoparticle-based therapeutics in humans: a survey. Nanotechnol. Law & Business 2008, 5(2), 135–155. 86. Emerich, D. F.; Thanos, C. G. The pinpoint promise of nanoparticle-based drug delivery and molecular diagnosis. Biomol. Eng. 2006, 23(4), 171–184. 87. Bawarski, W. E.; et al. Emerging nanopharmaceuticals. Nanomedicine 2008, 4(4), 273–282. 88. Devalapally, H.; Chakilam, A.; Amiji, M. M. Role of nanotechnology in pharmaceutical product development. J. Pharm. Sci. 2007, 96(10), 2547–2565. 89. Lipinski, C. A.; et al. Experimental and computational approaches to estimate solubility and permeability in drug discovery and development settings. Adv. Drug DelivRev 2001, 46(1-3), 3–26. 90. Svenson, S.; Chauhan, A. S. Dendrimers for enhanced drug solubilization. Nanomedicine 2008, 3(5), 679–702. 91. Torchilin, V. P. Targeted polymeric micelles for delivery of poorly soluble drugs. Cell Mol. Life Sci. 2004, 61(19-20), 2549–2559. 92. Fahr, A.; Liu, X. Drug delivery strategies for poorly water-soluble drugs. Expert Opin. Drug Deliv. 2007, 4(4), 403–416.
P1: OTA/XYZ P2: ABC c32 JWBS037-Chen January 25, 2011
13:56
Printer Name: Yet to Come
REFERENCES
809
93. Merisko-Liversidge, E.; Liversidge, G. G.; Cooper, E. R. Nanosizing: a formulation approach for poorly-water-soluble compounds. Eur. J. Pharm. Sci. 2003, 18(2), 113–120. 94. Alexis, F.; et al. Factors affecting the clearance and biodistribution of polymeric nanoparticles. Mol. Pharm. 2008, 5(4), 505–515. 95. Romberg, B.; Hennink, W. E.; Storm, G. Sheddable coatings for long-circulating nanoparticles. Pharm. Res. 2007, 25(1), 55–71. 96. Gref, R.; et al. Biodegradable long-circulating polymeric nanospheres. Science 1994, 263(5153), 1600–1603. 97. Sathe, T. R.; Agrawal, A.; Nie, S. Mesoporous silica beads embedded with semiconductor quantum dots and iron oxide nanocrystals: dual-function microcarriers for optical encoding and magnetic separation. Anal. Chem. 2006, 78(16), 5627–5632. 98. Pinaud, F.; et al. Bioactivation and cell targeting of semiconductor CdSe/ZnS nanocrystals with phytochelatin-related peptides. J. Am. Chem. Soc. 2004, 126(19), 6115–6123. 99. Gref, R.; et al. The controlled intravenous delivery of drugs using PEG-coated sterically stabilized nanospheres Adv. Drug Deliv. Rev. 1995, 16(2-3): 215–233. 100. Fang, C.; et al. In vivo tumor targeting of tumor necrosis factor-alpha-loaded stealth nanoparticles: effect of MePEG molecular weight and particle size. Eur. J. Pharm. Sci. 2006, 27(1), 27–36. 101. Liu, D.; Mori, A.; Huang, L. Role of liposome size and RES blockade in controlling biodistribution and tumor uptake of GM1-containing liposomes. Biochim. Biophys. Acta 1992, 1104(1), 95–101. 102. Kaminskas, L. M.; et al. Impact of surface derivatization of poly-L-lysine dendrimers with anionic arylsulfonate or succinate groups on intravenous pharmacokinetics and disposition. Mol. Pharm. 2007, 4(6), 949–961. 103. Zahr, A. S.; Davis, C. A.; Pishko, M. V. Macrophage uptake of core-shell nanoparticles surface modified with poly(ethylene glycol). Langmuir 2006, 22(19), 8178–8185. 104. Maeda, H. The enhanced permeability and retention (EPR) effect in tumor vasculature: the key role of tumor-selective macromolecular drug targeting. Adv. Enzyme Regul 2001, 41(1), 189–207. 105. Decuzzi, P.; et al. Intravascular delivery of particulate systems: does geometry really matter? Pharm. Res. 2000, 26(1), 235–243. 106. Steiniger, S. C.; et al. Chemotherapy of glioblastoma in rats using doxorubicin-loaded nanoparticles. Int. J. Cancer 2004, 109(5), 759–767. 107. Kreuter, J.; et al. Passage of peptides through the blood-brain barrier with colloidal polymer particles (nanoparticles). Brain Res. 1995, 674(1), 171–174. 108. Lu, W.; et al. Cationic albumin-conjugated pegylated nanoparticles allow gene delivery into brain tumors via intravenous administration. Cancer Res. 2006, 66(24), 11878– 11887. 109. Torchilin, V. P. Multifunctional nanocarriers. Adv. Drug Deliv. Rev 2006, 58(14), 1532– 1555. 110. Lukyanov, A. N.; et al. Tumor-targeted liposomes: doxorubicin-loaded long-circulating liposomes modified with anti-cancer antibody. J. Control Release 2004, 100(1), 135–144. 111. Yu, K. N.; et al. Multiplex targeting, tracking, and imaging of apoptosis by fluorescent surface enhanced Raman spectroscopic dots. Bioconjug. Chem. 2007, 18(4), 1155–1162. 112. Akerman, M. E.; et al. Nanocrystal targeting in vivo. Proc. Natl. Acad. Sci. U.S.A. 2002, 99(20), 12617–12621. 113. von Maltzahn, G.; et al. In vivo tumor cell targeting with “click” nanoparticles. Bioconjug. Chem. 2008, 19(8), 1570–1578.
P1: OTA/XYZ P2: ABC c32 JWBS037-Chen January 25, 2011
810
13:56
Printer Name: Yet to Come
CURRENT STATUS AND FUTURE PROSPECTS FOR NANOPARTICLE
114. Bartlett, D. W.; et al. Impact of tumor-specific targeting on the biodistribution and efficacy of siRNA nanoparticles measured by multimodality in vivo imaging. Proc. Natl. Acad. Sci. U.S.A. 2007, 104(39), 15549–15554. 115. Farokhzad, O. C.; et al. Targeted nanoparticle-aptamer bioconjugates for cancer chemotherapy in vivo. Proc. Natl. Acad. Sci. U.S.A. 2006, 103(16), 6315–6320. 116. Gu, F.; et al. Precise engineering of targeted nanoparticles by using self-assembled biointegrated block copolymers. Proc. Natl. Acad. Sci. U.S.A. 2007, 105(7), 2586–2591. 117. Zhu, J.; et al. Biomimetic glycoliposomes as nanocarriers for targeting P-selectin on activated platelets. Bioconjug. Chem. 2007, 18(5), 1366–1369. 118. Weissleder, R.; et al. Cell-specific targeting of nanoparticles by multivalent attachment of small molecules. Nat. Biotechnol. 2005, 23(11), 1418–1423. 119. Kukowska-Latallo, J. F.; et al. Nanoparticle targeting of anticancer drug improves therapeutic response in animal model of human epithelial cancer. Cancer Res. 2005, 65(12), 5317–5324. 120. McBain, S. C.; Yiu, H. H.; Dobson, J. Magnetic nanoparticles for gene and drug delivery. Int. J. Nanomed. 2008, 3(2), 169–180. 121. Tran, M. A.; et al. Targeting V600EB-Raf and Akt3 using nanoliposomal-small interfering RNA inhibits cutaneous melanocytic lesion development. Cancer Res. 2008, 68(18), 7638–7649. 122. Breunig, M.; Bauer, S.; Goepferich, A. Polymers and nanoparticles: intelligent tools for intracellular targeting? Eur. J. Pharm. Biopharm. 2008, 68(1), 112–128. 123. Duan, H.; Nie, S.; Cell-penetrating quantum dots based on multivalent and endosome-disrupting surface coatings. J. Am. Chem. Soc. 2007, 129(11), 3333–3338. 124. Derfus, A. M.; et al. Targeted quantum dot conjugates for siRNA delivery. Bioconjug. Chem. 2007, 18(5), 1391–1396. 125. Sasaki, K.; et al. An artificial virus-like nano carrier system: enhanced endosomal escape of nanoparticles via synergistic action of pH-sensitive fusogenic peptide derivatives. Anal. Bioanal. Chem. 2008, 391(8), 2717–2727. 126. Sawant, R. M.; et al. “SMART” drug delivery systems: double-targeted pH-responsive pharmaceutical nanocarriers. Bioconjug. Chem. 2006, 17(4), 943–949. 127. Stover, T. C.; et al. Thermoresponsive and biodegradable linear-dendritic nanoparticles for targeted and sustained release of a pro-apoptotic drug. Biomaterials 2008, 29(3), 359–369. 128. Saito, G.; Swanson, J. A.; Lee, K. D. Drug delivery strategy utilizing conjugation via reversible disulfide linkages: role and site of cellular reducing activities. Adv. Drug Deliv. Rev 2003, 55(2), 199–215. 129. Hu, S. H.; et al. Magnetic-sensitive silica nanospheres for controlled drug release. Langmuir 2008, 24(1), 239–244. 130. Skirtach, A. G.; et al. Laser-induced release of encapsulated materials inside living cells. Angew. Chem. Int. Ed. Engl. 2006, 45(28), 4612–4617. 131. Park, H.; et al. Multifunctional nanoparticles for photothermally controlled drug delivery and magnetic resonance imaging enhancement. Small 2008, 2(4), 192–196. 132. Bagalkot, V.; et al. Quantum dot–aptamer conjugates for synchronous cancer imaging, therapy, and sensing of drug delivery based on bi-fluorescence resonance energy transfer. Nano Lett. 2007, 7(10), 3065–3070. 133. Tosi, G.; et al. Polymeric nanoparticles for the drug delivery to the central nervous system. Expert Opin. Drug Deliv. 2008, 5(2), 155–174. 134. Yang, J.; et al. Multifunctional magneto-polymeric nanohybrids for targeted detection and synergistic therapeutic effects on breast cancer. Angew. Chem. Int. Ed. Engl. 2007, 46(46), 8836–8839.
P1: OTA/XYZ P2: ABC c32 JWBS037-Chen January 25, 2011
13:56
Printer Name: Yet to Come
REFERENCES
811
135. Tada, H.; et al. In vivo real-time tracking of single quantum dots conjugated with monoclonal anti-HER2 antibody in tumors of mice. Cancer Res. 2007, 67(3), 1138–1144. 136. Yezhelyev, M. V.; et al. Proton-sponge coated quantum dots for siRNA delivery and intracellular imaging. J. Am. Chem. Soc. 2008, 130(28), 9006–9012. 137. Al-Jamal, W. T.; et al. Functionalized-quantum-dot-liposome hybrids as multimodal nanoparticles for cancer. Small 2008, 4(9), 1406–1415. 138. Al-Jamal, W.; Kostarelos, K. Liposome-nanoparticle hybrids for multimodal diagnostic and therapeutic applications. Nanomedicine 2007, 2(1), 85–98. 139. Tyner, K. M.; Schiffman, S. R.; Giannelis, E. P. Nanobiohybrids as delivery vehicles for camptothecin. J. Control Release 2004, 95(3), 501–514. 140. Nasongkla, N.; et al. Multifunctional polymeric micelles as cancer-targeted, MRI-ultrasensitive drug delivery systems. Nano. Lett. 2006, 6(11), 2427–2430. 141. Kim, J.; et al. Multifunctional uniform nanoparticles composed of a magnetite nanocrystal core and a mesoporous silica shell for magnetic resonance and fluorescence imaging and for drug delivery. Angew. Chem. Int. Ed. Engl. 2008, 47(44), 8438–8441. 142. Park, J. H.; et al. Micellar hybrid nanoparticles for simultaneous magnetofluorescent imaging and drug delivery. Angew. Chem. Int. Ed. Engl. 2008, 47(38), 7284–7288. 143. Sanvicens, N.; Marco, M.-P. Multifunctional nanoparticles—properties and prospects for their use in human medicine. Trends Biotechnol. 2008, 26(8), 425–433. 144. Nel, A.; et al. Toxic potential of materials at the nanolevel. Science 2006, 311(5761), 622–667. 145. Lovri´c, J.; et al. Differences in subcellular distribution and toxicity of green and red emitting CdTe quantum dots. J. Mol. Med. 2005, 83(5), 377–386. 146. Jia, G.; et al. Cytotoxicity of carbon nanomaterials: single-wall nanotube, multi-wall nanotube, and fullerene. Environ. Sci. Technol. 2005, 39(5), 1378–1383. 147. Zhang, Y.; et al. In vitro and in vivo toxicity of CdTe nanoparticles. J. Nanosci. Nanotechnol. 2007, 7(2), 497–503. 148. Connor, E. E.; et al. Gold nanoparticles are taken up by human cells but do not cause acute cytotoxicity. Small 2005, 1(3), 325–327. 149. Patra, H. K.; et al. Cell selective response to gold nanoparticles. Nanomedicine 2007, 3(2), 111–119. 150. Chan, W. H.; Shiao, N. H.; Lu, P. Z. CdSe quantum dots induce apoptosis in human neuroblastoma cells via mitochondrial-dependent pathways and inhibition of survival signals. Toxicol. Lett. 2006, 167(3), 191–200. 151. Lovric, J.; et al. Unmodified cadmium telluride quantum dots induce reactive oxygen species formation leading to multiple organelle damage and cell death. Chem. Biol. 2005, 12(11), 1227–1234. 152. Kroll, A.; et al., Current in vitro methods in nanoparticle risk assessment: Limitations and challenges. Eur. J. Pharm. Biopharm. 2009, 72(2), 370–377. 153. Chen, H. T.; et al. Cytotoxicity, hemolysis, and acute in vivo toxicity of dendrimers based on melamine, candidate vehicles for drug delivery. J. Am. Chem. Soc. 2004, 126(32), 10044– 10048. 154. Qian, X.; et al. In vivo tumor targeting and spectroscopic detection with surface-enhanced Raman nanoparticle tags. Nat. Biotechnol. 2008, 26(1), 83–90. 155. Voura, E. B.; et al. Tracking metastatic tumor cell extravasation with quantum dot nanocrystals and fluorescence emission-scanning microscopy. Nat. Med. 2004, 10(9), 993–998. 156. Ballou, B.; et al. Noninvasive imaging of quantum dots in mice. Bioconjug. Chem. 2004, 15(1), 79–86.
P1: OTA/XYZ P2: ABC c32 JWBS037-Chen January 25, 2011
812
13:56
Printer Name: Yet to Come
CURRENT STATUS AND FUTURE PROSPECTS FOR NANOPARTICLE
157. Gopee, N. V.; et al. Migration of intradermally injected quantum dots to sentinel organs in mice. Toxicol. Sci. 2007, 98(1), 249–257. 158. Chen, L. D.; et al. The biocompatibility of quantum dot probes used for the targeted imaging of hepatocellular carcinoma metastasis. Biomaterials 2008, 29(31), 4170–4176. 159. Mancini, M. C.; et al. Oxidative quenching and degradation of polymer-encapsulated quantum dots: new insights into the long-term fate and toxicity of nanocrystals in vivo. J. Am. Chem. Soc. 2008, 130(33), 10836–10837. 160. Niidome, T.; et al. PEG-modified gold nanorods with a stealth character for in vivo applications. J. Control. Release 2006, 114(3), 343–347. 161. Takahashi, H.; et al. Modification of gold nanorods using phosphatidylcholine to reduce cytotoxicity. Langmuir 2006, 22(1), 2–5. 162. Patil, M. L.; et al. Surface-modified and internally cationic polyamidoamine dendrimers for efficient siRNA delivery. Bioconjug. Chem. 2008, 19(7), 1396–1403. 163. Liu, W.; et al. Compact cysteine-coated CdSe(ZnCdS) quantum dots for in vivo applications. J. Am. Chem. Soc. 2007, 129(47), 14530–14531. 164. Soo Choi, H.; et al. Renal clearance of quantum dots. Nat. Biotechnol. 2007, 25(10), 1165– 1170. 165. Mundargi, R. C.; et al. Nano/micro technologies for delivering macromolecular therapeutics using poly(D,L-lactide-co-glycolide) and its derivatives. J. Control Release 2008, 125(3), 193–209. 166. Potineni, A.; et al. Poly(ethylene oxide)-modified poly(beta-amino ester) nanoparticles as a pH-sensitive biodegradable system for paclitaxel delivery. J. Control Release 2003, 86(2-3), 223–234. 167. Pang, Z.; et al. Preparation and brain delivery property of biodegradable polymersomes conjugated with OX26. J. Control Release 2008, 128(2), 120–127. 168. Meng, W.; et al. Uptake and metabolism of novel biodegradable poly (glycerol-adipate) nanoparticles in DAOY monolayer. J. Control Release 2006, 116(3), 314–321. 169. Torchilin, V. P. Recent advances with liposomes as pharmaceutical carriers. Nat. Rev. Drug Discov. 2005, 4(2), 145–160. 170. Moghimi, S. M.; Szebeni, J. Stealth liposomes and long circulating nanoparticles: critical issues in pharmacokinetics, opsonization and protein-binding properties. Prog. Lipid Res. 2003, 42(6), 463–478. 171. Lee, C. C.; et al. Designing dendrimers for biological applications. Nat. Biotechnol. 2005, 23(12), 1517–1526. 172. Gaucher, G.; et al. Block copolymer micelles: preparation, characterization and application in drug delivery. J. Control Release 2005, 109(1-3), 169–188. 173. van Vlerken, L. E.; Amiji, M. M. Multi-functional polymeric nanoparticles for tumour-targeted drug delivery. Expert Opin. Drug Deliv. 2006, 3(2), 205–216. 174. Polizu, S.; et al. Applications of carbon nanotubes-based biomaterials in biomedical nanotechnology. J. Nanosci. Nanotechnol. 2006, 6(7), 1883–1904. 175. Lu, A. H.; Salabas, E. L.; Sch¨uth, F. Magnetic nanoparticles: synthesis, protection, functionalization, and application. Angew. Chem. Int. Ed. Engl. 2007, 46(8), 1222–1244. 176. Huang, X.; et al. Gold nanoparticles: interesting optical properties and recent applications in cancer diagnostics and therapy. Nanomedicine 2007, 2(5), 681–693. 177. Medarova, Z.; et al. In vivo imaging of siRNA delivery and silencing in tumors. Nat. Med. 2007, 13(3), 372–377. 178. Zhang, Y.; So, M. K.; Rao, J. Protease-modulated cellular uptake of quantum dots. Nano Lett. 2006, 6(9), 1988–1992.
P1: OTA/XYZ P2: ABC c32 JWBS037-Chen January 25, 2011
13:56
Printer Name: Yet to Come
REFERENCES
813
179. Nam, J. M.; Thaxton, C. S.; Mirkin, C. A. Nanoparticle-based bio-bar codes for the ultrasensitive detection of proteins. Science 2003, 301(5641), 1884–1886. 180. Yang, L.; et al. Single chain epidermal growth factor receptor antibody conjugated nanoparticles for in vivo tumor targeting and imaging. Small 2009, 5(2), 235–243. 181. Stoeva, S. I.; Lee, J. S.; Thaxton, C. S.; Mirkin, C. A. Multiplexed DNA detection with biobarcoded nanoparticle probes. Angew. Chem. Int. Ed. Engl. 2006, 45(20), 3303–3306. 182. Merkoc¸i, A. Electrochemical biosensing with nanoparticles. FEBS J. 2007, 274(2), 310–316.
P1: OSO ind JWBS037-Chen
February 1, 2011
0:46
Printer Name: Yet to Come
INDEX
Acoustic impedance, 237 Acoustic pressure, 265 Acoustic signal, 233 Adipic dihydrazide, 171 Akagenite (-FeOOH), 297 Albumin microsphere (see Albunex, quantison), 245 Alkaline hydrolysis–condensation, 349 Alkyne–azide click coupling reaction, 468 Alternating magnetic field (AMF), 772 Aminopropyltriethoxysilane (APTES), 349 A-mode (amplitude modulation), 233 Amorphous silica shell, 292 Ampere’s law, 271 Angiogenesis image-guided drug delivery, 729 Angiogenesis imaging, 415 Angiogenesis molecular imaging, 729 Angiography, 285 Annihilation, 402 Antiangiogenic theranostics, 731 Anti-Stokes Raman scattering, 198 Apoferritin, 310 Apolipoprotein A-I (apoA-I), 343 Apoprotein, 435 Apoptosis, 790 Arginine–glycine–aspartate (RGD), 352 Aspect ratio, 643 Atherogenesis, 741 Atherosclerotic plaque, 256, 740 Atherosclerotic restenosis, 740 ATP-binding cassette (ABC), 436 Atomic force microscopy (AFM), 482 Atorvastatin, 738 Au–Fe3 O4 dimer, 536 Auger electron emitters, 761 Autofluorescence background, 218 Bacterial lipopolysaccharides, 317 Bacteriophage, 463 Balloon overstretch injury, 741 Bandgap fluorescence modulation, 221 Bandgap semiconductor material, 163 Bandpass filter (BPF), 651 Bimetallic ferrite nanoparticle, 289
Biodistribution and pharmacokinetics of nanoprobes blood half-life, 87 effect of shape, 89 effect of size, 86, 566 effect of surface characteristics, 566 effect of surface coating and charge, 90 enhanced permeability and retention (EPR), 86 physiologically based pharmacokinetic (PBPK) model, 89 reticuloendothelial system (RES), 87 Bioluminescence resonance energy transfer (BRET), 170 Biopolymers, 184 natural polymers, 184 pseudosynthetic polymers, 184 synthetic polymers, 184 Blood–brain barrier (BBB), 285 Blood clot, 255 Blooming effect, 297 B-mode (brightness modulation), 234 Boron neutron capture therapy (BNCT), 188 Bovine serum albumin (BSA), 343 Bpy (2,2 -bypyridine), 350 Broadband radiofrequency, 265 Bromine radiolabeling, 413 Bubble oscillation, 234 Buoyancy, 242 B-Z transition, 225 Calf pulmonary artery endothelial (CPAE) cells, 352 Cancer theragnosis, 541 Capacitance chemical sensors, 485 Carbon quantum dot, 173 Carbon nanotubes (CNTs), 63–66, 78, 346, 784 direct adsorption, 65–66 DNA–CNT hybrids, 66 single strand DNA (ssDNA), 65 sugar–phosphate backbones, 65 direct surface modification, 63 ligand addition, 63–65 surfactant, 63 - interaction, 63 PEGlyated phospholipids, 65
Nanoplatform-Based Molecular Imaging Edited by Xiaoyuan Chen C 2011 John Wiley & Sons, Inc. Copyright
815
P1: OSO ind JWBS037-Chen
February 1, 2011
816
0:46
Printer Name: Yet to Come
INDEX
Carbon nanotubes (CNTs) (Continued ) multibeam filed emission X-ray (MBFEX), 144 single wall nanotubes (SWNTs), 63 Carcinoembryonic antigen (CEA), 295 Cardiovascular blood flow, 234 Catalytic growth, 296 CB-DO2A, 406 CB-TE2A, 406 Cell-penetrating peptide, 315 Cell therapy, 285 Cell tracking, 285 Central processing unit (CPU), 235 Cetyl trimethylammonium bromide (CTAB), 292, 346, 664 Chemical exchange dependent saturation transfer (CEST), 312 Chemical vapor deposition (CVD), 203, 482 Chemisorptions, 668 Chlorotoxin, 296 Cholesterol, 437 Cholesterol ester transfer protein (CETP), 437 Chylomicron, 434 Click chemistry, 49 alkyne, 49 azide, 49 azide–alkyne Huisgen cycloaddition, 49 1,3-dipolar cycloaddition, 49 1,2,3-triazole, 49 Clinical translation of nanoprobes Aurimune (CYT-6091), 85, 92 Combidex (Ferumoxtran-10, AMI-227, BMS-180549), 85 Feridex (Ferumoxide), 84 Ferumoxyltol, 85 Gastromark (Ferumoxsil, AMI-121), 85 Resovist (Ferucarbotran, SHU-555A), 85 Supravist (Ferucarbotran, SHU-555C), 85 VSOP-C184, 85 Cobalt ferrite, 292 Collimated lens assembly (CLA), 651 Color Doppler, 234 Combidex, 282 Combined photoacoustic and magnetoacoustic imaging, 273 Complementary DNA (cDNA), 221 Composite nanoparticles, 67–68 dumbbell, 68 nanocomposite, 67 Computed tomography (CT), 107–114 acceleration potential, 109 bremsstrahlung radiation, 109 Beer–Lambert law, 110 cardiac CT, 120–121 Compton scattering, 110 cone-beam CT (CBCT), 119–120 dual energy CT, 125–126 dual modality CT, 124–125
evolution of CT, 114–120 helical CT, 118 linear attenuation coefficient, 111 mass attenuation coefficient, 111 mathematics of CT, 113–114 medical applications, 131–134 CT angiography, 133–134 diagnosis, 131–132 digitally reconstructed radiograph (DRR), 132 Nyquist–Shannon criterion, 134 radiation therapy, 132–133 micro-CT, 121–124 multidetector CT (MDCT), 118–119 nanotube X-ray, 116–117 photoelectric absorption, 110 physics of X-ray, 108–112 power and data transmission, 118 P-transform, 112 random transform, 112, 113 Rayleigh scattering, 110 scattering removal and noise reduction in CBCT, 120 X-ray tube, 108 Computed tomography-contrast media and molecular CT, 126–131 aromatic iodinated contrast agents, 126 barium sulfate suspensions, 126 fullerenes, 127–128 bismuth sulfide nanoparticles (BPNPs), 128 dysprosium (Dy), 128 erbium (Er), 128 europium (Eu), 128 gadolinium (Gd), 128 lutetium (Lu), 128 N1177, 128 gold nanoparticles, 127 gold nanorods, 126 Composite organic–inorganic nanoparticles (COINs), 206 Conjugation with crosslinkers, 50 (bis(sulfosuccinimidyl) suberate)(BS3), 50 heterobifunctional crosslinker, 50 homobifunctional crosslinker, 50 4-maleimidobutyric acid N-hydroxysuccinimide ester, 50 sulfo-succinimidyl-(4-N-maleimidomethyl) cyclohexane-1-carboxylate (sulfo-SMCC), 50 Contrast, 280 Contrast agent, 280 Contrast enhance ultrasound, 237 Copper-64, 404 Core–satellite heterostructures, 536 Core–shell nanostructures, 534 Cowpea chlorotic mottle virus (CCMV), 463 Crosslinked iron oxide (CLIO), 283 Cross-wafer variation, 389 Cyclic–(arginine-glycine-aspartic acid) RGD, 188
P1: OSO ind JWBS037-Chen
February 1, 2011
0:46
Printer Name: Yet to Come
INDEX
Dark-field microscopy, 672 Decay half-life, 403 Dendrimers, 187, 759, 784 divergent synthesis, 187 convergent synthesis, 187 generation, 187 polyamidoamine (PAMAM), 187 Density of state (DOS), 199 Detoxification, 667 Dextran sulfate, 317 DiamSar, 406 Diethylenetriaminepentaacetic acid (DTPA), 334 Diffuse optical spectroscopy (DOS), 650 2,3-Dimercaptosuccinic acid (DMSA), 291 Dimyristoyl-sn-glycero-3-phosphoethanolamine (DMPE), 342 Dipalmitoylglycerosuccinate (DPSG), 344 Dipalmitoylphosphatidylethanolamine (DPPE), 344 Discrete dipole approximation (DDA), 624 Dispersion, 242 with aid of sonication, 242 high-shear mixing (amalgamation), 242 Dithiolated diethylenetriaminepentaacetic acid (DTDTPA), 533 DNA hybridization, 221 DNA sensor, 221 Dopamine, 291 Doped quantum dots (d-dots), 172 Doppler effect, 234 Doppler ultrasonography, 234 Dotarem, 334 DOTA (1,4,7,10-tetraazacyclododecane-1,4,7, 10-tetraacetic acid), 334 Double-disk structures, 381 Doxorubicin, 294 DPPC (1,2-sn-glycero-3-phosphatidylcholine), 342 Drug-eluting stents (DES), 740 Duplex ultrasonography, 234 Einzel type electrostatic lens, 145 Electrochemical immunosensor, 797 Electrochemical sensors, 484 Electron field emission, 142 Electronic transition energy, 199 Embryonic stem cells (ESCs), 320 Emitter, 402 ␥ -Emitting radionuclides, 403 Endocytosis, 317 Endogenous chromophores, 643 Endoventricular electromechanical mapping, 322 Endrem, 282 Enhanced permeability and retention (EPR) effect, 334, 641 Epichlorohydrin, 290 Epidermal growth factor receptor (EGFR), 296, 311 Epitaxial deposition, 174
817
Excited vibrational state, 198 Experimental autoimmune encephalomyelitis (EAE), 285, 319 Fast gradient echo (FGE), 321 Fe3 O4 –CdSe heterodimer, 297 FePt–Au, 296 FePt-Au dimmer, 535 Feridex, 282 Fermi level, 221 Ferritin (FT), 310 Ferritin cage, 463 Ferrofluid magnetic refrigeration, 286 Ferromagnetic coupling, 283 Ferroxidase activity, 310 Field-effect transistor (FET) sensors, 482 Figure of merit (FOM), 219 Flow sensor, 487 Fluorescein isothiocyanate (FITC), 188 Fluorescence recovery, 221 Fluorescence reflectance imaging (FRI), 602 Fluorescence resonance energy transfer (FRET). 169, 795 Fluorescent silica nanoparticles, 349 Fluorimetric assay, 218 Fluorimetric glucose sensor, 219 Fluorine radiolabeling, 414 Folate (or vitamin B9), 311 Folic acid (FA), 188 Four-dimensional (4D) ultrasound, 235 Fucose sulfate (fucoidan), 317 Full-width at half-maximum (FWHM), 197 Fumagillin, 730 Functional MRI (fMRI), 285 Gadofullerene, 340, 346 Gadolinium carbonate, 366 Gadolinium (Gd3+ ) chelate, 279 Gadolinium diethylenetriamine pentaacetic acid bismethylamide (gadodiamide), 344 Gadolinium fluoride (GdF3 ), 300, 365 Gadolinium oxide (Gd2 O3 ), 300 Gadolinium oxide nanoparticles, 361 Gadolinium phosphate (GdPO4), 300 Gadonanotube, 346 Gadophrin-2, 324 Gadoteridol, 350 Gallium radionuclide, 408 Galvanic replacement reaction, 616 Gas-filled contrast microbubble, 234 Gd-DTPA, 279 Genetic disorder, 223 Genomic mutation, 223 Glucose oxidase, 221 Glycol chitosan (GC), 185 Goeppert–Mayer units, 167 Gold nanocages, 60, 615
P1: OSO ind JWBS037-Chen
February 1, 2011
818
0:46
Printer Name: Yet to Come
INDEX
Gold nanocubes, 60, 616 Gold nanoparticles (GNPs), 60–62, 78, 760, 784 cetrimonium bromide (CTAB), 61 layer-by-layer deposition, 62 poly(diallyldimethyl-ammoniumchloride) (PDADMAC), 62 poly(sodium-4-styrenesulfonate) (PSS), 62 thiolated DNA oligos, 60 Gold nanoplates, 616 Gold nanorods (GNRs), 60, 659, 616 Gold nanospheres, 60, 616 Gold nanowires, 60, 616 GPIIb/IIIa receptors, 256 Gradient index (GRIN), 653 Green fluorescent protein (GFP), 321 Ground state, 198 Gruneisen coefficient, 266 HaloTag protein (HTP), 166 Hard-shell microbubble, 242 Harmonic imaging, 234 Harmonic frequency (see overtone, undertone), 234 Heat capacity, 265 Heat shock protein (Hsp) cage, 468 Heat-up process, 287 Hematopoietic progenitor cells, 320 Hematopoietic stem cells, 311 Heterostructural nanoparticle, 535 Hidden Markov modeling (HMM), 227 High-pressure carbon monoxide (Hipco), 200 High-resolution gratings, 221 Highest occupied molecular orbital (HOMO), 204 High-density lipoprotein (HDL), 343, 434 High performance liquid chromatography (HPLC), 344 HIV Tat protein, 315 Hollow magnetic cylinders, 381 Hollow oxide nanoparticles, 302 Hot-injection, 287 Huisgen 1,3-dipolar cycloaddition, 597 Human epidermal growth factor receptor-2 (Her-2), 188, 295 Human mesenchymal stem cells (hMSCs), 292 Hybridoma, 319 Hydrogel, 320 Hyperthermic radiosensitization, 646 Hyper-Rayleigh scattering, 662 Hypsochromic shift, 221 Iliac lymph node, 241 Imaging techniques, 6 computed tomography, 13 magnetic resonance imaging, 11–12 optical imaging, 8–9 radionuclide-based imaging techniques, 9–11 ultrasound imaging, 12–13 Indium-111, 407
Inducible NO synthase (iNOS), 225 Inductively coupled plasma mass spectrometry (ICP-MS), 649 Inner sphere, 345 Integrin ␣v 3 imaging, 416, 603 Intercellular adhesion molecule-1 (ICAM-1), 254 Interdigitation-fusion method, 698 Intimal endothelial healing, 742 Intravital fluorescence microscopy (IVFM), 602 Investigational new drug (IND), 726 In water in oil (W/O) microemulsion, 350 Iodine radiolabeling, 412 Iron oxide, 55–60 co-precipitation, 55–56 Combidex, 56 crosslinked iron oxide (CLIO), 56 dendrimer, 55 dextran, 55 Feridex, 56 PEG-silane copolymer, 56 PEGylated silane, 56 polyaniline, 55 polyaspartic acid (PASP), 56 poly(TMSMA-r-PEGMA), 56 polyvinylpyrrolidone (PVP), 55 Resovist, 56 pyrolysis, 56–60 Fe(acac)3, 56 Fe(CO)5, 56 Fe-oleateare, 56 ligand addition, 57 ligand exchange, 58 polymer and protein adsorption, 59 Iron oxide cell labeling, toxicity, 318 Isotopically modified SWNTs, 203 Keyboard cursor, 235 Kuppfer cells, 316 Larmor frequency, 280 Laser fluence, 265 Lattice-mismatch, 174 Layer-by-layer (LbL), 351 Lecithin-cholesterol acyltransferase (LCAT), 435 Lift-off resist (LOR), 391 Linear (resonant) oscillation, 234 Lipofectamine, 314 Lipoprotein, 433 for cancer imaging, 438 computed tomography, 449 drug interactions with, 437 metabolism, 436 MRI, 445 optical imaging, 444 probe incorporation, 438 covalent modification, 438 core-loading “reconstitution”, 440 intercalation, 440
P1: OSO ind JWBS037-Chen
February 1, 2011
0:46
Printer Name: Yet to Come
INDEX
radiolabeling, 441 rerouting, 451 Liposomal transfection reagent, 314 Liposomes, 189, 759, 784 incorporation of molecules, 189 RES, 189 Localized surface plasmon resonance (LSPR), 615 Longitudinal relaxation (see T1 relaxation), 280 Longitudinal resonance (LR), 660 Low-density lipoprotein (LDL), 316, 434 Lower critical solution temperature (LCST), 710 Lowest unoccupied molecular orbital (LUMO), 204 Luteinizing hormone/chorionic gonadotropin (LH/CG), 311 Luteinizing hormone releasing hormone (LHRH), 311 Lymphocyte trafficking, 319 MS2 capsid, 466 Macrophage, 285 Maghemite (␥ -Fe2 O3 ), 282 Magic bullet, 400 Magnetic flux, 269 Magnetic force, 269 Magnetic iron oxide nanoparticle (MION), 272 Magnetic labeling, 309 in vivo applications inflammation, 322 kidney injuries (nephropathy), 323 atherosclerotic plaque, 323 tumor, 323 Magnetic moment, 269 Magnetic nanoparticles, 79, 530, 758, 784 peptides, 598 small molecules, 599 targeting, 598 Magnetic relaxation switch, 286 Magnetic resonance imaging (MRI), 279 Magnetic resonance thermal imaging (MRTI), 643 Magnetic susceptibility, 270 Magnetite (Fe3 O4 ), 282 Magnetoacoustic imaging, 264 Magnetodendrimer (MD-100), 313 Magnetoelectroporation (MEP), 316 Magnetoferritin, 470 Magnevist, 334 Maltose binding protein (MBP), 169 Manganese-enhanced MRI (MEMRI), 299 Manganese ferrite (MnMEIO), 295 Manganese oxide (MnO), 300 Marrow stromal cells, 320 Matrix metalloproteinase-2 (MMP-2), 311 Mercaptopropyltrimethoxysilane (MPTMS), 355 Mesoporous nanoparticles, 346 Metallic nanomaterial, 33 Ag nanoparticles, 33–34 Au nanoparticles, 35
819
Rayleigh (Mie) scattering, 33 surface plasmon resonance (SPR), 33 Metallofullerenes, 344 Metal-oleate complex, 288 Metal organic framework (MOF), 534 Metal oxide nanoparticles, 28 gadolinium oxide, 29 iron oxide, 28–33 crystalline maghemite (␥ -Fe2 O3 ), 29 magnetite (Fe3 O4 ), 28 metal-doped iron oxide, 29 one-pot synthesis, 33 seed-mediated growth, 29 superparamagnetic, 28 thermal decomposition, 29 manganese oxide, 29 Methotrexate (MTX), 188 Microbubbles biodistribution, 245 synthesis, 242 Microcomputed tomography (micro-CT), 139 limitations, 140 signal-to-noise ratio (SNR), 140 Micro-CT, dynamic Charybdis scanner, 149 Cyclops system, 147 gating, 150–153 cardiac gating, 151 free-breathing mice, 150 respiratory gating, 150 Microdialysis capillary, 222 Micrometer-sized paramagnetic iron oxide (MPIO), 282 Molecular beacon, 184 Molecular imaging, definition, 5 Molecular imaging probes, 13 challenges for nanoparticle-based imaging probes, 19–21 nanoparticle-based probes, 15–19 Molecular magnetic resonance imaging (mMRI), 545 Molecular scale lithography bottom-up approach, 287 top-down approach, 287 Molecular ultrasound, 240 Molecular ultrasound imaging thrombus formation, 255 lymph node, 256 tumor angiogenesis, 250 VEGFR2, integrin, endoglin, leukocyte, P-selectin, MadCAM-1, VCAM-1, inflammation, 255 ischemia, arthrosclerosis, inflammatory bowel disease, ICAM-1, VCAM-1, 254 Monocyte, 285 Mononuclear phagocytic system (MPS), 764 Mucosal address in cellular adhesion molecule-1 (MAdCAM-1), 255
P1: OSO ind JWBS037-Chen
February 1, 2011
820
0:46
Printer Name: Yet to Come
INDEX
Multicolor Raman imaging, 203 Multifunctional nanomaterials, 35–41 core-shell, 36 heterodimer, 36 yolk-shell, 39 Multifunctional nanoparticles, 571, 801 clinical translation, 584 gold nanoparticles/nanoshells, 580 quantum dots, 576 nanoparticles embedded in polymers or liposomes, 581 superparamagnetic iron oxide nanoparticles (SPIONs), 571 surface modification, 595 conventional bioconjugation strategies, 597 orthogonal bioconjugation strategies, 597 synthesis, 595 Multimodal nanoparticle, 599 Multiphoton confocal microscopy, 672 Multispectral contrast agent, 378 Multiwalled carbon nanotube (MWNT), 482 Nanohybrid, 536, 802 Nanoparticles softness and hardness, 570 tumor targeting principles, 569 Nanoparticles for cancer theragnosis polymeric multifunctional, 547 multifunctional magnetic, 551 theragnostic inorganic, 553 Nanoparticle toxicity, 803 Nanoprobe toxicity carbon nanotubes, 83–84 gold nanoparticles, 82–83 magnetic nanoparticles, 84–85 quantum dots, 80–82 reactive oxygen species, 81 Narrowband imaging (NBI), 650 Neutron activation assay (NAA), 649 Nitric oxide (NO), 225 Nd:YAG laser, 698 NMRD, 345 Nobel metal nanoparticle, 533 Nominal sensor lifetime, 218 Noncomplementary DNA (nDNA), 221 Nonlinear (resonant) oscillation, 234 Nonphotobleaching sensor, 218 Nontargeted contrast-enhanced ultrasound, applications, 237 assessment of blood volume and perfusion, 238 diagnosing liver diseases, 238 edge delineation in the heart, 237 Notch filter, 221 Nuclear magnetic resonance (NMR), 280 Nuclear magnetic resonance dispersion (NMRD), 467
n-Octanol–water partition coefficient (logP), 437 Oligodendrocyte progenitors, 320 Optical absorption coefficient, 265 Optical coherence tomography (OCT), 615, 673 Optical properties, 161–162 animal imaging applications, 166 nontargeted animal imaging, 167 targeted animal imaging, 168 production, 162 biofunctionalization, 164 covalent binding, 165 direct linkage, 165 electrostatic interaction, 165 core synthesis, 162 water-solublization, 163 One-dimensional quantum wire, 199 Opsonin, 317 Organometalic precursor, 287 Outer sphere, 345 Overtone (see undertone, harmonic frequency), 234 Paramagnetic chemical exchange saturation transfer (PARACEST), 312 Paramagnetic complex, 279 Paramagnetic complexes in the membrane, 342 Paramagnetic gold nanoparticles, 358 Paramagnetic quantum dots, 354 Paramagnetic relaxation enhancement (PRE), 465 ␣-Particle emitters, 761 -Particle emitters, 761 ␥ -Particle emitters, 761 Pancreatic ductal adenocarcinoma (PDAC), 601 Pancreatic islets, 322 Perfluorocarbon (PFC), 725 Perfluorocarbon (PFC) emulsion, 246 PET/MRI, 531 PET radionuclide, 403 Phagocytic cells, 285 Phagocytosis, 317 Phosphatidyl ethanolamine, 342 Phosphine oxide, 291 Photoabsorber, 264 Photoacoustic imaging, 264 Photoacoustic tomography (PAT), 674 Photobleaching, 218 Photo-oxidation, 163 Photostability, 218 Photoswitchable, 185 Photothermal agents, 674 Photothermal therapy (PTT), 631, 642 Photoluminescence (PL), 660 pH-sensitive liposomes, 344 Plasmonic photothermal therapy (PPTT), 643 Plasmonic resonance, 267 Plasmon-resonant heating, 663 Plectin-1, 601 Polyactic acid (PLA) nanobubble, 248
P1: OSO ind JWBS037-Chen
February 1, 2011
0:46
Printer Name: Yet to Come
INDEX
Polymeric micelles, 784 Polymeric nanoparticles, 784 Polyamidoamine (PAMAM), 769 Polychelating amphiphilic polymer, 342 Polyethylene glycol (PEG), 291 Poly(ethylene glycol)-block-poly(d,l-lactide) (PEG-PLA), 294 Poly(glycolide) (PGA), 294 Polyinosinic acid, 317 Poly(d,l-lactide) (PLA), 294 Poly(d,l-lactide-co-glycolide) (PLGA), 294 Polystyrenesulfonate (PSS), 351 Polyvinyl alcohol (PVA), 273 Popliteal lymph node, 241 Porous polymerosomes, 342 Positron emitters, 761 Power Doppler, 234 Predisposition, 223 Pressure sensor, 487 Programmed cell death (PCD), 531 ProHance, 311 Prostate-specific membrane antigen (PSMA), 168, 311 Protease, 184 protease sensing, 184 proteolytic cleavage, 184 self-quenching NIR dye, 185 Protein cage, 463 atherosclerotic plaque imaging, 471 biofilm infection imaging, 473 T1 contrast agent, 465 T2 contrast agent, 469 Pulse inversion imaging, 234 Q-switched Nd:YAG laser, 274 Quadrupole-coupled paramagnetic doublet, 283 Quantum dots, 26, 51–55, 76–78, 531, 760, 784 cytotoxicity, 28 deeper tissue penetration, 175 dimethylcadmium (Cd(CH3 )2 ), 27 in vivo molecular imaging, 159 less RES uptake, 174–175 ligand addition, 53–54 ligand exchange, 52–53 mammalian cell delivery, 168 microinjection, 168 nonspecific pinocytosis, 168 peptide-induced transport, 168 multimodality, 174 non-Cd-based QDs, 55 organometallic procedure, 27–28 protein encapsulation, 55 selenium, 27 structure, 160–161 tri-n-octylphosphine (TOP), 27 tri-n-octylphosphine oxide (TOPO), 27 Quantum yield, 163, 218 Quartz crystal microbalance (QCM), 486
821
Radial breathing mode (RBM), 200 Radiation dose enhancement, 644 Radioimmunonanoparticle, 755 Radiolabeled nanoplatform, 415, 763 Radiometric assay. 402 Raman active dye, 206 fluorescent dyes, 206 4-mercaptobenzoic acid, 206 naphthalenethiol, 206 pyridine, 206 thiophenol, 206 Raman molecular imaging, 197 Rayleigh scattering, 198 RAW 264.7 macrophage cell, 225 Reactive oxygen species (ROS), 228 Receptor-mediated endocytosis, 310 Receptor saturation, 310 Reendothelialization, 740 Renishaw Raman microscope, 202 Resistance chemical sensors, 484 Resonance Raman scattering, 199 Resonant chemical sensors, 486 Resovist, 282 Reverse microemulsion, 350 RGD-4C, 472 Rhenium radiolabeling, 409 Rhodamine B isothiocyanate (RBITC), 349 RNA interference (RNAi), 490 Scanning electron microscopy (SEM), 482 Scanning tunneling microscopy (STM), 482 Scavenger receptor (SR), 316, 437 Second harmonic generation, 662 Selectin, 256 Self-illuminating QDs, 171 Sentinel lymph nodes (SLNs), 531 Shimming, 281 Silanol, 350 Silica nanoparticles, 66–67, 533 aminopropyltrimethoxysilane (APS), 66 mercaptopropylmethoxysilane (MPS), 66 tetraethyl orthosilicate (TEOS), 66 Silica sol-gel reaction, 292 Silver nanocubes, 617 Sinerem, 282 Single nucleotide polymorphism (SNP), 223 Single-walled carbon nanotubes (SWNTs), 197, 481, 760 biomedical instrumentation, 503 drug delivery, 491 fluorescence imaging, 497 gene delivery, 489 MRI, 495 photoacoustic tomography (PAT), 498 radiolabeling, 499 Raman spectroscopy, 497 tissue engineering, 500
P1: OSO ind JWBS037-Chen
February 1, 2011
822
0:46
Printer Name: Yet to Come
INDEX
Single-walled carbon nanotubes (SWNTs) (Continued ) cell based studies, 500 matrix enhancement, 501 thermotherapy, 494 toxicity, 505 vaccine delivery, 493 Sinusoidal echo, 234 Size exclusion chromatography, 283 Small interfering RNAs (siRNAs), 490 SNP sensor, 223 SOD, 406 Sodium cyanoborohydride, 49 reductive amination, 49 Sodium polystyrenesulfonate, 667 Soft-shell microbubble, 242 Solvothermal or thermal decomposition, 286 Spectroscopic coherence tomography (SOCT), 615 Spin dephasing, 281 Spin–spin relaxation, 281 Stable lipoprotein mimetic micelles, 343 Steady-state free precession (SSFP), 321 Stealth environment, 183 Stealth liposomes, 338 Stearylamine, 342 Stem cells, 285 Stent–intima interface, 740 Stimulated acoustic emission (SAE), 256 St¨ober process, 669 Stochastic biosensor, 227 Stokes Raman scattering, 198 Stokes Raman shift, 198 Streptavidin–Biotin interaction, 50–51 Avidin, 50 Biotin, 50 NeutrAvidin, 51 Streptavidin, 50 Superparamagnetic iron oxide (SPIO), 279, 282 Surface defect, 163 Surface enhanced Raman scattering (SERS), 197, 662 long-range electromagnetic effect, 204 short-range chemical effect, 204 Surface plasmon resonance (SPR), 684 Surface micromachining, 379 Susceptibility artifact, 297 T1-weighted image, 280 T2-weighted image, 280 Tangential G-band, 200 Targeted ultrasound contrast agent, 240 Tat peptide, 315 Technetium radiolabeling, 409 TETA, 406 Tetrabutylammonium hydroxide (TBAH), 344
Tetraethyl orthosilicate (TEOS), 346 Theragnostic nanoprobes, 92 5-(4-carboxyphenyl)-10,15,20-tryphenyl-2,3dihydroxychlorin (TPC), 93 hexanoyl-ceramide (Cer-6), 93 Thermal effect of ultrasound, 235 Thermal expansion coefficient, 265 Theranostic, 401 Thotoacoustic tomography (PAT), 615 Three-dimensional (3D) ultrasound, 235 Thrombosis, 256 Thrombus, 255 Time-of-flight angiography, 741 Tissue regeneration, 320 Tissue transparency, 218 TNP-470, 730 Top–down microfabrication, 378 Total heat dose (THD), 772 Transchelation, 406 Transducer pulse control, 235 Transferrin receptor (TfR), 310 Transmission electron microscopy (TEM), 482 Transplant rejection, 285 Transverse relaxation (see T2 relaxation), 280 Transverse resonance (TR), 660 Triacylglycerols (TAGs), 433 3-(trihydroxysilyl)propyl methylphosphonate (THPMP), 350 N-(trimethoxysilylpropyl) ethyldiamine (TSPETE), 350 Trioctylphosphine/trioctylphosphine oxide (TOP/TOPO), 355 Triblock copolymer, 185 Tumor vasculature imaging, 602 Two-photon autofluorescence (TPAF), 652 Two-photon excited luminescence (TPL), 659 Two-photon induced photoluminescence (TPIP), 652 Type 1 diabetes mellitus, 322 Type II QDs, 168 Ultrasmall superparamagnetic iron oxide (USPIO), 282 Ultrasound backscattering, 237 Ultrasound contrast agent solid nanoparticles, 248 liposomes, 249 Ultrasound imaging, 233 equipment, 235 safety and side effects, 235 general diagnostic indications, 236 obstetrics and gynecology, 236 cardiovascular imaging, 236 solid organs in the abdomen, 236 thyroid, 236 musculoskeletal radiology, 236 emergency room, 236
P1: OSO ind JWBS037-Chen
February 1, 2011
0:46
Printer Name: Yet to Come
INDEX
Ultrasound transducer, 235 Undertone (see overtone, harmonic frequency), 234
X-ray carbon nanotube field emission, 142–147 conventional thermionic, 141–142
Valence band, 221 van Hove singularities (VHS), 197 Vascular endothelial growth factor (VEGF), 188 Vascular cell adhesion molecule-1 (VCAM-1), 254 VEGFR-2 imaging, 419 Very low density lipoproteins (VLDL), 434 Virtual state, 198 Volumetric magnetization, 270
Yttrium radiolabeling, 409
Wavelength window, 219 Wrap–bake–peel process, 297
823
Zero-length crosslinking, 49–50 carboxylate-amine coupling, 49 N,N -dicyclohexylcarbodiimide (DCC), 49 1-ethyl-3-(3-dimethylaminopropyl) carbodiimide hydrochloride (EDC), 49 N-hydroxysuccinimide (NHS), 49 N-hydroxysulfosuccinimide (sulfo-NHS), 49 Schiff base, 49 Zeta potential, 313 z-Spectra, 384