Block Copolymers in Solution: Fundamentals and Applications
Block Copolymers in Solution: Fundamentals and Applications IAN HAMLEY University of Reading, Reading, UK
John Wiley & Sons, Ltd
Copyright © 2005
John Wiley & Sons Ltd, The Atrium, Southern Gate, Chichester, West Sussex PO19 8SQ, England Telephone
(+44) 1243 779777
Email (for orders and customer service enquiries):
[email protected] Visit our Home Page on www.wileyeurope.com or www.wiley.com All Rights Reserved. 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 under the terms of the Copyright, Designs and Patents Act 1988 or under the terms of a licence issued by the Copyright Licensing Agency Ltd, 90 Tottenham Court Road, London W1T 4LP, UK, without the permission in writing of the Publisher. Requests to the Publisher should be addressed to the Permissions Department, John Wiley & Sons Ltd, The Atrium, Southern Gate, Chichester, West Sussex PO19 8SQ, England, or emailed to
[email protected], or faxed to (+44) 1243 770620. Designations used by companies to distinguish their products are often claimed as trademarks. All brand names and product names used in this book are trade names, service marks, trademarks or registered trademarks of their respective owners. The Publisher is not associated with any product or vendor mentioned in this book. This publication is designed to provide accurate and authoritative information in regard to the subject matter covered. It is sold on the understanding that the Publisher is not engaged in rendering professional services. If professional advice or other expert assistance is required, the services of a competent professional should be sought. Other Wiley Editorial Offices John Wiley & Sons Inc., Ill River Street, Hoboken, NJ 07030, USA Jossey-Bass, 989 Market Street, San Francisco, CA 94103-1741, USA Wiley-VCH Verlag GmbH, Boschstr. 12, D-69469 Weinheim, Germany John Wiley & Sons Australia Ltd, 42, McDougall Street, Milton, Queensland 4064, Australia John Wiley & Sons (Asia) Pte Ltd, 2 Clementi Loop #02-01, Jin Xing Distripark, Singapore 129809 John Wiley & Sons Canada Ltd, 22 Worcester Road, Etobicoke, Ontario, Canada M9W 1L1 Wiley also publishes its books in a variety of electronic formats. Some content that appears in print may not be availabe in electronic books. Library of Congress Cataloging-in-Publication Data Hamley, Ian W. Block copolymers in solution : fundamentals and applications/Ian Hamley P. cm. Includes bibliographical references and index ISBN-13: 978-0-470-01557-5 (acid-free paper) ISBN-10: 0-470-01557-8 (acid-free paper) 1. Block copolymers. I. Title QD382.B5H355 2005 547'.84-dc22
2005005799
British Library Cataloguing in Publication Data A catalogue record for this book is available from the British Library ISBN: 0-470-01557-8 (HB); 9-78-0-470-01557-5 (HB) Typeset in 10/12pt Times by Thomson Press (India) Limited, New Delhi Printed and bound in Great Britain by TJ International Ltd., Padstow, Cornwall This book is printed on acid-free paper responsibly manufactured from sustainable forestry in which at least two trees are planted for each one used for paper production.
To Valeria and Lucas
Contents Preface
xi
1. Introduction
1
References
5
2. Neutral Block Copolymers in Dilute Solution
7
2.1 Introduction 2.2 Techniques for Studying Micellization 2.2.1 Cryo-TEM 2.2.2 Differential Scanning Calorimetry 2.2.3 Dynamic Light Scattering 2.2.4 Ellipsometry 2.2.5 Fluorescence Probe Experiments 2.2.6 Nuclear Magnetic Resonance 2.2.7 Rheology 2.2.8 Scanning Probe Microscopy 2.2.9 Small-angle X-ray and Neutron Scattering 2.2.10 Static Light Scattering 2.2.11 Surface Pressure-Area Isotherms 2.2.12 Surface Tensiometry 2.2.13 Viscometry 2.2.14 X-ray and Neutron Reflectivity 2.3 Micellization in PEO-based Block Copolymers 2.4 Micellization in Styrenic Block Copolymers 2.5 Determination of cmc 2.6 Thermodynamics of Micellization 2.6.1 Chain Length Dependence of Micellization 2.6.2 Effect of Architecture 2.6.3 Effect of Solvents and Salts on Micellization 2.7 Micellization and Micelle Dimensions: Theory and Simulation 2.7.1 Scaling Models 2.7.2 The Brush Model 2.7.3 The Self-consistent Mean Field Theory 2.7.4 The Model of Nagarajan and Ganesh 2.7.5 Computer Simulations
7 7 7 8 8 10 10 10 11 11 12 14 16 16 17 17 18 20 20 22 25 27 32 33 33 37 40 43 44
viii
Contents 2.7.6 Theory: ABC Triblock Micelles 2.8 Micelle Dimensions: Comparison Between Experiment and Theory 2.9 Interaction between Micelles 2.10 Dynamics of Micellization 2.11 Dynamic Modes 2.12 Specific Types of Micelles 2.12.1 Micelles from Telechelics 2.12.2 Micelles from ABC Triblocks 2.12.3 Micelles from Rod-Coil Copolymers 2.12.4 Cross-linked Micelles 2.12.5 Janus Micelles 2.12.6 Nonspherical Micelles 2.12.7 Micelles Formed due to Specific Interactions 2.13 Micellization in Mixed Solvents 2.14 Mixed Micelles 2.15 Block Copolymer/Surfactant Complexes 2.16 Complex Morphologies 2.17 Vesicles 2.18 Crystallization in Micelles References
3. Concentrated Solutions 3.1 Understanding Phase Diagrams 3.2 Phase Behaviour of PEO-containing Block Copolymers 3.3 Gelation 3.3.1 Rheology 3.3.2 Structure - Packing of Micelles 3.3.3 Thermodynamics of Gelation and Micellization in Concentrated Solution 3.3.4 Effect of Added Homopolymer, Salt or Surfactant 3.3.5 Influence of Architecture 3.4 Order-Disorder Phase Transition 3.5 Order-Order Phase Transitions 3.5.1 Structural Aspects 3.5.2 Ordering Kinetics 3.6 Domain Spacing Scaling, and Solvent Distribution Profiles 3.7 Semidilute Block Copolymer Solution Theory 3.8 Theoretical Understanding of Phase Diagrams 3.9 Flow Alignment 3.9.1 Lamellar Phase 3.9.2 Hexagonal Phase
45 47 51 52 56 60 60 62 66 68 71 71 74 75 75 76 79 83 90 91
105 105 111 117 117 124 126 127 129 132 135 135 139 140 143 146 149 149 151
Contents 3.9.3 Cubic Micellar Phases 3.10 Dynamics 3.10.1 Dynamic Modes 3.10.2 Dynamics of Gelation References
4. Polyelectrolyte Block Copolymers 4.1 Micellization 4.1.1 General Remarks 4.1.2 Micellization in Block Copolymers Containing Anionic Blocks 4.1.3 Micellization in Block Copolymers Containing Cationic Blocks 4.1.4 Micellization of Polyampholyte Block Copolymers 4.1.5 Micellization of Polyelectrolyte-containing ABC triblocks 4.1.6 Micellization of Block Copolymers Containing Grafted Polyelectrolytes 4.1.7 Micellization in Block Copolymers Containing Sulfonated Polyisoprene 4.2 Chain Conformation 4.3 Theory 4.4 Polyion Complexes 4.5 Copolymer-Surfactant Complexes 4.6 Complexation with other Molecules 4.7 Gelation 4.8 Hierarchical Order in Peptide Block Copolyelectrolyte Solutions 4.8.1 a Helix Structures 4.8.2 B Sheet Structures 4.8.3 Hydrogels 4.8.4 Polypeptide Block Copolymer-based Complexes References
5. Adsorption 5.1 Introduction 5.2 Adsorption at the Air-Water Interface 5.2.1 Adsorption of Neutral Block Copolymers 5.2.2 Adsorption of Polyelectrolyte Block Copolymers 5.3 Adsorption on Solid Substrates 5.3.1 Adsorption of Neutral Block Copolymers 5.3.2 Adsorption of Polyelectrolyte Block Copolymers
ix 152 159 159 160 164
173 173 173 175 179 182 182 183 183 184 188 195 198 199 200 200 202 204 206 207 208
215 215 215 215 221 222 222 225
x
Contents 5.3.3 Surface Micelles 5.4 Surface Forces Experiments 5.5 Modelling Adsorption References
6. Applications
226 231 234 236
241
6.1 Surfactancy/Detergency 6.2 Solubilization, Emulsification and Stabilization 6.2.1 Solubilization 6.2.2 Emulsification and Stabilization 6.3 Drug Delivery 6.4 Biodegradable Block Copolymer Micelles 6.5 Thermoresponsive Micellar Systems 6.6 Metal-containing Copolymer Micelles and Nanoreactors 6.7 Vesicles 6.8 Separation Media 6.9 Templating 6.10 Membranes 6.11 Other Applications References
241 241 241 245 247 253 254 255 261 268 268 274 275 276
Index
285
Preface I was inspired to write this book by developments in the field of block copolymer self-assembly in solution which have not been discussed and summarized in the form of a single convenient text. Aspects of the subject have been discussed in my previous book,1 in that by Hadjichristidis et al.,2 and in several chapters of a recent edited text.3 Recent advances have been stimulated in part by new synthetic methodologies (living polymerizations in particular) that have enabled the preparation of new materials with novel self-assembling structures, functionality and responsiveness. The present text covers the principles of self-assembly in both dilute and concentrated solution (micellization, mesophase formation, etc.) in Chapters 2 and 3, respectively. Chapter 4 covers polyelectrolyte block copolymers-these materials are just beginning to attract significant attention from researchers and a solid basis for understanding their physical chemistry is emerging, and this is discussed. Chapter 5 discusses adsorption of block copolymers from solution at liquid and solid interfaces. Chapter 6 concludes with a discussion of selected applications, focusing on several important new concepts rather than providing an account of commercial applications, which can be found elsewhere. I wish to thank several colleagues and collaborators for support and for helpful comments on several chapters: Colin Booth for Chapters 2 and 3, Steve Armes for Chapter 4, Harm-Anton Klok for Chapters 4 and 6. Tom Waigh also provided particularly insightful comments on Chapter 4. As usual I bear full responsibility for any errors and omissions, of which I would be grateful to be informed. I wish to thank Jenny Cossham for her continued support and attention in editing this book. I am also grateful to the Leverhulme Trust who provided a Leverhulme Research Fellowship which freed up time from some of my usual academic duties, enabling this book to be completed.
REFERENCES (1) Hamley, I. W. The Physics of Block Copolymers. Oxford University Press: Oxford, 1998. (2) Hadjichristidis, N.; Pispas, S.; Floudas, G. Block Copolymers. Synthetic Strategies, Physical Properties and Applications. John Wiley & Sons: New York, 2003. (3) Hamley, I. W. (Ed.) Developments in Block Copolymer Science and Technology. John Wiley & Sons, Ltd: Chichester, 2004.
1
Introduction
This book is concerned with the numerous aspects of the self-assembly of block copolymers in solution, and the diverse applications of this. Block copolymers in the melt, or in blends are not considered, and information on this can be found elsewhere.l An early review of micellization in block copolymers was presented by Tuzar and Kratochvfl,2 and these authors provided a further review of the literature up to 1992.3 Micellar properties of block copolymers were reviewed earlier by Price.4 A discussion of micellization was included in the general reviews on block copolymers by Riess et a/.5 and Brown et a/.6 Riess has recently published a very nice review specifically dedicated to micellization in block copolymers.7 Excellent reviews focused on the solution properties of a particular class of copolymer, i.e. copolymers of poly(oxyethylene) with poly(oxypropylene), have been presented by several groups.8-13 Micellization and micellar association in related poly(oxyethylene)/poly(oxybutylene) copolymers has been summarized by Booth et a/.14-16 The micellar properties of block copolymers in dilute solution, the properties of adsorbed block copolymers and ordered mesophase (lyotropic liquid crystal phase) formation in more concentrated solutions have been comprehensively discussed.1 Reviews on structure/rheology relationships in block copolymer gels,17 and on shear-alignment of ordered mesophases18'19 (the latter review incorporates work on block copolymer melts also) have also been provided. Liu and Armes20, Liu et a/.21 and Forster22'23 have reviewed the self-assembly of amphiphilic block copolymers, and the numerous applications of the resulting nanostructures. Applications of block copolymer surfactants have been the subject of a number of reviews by researchers from Dow in the United States.24-26 The texts edited by Nace27 and by Alexandridis and Lindman28 cover many aspects of the behaviour and properties of PEO-based amphiphilic block copolymers, with several chapters devoted to applications. A standard notation for block copolymers is becoming accepted whereby, for example, PX-b-PY denotes a diblock copolymer of polymer X and polymer Y.29 This convention is used here. In the case that a specific polymer with defined chain lengths is considered, the molecule is denoted Xm-b-Yn, where m and n are degrees of polymerization. This notation is somewhat more cumbersome than alternatives. For example, Booth and coworkers use single letters to indicate blocks in Block Copolymers in Solution: Fundamentals and Applications © 2005 John Wiley & Sons, Ltd.
I. W. Hamley
2
Block Copolymers in Solution: Fundamentals and Applications
PEO-based copolymers (E for poly(ethylene oxide), etc.), however this system breaks down when considering large numbers of distinct materials, as is the case here. Table 1.1 summarizes the abbreviations used. Note that throughout this book we have used the terms PEG and PEO according to the notation used in the original research-we have not attempted to distinguish carefully between them (PEG differs from PEO by hydroxyl termination as opposed to methyl termination). Table 1.1 Abbreviations used for polymers Abbreviation
Polymer/systematic name (where used alternatively)
OEGMA (see also PEGMA) PA PCsA PNaA PAA PAI PAM PAMS PAsp PB PBA PBLG PBMA PBO PBzMA PCEMA PCL PDAMA PDEA PDESCB PDLL PDMA PDMS PEB PEE PEGMA PEHA PEI PEMA PEO; PEG
Oligo(ethylene glycol) methacrylate
PEP PE4VP PFMA PFP
Poly(acrylate) Poly(caesium acrylate) Poly(sodium acrylate) Poly(acrylic acid) Poly[5- (N, N, N-diethyImethylammonium)]isoprene Poly(acrylamide) Poly(a-methyl styrene) Poly(a,b-L-aspartic acid) Poly(butadiene) Poly(butyl acrylate) Poly(7-benzyl L-glutamate) Poly(butyl methacrylate)/poly(n-butyl methacrylate) Poly(butylene oxide)/poly(oxybutylene) Poly(benzyl methacrylate) Poly(2-cinnamoyloxyethyl methacrylate) Poly(e-caprolactone) Poly[A^-(A^,A^-dicarboxymethylaminopropyl)methacrylamide] Poly [(2-diethylamino)ethyl methacrylate] Poly(diethylsilacyclobutane) Poly(D,L-lactide) Poly[(2-dimethylamino)ethyl methacrylate] Poly(dimethylsiloxane) poly(ethylene-co-butylene) Poly(ethylethylene) Poly(ethylene glycol) methacrylate Poly(ethylhexyl acrylate) Poly(ethyleneimine) Poly(2-phenylethyl methacrylate) Poly(ethylene oxide)/poly(oxyethylene); Poly(ethylene glycol) Poly(ethylene-co-propylene) Poly(/V-ethyl-4-vinylpyridinium) Poly(perfluorohexylethyl methacrylate) Poly(ferrocenylphenyl phosphine)
Introduction Table 1.1
3
(Continued)
Abbreviation
Polymer/systematic name (where used alternatively)
PFPO PFS P4FS PGMA PHEMA PhI PHIC PHOVE PHPMA PI sPI PIBVE PLGA PLLA PLMA PLys PMA PCsMA PNaMA PMAA PMDPS
Poly(perfluoropropylene oxide) Poly(ferrocenylphenyl silane) Poly(4-fluorostyrene) Poly(glyceryl monomethacrylate) Poly(hydroxyethyl methacrylate) Poly(hydrogenated isoprene) Poly(hexyl isocyanate) Poly(2-hydroxyethyl vinyl ether) Poly[/V-(2-hydroxypropyl)methacrylamide] Polyisoprene Sulfonated polyisoprene Poly(isobutyl vinyl ether) Poly(D,L-lactic acid-co-glycolic acid) Poly(L-lactic acid) Poly(lauryl methacrylate) Poly(L-lysine) Poly(methacrylate) Poly(caesium methacrylate) Poly(sodium methacrylate) Poly(methacrylic acid) Poly {3-[AK2-methacroyloylethyl)-A^W-dimethylammonio]propane sulfonate} Poly(methylene) Poly [2-(Ar-morpholino)ethyl methacrylate] Poly(methyl methacrylate) Poly(2-methoxyethyl vinyl ether) Poly(2-methyloxazoline) Poly(2-methacryloyloxy phosphorylcholine) Poly(methylphenyl silane) Poly(4-methyl styrene) Poly(methyltetracyclododecane) Poly[methyl tri(ethylene glycol) vinyl ether] Poly(methyl vinyl ether) Poly(n-butyl vinyl ether) Poly(N-isopropylacrylamide) Poly(2-phenoxyethyl vinyl ether) Poly(propylene oxide)/poly(oxypropylene); poly(propylene glycol) Poly(phenylquinoline) Poly(propylene sulfide) Polystyrene Poly(solketal methacrylate) Poly(styrene oxide)/poly(oxyphenylethylene) Poly(styrene sulfonate) (Continue)
PME PMEMA PMMA PMOVE PMOXA PMPC PMPS P4MS PMTD PMTEGVE PMVE PNBVE PNIPAM PPhOVE PPO; PPG PPQ PPS PS PSMA PSO PSS
4 Table 1.1
Block Copolymers in Solution: Fundamentals and Applications (Continued)
Abbreviation
Polymer/systematic name (where used alternatively)
PNaSS PSSA PTHF PtEA PtBS PTMEMS PVA PVBA PVP PVPh P2VP P4VP qP4VP PVPEA PVSO
Poly(sodium styrene sulfonate) Poly(styrene sulfonic acid) Poly(tetrahydrofuran) Poly(tert-butyl acrylate) Poly(tert-butyl styrene) Poly(trimethylammonium ethylacrylate methyl sulfate) Poly(vinyl alcohol) Poly[(4-vinyl)benzoic acid] Poly(vinyl pyridine) (position of substitution not stated) Poly(vinyl phenol) Poly(2-vinyl pyridine) Poly(4-vinyl pyridine) Quaternized P4VP Poly(vinylphenylethyl alcohol) Poly(phenylvinyl sulfoxide)
Abbreviations used for some common solvents and surfactants are listed in Table 1.2. Some technical terms are also abbreviated, but these can be crossreferenced using the index. Certain topics are omitted from the present text. Associative polymers which may be 'blocky' copolymers but are often random copolymers are generally not considered, although some aspects of the self-assembly of telechelic chains is discussed. Texts on this subject are available elsewhere.30-32 It should be noted that a telechelic polymer is defined by IUPAC as a 'prepolymer capable of entering into further polymerization via its reactive endgroups'.33 We follow common usage here, and use telechelic to refer to an ABA triblock with short endblocks that can undergo physical as well as chemical cross-linking, for example due to association of hydrophobes. The behaviour of block copolymers in blends with homopolymer 'solvent' is also not considered (good reviews on this can be found elsewhere1). Table 1.2 Abbreviations used for solvents CPC1 CTAB DBP DEP DMF DMP DOP DTAB SDS THF
Cetyl pyridinium chloride Cetyl trimethylammonium bromide Di-n-butyl phthalate Di-n-ethyl phthalate Dimethylformamide Di-tt-methyl phthalate Di-n-octyl phthalate Dodecyl trimethylammonium bromide Sodium dodecyl sulfate Tetrahydrofuran
Introduction
5
Here we consider self-assembly of block copolymers in low molecular weight solvents. The behaviour of block copolymer melts and nanostructure formation in thin films are also outside the scope of the present volume.
REFERENCES 1. Hamley, I. W. The Physics of Block Copolymers. Oxford University Press: Oxford, 1998. 2. Tuzar, Z.; Kratochvil, P. Adv. Colloid Interface Sci. 1976, 6, 201. 3. Tuzar, Z.; Kratochvil, P. Micelles of Block and Graft Copolymers in Solutions. In Surface Colloid Science; Matijevic, E., Ed. Plenum: New York, 1993; Vol. 15; pp 1. 4. Price, C. Colloidal Properties of Block Copolymers. In Developments in Block Copolymers; Goodman, I., Ed. Applied Science: London, 1982; Vol. 1; p 39. 5. Riess, G.; Hurtrez, G.; Bahadur, P. Block Copolymers. In Encyclopedia of Polymer Science and Engineering; Mark, H. E, Kroschwitz, J. I., Eds. Wiley: New York, 1985; Vol. 2; p 324. 6. Brown, R. A.; Masters, A. J.; Price, C.; Yuan, X. F. Chain Segregation in Block Copolymers. In Comprehensive Polymer Science; Booth, C., Price, C., Eds. Pergamon: Oxford, 1989; Vol. 2; p 155. 7. Riess, G. Prog. Polym.Sci. 2004, 28, 1107. 8. Almgren, M.; Brown, W.; Hvidt, S. Colloid Polym. Sci. 1995, 273, 2. 9. Alexandridis, P. A.; Hatton, T. A. Coll. Surf. A 1995, 96, 1. 10. Alexandridis, P. Curr. Opin. Colloid Interface Sci. 1997, 2, 478. 11. Chu, B. Langmuir 1995, 11, 414. 12. Chu, B.; Zhou, Z. Physical Chemistry of Polyoxyalkylene Block Copolymer Surfactants. In Nonionic Surfactants: Polyoxyalkylene Block Copolymers; Nace, V M., Ed. Marcel Dekker: New York, 1996; Vol. 60. 13. Mortensen, K. Coll. Surf. A 2001, 183-185, 277. 14. Booth, C.; Yu, G.-E.; Nace, V. M. Block Copolymers of Ethylene Oxide and 1,2-Butylene Oxide. In Amphiphilic Block Copolymers: Self-Assembly and Applications; Alexandridis, P., Lindman, B., Eds. Elsevier: Amsterdam, 2000; p 57. 15. Booth, C.; Attwood, D. Macromol. Rapid Commun. 2000, 21, 501. 16. Hamley, I. W.; Mai, S.-M.; Ryan, A. J.; Fairclough, J. P. A.; Booth, C. Phys. Chem., Chem. Phys. 2001, 3, 2972. 17. Hamley, I. W. Phil. Trans. R. Soc. Lond. 2001, 359, 1017. 18. Hamley, I. W. Curr. Opin. Colloid Interface Sci. 2000, 5, 342. 19. Hamley, I. W. J. Phys.: Condens. Matter 2001, 13, R643. 20. Liu, S.; Armes, S. P. Curr. Opin. Colloid Interface Sci. 2001, 6, 249. 21. Liu, T; Burger, C.; Chu, B. Prog. Polym.Sci. 2003, 28, 5. 22. Forster, S.; Antonietti, M. Adv. Mater. 1998, 10, 195. 23. Forster, S.; Plantenberg, T. Angew. Chem., Int. Ed. Engl. 2002, 41, 688. 24. Nace, V. N. Properties of Polyoxyalkylene Block Copolymers. In Nonionic Surfactants. Polyoxyalkylene Block Copolymers; Nace, V. N., Ed. Marcel Dekker: New York, 1996; Vol. 60; p 145. 25. Edens, M. W. Applications of Polyoxyethylene Block Copolymer Surfactants. In Nonionic Surfactants. Polyoxyalkylene Block Copolymers; Nace, V. N., Ed. Marcel Dekker: New York, 1996; Vol. 60; p 185.
6
Block Copolymers in Solution: Fundamentals and Applications
26. Edens, M. W.; Whitmarsh, R. H. Applications of Block Copolymer Surfactants. In Developments in Block Copolymer Science and Technology, Hamley, I. W., Ed. John Wiley & Sons, Ltd: Chichester, 2004; p 325. 27. Nace, V. N. (Ed.) Nonionic Surfactants. Polyoxyalkylene Block Copolymers. Marcel Dekker: New York, 1996; Vol. 60. 28. Alexandridis, P.; Lindman, B. (Eds) Amphiphilic Block Copolymers: Self-assembly and Applications. Elsevier: Amsterdam, 2000. 29. Hamley, I. W. (Ed.) Introduction to Block Copolymers. In Developments in Block Copolymer Science and Technology. John Wiley & Sons, Ltd: Chichester, 2004. 30. Glass, J. E. (Ed.) Polymers in Aqueous Media: Performance Through Association. American Chemical Society: Washington, DC, 1989; Vol. 223. 31. Shalaby, S. W.; McCormick, C. L.; Butler, G. B. (Eds) Water-soluble Polymers. Synthesis, Solution Properties and Applications. American Chemical Society: Washington, DC, 1991; Vol. 467. 32. Schulz, D. N.; Glass, D. E. (Eds) Polymers as Rheology Modifiers. American Chemical Society: Washington, DC, 1991; Vol. 462. 33. Odian, G. Principles of Polymerization. John Wiley & Sons, Ltd: New York, 2004.
2 Neutral Block Copolymers in Dilute Solution 2.1 INTRODUCTION Block copolymers in a dilute solution of a solvent selective for one block usually tend to form spherical micelles. This is now established for so many copolymer systems that to attempt to discuss every publication on this would be foolish. In the following, the salient features are highlighted.
2.2 TECHNIQUES FOR STUDYING MICELLIZATION In the following, the main techniques that are used to characterize block copolymers in solution are discussed - these include methods for characterizing lyotropic mesophases and transitions between them (the subject of Chapter 3) as well as classical methods for studying micelle dimensions and the influence of micellization on solution properties. Characterization methods for adsoption are also introduced, in anticipation of the discussion of this in Chapter 5. The following are listed in alphabetical order, not order of importance. 2.2.1
CRYO-TEM
Cryo-TEM is an abbreviation for cryogenic transmission electron microscopy. It is a technique where transmission electron microscopy (TEM) is used to image cryogenically cooled samples. Rapid cooling into cryogenic liquids is intended to 'trap' structures formed in solution, by vitrifying the sample and avoiding crystallization in the solvent. TEM relies on electron density contrast within a thin film of a sample to provide an image due to spatial variations in transmission of the electron beam. In the case of block copolymer solutions, the sample is usually prepared by coating directly onto a carbon-coated TEM grid (by spin or dip coating). Figure 2.1 shows a representative cryo-TEM image from an array of PSO-b-PEO diblock micelles.
Block Copolymers in Solution: Fundamentals and Applications © 2005 John Wiley & Sons, Ltd.
I. W. Hamley
8
Block Copolymers in Solution: Fundamentals and Applications
Figure 2.1 Cryo-TEM image of micelles formed by a PSO-b-PEO diblock in aqueous solution.429 Reproduced by permission of Springer Verlag.
Cryo-TEM as applied to imaging micellar structures is discussed in reviews by Talmon and coworkers.1,2 An excellent account of TEM is provided by Brydson and Hammond.3 2.2.2
DIFFERENTIAL SCANNING CALORIMETRY
As its name suggests, this technique involves measuring the differential power necessary to maintain a given temperature for two pans containing the polymer and a reference sample. Single pan differential scanning calorimetry (DSC) instruments are also available in which the reference sample is run prior to the sample to be studied. In DSC, a phase transition is indicated by a sharp endotherm or exotherm which causes changes in the differential power supplied to the sample. It is used to investigate the enthalpy of micellization (Section 2.6) and to detect the critical micelle concentration (cmc). It can also be used to detect gelation, as described further in Section 3.3.3. Since the enthalpy associated with these transitions (especially the latter) can be small, high sensitivity instrumentation is sometimes required. The technique is discussed in more detail elsewhere.4 2.2.3
DYNAMIC LIGHT SCATTERING
Dynamic light scattering (DLS) is also known as photon correlation spectroscopy (PCS) or quasi-elastic light scattering (QELS). It involves measuring the temporal fluctuations of the intensity of scattered light. The number of photons entering a
Neutral Block Copolymers in Dilute Solution
9
detector are recorded and analysed by a digital correlator. The correlation between counts measured at angle 9 over an interval t is computed:
Laplace transformation of Equation (2.1) (often using the CONTIN program5) yields the distribution of relaxation times, A(t). The decay rates of the relaxation modes provide translational diffusion coefficients. The measured intensity correlation function is related to the field correlation function, g(1)(o,t), by the Siegert relationship:6
Here c is an experimental constant proportional to the ratio between the coherence area and the detector area. In polymer solutions, DLS is used to determine the hydrodynamic radius of the constituent particles using the Stokes-Einstein equation:
where kB is the Boltzmann constant, T is the absolute temperature, 77 is the solvent viscosity and D is the diffusion coefficient. DLS has also been exploited to study diffusion in polymer solutions, and details of experimental work are provided in Sections 2.11 and 3.10. Because the intensity of scattered light is z-weighted (z a cMw, where c is mass concentration and Mw is mass-average molar mass), DLS is sensitive to low levels of high molar mass solutes. The concentration dependence of the mutual diffusion coefficient, D, in binary solution can be expressed as:
Here DO is the infinite dilution diffusion coefficient, kd is the concentration coefficient and c is the concentration. The concentration coefficient is given by:7
where A2 is the second virial coefficient, Mw is the weight-average molar mass, and V is the partial specific volume, which is generally small compared with the other two terms on the right-hand side of Equation (2.4). The first term in this equation accounts for thermodynamic interactions, and kf accounts for hydrodynamic interactions.
10
Block Copolymers in Solution: Fundamentals and Applications
There is a substantial body of work using DLS to probe the hydrodynamic properties of block copolymers containing PEO in aqueous solution, as discussed o elsewhere. The technique of DLS is the subject of the book by Berne and Pecora.6 2.2.4
ELLIPSOMETRY
This technique has been used to measure the thickness of adsorbed polymer films, and hence the adsorption isotherm. It relies on measurements on the angular dependence of the intensity of reflected s- and p-polarized light. The data are modelled based on the thickness and refractive index of the layer. Further details on the technique can be found elsewhere.9 Surface plasmon resonance has also been used to measure adsorbed layer thicknesses. Surface plasmons are electromagnetic surface waves propagating at the interface between a metal and a dielectric material. The angular dependence of the reflected p-polarized light exhibits a minimum at a resonance condition for an evanescent wave established in the electron gas in the metal near the interface. The position of the resonance depends on the dielectric properties of the medium which can be modelled using formalisms from optics. The experiments are normally performed using the so-called Kretschmann configuration where the light is incident through an index-matched prism placed over a gold-plated slide onto which the polymer is adsorbed. The method is described in detail in a thorough review.10 Further details on the application of the technique to block copolymer adsorbed films are available.11-13 2.2.5
FLUORESCENCE PROBE EXPERIMENTS
This method relies on changes in the fluorescence of free probe molecules or probes tagged to copolymer chains. In the former case, the fluorescence changes depending on the environment of the probe. For example, for the commonly used probe pyrene the intensity of the first and third vibronic peaks changes depending on the local polarity. Pyrene is used due to its low solubility in water, its long fluorescence lifetime and its sensitivity to the polarity of its environment. Fluorescence quenchers are sometimes used as an alternative (donor-acceptor systems). The technique of time-resolved fluorescence quenching is used to study kinetic processes. An excellent review provides more detailed information on all aspects of fluorescence probe experiments on block copolymer solutions.14 2.2.6
NUCLEAR MAGNETIC RESONANCE
Nuclear magnetic resonance (NMR) has been widely used to probe micelle structure. Proton NMR on copolymers in D2O is employed to monitor the presence or
Neutral Block Copolymers in Dilute Solution
11
absence of micellization. For example, Wanka et al. used this technique to locate the critical micelle temperatures of Pluronic block copolymers.15 The fine structure associated with PO units present for molecularly dissolved unimers disappears above the critical micelle temperature (cmt) as the mobility of the PO units is reduced in the hydrophobic micellar core. Armes et al. have used NMR extensively to probe micellization in their tertiary amine methacrylate block copolymers (see Section 4.1.3). Pulsed field gradient NMR can be used to measure self-diffusion coefficients in polymeric systems, and has been employed to determine this quantity for several types of poly(oxyethylene)-based block copolymer in aqueous solution.16-18 A difference in self-diffusion coefficients (and hydrodynamic radii derived from these) in H2O and D2O was noted for solutions of Pluronic F88.18 The method has also been used to examine diffusion in micellar solutions of PS-b-PEB-£-PS in the midblock selective sovent, n-heptane.19 The technique has been used to probe gelation, for example in PEO/PBO block copolymers in aqueous solution.20 2.2.7
RHEOLOGY
The flow properties of block copolymer solutions depend on the state of order in the system, and this has been exploited to locate sol - gel transitions in concentrated block copolymer solutions. Gels exhibit a finite yield stress (i.e. they are Bingham fluids), which can be measured in steady shear experiments. Details of the linear and nonlinear viscoelasticity are provided in Section 3.3.1. Experimentally, the dynamic shear moduli are usually measured by applying sinusoidal oscillatory shear using constant stress or constant strain rheometers. This can be in parallel plate, cone-and-plate or concentric cylinder (Couette) geometries. An excellent monograph on rheology, including its application to polymers, is available.21 The related technique of viscometry is discussed in Section 2.2.13. 2.2.8
SCANNING PROBE MICROSCOPY
Scanning probe microscopy (SPM) is a general term for methods where the deflection of a scanning probe is used to build up an image of the sample surface. As applied to polymers, the SPM method usually used is often termed atomic force microscopy (AFM). This is a technique for imaging surfaces to near 1 A resolution. The method depends on the interaction force between a sharp tip (often made from silicon nitride) and the substrate. The deflection of a cantilever to which the tip is attached due to the force it experiences as it approaches the surface is measured using a reflected laser beam or the interference pattern of a light beam from an optical fibre. For polymeric systems, the SPM experiment is usually conducted in a noncontact 'tapping mode', where the tip oscillates in proximity to the sample
12
Block Copolymers in Solution: Fundamentals and Applications
surface. This avoids damage to the sample surface. The sample or tip is then moved so that the tip rasters over the surface to build up an image. This image contains information on surface topography and phase contrast, which measures the dissipation of energy in regions of the surface with different stiffness. A further variant of this is lateral force microscopy, in which the displacement of the cantilever is resolved in-plane as well as perpendicular to the substrate. Many texts on nanotechnology describe the principles of AFM in more detail.22,23 To date, AFM has largely been used to image dried films of block copolymer surfaces. In recent work, the technique has been applied to investigate adsorption of block copolymer micelles in situ. The AFM tip is placed directly into a cell containing the liquid covering the substrate onto which adsorption occurs. A representative image is shown in Figure 2.2. Further details can be found in Section 5.2.3.
Figure 2.2 AFM topography image of micelles of a PPO-b-PEO diblock adsorbed from a 1% aqueous solution onto silica.430 Reproduced by permission of American Chemical Society.
2.2.9
SMALL-ANGLE X-RAY AND NEUTRON SCATTERING
Small-angle scattering is a powerful technique to determine micelle dimensions and via suitable models can provide detailed information on intra-micellar structure. Small-angle scattering methods are well suited to investigate the structure of micelles because their size is typically ~5-100 nm, which leads to scattering at small angles. Both small-angle X-ray scattering (SAXS) and small-angle neutron scattering (SANS) may be employed. In very dilute solution, it is possible to measure only intra-micellar scattering, the so-called form factor. However, in most cases the inter-micellar scattering contributes to the intensity, especially at low
Neutral Block Copolymers in Dilute Solution
13
wave vector q, and to a greater extent as the concentration increases. For lyotropic mesophases, the relative positions of a sufficient number of reflections arising from microstructural periodicities enable unambiguous identification of morphology. Further information can be obtained by preparing oriented specimens, and obtaining diffraction patterns for different orientations.8,24,25 Scattering data are presented as a function of the scattering vector q or its magnitude, where:
Here 29 is the scattering angle and A is the wavelength. SAXS is appropriate where the electron density contrast (between micelle and solvent, for example) is sufficient for the system to diffract X-rays.26 This is often possible with an intense source of X-rays, such as a rotating anode generator or a synchrotron source. SANS is valuable for studies of polymer structure27 because of the opportunity for contrast variation via isotope labelling. Typically, hydrogen atoms are selectively replaced by deuterium. This changes the scattering contrast and can be used to obtain local information on chain conformation or intra-micellar structure, for example. Neutron scattering has also been extensively used to enhance the scattering contrast of the solvent and/or the block copolymer. The radius of gyration, Rg, of block copolymer micelles in dilute solution can be obtained from SAXS and SANS using the Guinier equation:28
This is valid for small scattering angles, qRg <
14
Block Copolymers in Solution: Fundamentals and Applications
Figure 2.3 SAXS data for 2 wt % PEO87-/?-PBO18 in water at 75 °C.213 The full line is a fit using the Pedersen-Gerstenberg form factor for a spherical micelle with hard core and attached Gaussian chains and the hard sphere structure factor. Reproduced by permission of American Institute of Physics.
The structure factor is often taken to be that for hard spheres.32,35,41 For interacting micelles, use of the structure factor from Baxter's sticky hard sphere model42 may be more appropriate. Expressions for Baxter's sticky hard sphere structure factor have been used by several groups in their analysis of SANS data for Pluronic43 and reverse Pluronic44 block copolymers. 2.2.10
STATIC LIGHT SCATTERING
Here, the intensity of elastically scattered light is measured as a function of scattering angle. Many micelles, particularly those of water-soluble block copolymer surfactants are small compared with the wavelength of light and act as point scatterers, hence the scattered intensity does not depend on scattering angle 9 (when corrected for geometrical effects). Micelles in organic solvents (and some in water) are larger, and intensity is sensitive to 0. In this case extrapolation to zero angle gives values of weight-average molecular weight (Mw) corrected for intramolecular interference, and the angular dependence gives values of the z-average radius of gyration (Rg)z. Alternatively, scattering is measured at small angle (small-angle light scattering, SALS) to obtain Mw directly. In all cases, the variation of scattering intensity with concentration gives information about the interactions of the micelles in solution (e.g. via the second virial coefficient) and so the excluded volume of one micelle for another, which in turn relates to the effective volume fraction of
Neutral Block Copolymers in Dilute Solution
15
micelles in solution. In the analysis of light scattering data, only the scattering from solute is required. The contribution from local solvent concentration fluctuations is accounted for by using the excess intensity of scattered light / — 7S, where / is the total intensity and IS is that of the solvent background. This can be used to calculate the Rayleigh ratio, Ro, which for vertically polarized light is defined as Rg = Ivd 2//v,0. Here 7V is the intensity of scattered light, Iv,o is the incident intensity, and d is the distance from the particle to the detector. The usual procedure for obtaining Mw and A2 involves Debye plots. At small angles, the inverse scattered intensity is given by:45,46
where c is the concentration of polymer, K is an optical constant depending on refractive index, wavelength and polarization of the light:47
Here A is the wavelength of the light, ns and 7?s are the refractive index and Rayleigh ratio, respectively, of the solvent background and dn/dc is the refractive index increment. In practice, plots of Kc/Rg often cannot be fitted to the Debye equation taken to the second term, since they are strongly curved - see for example Figure 2.4. This
Figure 2.4 Debye plots of light scattering data for PEO-b-PSO-b-PEO triblock copolymer in aqueous solution at two temperatures.74 The curves are fits to Equation (2.10). Reproduced by permission of American Chemical Society.
16
Block Copolymers in Solution: Fundamentals and Applications
indicates interactions between micelles (beyond a second virial coefficient description). The data can be fitted using a procedure due to Vrij,48 who used the Carnahan-Starling equation49 (equivalent to the hard sphere structure factor taken to the seventh term) to obtain the following expression:
Here the inverse structure is given by:
with 0e the volume fraction of equivalent uniform spheres, which may be used to compute the swollen volume, knowing the 'dry' volume calculated from the intercept (which provides Mw).47 Light scattering from block copolymer micelles has been reviewed.45 2.2.11
SURFACE PRESSURE-AREA ISOTHERMS
A Langmuir trough is used to measure the surface pressure (TT) of an adsorbed insoluble copolymer layer as it is compressed using a barrier. Compression reduces the surface area per molecule, A, and n—A isotherms are determined. The copolymer is spread as a film from a volatile organic solvent. A Wilhelmy plate (see following section) can be used to measure the surface force and hence pressure. The compression (or decompression) rate needs to be controlled to eliminate possible kinetic effects. A description of Langmuir film techniques can be found in many physical chemistry textbooks.
2.2.12
SURFACE TENSIOMETRY
The critical micelle concentration (cmc) in block copolymer solutions can be determined by measurement of the surface tension (7) as a function of concentration. The method detects completion of the Gibbs monolayer at the air-water interface, and is a secondary indicator of the onset of micellization. The cmc for solutions of monodisperse polymers is indicated by a fairly sharp decrease in 7 versus log(c) (this plot being known as the Gibbs adsorption isotherm). The most common methods used to measure surface tension of surfactant solutions using commercial instruments involve Du Noiiy ring and Wilhelmy plate techniques. In the former, the force necessary to detach a ring or wire loop from a liquid surface is measured (using a balance). This detachment force is proportional to surface tension via Young's equation. The Wilhelmy plate method works similarly, in the detachment mode. Here, a glass plate or slide is pulled from the surface. The weight
Neutral Block Copolymers in Dilute Solution
17
of the meniscus formed is measured with a force balance, and this is equal to the vertical component of the surface tension. In practice, the Wilhelmy plate method usually works by immersing the slide in the liquid by raising the liquid, and the corresponding change in weight due to the meniscus is measured. Both Du Noiiy ring and Wilhelmy plate methods work best when the liquid wets the immersed solid (i.e. ring or plate). Further details can be found elsewhere.50 Surface tension measurements have been used by many researchers to determine the cmc of Pluronic copolymers.51'52 In some cases, two breaks in the curve of surface tension against concentration are observed.51 The first is ascribed to rearrangements of copolymer molecules at the air/water interface and the second is identified as the cmc. The method has also been used to investigate block copolymers in organic solvents. C'l
2.2.13
VISCOMETRY
Viscometry has been used extensively to provide information on the hydrodynamic properties of solutions of block copolymer micelles. The viscosities of polymer solutions are measured in capillary flow viscometers, which are described in detail elsewhere.21 Dynamic viscosity can be measured using oscillatory shear. The specific viscosity, nsp, divided by the concentration can be used to determine the intrinsic viscosity, [77], by extrapolation via:
Here kH is the Huggins coefficient. The intrinsic viscosity decreases and the Huggins coefficient increases, as micelles become smaller. Specifically, intrinsic viscosity is related to hydrodynamic volume, hence to the hydrodynamic radius. On micellization, nsp/c has been observed to increase for some systems but to decrease for others, and unfortunately there are no firm rules governing which case will prevail for a given block copolymer solution.
2.2.14
X-RAY AND NEUTRON REFLECTIVITY
These methods are used to probe the structure of block copolymers adsorbed at the liquid-air or liquid-solid interface. The intensity of X-rays or neutrons reflected from a block copolymer film is measured as a function of angle of incidence [or wavevector magnitude q, Equation (2.6)]. The case where the angle of the reflected beam is equal to the incident angle, and the plane of reflection is normal to the surface, is termed specular reflectivity. It provides the scattering density profile normal to the film surface (z direction). Below a critical angle for reflection (which is proportional to the average scattering density in the material), all X-rays or
18
Block Copolymers in Solution: Fundamentals and Applications
neutrons are reflected. Above (but close to) the critical angle for reflection, the ratio of reflected to incident intensity is given approximately by:54,55
where p(z) is the scattering density profile. The reflectivity for a film of uniform density falls off as q~4 (Fresnel relectivity). Modulations that are superimposed on this result from the film structure. An exact calculation of reflectivities requires allowance for refraction within the block copolymer film, and can be achieved using methods developed from optics, where the film density profile is considered to be divided up into a finite number of slices.56 As in all scattering techniques, information on the scattering density cannot be obtained directly, however in the case of specular reflectivity the information content is further limited by the onedimensional nature of the profile. Modelling is required to extract quantitative information. Off-specular reflection is, in principle, a powerful method for determining inplane structure in block copolymer films but is not yet widely used.
2.3 MICELLIZATION IN PEO-BASED BLOCK COPOLYMERS The most widely studied and commercially important amphiphilic block copolymers are the PEO-b-PPO-b-PEO triblocks. They are manufactured by a number of companies,57,58 but best known by their BASF tradename Pluronics or as poloxamers. Pluronic sample codes contain a letter prefix followed by two or three numbers (poloxamer codes are listed elsewhere59). The letters used are L for liquid, P for paste and F for flakes. The last number indicates the mass fraction of PEO in the copolymer, the others code the molar mass of the PPO block. The properties of these materials have been extensively reviewed.59-62 Applications are discussed in Chapter 6. At low temperature (< ca. 15 °C) both PEO and PPO are water soluble. As temperature is increased PPO becomes increasingly hydrophobic and micelles form. As temperature is increased further the PEO becomes less soluble, leading to shrinkage of the corona and above a critical solution temperature to phase separation and clouding. Table 2.1 includes information on composition and molar mass for these block copolymers. There is variation in these values from one manufacturer to another, and from batch to batch. Different experimental techniques used to determine these quantities also contribute to uncertainties in these quantities. The compositions quoted are typical values - the degrees of polymerization may vary by up to ±3 units. The values in Table 2.1 are compiled from those provided from BASF technical data (as also quoted by Alexandridis59) and a review article by Chu and Zhou.62 The molar mass is accurate to approximately 10%. Also included in Table 2.1 are cmc values determined at the temperatures indicated, together with
Neutral Block Copolymers in Dilute Solution
19
Table 2.1 Composition and cmc values for Pluronic-type block copolymers.
Pluronic
Composition
L35 L43 L44 L61 L62 L64
EO11P017EO11 E06P023E06 E010P024E010 EO2PO32EO2 E06P034E06 E013P030E013
Molar mass (g moP1)
1900 1850 2200 2000 2500 2900
cmc (g dm 3) (20°C, except where stated)
330a 190a 500a
25a 22 L81 L92
P65
EO3PO22EO3 E08P047E08 EO4PO61EO4 EO5PO70EO5 EOnP070EO11 E020P030E02o
2750 3650 3800 4400 5000 3400
P84 P85
EO19PO43EO19 E026P041E026
4200 4600
2.7b 410a 410a 120* 200°
P94
E021P047E021
4600
0.02 (40°C)
P103 P104
E017P062E017 E027P063E027
4950 5900
P105 P123
E037P058E037 E020P072E020
6500 5750
72 7 20 15 22 1.8
F68
E080P030E080
8750
0.89* 290a
F77 F87 F88
E053P035E053 E061P041E061 EO104PO41EO104
6600 7700 11400
E0118P046E0118 E0132P052E0132 EO99PO69EO99
13000 14600 12600
L101 L121 L122
F98 F108 F127
16
100 190a
6 130b
40 2.6b
1.4 aObtained by extrapolation from log(cmc) versus l/T or log(c) versus 1/cmt. bObtained by interpolation in the same way.
Reference
60 63 15 64 65
15 60 15 60 60 65 66 67 60 60 15 60 60 15 60 68 60 65 60 60 15 69
20
Block Copolymers in Solution: Fundamentals and Applications
references to the relevant original literature. The variation is large, this in part may reflect the different methods used (see the original source for this information62) as the cmc is not a true colligative property and the value obtained can be expected to depend on the measurement technique. Other parameters including viscosity and surface tension under reference conditions and cloud points can be found elsewhere.59,60 Information for reverse Pluronics (PPO-b-PEO-b-PPO architecture) is excluded from this table but can be found elsewhere.58,59,61 Tables listing the composition and properties of similar PEO-b-PPO,70 PEO-bPBO diblock46'70 and PEO-b-PBO-6-PEO46,70and PBO-b-PEO-b-PBO triblocks46' 70,71 have also been compiled. Micellization of PEO block copolymers with PDLL,72 and PSO have also been investigated.73,74 The micellization thermodynamics of PEO-based copolymers are discussed in detail in Section 2.6. Sources for other important physico-chemical data are briefly summarized here. Specific volumes for PEO and PBO in the melt as a function of temperature are presented by Mai et al.75 and for these two polymers and also PPO via accurate density measurements on EO92-b-BO18 and Pluronics P85 and P94.76 Flory-Huggins interaction parameters x have been measured for PEO in water77-79 and for melts of PEO-b-PEO80 and PEO-b-PPO81 diblocks. Dormidontova has reported that use of x = —0.209 + 93.5/T accounts very well for the observed phase diagrams of PEO in water.82,83 The latter reference presents the most refined model, in which the influence of end groups on hydrogen bonding is incorporated into the model. Karlstrom84 and later Malmsten et a/.85 analysed contributions to x for PEO in water in detail. These parameters have also been obtained by fitting adsorption isotherm data, for use in self-consistent field theory calculations. Interaction parameters for the three pairs of components in either the PEO/PBO/ water86 or PEO/PPO/water85 systems have been tabulated.
2.4 MICELLIZATION IN STYRENIC BLOCK COPOLYMERS Micellization of styrenic block copolymers (generally in copolymers with PB or PI, or alternatively with hydrogenated midblocks, PEB or PEP) has been extensively studied in solvents selective for either component. The relevant literature was summarized in an earlier review.8 Due to the large number of studies this is not reiterated here. For all PS-containing block copolymers in aqueous solution, it is important to note that vitrification of the PS core can lead to kinetically trapped nonequilibrium structures. Therefore a standard procedure is to dissolve in a PS-selective solvent and then dialyse against water.87
2.5
DETERMINATION OF CMC
The cmc is generally determined from the concentration dependence of surface tension - the surface tension is independent of concentration above the cmc
Neutral Block Copolymers in Dilute Solution
21
Figure 2.5 Surface tension versus log(c) for aqueous solutions of EO41-b-BO8 diblock at 30 °C (o), 40 °C (•) and 50 °C (D).88 Reproduced by permission of Royal Society of Chemistry.
(Section 2.2.12). Figure 2.5 shows good quality data obtained for a model PEO-bPBO diblock in aqueous solution.88 The cmc is generally less clearcut for commercial Pluronic PEO-b-PPO-b-PEO triblocks due to the presence of impurities, in particular diblock copolymers, resulting from a transfer reaction.70,89 In fact, two apparent breaks in the concentration dependence of surface tension may be observed for these copolymers51 - the first is ascribed to collapse of copolymer molecules at the air-water interface and the second to the formation of micelles. The temperature dependence of the cmc evident in Figure 2.5 will be discussed in Section 2.6. The compositional polydispersity of commercial Pluronics may be the origin of the so-called 'anomalous micellization' discussed in detail elsewhere.8,90 Similarly, for PS-b-PI diblocks in Pi-selective solvents, removal of PS homopolymer suppresses anomalous micellization (and addition of PI homopolymer causes it to appear in PS-selective solvents), indicating again that the effect is due to • • 90 impurities. The surface activity and surface adsorption of PEO-b-PPO-b-PEO triblocks have been summarized.62 The physical characteristics of poly(oxyalkylene) block copolymers in solution, including wetting and foaming properties have been reviewed by Nace,91 and Edens has summarized applications of these materials.61 Nace has compared the surface activity and surfactant effectiveness of PEO-b-PPjQ diblocks and analogous PEO-b-PPO diblocks, the former being found to lower the surface tension more than the latter due to a lower cmc for the PEO-b-PBO copolymers.92 The surface activity reflects the ability of the copolymer to reduce the surface tension of water, and surface adsorption relates to the tendency for
22
Block Copolymers in Solution: Fundamentals and Applications
solute to concentrate at the air-water interface. The adsorbed amount and surface area per surfactant molecule at the air-water interface are provided via surface tension measurements as a function of the bulk concentration. This technique yields (i) the surface tension at the cmc, 7cmc; (ii) the cmc/c2o ratio (where c2o is the copolymer concentration at which the solvent surface tension is lowered by 20 mN m"1) which reflects the competition between micellization and adsorption, and (iii) the surface concentration at complete coverage or the surface area per copolymer molecule, A, at the cmc. The surface tension profile of Pluronic copolymers in aqueous solution is complex, particularly when a wide range of concentrations and a broad temperature range are covered. Extensive surface tension measurements were performed on seven Pluronic copolymers at 20° C and two breaks were observed in the surface tension curves,65 as mentioned above. Hoffman and coworkers performed surface tension measurements also at different temperatures.15 They found that the temperature has a significant influence on the cmc and the area, A, at the cmc, whereas the surface tension value at the cmc shows little temperature dependence. Alexandridis et al. also observed two breaks in the concentration curve of the surface tension, the lower concentration one being ascribed to collapse of the copolymer molecules at the air-water interface and the second to the formation of polymolecular micelles.15 Chu and Zhou summarize the surface-active behaviour and surface adsorption of Pluronic triblocks as follows:62 (i) At complete surface coverage, the area occupied by a copolymer molecule increases with increasing temperature, (ii) The area per molecule increases with the number of EO units in the chain, showing a scaling relationship A ex A^o, with v = 0.43.60 (iii) An increase in the PPO block length at fixed PEO block length leads to a decrease in the area occupied by each copolymer molecule, (iv) An increase in the cmc/c2o ratio indicates that micellization is inhibited more than adsorption, or that adsorption is facilitated more than micellization. This ratio ranges between 100 and 40000 for Pluronic surfactants at 25 °C, increasing with increasing PEO length but decreasing with increasing PPO block length and increasing temperature, (v) For Pluronics with two breaks in the 7-log(c) curve, the separation between them decreases with increasing temperature and finally vanishes at about 40 °C. This appears to be associated with the enhanced micellization tendency at high temperature.
2.6
THERMODYNAMICS OF MICELLIZATION
There are two possible models for the association of molecules into micelles.93"95 In the first, termed open association, there is a continuous distribution of micelles containing 1 , 2 , 3 , . . . , n molecules, with an associated continuous series of equilibrium constants. However, the model of open association does not lead to a cmc. Since a cmc is observed for block copolymer micelles, the model of closed association is applicable. However, the cmc does not correspond to a thermodynamic property of the system, it can simply be defined phenomenologically as the
Neutral Block Copolymers in Dilute Solution
23
concentration at which a sufficient number of micelles is formed to be detected by a given method. Thermodynamically, closed association corresponds to an equilibrium between molecules (unimers) and micelles containing p molecules:
with an associated equilibrium constant:
For an advancement of the equilibrium from left to right by a fractional extent a, K is given by:70'96
where (3 = 1 — a + a /p. If the association number is large (1 /p —» 0) and independent of temperature, then K « [A]"1 97'98 and the standard Gibbs energy of association is:
Under this condition, for molecules in their standard state of ideally dilute solution at unit concentration (1 mol dm~3) in equilibrium with micelles:99
Otherwise Equation (2.16) must be used to obtain K. For a small association number, say N = 10, the error in approximating K by 1/cmc is large. Calculation of AmicG0 via Equation (2.18) requires/? >« 50.97 For example, for a = 0.1 and cmc values of KT1 and 10~6 mol dirT3, the error in AmicG° is 30% and 15%, respectively. For a large association number, the standard enthalpy of micelle formation is then:
However, AmjC//° and rAmic5°, determined from the temperature dependence of the Gibbs energy, are less sensitive to the association number than is A m j C G° itself.96 Assuming that AmiC//° is approximately constant within a certain temperature range, Equation (2.19) can be integrated to yield:
24
Block Copolymers in Solution: Fundamentals and Applications
Figure 2.6 Temperature dependence of cmt (plotted against reciprocal temperature) for aqueous solutions of EC^r^-BOg-b-EC^i.46'100 Reproduced by permission of American Chemical Society.
Thus the logarithm of the cmc can be plotted against inverse temperature to extract information on the micellization enthalpy. Equivalently, the logarithmic concentration can be plotted against the inverse cmt, as done for a number of PEO-b-PPO-bPEO triblocks,51 and PEO-b-PBO-/7-PEO triblocks.100 Figure 2.6 shows representative data for a PEO-b-PBO-b-PEO triblock. Micellization in PEO-based block copolymers is driven by changes in solvation upon increasing temperature. The solvent quality for PEO decreases whilst that for hydrophobic blocks such as PBO or PPO increases. This leads to dehydration upon increasing temperature. This is the origin of the strong temperature dependence of cmc seen for these systems. It also causes clouding commonly observed for this type of copolymer due to aggregation of micelles at high temperature, which ultimately causes phase separation. Via Equation (2.20), the corresponding enthalpy of micellization is large and positive. The various contributions to Amic//° have been described in detail.101 Typical values of Amic//° for Pluronic block copolymers at the cmc lie in the range 170-340 kJ mol"1.60'62'70 Values of AmicG° range typically from -24.5 to -28.8 kJ mol"1.60'62 As this quantity is the difference of two large quantities, precise determination can be difficult. It is thus evident that although the unfavourable enthalpy of micellization is large, micellization is entropy driven. The large endothermic enthalpy of micellization is responsible for the significant negative temperature dependence of the cmc of these systems. The traditional view is that this is due to the hydrophobic effect.102 Unassociated hydrocarbon chains break up the hydrogen bonds between water molecules and
Neutral Block Copolymers in Dilute Solution
25
impose a locally more ordered structure that is entropically unfavourable. Because this disruption of water structure is reduced when micelles are formed, they are entropically favoured compared with unassociated molecules.103 Values of AmjC//° for PEO-b-PBO-b-PEO triblocks with PBO midblock lengths 8-12 range between 87 kJ moP1 and 121 kJ mol" 1.46>7° However, this is a factor of 3-5 smaller than the hydrophobe length in Pluronics for which Amjc//° was also measured. Nevertheless, values of AmicG° are comparable.46 Each BO unit contributes 2.00 ± 0.02 kJ m o l 1 to AmicG°.96 Nearly athermal micellization (i.e. the cmc is independent of temperature) has been reported in aqueous solutions of diblocks PEOio6-^-PBOi6 and PEO2io-bPBOi6,104 as discussed further in the following section. 2.6.1
CHAIN LENGTH DEPENDENCE OF MICELLIZATION
Micellization is mainly controlled by the length of the hydrophobic block. This is shown for example in Figure 2.7, which shows the exponential decrease in cmc with the length of the hydrophobic midblock for EOm-b-POn-b-EOm and POm-b-EOf,-bPO,n triblocks, with a similar behaviour for the analogous polymers in which BO replaces PO.70 A similar trend has been noted for EOm-b-EOn-b-BOm triblocks with short BO blocks.71 The decrement in cmc with hydrophobic block length n for various systems is discussed quantitatively by Yang et al.96 The PEO block has a smaller influence on micellization. An increase in the number of PEO units leads to a small increase in cmc and cmt.62 The cmt and cmc decrease with increasing total
Figure 2.7 Dependence of cmc (plotted as logarithm) on hydrophobe length, n, for diblock and triblock copolymers containing PEO and PBO or PPO.70 Reproduced by permission of Wiley-VCH.
26
Block Copolymers in Solution: Fundamentals and Applications
Figure 2.8 Standard enthalpy of micellization per BO unit for aqueous solutions of EOmb-BOn diblocks. The dashed curve emphasizes the asymptotic approach of Amjc//°/« to zero.70'104 Reproduced by permission of Wiley-VCH and American Chemical Society.
molecular weight of the copolymer, if comparisons are made for a constant PPO/ PEO ratio. The lower the relative PEO content, the larger is the influence of the total molar mass.62 The value of A m j C //°/n for EOm-b-BOn-b-EOm decreases to zero as the PBO block length increases, as shown in Figure 2.8.104 This is ascribed to the formation of 'unimolecular micelles', whereby lengthy PBO blocks are already shielded in water in the unassociated chains.70'105 The transfer of PBO blocks to the micelle core would then involve only a small standard enthalpy change. For Pluronics on the other hand A m j c //°/(n + m) decreases from about 3 kJ mol'1 down to 1 kJ mol"1 on decreasing the relative hydrophobe content (PPO/PEO ratio).15'60 Booth has shown that the hydrophobe block length dependence of cmc for PEO-b-PBO-bPEO and PEO-b-PPO-b-PEO triblocks coincide if one PBO unit is considered equivalent to four PPO units.46'96 Based on similar considerations, the hydrophobicities for common hydrophobic units in PEO-based block copolymers have been ranked in the ratio PO:DLL:ME:BO 1:4:5:6.70 Association numbers for Pluronic copolymers in the temperature range 20-50 °C vary between about 10 and 400, increasing with hydrophobe length.62 This effect has been examined for a broader range of copolymer architectures and compositions for PEO/PBO copolymers. Figure 2.9 shows the increase in p with n for a range of PEO/PBO copolymers. The dependence on architecture is discussed in the next section.
Neutral Block Copolymers in Dilute Solution
27
Figure 2.9 Association number (p) as a function of number of BO units (n) in the copolymer for aqueous solutions of PEO/PBO block copolymers at 30 °C: (•) EOm-b-BOn; (•) BOn,2-b-EOm-b-EOnt2', (o) EOm/2-b~BOn-b-EOm/2.46 Reproduced by persmission of Elsevier.
Booth and Attwood have summarized the scaling of micelle radius and association number for PEO/PPO and PEO/PBO systems as a function of hydrophilic and hydrophobic block length70. Both p and radius R (hydrodynamic radius, Rh or thermodynamic radius, Rt) increase with hydrophobic block length. There is also a surprisingly strong decrease in p with increasing PEO block length, and a small increase in R^, where the comparison is made for fixed n. Scaling relationships were derived with reference to the critical value of n for micelle formation (or the critical value of m related to the solubility limit of the copolymer). This is discussed further in Section 2.8. The observed increase in Pluronic micelle radius with temperature was quantified by SANS.41'106 Chu and Zhuo tabulate other data (up to 1996) on the radii of Pluronic block copolymers.62 2.6.2
EFFECT OF ARCHITECTURE
By custom synthesis of PEO/PBO block copolymers, Booth and coworkers have been able to examine micellization in a far wider range of architectures than for related PEO/PPO copolymers. The following discussion therefore focuses on these systems. A schematic illustration of micelles formed from copolymers with different architectures is shown in Figure 2.10.46 Compared with diblock copolymers, the triblock and cyclic copolymers are entropically disfavoured in the
28
Block Copolymers in Solution: Fundamentals and Applications
Figure 2.10 Schematic representation of chain conformation in micelles formed from PEO/ PBO block copolymers.46 Reproduced by permission of Elsevier.
micellar state. For cyclic copolymers, two block junctions per molecule must reside in the core/fringe interface and for triblocks a fraction of chains will be looped. Making comparisons at constant composition and n, this means that the cmc of a linear diblock must be significantly lower than those of other architectures (see Figure 2.11). In addition, the maximum possible radius of spherical micelles formed from a linear diblock will be up to two times that of micelles formed from copolymers of the other architectures, i.e. the association number and radius of the diblock will be large. Compared with fully looped triblocks p is larger by a factor of 23 = 8 and R^ (or Rt) by a factor of 2. These trends are confirmed by experimental results for PEO-/.-PBO,96'100'107'108 PEO-W>BO-/>-PEO96-100'109'110 and PBO-b-PEO-b-PBO100'111 copolymers. A cyclic copolymer in a micelle experiences the greatest conformational restriction, since both of its blocks must loop, producing a flower micelle. However, because a cyclic copolymer is restricted in both unassociated (solution) and micellar states, the additional contribution to the Gibbs free energy from this effect is likely to be small.46 Comparisons of experimental results for cyclic PBO-&-PEO and PEO-b-PBO-b-PEO copolymers show that the cyclic copolymers form larger micelles than their linear triblock counterparts, and have lower cmc values.20'47
Neutral Block Copolymers in Dilute Solution
29
Figure 2.11 Critical micelle concentration (plotted as logarithm) versus total number of BO units for PEO/PBO diblock and triblock copolymers in aqueous solutions at 30 °C: (•) EOm-b-BO,,\ (•) BOn/2-b-EOm-b-EOn/2; (o) EOm/2-b-BOn-b-EOm/246 Reproduced by permission of Elsevier.
The effect of copolymer architecture is also manifested in the bridging of chains between micelles that is possible for PBO-&-PEO-&-PBO copolymers, as shown for example, by DLS.46'111"115 Zhou et al. performed light scattering experiments on BOi2-^-EO26o-^-BOi2 m water and observed that closed association of flower-like micelles occurs in dilute solution.115 However, even in relatively dilute solutions (1-2% w/v), secondary associates were formed, as shown by DLS which revealed a mode coresponding to aggregates with a size up to 100 nm at a concentration of 20 mg cm~3. The secondary aggregates increase in both number and size with increasing concentration and temperatures and were ascribed to the formation of open, branched structures due to bridging of micelles by the long PEO blocks. This leads to highly viscous solution at concentrations as low as 4%, and a marked increase of viscosity with temperature. Gelation due to bridging is discussed further in Section 3.3.5. The cmc data for PEO/PBO diblocks and triblocks as a function of hydrophobe block length are compared in Figure 2.11. The cmc at a given value of n is approximately two orders of magnitude lower for the diblock due to the entropic penalty discussed above. The slopes, however are similar and correspond to an increment in AmicG° of approximately 2.7 kJmol"1 (B units). The cmc for EOTO/2b-EOn-b-EOm/2 diblocks is intermediate between values for the other two architectures, although the situation is complicated by the broader E-block length distribution for these molecules. That p directly reflects the core volume, which is expected to be in the ratio 8:1 for a diblock versus triblock copolymer, is
30
Block Copolymers in Solution: Fundamentals and Applications
illustrated by Figure 2.9, although p values for EOm/2-b-BOn-b-EOm/2 copolymers with n < 10 are higher than expected, which reflects the wide PEO block length distribution. Comparing cyclic copolymers with linear triblock analogues, Booth et al. conclude that micellization is favoured in the cyclic copolymer compared with the triblock.46 Micellization of pentablock copolymers has been investigated.116 The copolymers were prepared using Pluronic F127 as a difunctional initiator for the polymerization of PDEA. The latter is pH sensitive, thus facilitating stimuli responsiveness in the copolymer. The cmt and enthalpy/entropy of micellization were similar to those of the Pluronic precursor. The micelle properties and surface activity of copolymers containing a branched block have been studied. Diblocks comprising a PPO hydrophobic block and an OEGMA hydrophilic block were prepared by atom transfer radical polymerization (ATRP).117 The oligo-PEO chains form the branches in the copolymer, which are attached to each repeat unit, leading to a densely grafted 'bottle brush'. Micellization of cyclic diblock copolymers has been examined for PS-b-PI copolymers.118'119 Borsali and coworkers have compared the micellization in heptane of cyclic PS-&-PI diblocks and their linear precursors (the cylic copolymer was prepared via direct coupling of a heterodifunctional precursor).118'119 In contrast to the conventional spherical shape of the micelles of the linear diblock, the double constraint of looped PS in the core and looped PI in the corona leads to flower micelles which further stack into giant wormlike micelles, as concentration increases (Figure 2.12).
Figure 2.12 Comparison of morphologies of linear and cyclic PS-£-PI diblocks in a Piselective solvent proposed by Minatti et a/.348 Reproduced by permission of American Chemical Society.
Neutral Block Copolymers in Dilute Solution
31
Hadjichristidis and coworkers have synthesized a vast range of copolymers with complex architectures over the last decade, and have investigated micellization of many of these materials. For example, latrou et al. investigated the micellization of Pl^-h-dPS-h-Pli 'super-H' shaped molecules in n-decane, a selective solvent for the PI stars at each end.120 The aggregation number was found to be greatly reduced compared with a PS-b-PI diblock, and the solubility was increased. The micellar dimensions and aggregation number as a function of PS/PI chain length were compared with scaling theories, developed to allow for the H-shaped architecture. The structure of micelles formed by PEnPEPw mixed arm star copolymers (n,m= 1,2 with m + n = 3,4) in decane was investigated by detailed SANS experiments.121 Platelet structures were observed, comprising crystalline PE coated by solvated PEP layers (this type of structure is discussed further in Section 2.18). Micellization of PI2PS and PS2PI mixed arm star copolymers has been compared with that of linear PS-b-PI diblocks in n-decane.122 Association numbers were found to lie in the order PI2PS < PS2PI < PS-b-PI. The association number is lower for PI2PS due to the larger core-corona interfacial area per chain, as also applies for PS2PI, compared with the linear diblock although to a lesser extent than for PI2PS since the PS is on the concave side of the interface. A scaling theory was presented to account for these observations. The same group has investigated micellization of more densely crowded PS8PI8 mixed arm star copolymers in n-decane.123 Compared with diblocks with the same block length, the association number of micelles formed by the mixed arm star copolymers was lower (due to a larger interfacial area per copolymer), and the micelles were larger. A symmetric sample formed elongated micelles, but as the soluble block length increased, the micelles became more spherical. Micellization of tapered block copolymers has been investigated for PEO-fr-PBO diblocks124 and PBO-^-PEO-6-PBO triblocks,125 and also for PEO-^-PPO diblocks.126 In each case, the gradient of PEO content led to tapered statistical block copolymers. The PBO-b-PEO-b-PBO diblocks showed low cloud points, and comparison was made to related untapered block copolymers. Micellization of tapered PS/PI block copolymers of various architectures (see Figure 2.13) has been compared, using n-hexane as a selective solvent for the PI.127 The cmt, micelle dimensions, association numbers and second virial coefficients were obtained from static and dynamic light scattering experiments. Triblock copolymers with a normal tapered midblock were found to have similar properties to those of the corresponding diblock. In contrast, a large decrease in aggregation number and a decrease in the cmt were found for triblock copolymers with an inverse tapered midblock. A similar decrease in aggregation number was also observed for a normal tapered copolymer. Understanding the role of composition gradients and block sequence on the micellar properties remains a challenge for theory. It has been shown that dumbbell-shaped triblock copolymers can form aggregate structures in solution.128 Small aggregates containing five or six molecules were observed for triblocks comprising a polythiophene midblock and wedge-shaped poly(benzyl ether) dendrons when dispersed in dichloromethane. The small
32
Block Copolymers in Solution: Fundamentals and Applications
Figure 2.13 Schematic sequence distribution in tapered PS/PI copolymers and block copolymers.127 TBC, tapered block copolymer; ITMB, inverse tapered midblock; TMB, tapered midblock; RBC, reverse block copolymer; RMB, reverse midblock. Reproduced by permission of American Chemical Society.
aggregation number is probably a consequence of the restriction on packing wedgeshaped blocks into a 'micelle'. Strong concentration-dependent changes in thermochroism (shift in optical absorption spectrum) were ascribed to intermolecular TT-TT stacking interactions within the aggregates. Micellization of diblocks containing a PEO linear block and a hydrophobic dendritic unit has also been examined.129 The carbosilane dendrimer was synthesized using a divergent approach. Micelles were observed in aqueous solution of copolymers containing first and second generation dendrimers, but not for a copolymer with a third generation dendrimer which was insoluble in water due to the significant increase in hydrophobic content. Micellization of rod-coil dendrons is discussed in Section 2.12.3. 2.6.3
EFFECT OF SOLVENTS AND SALTS ON MICELLIZATION
Different solvents have a distinct effect on micellization in PEO-based block copolymers, as exemplified by DSC studies on PEO/PPO block copolymers.59 For example, methanol, ethanol, urea and formamide suppress the micellization of Pluronic F87 (and decrease both the cmt and AmiCH°) whereas butanol and hydrazine favour micelle formation (and increase the cmc and Amic//°).130 Ethanol increases the cmc of F87130 and F127,131 and increases the cmt of the latter.132
Neutral Block Copolymers in Dilute Solution
33
This indicates that aqueous ethanol behaves as a better solvent than water alone. The influence of other solvents has also been investigated.133 Addition of electrolytes having anions and cations with different sizes and polarizabilities has a significant influence on micellization.59'70 Either an increase or decrease in cloud point is observed, and has been discussed in terms of 'salting in' and 'salting out'. The influence of a particular ion generally follows the Hofmeister series.134'135 In this picture, salts affect the solvent quality of water, acting as structure makers or breakers. As an example, the presence of the alkali halide salts LiCl, KC1, NaCl and NaBr decreases both the cmt and the cloud point of Pluronic L64, wheras addition of sodium thiocyanate and urea results in an increase in cmt and cloud point.135 Other studies have been performed on the effects of added potassium halides136'137 and urea138 on the cmt and micelle properties of Pluronics. The reduction in cmt of Pluronic copolymers upon addition of NaCl has been determined via high sensitivity DSC.139 The cloud point is also reduced upon addition of salt.139 The effect of salt on cloud point is discussed further in Section 3.3.4. Since addition of NaCl causes a reduction in 0 temperature, it can be considered to have the same effect as an increase in temperature. In this way, high temperature effects such as the sphere-to-cylinder micelle transition137'140 can be observed in a more accessible temperature range.70 The effect of adding salt to aqueous solutions of an PEO/PBO copolymer has been investigated.107 Static and dynamic light scattering data for EO^-^-BOio indicate that adding salt reduces the excluded volume of the micelles. The concentration dependencies are consistent with the PEO block fringe approaching the 0 condition as the concentration of salt is increased.107 On the other hand, for Pluronics, several reports have noted little change in micelle dimensions upon addition of salt.136'138
2.7 MICELLIZATION AND MICELLE DIMENSIONS: THEORY AND SIMULATION 2.7.1
SCALING MODELS
A simple scaling model of block copolymer micelles was derived by de Gennes.141 He obtained scaling relations assuming uniformly stretched chains for the core, radius RE- This model can be viewed as a development of the Alexander-de Gennes theory 142~144 for polymer brushes at a planar interface, where the density profile normal to the interface is a step function. In the limit of short coronal (A) chains (crew-cut micelles) de Gennes predicted:141
where 7 is the A-B interfacial tension and a is a segment length. The number of chains per micelle was found to be linear in core chain length N#:
34
Block Copolymers in Solution: Fundamentals and Applications
Figure 2.14 Star polymer model for chains tethered to a curved surface. Chains are represented as a string of blobs extending radially from the core.
The Daoud-Cotton model145 for star-like polymers in a good solvent is based on a simple geometric picture of polymer 'blobs', with chain ends confined on a spherical surface (Figure 2.14), an extension of the original Alexander-de Gennes theory. It has been applied to analyse experimental results for the scaling of micellar dimensions in block copolymer solutions. For block copolymer micelles the number of arms/is replaced by the association number p. Daoud and Cotton145 identified three regions with characteristic density profiles: (i) an inner melt-like core; (ii) an intermediate concentration region (dense brush); and (iii) a swollen brush in the outer semidilute region. In this model, the density profile of a star with p arms and Flory exponent v (Table 2.2) and a fixed monomer excluded volume
Table 2.2 Flory exponent v and exponent a in radial density profile for polymer chains in different thermodynamic states. V
i
3 1 2
a
Remarks
0
Collapsed polymer, core domain
1
Polymer in O-solvent, semidilute solution
3 5
4 3
Polymer in good solvent
0.588 1
1.299 2
Exact value Polymer in stretched conformation e.g. polyelectrolyte
Neutral Block Copolymers in Dilute Solution
35
parameter decays as:
This yields: (i) >(r) ~ r° for the core; (ii) 0(r) ~ p ! / 2 r l for the unswollen brush; and (iii) <j)(r) ~^, 2 / 3 r - 4 / 3 for the outer swollen brush (Figure 2.15). Thus, for a
Figure 2.15 The density profile predicted for a micelle containing p chains.
fixed number of arms, in a good solvent, the profile should decay as llr. The average brush height referenced to the end-to-end distance of the chains (RQ) should scale as L/Ro ~ p^~v^2, which gives a scaling (i) /?1//3 in the core, (ii) pl/4 in the unswollen brush, and (iii) pu5 in the swollen brush. The Daoud-Cotton model predicts that in a good solvent the micelle radius in the limit of very long chains scales as: for a given statistical segment length and monomeric volume.145 Scattering from star polymers has been analysed using the Daoud-Cotton model to describe the density profile.146 Scattering from overlapping stars in the semidilute solutions was considered, in addition to that from isolated stars in the dilute regime. The model has been applied also to analyse scattering from block copolymer micelles. Scaling theory was used by Zhulina and Birshtein for the specfic case of a micelle formed by an AB diblock in a solvent selective for the A block.147 Four regions were identified depending on the copolymer composition (Figure 2.16). In region I, A blocks behave like isolated chains, attached at one end to the core. The core/shell interfacial area per monomer (
36
Figure 2.16 Birshtein.147
Block Copolymers in Solution: Fundamentals and Applications
Schematic of micellar structures in regimes I-IV defined by Zhulina and
increasing R&. With further increase in A/A (region IV), the curvature of the core enables each A block to occupy a larger area of the interface. Consequently, the dependence of the micellar parameters RE, p and a on A/A disappears. It was also found that the core chains are stretched in block copolymer micelles compared with the unperturbed value, the stretching relaxing from region I toward IV.147 The scaling of micellar parameters in the four regimes is summarized in Table 2.3. Table 2.3 Predictions of Zhulina and Birshtein147 for micellar characteristics in an A-B diblock in a solvent selective for the A block" Region Copolymer composition
RQ
LA
p
a
I II III IV "Here LA = R — RB is the thickness of the corona layer and v is the scaling exponent for the radius of gyration of linear polymers, Rg ~ Mv'.
Neutral Block Copolymers in Dilute Solution
37
Halperin also developed a model for regime IV,148 and derived the same scaling relationships for/? and R^ as Zhulina and Birshtein. For a large core micelle (regime I), it was pointed out that the tendency for the system to lower its interfacial free energy is countered by the stretching of the core-forming blocks which increases the free energy. For a small-core micelle (regime IV), micellar growth was found to be limited by the confinement free energy associated with the corona. Halperin also obtained a scaling relationship for the total micellar radius in the small core limit (A^B«^A):
It is worth emphasizing that all scaling theories (due to de Gennes, Daoud and Cotton, Zhulina and Birshtein, and Halperin) for block copolymer micelles with a small core and large corona predict that the association number and core radius are independent of the coronal chain length. 2.7.2
THE BRUSH MODEL
A simple mean field theory for micelle formation by a diblock copolymer in a homopolymeric solvent was developed by Leibler and coworkers.149 This model enables the calculation of the size and number of chains in a micelle and its free energy of formation. The fraction of copolymer chains aggregating into micelles can also be obtained. A cmc was found at a low copolymer concentration even for weak incompatibilities between components. Leibler et al. emphasize that for a finite aggregation number /?, the cmc is a region rather than a well-defined concentration and some arbitrariness is involved in its definition.149 In the model, spherical micelles of radius R comprising a core of B monomers (radius R#) and a corona of A monomers (thickness LA = R — RB)> are considered.149 Only a fraction 77 of A monomers in the corona belong to the copolymer, the rest belonging to the homopolymer chains penetrating the corona. The homopolymer chains are shorter than the copolymer, the parameter a = N/Nu (where /VH is the homopolymer degree of polymerization) being greater than 2. The case of relatively incompatible components was considered, and in this case the A-B interfacial region is fairly narrow. Thus, the interface was taken to have the same properties as that between two incompatible homopolymers. The free energy of a micelle containing p copolymer chains and ap(\ — r]}/2rj homopolymer chains can than be written as:
Here Fcore is the free energy due to deformation of chains in the micelle core:
38
Block Copolymers in Solution: Fundamentals and Applications
This equation assumes that the blocks have an end-to-end distance /?B that is undeformed with respect to the unperturbed dimension (Afe) a. The free energy of the corona contains contributions from the deformation energy [cf. Equation (2.27)] and from the entropy of mixing of homopolymer chains and A blocks in the micellar corona.141'150'151 Summing these two terms gives: 149
The interfacial free energy is:
where the term in brackets is the Helfand-Tagami152'153 expression for the interfacial free energy between A and B homopolymers. Figure 2.17 shows
Figure 2.17 Fraction, £, of copolymer chains aggregated in micelles as a function of the overall copolymer volume fraction 0, for different values of the incompatibility parameter xN(a — 5, N = 200).149 The number of copolymer chains in a micelle, p, and the volume fraction of copolymer monomers in the micelle corona, 77, depend weakly on 0. Solid line: XN = 20, for 0 = 0.1, p w 79.9,77 = 0.194; for 0 = 0.009, p fa 77.0,77 = 0.19. Dashed line: XN = 17.5, for 0 = 0.16, p « 78.3,77 = 0.194; for 0 = 0.026, p « 75.0,77 = 0.189. Dotteddashed line: x# = 16, for 0 = 0.18, p w 77.8, 77 = 0.196; for 0 = 0.054, p « 74.5, 77 = 0.191. Dotted line: x^ = 15, for 0 = 0.19, p»78.6, 77 = 0.201; for 0 = 0.1, p Ki 76.0, 77 = 0.197. Here x = XAB- Reproduced by permission of/. Chem. Phys.
Neutral Block Copolymers in Dilute Solution
39
representative results from this theory for the fraction of copolymer chains aggregated in micelles, (, as a function of copolymer volume fraction. The total free energy of the micellar phase contains contributions from the translational entropy of the micelles, the entropy of mixing of homopolymers and copolymers and their interaction outside the micelles. It can be written as:
where/ = N^/2NB. The first term on the right-hand side is equal to the free energy of the micelles in solution, the second represents the free energy of mixing copolymer and solvent outside the micelles, and the last term accounts for the translational entropy of the micelles. 4>\ is the volume fraction of copolymer chains outside the micelles and is given by:
and
Ftotai can be expressed in terms of three independent variables p, (, and 77 and requires the specification of model parameters N&, N#, NH, XAB and/. Calculations showed that for small incompatibility, p scales as N 06 , and R# ~ N 053 . 149 Under conditions of strong segregation, for the case of symmetric diblocks considered by them, Leibler et al. found that the cmc depends exponentially on X^A — X^B, for a fixed homopolymer degree of polymerization.149 An exponential dependence on interfacial energy is expected for a generalized model of micellization.15 It may be noted that incompressibility conditions give simple relations between /?,/? B ,pandr;: 1 4 1 ' 1 4 9
Analogous equations to (2.26)-(2.30) can be written for triblock copolymer micelles in a homopolymeric solvent.155'156 However, in a BAB triblock copolymer where the solvent is selective for the A block, the A block must be looped. Then each chain enters the core twice, and Equation (2.27) must be multiplied by 2, with a similar multiplier of the analogous term in Equation (2.28). An additional
40
Block Copolymers in Solution: Fundamentals and Applications
contribution must be added to the free energy of the corona due to looping. Balsara et al. estimated this to be:156 where q is the fraction of chains that end up in the A-B interfacial shell. This estimate is significantly lower than the one obtained earlier by ten Brinke and Hadziioannou,l 5 F\oop = /3^pk^T\nN^, where (3 is a correction factor introduced in the cyclization approximation that accounts for the entropy loss due to looping. This difference has a substantial effect on the predictions of the theory such as the fraction of copolymer that forms micelles, as shown in Figure 2.18 which shows a prediction for the fraction of associated copolymer chains as a function of the overall volume fraction occupied by the copolymer in the system.
Figure 2.18 Theoretical predictions for the dependence of £ on (/>. Dashed lines show the calculations of ten Brinke and Hadziioannou155 for various values of ft. The solid line represents the results of Balsara et al.156 NA = 200, NB = 100, WH = 40, c = 0.1. (The number of copolymer chains per micelle, p, and the coronal concentration, 77, depend weakly on cj). For 0 = 2.0 x 10~3, C = 0.394, p = 38.3 and r? = 0.189. For 0 = 2.0 x 10~2, C = 0.938, p = 39.2 and r] — 0.191.) Reproduced by permission of American Chemical Society.
Mayes and Olvera de la Cruz applied the model of Leibler et al. to investigate micelle shape transitions.157 The cmc was determined for cylindrical and spherical micelles. A trend towards the formation of cylinders was observed with increasing B block fraction and increasing homopolymer molecular weight. 2.7.3
THE SELF-CONSISTENT MEAN FIELD THEORY
Self-consistent mean field theory has been used to investigate the structure of AB diblock copolymer micelles in solution. The method is also known as self-consistent
Neutral Block Copolymers in Dilute Solution
41
field theory. In the earliest model, it was assumed that the insoluble B block formed a uniform core and that the density of A blocks in the corona was also uniform.158 Using a simple approximation for the interfacial tension, together with the known block copolymer composition, molecular weight and concentration in solution, the equilibrium size of micelles could be obtained. The scaling of micelle size, association number and radius with copolymer degree of polymerization were found to be in general agreement with the earlier scaling theory of de Gennes.141 The core radius was found to scale as /?B ~ A^'64, independent of the coronal chain length and the association number was found to scale asp ~ ./VJ]9.158 Whitmore and Noolandi159 developed this approach to analyse other micelle dimension scaling relationships:
They found 0.67 < /3 < 0.76, -0.1 < \L < 0, and 0.5 < u; < 0.86. These authors also found that the cmc is dominated by an exponential dependence on the product XAB^B (as for the brush model discussed in the preceding section) and that the fraction of copolymer in a micelle increases exponentially with decreasing temperature. The scaling of micellar dimensions was compared with SANS results on PS-b-PB diblocks in PB homopolymer, where PS forms the core.160 Scaling exponents in the range -0.19 < p, < —0.14 and 0.75 < (3 < 0.81 were obtained for a number of copolymers of different molecular weights in solution with homopolymers of varying Mw. The theoretical exponents are in reasonable agreement with these experimental values. Bluhm and Whitmore applied the Noolandi-Hong model in detail to micelles with a PS core and a PB shell in heptane.161 They obtained:
It is notable that R& scales with N% with a power close to 2/3, as obtained in the simple de Gennes model. This scaling is also obtained for strongly segregated block copolymer melts, illustrating that the core chains are effectively in a melt-like state. The power law scaling of/? A with N& has an exponent slightly greater than 1/2, the exponent for the scaling of the radius of gyration with N for a 6 solvent (Table 2.2). Both these scalings were found to be consistent with the observed dependence of the micellar radius of gyration, Rg, on the number-average molar mass for PS-&-PB diblocks in heptane:161
An exponent 0.5, close to the value in Equation (2.39) was observed using SAXS on PS-b-PI diblocks in heptane, using copolymers with a wider range of Mn.162 Further information on scaling of micelle dimensions can be found in Section 2.8.
42
Block Copolymers in Solution: Fundamentals and Applications
Pepin and Whitmore163 employed the earlier model of Whitmore, Noolandi and coworkers,159'161'164 with small modifications whilst retaining the central assumptions of a uniform corona surrounding a uniform core. Pepin and Whitmore also performed Monte Carlo (MC) simulations, which revealed a smaller exponent j3 than predicted by self-consistent mean field theory, due to nonequilibrium effects (such as 'freezing' of chains in micelles) in better agreement with experiment. An important development in the application of self-consistent field theory to the analysis of block copolymers in solution was the extension of Scheutjens-Fleer theory to such systems. Self-consistent equations were solved numerically on a lattice, extending an earlier model developed for adsorption of homopolymers from solution.165'166 The formation of micelles of a diblock copolymer in a selective solvent was considered by van Lent and Scheutjens.167 These authors found that the cmc depends most strongly on the length of the block forming the micellar core and on the solvent quality. A strong repulsion between A and B segments (large XAB) was found to slightly disfavour association. The micelles were found to be usually spherical, but when the A blocks are much longer than the B blocks, a lamellar bilayer was predicted to be the preferred structure. They also obtained isotherms for adsorption at a solid-liquid interface. The lattice self-consistent field theory of Scheutjens and Fleer has been applied specifically to PEO/PPO di- and triblock copolymers in aqueous solution by Linse.168"172 The cmc, association number and hydrodynamic radius were determined for a number of copolymers,169'170 and a semiquantitative description of the temperature dependence of these quantities was compared with experimental results for Pluronic F127 (EO99-&-PO69-&-EO99).168'169 An increase in molecular mass or a decrease in the PEO/PPO ratio was found to reduce the cmt at a given polymer concentration, and to decrease the cmc at a given temperature. A similar trend was found on going from a triblock to a diblock. At high temperatures, a transition from spherical to rod-like micelles was observed, in agreement with the experimental phase diagram for several Pluronics.170 Linse also considered the effect of polymer impurities171 and copolymer polydispersity.172 For a PEO-b-PPO-b-PEO triblock solution, a PEO-b-PPO diblock copolymer and PEO and PPO homopolymers were considered as impurities.171 The presence of a diblock impurity was found to reduce the cmc and increase the aggregation number. Because the diblock associates more readily than the triblock, an enhancement of diblock copolymer content was observed at the cmc. Homopolymer PPO was also observed to reduce the cmc, but to a lesser extent. Solubilization of PPO in the hydrophobic micellar core was observed, leading to an enhancement of solubility by several orders of magnitude. This solubilization was found to increase with temperature, and with the volume fraction of free PPO. PEO was found to have a negligible effect on micellization. The effect of molecular mass polydispersity was modelled by Linse172 using a Schulz-Zimm distribution. It was found that polydispersity leads to a reduction of the cmc by several orders of magnitude, bringing the theory into closer agreement with experiment. The cmt is also reduced. The precise value of the cmc was found to depend strongly on the criterion used to define it. Polydispersity also leads to an
Neutral Block Copolymers in Dilute Solution
43
increase in micellar size, a greater separation of the EO and PO segments, and a marked decrease in micellar size with increasing polymer concentration. Close to the cmc the micelles are predominantly formed by the longest components. The strong temperature dependence of the cmc and association number was not significantly changed by polydispersity. Predictions of the poly disperse model were compared with experimental results for Pluronics (Table 2.1) L64515'51'63>64 P105 51 and pl27.15'51'69'168'173 Limited studies of chemical polydispersity were also performed by Linse,172 and this was found to have a similar effect to mass polydispersity. The Scheutjens-Fleer lattice self-consistent field theory has also been applied to model micelle formation and solubilization in PEO/PPO block copolymer solutions by Hurter et a/.174'175 They performed detailed calculations of the cmc, volume fraction profiles, association numbers and partition coefficients for Pluronic triblocks and Tetronic (PEO-fr-PPO)4 star blocks in aqueous solution. Building on this work, the model has been used to calculate the interaction parameters relevant to solutions of Pluronic triblocks in aqueous solution.176 This was used to calculate association numbers, micelle dimensions, radial density profiles, cmc values, etc. The Scheutjens-Fleer numerical self-consistent field model has additionally been employed to analyse interactions between block copolymer micelles and homopolymers, in particular it was applied to PEO-b-PPO-b-PEO triblocks in aqueous solution with dextran.177 It was concluded that the Flory-Huggins random mixing approximation restricts the model to the limit of weak concentration gradients within the phases. The x parameters were found to be strongly influenced by interactions among the components in the system.
2.7.4
THE MODEL OF NAGARAJAN AND GANESH
A theory for the self assembly of block copolymers into micelles in a selective solvent was developed by Nagarajan and Ganesh,178 following their earlier treatment of micellization in low molecular weight surfactants. In the theory, the copolymer-solvent system is treated as a multicomponent system consisting of solvent molecules, singly dispersed copolymer molecules and micelles of all possible sizes, each treated as a distinct chemical component. An expression for the difference in the reference state free energy between a copolymer in its micellized state and that in its singly dispersed state was derived. An expression for the equilibrium size distribution of micelles was also obtained. Illustrative calculations of micellar properties were carried out for a range of block copolymersolvent systems studied experimentally. In contrast to earlier theories,141'149'15 Nagarajan and Ganesh found that the solvent compatible coronal A block can have a strong influence on the micellar properties, especially when the solvent is very good for the A block.178 To compare their results with systems studied experimentally, they obtained scaling relations for PEO-£-PPO block copolymers in water (i.e.
44
Block Copolymers in Solution: Fundamentals and Applications
a system of a diblock in a good solvent), using interaction parameters estimated from activity data in the literature:
For PS-&-PB diblocks in the near-9 solvent n-heptane, the scalings:
were obtained.178 Clearly this model predicts a significant dependence of /?B and p on the size of the coronal block NA, in contrast to the earlier theoretical work141'149'158 and the predictions of Zhulina and Birshtein (except in regime III, Table 2.3). By combining numerical results for the systems PS-£-PB/heptane, PS-£-PI/heptane, PPO-b-PEO/ water and two model systems, Nagarajan and Ganesh obtained 'universal' scaling relations:
Here 7Bs is the core-solvent interfacial tension and XAS is the coronal blocksolvent interaction parameter.
2.7.5
COMPUTER SIMULATIONS
Monte Carlo (MC) simulations of micellization of block copolymers have to date been limited to short chain length (N < 32) model amphiphiles. The formation of block copolymer micelles has been investigated by MC simulations using a lattice model.179"182 Chain dynamics have also been examined.183'184 Further details are provided elsewhere.8 Binder and coworkers have investigated the micellization of short A2B2 symmetric185 and asymmetric diblocks with / = 0.25 and N = 4 to 32 ise T^y use(j off-lattice MC simulations with a bead-spring chain to explore the chain length dependence of aggregation number, density profiles and dimensions. In studying the dynamics of micellization for the symmetric diblock, an exponential increase in relaxation time with the strength of core block interaction energy was noted.185 Micellar size and shape distributions for a similar A2B2 molecule were
Neutral Block Copolymers in Dilute Solution
45
computed via lattice MC techniques by Hatton and coworkers.187 The self-assembly of this molecule was also investigated using stochastic dynamics simulations,188 in which solvent molecules are not included explicitly in the simulation which otherwise resembles molecular dynamics. Their effect is retained by incorporating random forces to account for uncorrelated solvent motions and by introducing a modified force field. The micellization of A10B10A10 and B5A2oB5 in a solvent selective for A were compared using lattice MC simulations.189 The additional entropy loss due to looping of the B5A2oB5 chains in flower micelles was shown to lead to micelles with a larger size and a broader size distribution compared with the Ai 0 B 10 A 10 micelles. The chain conformation in block copolymer micelles can also be modelled via computer simulations. An image showing the conformation of corona chains from a MC simulation of semiflexible chains grafted to a hard core, which represents the micelle core is shown in Figure 2.19.190 It is evident that the chains are rather dilute at the extremity of the corona, and that there are large variations in density. In fact, MC simulations of polymers tethered to a spherical surface indicate that block copolymer micelle coronas modelled in this way can be considered to be quasi twodimensional polymer solutions.191 The reduced surface coverage a = N-jrR';A/ [47r(/?B + RS.A) } (where RgA is the corona chain radius of gyration) is analogous to the concentration relative to the overlap concentration in bulk, c/c*.
Figure 2.19 Image from a Monte Carlo simulation showing chain conformation in the corona of a symmetric diblock copolymer.190 Reproduced by permission of American Chemical Society.
2.7.6
THEORY: ABC TRIBLOCK MICELLES
The formation of ABC triblock copolymer micelles in the strong and superstrong segregation limits was studied for melts by Dormidontova and Khokhlov.192 The
46
Block Copolymers in Solution: Fundamentals and Applications
Figure 2.20 Schematic showing micelle structure for ABC triblocks in the strong segregation limit.192 (a) Case where A block is much more incompatible with core-forming B block than corona-forming C block, (b) General case where multiple A aggregates formed. Reproduced by permission of American Chemical Society.
concepts are relevant also to ABC triblocks in solution, and consequently are discussed here. A particular focus was on the formation of micelles with internal structure due to the association of short A blocks (see, for example, Figure 2.20). Various arrangements of the A block aggregates inside the B block core were considered, and also spherical and disk-like shapes. The formation of internal structures also influences the association number compared with the BC diblock system. If the A block is strongly incompatible with the B block and less so with the C block, it forms a single aggregate in the centre of the micelle which has a lower association number than that of the BC diblock. Otherwise multiple A aggregates are formed inside the core. The A aggregates are predicted to be disk-like in the
47
Neutral Block Copolymers in Dilute Solution
superstrong segregation limit. Such structures have been observed for linear193 and mixed arm star194 ABC copolymer micelles by Lodge and coworkers (see Section 2.12.2).
2.8 MICELLE DIMENSIONS: COMPARISON BETWEEN EXPERIMENT AND THEORY The scaling of micelle dimensions with copolymer composition has been investigated by many researchers. Table 2.4 summarizes relevant results. Several of these papers also discuss the scaling of association number with chain length, as discussed in the following. Forster et al. also provide a useful table summarizing the scaling relationships from theory.195 The structure of micelles from two PS-&-PEO diblocks in the PS-selective solvents cyclopentane and d-cyclohexane determined by SAXS and SANS has
Table 2.4 Experimental studies on scaling of micelle dimensions with chain length. Adapted and extended from Forster et al.]95 Here N denotes the total degree of polymerization of the block copolymer, and Rg, R^ and 7?t indicate, respectively, the radius of gyration, the hydrodynamic radius and the thermodynamic radius of the whole micelle. System"
Scaling*
Method
PS-/?-PI/heptane PS-6-PI/DMA PS-£-PB/heptane PS-6-PB/heptane PS-6-PB/DMF PS-6-PAA/toluene PS-b-PMA and PMAb-PS-bPMA/dioxane + water PS-fcPCEMA/cyclopentane PS-6-P4VP/toluene PS-6-PEO/water PEO-£-PPO/water
/?g - N°-5
Rh~N°Rh~N°'5] RB ~ A^75 Diblocks: Rh - A^09W£71
SAXS DLS SAXS DLS DLS SAXS DLS
162 196 161 197 197 198 199
Triblocks: Rh - A^30A^56 Consistent with RA - p^5N^5
SLS/DLS
200
SLS/DLS SLS/DLS DLS (different exponent obtained for Rt from SLS) DLS (different exponent obtained for Rt from SLS)
195 201 70
PEO-£-PBO/water
Rh ~ yv°-5244
/?g ~ yv°-53
*A~p a 2 1 < 6 3
Consistent with R ~ N^N^5 R - A^:22
R - A^-2X'2°
Reference
70
"As noted in Section 2.4, micelles with a PS core may not be in equilibrium, with obvious effect on observed scaling relationships. ^Strictly scaling with N can only be considered for copolymers with constant composition (not the case for entries 2-6 in this table). Bluhm and Malhotra studied near-symmetric diblocks.162
48
Block Copolymers in Solution: Fundamentals and Applications
been compared with the predictions of the star-like polymer model.202 The scattering in the intermediate q range was analysed to provide the ^-average radius of gyration of the corona chains. The corona chains were considered as twodimensional semidilute polymer solutions (Section 2.7.5), with a blob size that decreases with increasing association number. Micelles with large PS blocks exhibit this behaviour under 9 conditions. In micelles with shorter PS blocks, the PS blocks are stretched and the scaling does not follow the Daoud-Cotton model. The scaling theory for spherical polymer brushes has been applied to analyse the coronal density profile of block copolymer micelles.203 If the density profile is of the form of a power law r~a, the brush height scales as:
For the PS-&-P4VP micelles studied by them, Forster et al.203 determined that the P4VP coronal density profile can be modelled as a power law with an exponent a between 1.05 and 1.35. Results from TEM experiments on solutions of a series of PS-b-PCEMA diblocks with short PS blocks and long PCEMA blocks have been compared200 with the theories for block copolymer micelles described above. Micelles of type IV in the Zhulina-Birshtein classification formed in cyclopentane, which is a selective solvent for PCEMA (coronal A block), when N\/NB > 9.200 Assuming that the association number is independent of N& (as predicted by scaling theories and the theory of Noolandi and coworkers), it was found to scale as p ~ A^ 92 , the exponent being in good agreement with these theories. The scaling of the core radius, R# ~ A/^ 63 , was also in good agreement with the theories of de Gennes, Zhulina and Birshtein, Halperin, Noolandi and Hong and Nagarajan and Ganesh (Section 2.7). Tao et al. also compared the scaling of the coronal thickness with the predictions of the Daoud and Cotton model for star polymers in a good solvent.200 Excellent agreement with the scaling [Equation (2.24)] was obtained (see Figure 2.21). Xu et al.201 used light scattering to characterize micelles formed by a wide range of PS/PEO di- and triblock copolymers in dilute solution in water. Although full analysis of the data was complicated by the tendency of the micelles to undergo secondary association, they did find that the micellar radius scaled as Equation (2.24). With values of p and R# from the star-like micelle model, it was possible to compute x parameters for the interactions of PEO with water and with PS, in good agreement with values obtained from independent measurements. Antonietti et al.204 determined association numbers for micelles formed by asymmetric PS-&-P4VP diblocks in solvents selective for PS. The association number was found to scale with N£, with a between 1.5 and 2, a much stronger dependence than anticipated by Halperin's theory, however in agreement with the predictions of the Zhulina-Birshtein theory for type III micelles,147 for which a — 2 (Table 2.3). Evidently the block copolymers studied by Antonietti et al. were not sufficiently asymmetric to lead to micelles in regime IV, i.e. the type of micelle also
Neutral Block Copolymers in Dilute Solution
49
Figure 2.21 Coronal thickness, LA, plotted as a function of A^ p1(/5 for micelles formed by PS-&-PCEMA diblocks in cyclopentane.200 This yields a straight line in accord with the predictions of scaling theory for the micellar radius (the core radius for these micelles was small enough to be neglected then LA = /?A)- Reproduced by permission of American Chemical Society.
considered by Halperin. This system was later studied by Forster et al., as discussed shortly. Booth and coworkers70'73'74'205'206 compared their results for the scaling of p and R with NA and A/B to the predictions of the Nagarajan-Ganesh model.178 For PEOb-PPO diblocks (and related PEO/PBO205'206 and PEO/PSO73'74 diblocks and triblocks) they found that p scales approximately as predicted, i.e. p ~ A^05^ ° [cf. Equation (2.40)] (here primes indicate that values of WA and NB were expressed relative to the critical values for micellization). The observed dependence of Rt on A/B was much steeper than predicted (the scaling of /?h is in better agreement, Table 2.4). The scaling of R^ ~ A/^0-3 was characterized by a significantly lower exponent than predicted by either the Nagarajan-Ganesh or Halperin models. As shown in Figure 2.22, Forster and coworkers have found that for a wide variety of amphiphilic block copolymers, the association number scales according to. 195,207
Here po is related to the interaction parameter x and the monomer volume and has a direct relation to the packing parameter for surfactants introduced by Israelachvili 7O8 et al. Explicity, it is given by:
50
Block Copolymers in Solution: Fundamentals and Applications
Figure 2.22 Association number p as a function of AfB and 7VA for PS-&-P4VP in toluene (o), PS-b-PMA in dioxane/water (D), PMA-b-PS-b-PMA in dioxane/water (A), poly(styrene-co-maleic acid-g-ethylene oxide) in water (V), alkyl ethylene glycol/water (0), alkylammonium bromides in water(H), alkylsulfonates in water (AX alkylsulfates in water (V)-195'207 Reproduced by permission of Wiley-VCH and American Institute of Physics.
Here Vis the molar volume, a the area per head group, / the contour length and e is related to the scaling of interchain distance with corona block length.195 It has to be noted that since PQ scales also with N% according to Equation (2.47), these relationships suggest that p ~ N#+6£, the exponent of which is higher than any theoretical prediction (Section 2.7). Furthermore, many of the systems included in the 'universal' plot are not diblock copolymers. The generality of this scaling therefore needs to be critically examined. The scaling with core block length Afe in Equation (2.46) agrees with the predictions of the Zhulina-Birstein model for micelles in regime III (Table 2.3), i.e. away from the limits of crew cut and star-like micelles. It applies to strongly segregated block copolymers, as emphasized by Forster et a/.195 The scaling with A^A does not agree with the Zhulina-Birshtein model, either in regime III or the other regimes where no dependence on N& is expected, as is also the case for starlike micelles in the Daoud-Cotton model [Equation (2.24)]. The dependence on NB is also much stronger than predicted by the Nagarajan-Ganesh model (Section 2.7.4). The results of Booth and Attwood also appear not to fit with this behaviour. Equation (2.46) was originally reported based on light scattering experiments on PS-&-P4VP diblocks in toluene, a PS-selective solvent.195 In the same work it was also reported that the corona dimensions (obtained via the hydrodynamic radius) scale as /?A ~p° 21 A^ A 63 , in good agreement with predictions for star-like micelles in a good solvent [Equation (2.24)].
Neutral Block Copolymers in Dilute Solution
51
A systematic study of the influence of the length of the soluble PEO block on the micellar structure was undertaken for PEP-b-PEO micelles modelling SANS curves using a Fermi function to describe the density profile of the shell.209'210 A related density profile with a power-law derived from scaling theory for star-like polymers has been used in the shell component of the form factor of spherical polymer brushes.203
2.9
INTERACTION BETWEEN MICELLES
The potential between polymer-coated colloidal beads211 has been used to analyse the intermicellar potential of block copolymer micelles.212 In three dimensions, the intermolecular potential of mean force is predicted211 to be:
where/is the number of attached polymer chains (equal to the association number/? in the case of micelles) and Rg is radius of gyration of the polymer chains. The scaling behaviour of the structure factor was also analysed and it was predicted that a peak would develop for concentrations near the overlap concentration, and that the peak height should scale with the same /3/2 dependence. An approximate interaction potential has been used to describe the structure factor of interacting block copolymer micelles. The potential for hard spheres with surface adhesion introduced by Baxter42 takes the form:
Here R — R' is the thickness of the adhesive surface layer (micelle corona). This potential has been used in the analysis of small-angle scattering data by a number of authors.43'213 Several attempts have been made to extract the interaction potential between block copolymer micelles. Gast and coworkers used self consistent mean field theory to study interactions between spherical diblock copolymer micelles in solution.214'215 The theory was used to calculate intermicellar pair potentials and combined with liquid state theory this enabled a comparison with the static structure factor determined from SANS experiments on PS-&-PI diblocks in decane, a selective solvent for PI. Further details are provided in Section 3.5.1. Buitenhuis and Forster have used rheology to obtain the interaction potential between PS-£-P4VP diblocks in toluene, in the solid gel phase.216 They used the relationship between the high frequency shear modulus and the interaction potential obtained by Zwanzig and Mountain from liquid state theory, which had earlier been applied to colloidal dispersions.
52
Block Copolymers in Solution: Fundamentals and Applications
As discussed in Section 2.2.9, the hard sphere structure factor has been used extensively to model interactions among block copolymer micelles, and generally provides an excellent fit to the structure factor data. Brown et al. used SANS to obtain the structure factor of PS-fr-PEO micelles in aqueous solution.217 Analysis using the Yukawa potential provided the interaction potential, which was explored as a function of concentration, i.e. micellar overlap. This potential is purely repulsive, but the soft tail increases as the overlap increases (Figure 2.23).
Figure 2.23 Yukawa potential for dPS-b-PEO diblocks deduced from SANS measurements of the structure factor, at the concentrations indicated (units 1(T2 g ml"1).217 Reproduced by permission of Elsevier.
SANS has been used to investigate the percolation transition due to the attractive intermicellar interactions of EOl9-b-PO4i,-b-EOl9 micelles in aqueous solution, where micellar interactions were described by the sticky hard sphere model for the attractive interaction potential.43'218"220 The second and third virial coefficients expressing two- and three-body interactions among diblock and triblock copolymers have been evaluated using mean field theory.221 Conditions under which phase separation occurs were identified.
2.10
DYNAMICS OF MICELLIZATION
The dynamics of the micellization process have been studied by laser temperature jump and stopped flow techniques, using light scattering or fluorescence measurements to probe changes in micellar dimensions or association number. Most studies
Neutral Block Copolymers in Dilute Solution
53
have been performed within the micellar state, i.e. experiments are performed where the conditions are changed from one state of micellar equilibrium to another. However, some experiments have investigated the transition between unimer and micellar states. Honda and coworkers have investigated the kinetics of micelle relaxation following a temperature jump or quench via light scattering experiments.222'223 Changes in apparent micelle molar mass, Mw,app, and radius of gyration, Rs,app, were monitored. For a quench to an increasing depth from the cmt, it was found that the micelle mass increased monotonically whereas the radius of gyration first decreased and then increased slightly. This is consistent with the association of existing unimers into new small micelles, which grow as fusion and exchange between old and new micelles occurs. In contrast, for a temperature jump (T-jump), the molar mass and radius of gyration decreased in parallel initially indicating an increase in the number of micelles accompanied by a decrease in association number. This was ascribed to micelle dissociation into unimers, a process that occurs faster than micelle breakup, even though the most efficient way to decrease the free energy of the system is breakup into small micelles.224 Both Mw>app and Rg?app then increase to their equilibrium values, the former increasing more rapidly. This indicates a decrease in the number of micelles with an increase in aggregation number.223 This is probably due to the coupling of small micelles. Micellar requilibration is slower at low concentration than micelle formation directly from unimers,223 consistent with a micelle formation/breakup mechanism. The same group have also investigated co-micellization induced by T-jumps in solutions of binary mixtures of diblocks with different soluble block lengths.225 They also investigated the kinetics of vesicle formation by PS-b-PDMS dibocks in a mixed selective solvent.226'227 The process occurred via a transient structural intermediate, comprising hollow cylinders. The kinetics of micellar re-equilibration following a laser-induced T-jump (1 °C heating in 2.4 (is) have been studied for a number of Pluronic block copolymers.228 The scattered light intensity was monitored during the relaxation - Figure 2.24 shows representative data. The study was performed above the cmt, in the region where significant amounts of unimer coexist with micelles. Two processes were observed. The fast process was characterized by an increase in scattered light intensity (timescale 10 us - 10 ms) and was associated with unimer insertion into micelles. The second slower process was characterized by a decrease in scattered light intensity over a 1-100 ms timescale, and was attributed to the rearrangement of the micelle size distribution. The Aniansson-Wall theory229'230 was used to interpret the experimental results. Kositza et al. have investigated micellization dynamics in solutions of Pluronics L64, 231~233 pg4 231 and P104 231 using T-jump and stopped flow techniques. The relaxation from the nonequilibrium state was monitored in the following 1.5 s, either through the intensity of scattered light (in a certain wavelength interval) or using a fluorescent probe, the fluorescence from which increases in a hydrophobic environment.231'232 In the stopped flow experiments, NaCl solution in one syringe was mixed with the aqueous copolymer
54
Block Copolymers in Solution: Fundamentals and Applications
Figure 2.24 Representative data from a laser T-jump experiment showing the scattered light intensity as a function of time during the relaxation process for Pluronic triblock P84.228 Reproduced by permission of American Chemical Society.
solution in another syringe.231'233 Increasing NaCl concentration corresponds to an increase in temperature (Section 2.6.3) or an increase in copolymer concentration.231 With this method it was possible to study the kinetics of equilibration between two states with different micelle association numbers, or alternatively the transition from unimers to micelles. From both T-jump and stopped flow experiments, three distinct relaxation processes were identified. A fast relaxation process was ascribed to the exchange of unimers between solution and micelle (this was observed in the experiments from one micellar state to another and for the stopped flow unimer-micelle experiments). The amplitude (change in light scattering intensity) is positive for this process. The associated lifetime decreased as temperature or copolymer concentration increased. This process is equivalent to the first step in the Aniansson-Wall234 mechanism of micelle formation. The second process, with a negative amplitude, was ascribed to the redistribution of molecules in micelles, i.e. to changes in association number. The negative amplitude arises from a decrease in dimensions of micelles due to dehydration. A third relaxation process was related to the clustering of micelles. It was noted that it is necessary to be cautious in comparing results from the two techniques. The perturbation caused by dilution with salt solution in the stopped flow experiments is much greater than that in the T-jump experiments, since there is a large change in solution concentration and the salt itself may alter the micellar structure as well as the micellization point. Waton et al. 235 have suggested that the second and third relaxation processes observed by Kositza et al. are in fact associated with the same process of micellar equilibration. In fact, they were able to show that the relaxation times and associated amplitudes of these processes lie on a common curve as a function of temperature (this was not noticed by Kositza et al. since the two processes were never observed simultaneously). The timescales of the two processes are typically
Neutral Block Copolymers in Dilute Solution
55
10 |is to 10 ms, depending on temperature and concentration, for the fast relaxation process (exchange of unimers between solution and micelle) and 0.2 to 100 ms for the slow process ascribed to micelle formation and breakup.232 Waton et al. have investigated the kinetics of formation and breakup of micelles of Pluronic block copolymers (L64 and PF80) by T-jump light scattering experiments,235'236 and ultrasonic absorption spectroscopy.236 A fast process observed via laser T-jump experiments (monitored by scattered light intensity measurements) was ascribed to the exchange of molecules between micelles. This process was also observed via ultrasound absorption. The slow process was ascribed to micellar re-equilibration due to an increase in the number of micelles, i.e. micelle formation and breakup. These processes follow the Aniansson-Wall model.229'230 The amplitude of the relaxation process associated with the micellar formation-breakup mechanism (expressed as the relative change in intensity of the scattered light) was found to go from negative to positive on increasing temperature.235 This conclusion was supported by a model based on the temperature dependence of the moleculemicelle equilibrium. The relative importance of unimer exchange and micellar fusion during micellization has been investigated theoretically, by combining Kramers theory calculations of association/dissociation rates with scaling theory for micellar relaxation times and activation energies.224 It was found that early in the micellization process, following the initial association of unimers, the fusion of micelles becomes the dominant process, because unimer exchange is penalized by a high activation energy. In the later stages, both unimer exchange and fusion play a role, the latter slowing down as micelle size increases. Micelle fission is also relatively slow, although important in the re-equilibration associated with a decrease in association number (brought about for example by a T-jump). The results were compared with the results of T-jump experiments.223'236'238 The kinetics of exchange between micelles of deuterated and protonated PEP-bPEO block copolymers was investigated by SANS, and a two-stage intensity decay was observed.237 The fast process was associated to the exchange of single unimers. The interpretation of the slow process was more ambiguous. Exhange kinetics of PDMA-£>-poly(sodium alkyl acrylate) diblocks and the two corresponding types of triblock [PDMA midblock or poly(sodium alkyl acrylate) midblock] chains forming micelles in aqueous solution have been investigated by steady state fluorescence spectroscopy.238 Molecules containing naphthalene covalently bound to the hydrophobic poly(sodium alkyl acrylate) block formed micelles that coexisted with micelles of unlabelled copolymers containing soluble pyrene (fluorescence acceptor). Slow unimer exhange with a rate constant ~10~3 s ] was observed, decreasing on making the alkyl chain more hydrophobic or going from a diblock to a triblock architecture. Molecular exchange in block copolymer micelles is arrested if the core block is glassy, as confirmed for micelles containing ps239"241 or PMMA242'243 cores. A study of the redistribution of chains in micelles formed by two PS-&-P2VP diblocks with different composition in solution revealed that the equilibrium state with a
56
Block Copolymers in Solution: Fundamentals and Applications
bimodal size distribution was only reached after many months, starting from the initial state of mixed micelles.240'241 A theoretical model for chain distribution was also developed.240 Steady-state fluorescence measurements also revealed no exchange for PS- and P/BS-containing diblocks [with a water-soluble PNaMA block] at room temperature, although some exchange was observed at 60 °C, this being ascribed to a reduction in glass transition temperature (rg) due to the low molar mass of the hydrophobic blocks.243 The influence of core block Tg was also confirmed by experiments which revealed that exchange kinetics were accelerated when solvents miscible with the core ('plasticizers') were added.243 The kinetics of molecular exchange in micelles has also been investigated for systems with a nonglassy core block, specifically PB-&-PEO diblocks by SANS on selectively labelled copolymers.244 It was found that even in this case, the micelle structures formed upon dissolution are completely locked in. This was ascribed to the level of amphiphilicity, i.e. due to the presence of a highly hydrophobic polybutadiene core. Jain and Bates note that the timescale for micelle equilibration can be extremely long, perhaps many years.245 They studied micellization using pairs of PB-^-PEO diblocks and observed completely different morphologies, depending on whether the copolymers were pre- or post-mixed prior to dispersion in solution. Despite this global nonergodicity, they argue that a state of local equilibrium can be reached via the redistribution of copolymer chains within the topology established during dispersion. Webber and coworkers have investigated the kinetics of release of fluorescent probes from micelles formed by three types of diblock copolymer in aqueous solution.246 The micelle cores were formed by PS, PtBA or P2VP. The former two are glassy and the diffusion constant for release was very small (D = 10~1810~16 cm2 s"1) whereas for P2VP the release kinetics were too fast to be measured. On the basis of fluorescence experiments, it was possible to distinguish inner and outer corona regions. The ionization of the inner corona was suppressed, whereas in the outer corona there was a much higher charge density, and the chains were much less crowded. Confined impinging jet mixing has been used to induce the rapid self-assembly of PBA-b-PAA micelles in aqueous solution.247 A jet of the polymer dissolved in methanol was mixed with a stream of water to selectively precipitate the PB A block. The supersaturation ratio was large and the change in supersaturation rate was very rapid. The micelles self-assemble into a nonequilibrium structure via a nucleation and diffusion limited growth process in which pre-micellar fusion occurs until a critical size is reached, at which polymer brushes overlap. The micelle size or aggregation number depends on the rate and magnitude of the solvent quality change.
2.11
DYNAMIC MODES
DLS has also been used to probe the dynamics of individual blocks. For example, it has been used to investigate the dynamics of PS chains in a PS-b-PMMA
Neutral Block Copolymers in Dilute Solution
57
diblock in a mutual good solvent, isorefractive for PMMA.248 Two modes were identified. The slow mode was ascribed to diffusive motion of the PS chain, which was however strongly coupled to internal modes of motion which deform the block copolymer chain during diffusion. The fast mode provided information on internal modes of motion (the so-called 'copolymer mode'). The corresponding decay rate scaled as F2 ~ q3, which is characteristic of chains with nondraining hydrodynamic interactions. The intermolecular interactions were examined through the concentration dependence of the hydrodynamic virial coefficient [Equation (2.4)] which was described by the Ackasu-Benmouna formula, originally derived for flexible linear polymers in good solvents.249 Further information on diffusive dynamics (also in semidilute and concentrated solution) is provided in Section 3.10. The dynamic structure factor in dilute homogeneous solutions of very high molar mass PS-b-PI diblocks in a neutral solvent has been investigated.250 The polymers with molar masses Mw « 2 x 106 were weakly entangled. The effects of composition and polydispersity on the intermediate scattering function were examined. This function obtained from DLS revealed three relaxation processes. Two modes are due to chain conformational motions assigned to Rouse-like backbone modes of reptating chains and one is due to centre of mass chain diffusion. The latter is inactive for compositionally monodisperse diblocks. Inelastic (Brillouin) light scattering has also been performed on these systems, this providing the phonon dipersion spectrum which was found to be sensitive to the orientation of the structure with respect to the scattering 9S1 vector. NMR has been used to probe the dynamics of chains in block copolymer micelles. The dynamics of PS-&-PEP chains in micelles formed in paraffmic solvents was studied by *H and 13C NMR, in tandem with DLS and TEM which provided direct evidence for spherical micelles.252 In dilute solutions in n-octane below 50 °C, the PS core was found to have two components, a 'rigid' component with the same NMR linewidth as bulk glassy PS and a 'mobile' component atrributed to a plasticized surface layer. The 'rigid' component underwent a sharp linewidth transition at 50 °C attributed to segmental motions. In n-octane, the micelles were found to dissociate above 80 °C. In higher alkanes, the dissociation occurred at increasing temperature with increasing solvent molecular weight. High resolution NMR spectroscopy was used to investigate chain dynamics in micellar solutions of a PS-&-PB diblock in heptane (selective for PB) and a PS-£-PB-£-PS triblock in 1,4-dioxane and 1,4-dioxane/D2O mixtures (precipitant for PB).253 An increase in mobility of PS units above about 40 °C was observed in the PS-&-PB solution, caused by a gradual swelling of the PS core. For the PS-b-PB-£>-PS solution, considerable mobility of the PB was noted and entanglements between PB in the core could not be detected, which was ascribed to the effect of isotropic Brownian motion averaging the dipolar interactions. The order of magnitude difference in segmental mobilites of PS and PB explained why this effect was not relevant for the system with PS-core micelles.
58
Block Copolymers in Solution: Fundamentals and Applications
Field gradient NMR has been employed to determine the self-diffusion coefficient of Pluronic P85, and the hydrodynamic radius has been compared with DLS measurements on the same system.254 NMR was found to give a somewhat lower value for the hydrodynamic radius than DLS. However, at infinite dilution the values obtained from the two techniques are the same. A similar observation has been made for cyclo-PBO27PEO144 in aqueous solution.20 This effect has been attributed to the difference in dynamic averaging for the DLS and NMR experiments.255 In DLS, monomers and micelles are distinguished from their relaxation times, which depend on the z-average dimensions of each. In contrast, the NMR diffusion coefficients are close to a number average value. Thus, the NMR measurement reflects averaging of both monomeric and micellar motion and hence provides lower values for diffusion coefficients. Proton NMR longitudinal and transverse relaxation time measurements have been used to determine changes in segmental dynamics upon micellization of Pluronic F127.256 At the cmt, there was a marked transition in the relaxation times of the hydrophobic PPO block attributed to a change from well solvated mobile chains below the cmt to a more restricted, concentrated micelle-core environment above the cmt. However, the dynamics of PEO segments were not changed at the cmt, indicating the persistence of a solvated, mobile chain structure. Transient nuclear Overhauser effects indicated considerable interpenetration of PEO and PPO blocks at the interface and relaxation times analysed using a two mode correlation function supported the conclusions from relaxation time measurements regarding segmental mobilities. Chain dynamics at the segmental level in the corona of micelles formed by ionic block copolymers or in the area surrounding the ionic cores have been investigated using 2H NMR experiments. Segmental dynamics were probed in micelles formed in CC\4 by PS-£>-PNaA (and also sodium carboxylate terminated polystyrene), where part of the PS block was deuterium labelled (ca. 3 units). The distance between the 2H-labelled segments and the ionic cores was controlled by the number of styrene units separating the labeled segment from the nonionic-ionic block junction. NMR line widths, signal intensity, and relaxation times of the block ionomers and their nonionic precursors (PS-&-P/BA) indicated that the mobility of the soluble segments near the ionic cores was dramatically reduced. At a distance of 25 repeat units from the nonionic-ionic block junctions, the mobility was still significantly lower than in single chains, while at a distance of 50 repeat units from the junction, the mobility was essentially the same as that in the single chains. Even for sodium carboxylate- terminated PS, where there is only a single ionic group, the 2 H-labelled styrene segments 14 repeat units away from the block junction experienced restricted mobility due to the ionic association. Deuterium-labelled ionomers with the same PS block but different lengths of ionic blocks were also examined, and it was found that the longer the ionic block, the slower the motion in the coronas, but the effect was diminished for ionic blocks with more than six repeat units.257
Neutral Block Copolymers in Dilute Solution
59
Kfiz et al. have performed single and multi-quantum !H NMR and also magicangle spinning (MAS) NMR on a range of poly(alkyl methacrylate)-b-poly(sodium acrylate) diblocks in D2O.258'259 From quantitative measurements of transverse relaxation times, they found a dramatic reduction in mobility of the alkyl side chains, as well as of the backbone, when micelles formed. As expected, the most immobilized groups are those near the block junction (micelle interface). Segmental motion in the inner part of the shell was also found to be hindered.258 The motion in micelles containing cores swollen with chlorobenzene was also analysed.25 High molecular weight PEO was also added to the solvent to decrease micellar tumbling to improve the resolution of the spectra.259 Segmental motion in homopolymer corresponding to the core block was found to be reduced compared with the bulk, although the motion of the 'free' chains was not as restricted as the tethered chains in the copolymer.258 X-ray photon correlation spectroscopy (XPCS), has been applied to investigate the dynamics of block copolymer micelles.260 The principle of this technique is identical to that of DLS (or PCS). Fluctuations in the intensity of a coherent X-ray beam are analysed to provide information on collective dynamics, at X-ray wavelengths. A wavelength-dependent diffusion coefficient was found for the PSb-Pl micelles, with a weak peak at the same wave vector as the peak in the static structure factor. This was ascribed to hydrodynamic interations between the particles, mediated by the solvent. XPCS has also been used to investigate the dynamics of poly electrolyte block copolymers below and above the overlap concentration.261 The PS-b-PCsA diblocks in toluene remained liquid despite the overlap of coronal chains, as confirmed by the static structure factor and the ergodicity of the intensity correlation function. The diffusion was greatly slowed down (by a factor ~106 to 1.3-7.9 x 10~14 cm2 s"1) in the gel compared with the liquid. Two diffusive modes were ascribed to collective motions of micelles within their cages and their motion between cages. The dynamics of chains in block copolymer micelles has been investigated by neutron spin echo (NSE) by several groups.262"264 In recent work,263'264 the decay of S(q,t) has been represented by a simple two mode relaxation function (see for example Figure 2.25), in contrast to a previous (NSE) study262 on block copolymer micelles where the normalized dynamic structure factor for several values of q was fitted to a complicated model for the 'breathing modes' of tethered polymer chains265 involving a large number of coefficients. Despite the complicated equation used (involving a summation over 78 coefficients in an expansion containing an integral of a Bessel function) the model did not fit the data at low q. For the PEO-b-PBO diblocks in aqueous solution, two modes were observed - a slow mode with a diffusion coefficient that corresponds to the translation of the micelles and a fast diffusive mode due to internal 'blob scattering', although the decay rate at q — 0 is nonzero. In contrast to Matsuoka et a/.,263 Castelletto et a/.264 also investigated the micellar dynamics in the gel phase, as well as in the micellar solution.
60
Block Copolymers in Solution: Fundamentals and Applications
Figure 2.25 NSE normalized dynamic structure factor S(q,t)/S(q,Q) for a 12.5 wt% of diblock EO92-£-BO18 measured at 30 °C, for wavectors q = (A) 0.037, O) 0.049, (Q) 0.066, (•) 0.075, (O) 0.099 and (•) 0.15 A"1.264 Reproduced by permission of J. Chem. Phys.
2.12 SPECIFIC TYPES OF MICELLES 2.12.1
MICELLES FROM TELECHELICS
Triblock copolymers with short solvophobic end blocks (so-called telechelics, Section 1.1) can self-assemble into flower micelles, in which the midblocks are looped (Figure 2.26). Bridging between micelles occurs when the midblock spans the space between different micelles. This type of copolymer system is also a model for associating polymers. These are of great industrial importance due to applications where they are used as associative thickeners or hydrogels in biomedical materials or in separation media. A particularly important class are telechelic endcapped polymers such as the HEUR (hydrophobic ethyoxylated urethane) copolymers, which comprise poly(ethylene glycol), chain extended by diisocyanates, and end-capped by long-chain alcohols.266'269 Analogous F-HEUR polymers are terminated at both ends with hydrophobic fluoroalkyl segments.270"276 These are much more effective thickeners compared with the corresponding hydrocarbon derivatives.270'271'273 Since the urethane linking group can influence the association properties (especially for copolymers with short hydrophobes), effort has been devoted to the synthesis of polymers in which these linkers are absent.277"281 In telechelic copolymers, bridging of flower micelles can produce a necklace or string configuration282 (also known as a superbridge structure267'283'284) as sketched in Figure 2.26. The relationship between structure and rheology for associative
Neutral Block Copolymers in Dilute Solution
Figure 2.26 copolymers.
61
Flower micelle, superbridge and extended network formed by telechelic
triblock copolymers fully end capped with alkyl chains has recently been interpreted in terms of the structure and interactions of flower micelles.285 The micelles were treated as adhesive hard spheres. The 'stickiness parameter' was calculated using a model due to Semenov that enabled the second virial coefficient to be computed, given the micellar radius and association number, which were determined by DLS.286 It was found285 that at low concentrations the rheology is dominated by pairwise interactions between associating spherical micelles. As the concentration increases, the density of bridging chains increases and the behaviour approaches the ideal transient network limit. The relaxation time scales as the diffusion time, rjR^/k^T, for a hydrophobe escaping from a micellar core of radius /?B (77 is the solvent viscosity), multiplied by a Boltzmann factor accounting for an association energy that increases linearly with hydrophobe length. If the bridges span the system, gelation occurs due to the formation of an extended network (Figure 2.26). The influence of bridge chain density on the modulus of gels has been examined by mixing PBO-b-PEO-£>-PEO triblock with a PEO-b-PBO diblock (which cannot form bridges) in aqueous solution.287'288 This is discussed further in Section 3.3.5. The micelle structure, and interactions, of a diblock and triblock of PEO with one or two hexadecyl hydrophobes has been probed by SANS.280 In dilute solution, the hard sphere structure factor is sufficient, and the micelles can be modelled using expressions obtained for star polymers. Normal and flower-like micelles were not
62
Block Copolymers in Solution: Fundamentals and Applications
distinguished. In semidilute solution, there is a liquid-like arrangement of micelles, the effective interaction radius increasing with concentration. However, the influence of attractive interactions that must result from bridging of the telechelic chains was not investigated. This group has more recently investigated the effect of copolymer architecture (mono- and difunctionalized chains), PEO chain length and hydrophobe size on the critical aggregation concentration (cac), using pyrene fluorescence probe measurements.281 The cac was similar for diblocks and triblocks (making the comparison at a fixed ratio of hydrophobic/hydrophilic units) and increased strongly with increasing PEO chain length. It has recently been shown by SANS that at high concentration, the hydrophobic aggregates in F-HEUR copolymers with short fluoroalkyl hydrophobes, form a body-centred cubic (BCC) structure.276 This ordering may result simply from the packing of soft spheres at sufficiently high density. The dynamics of networks formed by telechelic associative thickeners284 are predicted to be dominated by the detachment of hydrophobic end groups from the clusters, such that the motion of the chain is governed by a detachment time plus the Rouse motion of the detached chain.284 For HEUR associative polymers, a relaxation time related to the network relaxation has been noted, in addition to the relaxation time related to the lifetime of the hydrophobe in the micellar junction. 289 The aggregation behaviour of PEO end capped at one or both ends by the fullerene C60 has been examined.290 Very large association numbers (p > 104) were reported due to the highly hydrophobic character of C60. 2.12.2
MICELLES FROM ABC TRIBLOCKS
Liu and coworkers have observed micellization of PBMA-b-PCEMA-&-P?BA triblocks.291 The PCEMA block forms the core, which was photo cross-linked. Hydrolysis of the PfBA yields an amphiphilic poly(acrylic acid) corona. Micellization in a coil-crystalline-coil ABC triblock was examined by Manners and coworkers.292 The PFP-b-PFS-b-PDMS copolymer was dissolved in hexane, a selective solvent for PDMS. This led to the formation of micelles with an organometallic core. The morphology varied from rod-like for short PFP blocks to spherical for longer PFP blocks, which was ascribed to reduced crystallinity as probed by WAXS. The formation of spherical core-shell-corona micelles has been observed for PS-&-P2VP-&-PEO triblocks in water.293'294 The P2VP shell is pH responsive.293 At pH > 5, the P2VP shell is neutral, hydrophobic and collapsed on the PS core, whereas at pH < 5, the P2VP shell is protonated, water soluble and has an extended conformation. A transition to a rod-like shape was driven by addition of the PS-selective solvent benzene to the aqueous solution. Furthermore it was possible to produce gold nanoparticles in the P2VP shell. The conformation of the P2VP chains in the inner shell of micelles of a PI-6-P2VP-6-PEO triblock in aqueous solution has been shown to expand in acidic solution as the P2VP becomes
Neutral Block Copolymers in Dilute Solution
63
charged, although in this case the core-shell structure of spherical micelles was retained.295 Multiple aggregate structures were observed for a PS-&-PMMA-/?-P?BA triblock in mixed solvents.296 Following the usual procedure adopted by the Eisenberg group (other work is described in much more detail in Section 4.1.2) the polymer was first dissolved in an organic solvent (dioxane, THF or DMF) and then dialysed against water. Spherical, rod-like and vesicular (nonequlilibrium) structures were observed. In solutions of this double hydrophobic ABC triblock, the cmc was found to be controlled by the PS block. Patrickios and coworkers have investigated micellization in PDMA-&-PEMA-&PMAA and PDMA-b-PMMA-£-PMAA ABC triblocks (and also diblocks containing combinations of these polymers).297 These copolymers are polyampholytes containing a hydrophobic midblock and oppositely charged endblocks in an appropriate pH range. Isoelectric points were determined by ti trad on. The precipitation of an oligomeric PDMA-b-PMMA-&-PMAA polyampholyte at the isoelectric point was investigated by turbidimetry. This method to induce precipitation could be relevant to the extraction of solutes via electrostatic complexation.298 Structures analogous to those formed by ABC triblocks can also be formed by coprecipitation of AB and BC diblocks. This leads to the formation of so-called onion micelles (Figure 2.27). SANS has been used to probe the structure of onion
Figure 2.27 Onion-type micelle formed by co-precipitation of P2VP when an alkaline solution of P2VP-&-PEO is added to a solution of protonated PS-6-P2VP micelles.431 Reproduced by permission of American Chemical Society.
micelles formed from PS-&-P2VP combined with P2VP-6-PEO to produce a threeshell structure with a PS core and PEO corona.299 Despite the complexity of the onion micelles the scattering data could be fitted using simple models for the form factor: either that for polydisperse uniform spheres (so-called 'bare core' approximation) or the Pedersen-Gerstenberg model (Section 2.2.9). Three-layer nanoparticles have also been prepared by 7-radiation-induced polymerization of methyl methacrylate (MMA) around the PS core of PS-b-PMAA micelles in aqueous solution.300'301 The structure of micelles of a PEHA-£-PMMA-fc-PAA ABC triblock in H2O/D2O mixtures have been studied via SANS using contrast variation by
64
Block Copolymers in Solution: Fundamentals and Applications
selective swelling of PEHA blocks with d-cyclohexane.302 Modelling of the scattering data showed that the three blocks are segregated in a micelle, PEHA forming the inner and PMMA the outer layer of the core. Lodge and coworkers have recently shown that it is possible to access the superstrong segregation limit192'303 using linear ABC triblocks that self-assemble into core-shell disk-shaped micelles.193'304 The micelles in aqueous solution comprised a fluoropolymer block core, a PS shell and a PEO corona. The corresponding block copolymer with a nonfluoro-derivatized polybutadiene core block formed a regular spherical core-shell-corona structure. The formation of the disk-like micelle in the system with more strongly segregated core and shell blocks was driven by the increase in interfacial tension, which overwhelmed the entropic penalty associated with chain crowding. This was rationalized on the basis of the free energy of micellization, calculated using a polymer brush model. Even more remarkable aggregate structures were observed for PEE/PEO/PFPO [poly(perfluoropropylene oxide)] mixed arm ABC triblocks.194 Polymers are denoted with molecular weights (in kDa) x-y-z representing PEE-PEO-PFPO. Copolymers 2-13-2 and 2-13-3 formed mainly spherical micelles, with some three- and four-lobe micellar structures (Figure 2.28). Copolymers with shorter PEO blocks formed multicompartment structures - strings of micelles or segmented worm micelles
Figure 2.28 Cryo-TEM image showing three- and four-lobe micelle structures formed by a PEE/PEO/PFPO mixed arm star terpolymer (2-13-3) in 1% solution in water.194 The scale bar indicates 50 nm. Reproduced by permission of Science.
(Figure 2.29). The self-assembly into these structures (which may not be in global equilibrium) is possible since they facilitate reduced crowding of corona chains. This was again rationalized on the basis of strong incompatibility between the blocks (superstrong segregation regime), in particular between PEO and PFPO. There is then a preference for flat interfaces, which is penalized since three blocks meet in one junction point. The PEE and PFPO domain sizes are also restricted by chain length - in fact the PFPO blocks were found to be fully extended. Models for the self-assembly are also shown in Figure 2.29. The formation of multicompartment micelles by rational self-assembly using synthetic block copolymers is interesting also in relation to the biological self-assembly of compartmentalized systems, such as eukaryotic cells. Wooley and coworkers have used cross-linking to trap intermediate structures such as strings of micelles ('pearl necklace' structures305) across the rod-to-sphere
Neutral Block Copolymers in Dilute Solution
65
Figure 2.29 (a) Cryo-TEM images showing a string of micelle (A) and segmented worm (B,C,D) morphologies of PEE/PEO/PFPO mixed arm star terpolymers 2-7-2 (A) and 2-9-2 (B,C,D).194 (b) Models for an individual 2-7-2 micelle and stacking into a string as in (a)A. (c) End-on and side-on views of segmented worm micelles formed by 2-9-2 as in (a) B,C,D. Reproduced by permission of Science.
66
Block Copolymers in Solution: Fundamentals and Applications
micelle transition in a PAA-b-PMA-b-PS triblock.306 The PAA shell was crosslinked as described in Section 2.12.4. The PS core domain was solvated using THF to drive the morphology transition. Alternatively, rod-shaped micelles could be preserved by using a cross-linking chemistry that proceeded faster than the morphological transition. ABC triblocks offer the possibility to synthesize so-called Janus micelles, in which the two faces of the nanosphere comprise distinct chains, as discussed in Section 2.12.5. 2.12.3
MICELLES FROM ROD-COIL COPOLYMERS
The main category of rod-coil block copolymers for which solution nanostructures have been investigated are those based on polypeptide rod blocks. This is discussed in Section 4.8. Nanostructure formation by synthetic PPQ-&-PS diblocks has been investigated, with a view to exploiting the photoactive and electroactive properties of the TTconjugated PPQ block.307 The self-assembly of PPQ-£-PS rod-coil diblocks in selective solvents for the PPQ rod block307 or the PS308 coil block results in very large aggregates (typically 3-5 um across). The large size resulted from the formation of hollow structures (the fully extended length of the polymer chains was much smaller than the observed aggregate dimensions). For example, hollow micelles with a monolayer shell (Figure 2.30), instead of a bilayer as in a vesicle, were formed as a result of the molecular shape. In a PPQ-selective solvent, a range of aggregate structures was observed, their size decreasing with decreasing fraction of rod block. Spherical aggregates were able to solubilize large molecules such as fullerenes, as discussed further in Section 6.2.1. In a selective solvent for the PS block, hollow spherical aggregates were observed which formed ordered arrays in two and three dimensions when dried.308 Due to the size scale of the particles, iridescence was observed as a result of the spatial variation of refractive index. Potential applications as photonic band gap materials were highlighted. The solution self-assembly of rod-coil diblocks capped with a small dendritic unit at the end of the rod block has been investigated by DLS, NMR and TEM on dried films.309'310 The copolymers comprise a polyisoprene coil block linked to a short p-biphenyl ester trimer which is capped with a first generation dendritic unit. The dendritic units are capable of hydrogen bonding and this, together with TT-TT stacking of the biphenyl ester units leads to the formation of a ribbon-like superstructure from stacked rod-coil dendron tetramers. The influence of hydrogen bonding on gelation in organic solvents was studied by controlling the number of hydroxyl units in the dendritic unit. The polarity of the solvent was found to be important in controlling aggregation. In solutions, DLS was used to investigate the kinetics of the self-assembly in two different organic solvents.310 Nanostructure formation in aqueous solution of a different class of rod-coil dendron copolymer was also examined. A cholesteryl terminated L-lactic acid block was linked to an
Neutral Block Copolymers in Dilute Solution
67
Figure 2.30 Schematic showing self-assembly of a PPQ-&-PS diblock into a hollow microsphere.307 Reproduced by permission of Science.
68
Block Copolymers in Solution: Fundamentals and Applications
L-lysine dendrimer (Gl, G2 or G3). A lamellar structure was seen in concentrated aqueous solution. Upon dilution, discrete nanosized aggregates were observed for the copolymer containing a G3 dendron. It has been observed that well-defined solution aggregate structures can be observed for multiblock copolymers comprising rod-like PMPS segments interspersed with flexible PEO segments, despite the fact that the PMPS is relatively polydisperse (the PEO was nearly monodisperse).311 PMPS is interesting due to conjugation of s electrons that gives rise to electronic properties with potential applications in materials with semiconducting, electroluminescence and nonlinear optical properties. Vesicles and micellar rods were observed, together with a novel helical structure resulting from pairwise association of helices formed by the polysilane groups. Superhelical structures have also been observed for peptidebased block copolymers in solution, as discussed in Section 4.8. Tu et al. have observed spherical micelles with a core of the rod-like poly {2,5-bis-[(4-methoxyphenyl)oxycarbonyljstyrene} block and a corona of PS in a selective solvent for PS.312 The micellization of PS-b-PHIC block copolymers in a selective solvent for the rod-like PHIC block has been examined by light scattering.313 Planar disk-like micelles with a very large diameter (0.9 um) yet small thickness (20 nm) were observed. The micelles were large enough to image by polarized light microscopy. A 'hockey puck' structure of disk-like micelles was first anticipated theoretically by Williams and Fredrickson for a melt of rod-coil diblocks. They used strong segregation limit theory for isolated aggregates to predict that such phases would occupy most of the lamellar part of the phase diagram.314 Micelles containing the conducting polymer polyacetylene have been prepared from a P4MS-&-PVSO precursor dissolved in a selective solvent for the P4MS.315 The PVSO was converted into the conducting polymer acetylene (conductivity was not assessed in the micelles) by heating. 2.12.4
CROSS-LINKED MICELLES
Wooley and coworkers have investigated shell cross-linked nanoparticles prepared from diblock copolymer micelles (Figure 2.31), as discussed in several reviews.316"318 A number of approaches to cross-link the corona to form the socalled shell cross-linked knedel (SCK) nanoparticles have been successfully employed. Cross-linking can be achieved by direct reaction between polymer chain segments in the corona, or via the introduction of multifunctional linkers, as discussed in an early review.316 Considering first direct cross-linking, PS-&-PV4P diblocks have been used for this purpose. Approximately 40% of the P4VP units were quaternized with /?-chloromethylstyrene, which enabled polymerization of the styrenyl side groups.319'320 The micelles formed have a glassy PS core and a charged corona. Packaging of DNA using these nanoparticles has been demonstrated.321 The binding of DNA in the complexes is facilitated via electrostatic interactions with the positively charged corona. The DNA was found to be compacted as a result of these interactions.
Neutral Block Copolymers in Dilute Solution
69
Figure 2.31 Schematic showing preparation of shell cross-linked knedel nanoparticles.316 Reproduced by permission of Elsevier.
The use of condensation reactions using multifunctional cross-linkers has also been explored. Diblocks containing PAA were cross-linked using multifunctional amines. The PAA block was derived from a PtBS precursor by selective hydrolysis.322'323 The effect of the core block on the micelle properties has been investigated using amorphous polyisoprene (PI), glassy PS, crystalline PCL and hollow (water-filled) nanocage structures.316'318 SCKs containing PCL were prepared from PCL-b-PAA diblocks via diamino cross-linking of the carboxylic acid units in PAA.322 Dried nanoparticles were imaged on mica by AFM, and changes in dimension as a function of copolymer composition and extent of cross-linking were examined. The PCL melting temperature increased as a function of core volume. The frustration imposed on crystalline PCL chain packing imposed by the cross-linked shell was shown to lead to the formation of disk-shaped micelles, as imaged when adsorbed onto mica.324 A reversible transition between disk-shaped and spherical micelles on melting/crystallization of the PCL block in bulk solution was also proposed, on the basis of DLS measurements of micelle dimensions. The PCL is also susceptible to hydrolytic degradation enabling the preparation of hollow nanoparticles. Nanocage structures with pH responsiveness have been prepared from SCKs with poly(acrylic acid-co-acrylamide) shells and PMA cores, which were hydrolytic ally degraded to produce methanol, extracted by dialysis. Reversible changes in the adsorbed structure were observed upon changing solution pH by in situ AFM in a liquid cell.325 Two dimensional arrays of the SCKs were observed when adsorbed onto mica. Nanocages have also been prepared from Pl-b-PAA diblocks, followed by cross-linking of the PAA and removal of the PI by ozonolysis, and subsequent reduction of the resulting ozonides and removal of the fragments by dialysis.3 This chemistry has also been used by Liu's group as discussed below. When the micelle core is PI and the corona is PAA, the corona chains may be cross-linked using a diamino linker.320 The PI core block has a very low glass transition temperature, so the micelles are fluid. A range of nanospheres can be prepared from the same diblock by varying the composition and amount of added
70
Block Copolymers in Solution: Fundamentals and Applications
cross-linker to control the shell properties. For example, the use of different diamine and tetramine cross-linkers has been explored, leading to small changes in the diameter of the SCK nanospheres. In the case of a 2,2'-(ethylenedioxy)bis(ethylamine) cross-linker, hydrogel-like behaviour was observed as the shell swelled substantially (2- to 3-fold thickness increase) in water.327 Increasing cross-linker chain length using, for instance, diamino PEO led to even greater swelling. The effect of increasing the glass transition temperature of PI (through hydrochlorination) was examined328 - as expected these micelles were much more rigid, as determined from AFM on micelles adsorbed onto mica. Controlled release nanoparticles may be prepared from SCKs with a cleavable linkage between shell and core block.329 PS-b-PAA diblocks with a thermally labile C—ON bond were used for this purpose. The PS in the core could be released upon cleavage due to permeation through the hydrogel-like shell membrane. Sugar-coated SCKs have been prepared from PAA-b-PMA diblocks containing a terminal mannoside group (which resulted from the use of a mannoside-functionalized initiator for ATRP).330 Interaction of the functionalized nanoparticles with lectins, red blood cells and bacterial cells was examined. Related work on sugarcoated micelles is discussed in Section 6.3. In later work, the cross-linked shell in PI-&-PAA diblock nanospheres was functionalized with folic acid which is a high affinity ligand for the folate receptor.331 This receptor has been identified as a tumour marker, expressed at elevated levels on cancerous growths, relative to normal tissue.331 SCKs containing a fluorinated core have been prepared from P4FS-£-PAA diblocks in which the PAA shell was crosslinked.332 Incorporation of 19 F is useful for spin-labelled NMR studies on micellization. Wooley's group have also prepared SCK nanoparticles from ABC triblock micelles. A PAA-&-PMA-&-PS triblock formed micelles in aqueous solution.323 The PAA outer shell layers were then cross-linked. A rod-to-sphere transition was driven by addition of THF to fluidize the PS core. Intermediate structures could be accessed by subsequent cross-linking of the shell.306 Armes and coworkers have reported the preparation of shell cross-linked micelles in aqueous salt solutions of a PEO-b-PDMA-&-PMEMA triblock prepared by ATRP.333 The PMEMA block forms the core, the hydrated PDMA block the inner shell and the PEO block the outer shell. The inner shell was then cross-linked. The PEO outer shell provided steric stabilization, enabling cross-linking at relatively high copolymer concentration (10% solids). Shell cross-linked micelles with 'normal' and 'inverted' core-shell structures have been prepared from PMPCb-PGMA-b-PDEA triblocks, depending on pH/solvent conditions (cf. Figure 4.6).334 The PGMA shell was cross-linked using divinyl sulfone. The structure of the crosslinked micelles was probed by X-ray photoelectron spectroscopy (XPS). MPCcoated particles are of interest in biomedical applications and are used as implant coatings due to their biocompatibility. In related work,335 shell cross-linked micelles were prepared from PEO-6-PGMA-b-PDEA and PEO-b-PHEMA-bPDEA triblocks. These dissolved molecularly at low pH, but micellization occurred above pH 7-8 to form onion-like micelles comprising PDEA cores, PGMA (or PHEMA) inner shells and PEO outer coronas. Selective cross-linking of the PGMA
Neutral Block Copolymers in Dilute Solution
71
(or PHEMA) inner shell was achieved using divinyl sulfone. The resulting SCK micelles exhibited reversible swelling behaviour as a function of solution pH due to protonation of the PDEA cores. It was shown that the SCK micelles could be used as nanoreactors for the synthesis of gold nanoparticles, via reduction of solubilized HAuCU. Cross-linked micelles with a PPO core, cross-linked PDMA inner shell and OEGMA corona were prepared in a similar fashion, in a one-pot synthesis at high solids content.336 Shell cross-linked diblock copolymer micelles have also been prepared. Partially quaternized PDMA-&-PMEMA diblock micelles were formed with chemically cross-linked gPDMA coronas.337 Two types of shell cross-linked zwitterionic micelle (anionic or cationic corona block) were also developed.338 Liu and coworkers have prepared cross-linked micelles from PS-&-PCEMA where the PCEMA shell was radiatively cross-linked.339 The resulting nanospheres were hydrophobic. Water soluble nanospheres based on PI-b-PCEMA have also been prepared, these have potential for drug delivery applications.340 After crosslinking of the PCEMA, the PI was hydroxylated to make water-soluble vesicles. Further details are provided in Section 6.7. Later, this group prepared nanofibres by cross-linking wormlike micelles formed by PS-fr-PI diblocks in a PS-selective solvent.341 The PI core was cross-linked using S2C12 to yield fibres with a diameter of 30-70 nm and a length 0.9-3 um. This group has also explored a number of routes to the preparation of porous nanospheres, and this is discussed further in Section 6.7. 2.12.5
JANUS MICELLES
An ingenious method to produce so-called 'Janus' micelles has been developed (Figure 2.32).342 These are micelles with two different faces. Spherical Janus micelles were prepared from an ABC triblock microphase separated in the melt into a so-called 'ball at the wall' morphology, comprising spheres of the minority PB block at the interface between PS and PMMA lamellae. The PB domain was crosslinked, allowing the structure of PS and PMMA hemispheres in the corona to be retained after dissolution.342 It was shown that above a cac, so-called supermicelles are formed by this system, comprising aggregates of the cross-linked micelles (Figure 2.32). Janus micelles containing a polyelectrolyte block were prepared by alkaline hydrolysis of the PMMA to give PMAA. The size of the supermicelles was then strongly pH-dependent.343 In a similar vein, cylindrical Janus micelles in which the blocks are segregated in a plane parallel to the cylinder axis have been prepared, starting from a morphology of B cylinders at the interface of AC lamellae in an ABC triblock.344 The polybutadiene cylinders were cross-linked, followed by dissolution of the matrix. 2.12.6
NONSPHERICAL MICELLES
Micelles with an elongated shape, i.e. cylindrical, wormlike or rod-like micelles are much less commonly observed in dilute block copolymer solutions than spherical
72
Block Copolymers in Solution: Fundamentals and Applications
Figure 2.32 Schematic showing synthesis and self-assembly of Janus micelles from PSb-PB-b-PMMA triblocks.343 The initial PS-6-PB-&-PMMA micelle was prepared by dissolution of a bulk 'ball at the wall' morphology, in which the PB 'ball' at the lamellar interface 'wall' was cross-linked. Reproduced by permission of American Chemical Society.
ones. In addition, the mechanisms of interaction between these micelles are poorly understood. There are however several SANS studies of elongated micelles formed by block copolymers.106'118'345"351 Expressions for form and structure factors have been detailed in published reviews.32'37 For several Pluronics, a transition from spherical to cylindrical micelles is observed on raising the temperature. We consider here a change in micellar shape in dilute solution, and not the formation of a hexagonal packed cylindrical micelle structure at high concentration. The effect is due to an increase in association number with increasing temperature, the transition occurring when the radius of the micelle core exceeds the stretched length of the hydrophobic block (or half-block for a triblock with a hydrophobic midblock). The has been most extensively studied for Pluronic pgs 41.62,106,137,140,255,347,352,353 Qther studies are summarized by Booth and Attwood.70 A good example is the SANS study on Pluronic P85 micelles.354 Simultaneous SANS-shear flow experiments showed that cylindrical micelles are easily oriented under shear flow.354 Dilute aqueous solutions of Pluronic F88 have been studied by SANS. It was observed that the
Neutral Block Copolymers in Dilute Solution
73
transition temperature between spherical and cylindrical micelles decreased with increasing salt (potassium carbonate) concentration.349 However, structure factor effects in the SANS curves prevented the modelling of the cylindrical micelle form factor. As another example, cylindrical micelles have been observed at high temperature for EO]8-b-BOio in D2O.350 It was found that the intensity at high q scaled as g~ 1 7 , in agreement with the calculations on the form factor for semiflexible chains with excluded volume.355 This suggested that the micelles were wormlike rather than straight cylinders. This was confirmed350 by fits to the intensity (Figure 2.33) calculated using numerical interpolation formulas. Changes in the molecular weight of the polymer can also induce a spherical to cylindrical
Figure 2.33 SANS intensity profiles from solutions (1 wt %) of EO18-£-BO10 (A) and EO40-b-BOio (°) in D2O at T = 60 °C.350 The solid line is a fit of the data for EO]8-£-BOi0 to a model for the form factor of wormlike micelles, and the dashed line is a fit of the data for EO40-b-EOiQ to the form factor of a spherical micelle. Reproduced by permission of American Chemical Society.
micellar shape transition. The micellar behaviour of several PEP-&-PEO diblocks has been studied by SANS.351 The scattering of the solutions showed that a morphological transition takes place upon lowering the molecular weight. The high molecular weight block copolymers all formed spherical micelles while cylindrical micelles were observed at low molecular weights.351 A SANS study of the temperature dependence of the micellar structure has been performed on dilute solutions of a PEO-b-PPO-b-PEO block copolymer using a
74
Block Copolymers in Solution: Fundamentals and Applications
Poiseuille-geometry shear flow apparatus.347 Ellipsoidal micelles were observed at low temperatures, while a cylindrical shape was attained at higher temperatures. The influence of a flat interface on the isotropic to nematic transition was investigated using SANS for cylindrical micelles of PB-&-PEO block copolymers in aqueous solution.346 The form factor of the micelles was modelled using the product of Koyama's356 form factor of a wormlike micelle with the cross section of an infinite cylinder. It was found that the interface induced the formation of a condensed nematic ordered layer below the bulk isotropic to nematic phase transition. Wormlike micelles are formed by asymmetric PS-b-PI diblocks in heptane, as imaged by AFM and also probed by light scattering.357 SAXS has been used to study the shear flow behaviour of dilute solutions in decane of cyclic and linear PSb-Pl diblocks.118 Application of shear revealed that only the micelles of cyclic chains adopted a cylindrical shape, leading to the conclusion that cyclization of diblock copolymer chains might be a method to control micellar morphology. Ellipsoidal micelles have also been observed for asymmetric PS-/?-PI and PS-b-PIb-PS triblocks in DBP (selective for PS).358 Systematic contrast variation SANS studies on an asymmetric dPS-b-PI diblock in DBP, a slightly selective solvent for the dPS block revealed an anisotropic shape that could be modelled using the form factor of micelles with ellipsoidal or cylindrical cores surrounded by solvated Gaussian chains.35 Micelles formed by amphiphilic block copolymers of vinyl ethers containing PHOVE and partially deuterated PNBVE have been studied by SANS.360 Four block copolymers with the same hydrophilic block length but different hydrophobic chain lengths were prepared and the micellar structures formed by these copolymers in aqueous solution were investigated. The polymer with the shortest hydrophobic chain was suggested to form spherical micelles, whereas the scattering curves of the polymers containing longer hydrophobic chains reflected the formation of rod-like micelles. The volume fraction of the rod-like micelle was found to increase with increasing hydrophobic chain length. Addition of a cationic surfactant has been shown to drive a shape transition even for micelles formed by neutral block copolymers (for a discussion of the interaction between surfactants and charged block copolymers see Section 2.15).361 A transition from predominantly cylindrical micelles to smaller spherical mixed micelles was observed via SANS and light scattering on addition of the cationic surfactant CTAB to an aqueous solution of a PB-b-PEO diblock. These results were also supported by cryo-TEM. This behaviour is in contrast to that usually observed, whereby 'micelles' composed of a single polymer chain decorated with surfactant molecules are observed. 2.12.7
MICELLES FORMED DUE TO SPECIFIC INTERACTIONS
Micellization can be driven by specific interactions, for example hydrogen bonding. Yoshida and Kunugi showed that addition of a solvent that promoted hydrogen
Neutral Block Copolymers in Dilute Solution
75
bonding to a nonselective solvent, led to micellization in a PVPh-£-PS diblock.362 This occurred due to hydrogen bonding of the added solvent (1,4-butanediamine) with the PVPh block which caused this block to become solvophobic. Other types of noncovalent interactions have also been exploited to prepare diblocks which undergo micellization, for example metal coordination chemistry, as discussed further in Section 6.6.
2.13
MICELLIZATION IN MIXED SOLVENTS
By varying the composition of a mixed solvent, the selectivity for a particular block can be controlled and it is possible to invert the composition of micelles in this way.363 The formation of micelles of PS-b-PEO diblocks in toluene/propan-2-ol mixtures has been investigated by DLS.364 Added water was solubilized in the PEO corona. Kurata and coworkers investigated the micellization of a PS-b-PMMA diblock in a mixture of PMMA-selective solvents.365 Alexandris et al. have investigated association properties of numerous Pluronics in binary water/oil solvent mixtures as listed in Table 3.1. Here the oil is solubilized in the mesophase. A few studies have focused on the effect of cosolvents in dilute solution as discussed in Section 2.6.3.
2.14
MIXED MICELLES
Chu and coworkers have investigated the micellization of Pluronic F127 (EO99-£>PO69-£-EO99) with the related triblock EO45-£-BO14-b-EO45 in which the central block is more hydrophobic.366 At low temperature, the latter copolymer has a much lower cmc and addition of Pluronic F127 leads to incorporation of the chains into mixed micelles. The temperature dependence of the cmc of Pluronic F127 is greater than that of EO45-&-BOi3-£-EO45 and at high temperature Pluronic F127 has a lower cmc. Added EO45-^-BO13-^-EO45 is therefore incorporated into the preexisting F127 micelles. However, due to polydispersity there is a bimodal distribution of average micelle sizes. At an intermediate temperature where both copolymers have the same cmc, a single distribution of mixed micelles containing equal proportions of each copolymer is formed. A change in cubic micellar gel structure was also observed upon varying the composition of mixed micelles in this system.367 Mixed micelles have also been observed in aqueous solution of diblock EO60-fc-BOi2 and triblock EO55-£-BO20-&-EO55.368 Since the EO block length is essentially the same, the dimensions of the mixed micelles were unaffected by mixture composition. The association number was found to be proportional to composition, when allowance was made for coronal chain looping of the triblock. The size distribution of mixed micelles was found to be narrower than that of the individual copolymer micelles, leading to a harder interaction potential.
76
Block Copolymers in Solution: Fundamentals and Applications
The micellization of binary mixtures of PaMS-b-PVPEA diblocks in a selective solvent has been studied via static and dynamic light scattering following a T-jump.225 Two diblocks having approximately the same molar mass of soluble PVPEA block but differing in the molar mass of the insoluble PaMS block were studied. Co-micellization was observed in the region where the diblock with longer hydrophobes formed micelles but where the shorter chain diblock was unassociated. Some of the shorter chains were incorporated into the micelles, but the extent of mixing was limited. In contrast, following a direct jump to a regime where both diblocks could form micelles in isolation, co-micellization was observed with nearly complete mixing of chains. In a double step T-jump, some of the shorter chains mix with the micelles formed by the longer copolymers in the first step, whilst the excess associate into 'pure' micelles. A model for micellization in dilute solutions of mixtures of block copolymers with a common lengthy corona block but different solvophobic block lengths predicts regimes of pure micelle or mixed micelle formation.369 The model is based on Flory lattice theory together with expressions for the free energy of polymer brushes. The block copolymers with longer hydrophobes associate first into micelles, as expected. These micelles are enriched with the shorter chain diblocks at concentrations below the cmc for the short chains alone. These findings are in 99S agreement with the results of Honda et al. Micellization in a similar system of a pair of diblocks with different asymmetries was analysed on the basis of a generalized model for the thermodynamics of micellization combined with a model for the osmotic and interfacial tension contributions from the copolymer micelle.370 It was found that if the asymmetry ratio is not too large, the short copolymers can be incorporated into mixed micelles. If the asymmetry is larger, mixed micelles coexist with micelles containing only the longer diblock. These results were compared with those of Hecht et al. who studied mixtures of Pluronic triblocks and SDS, as discussed in the following section. Onion-type micelles (Figure 2.27) have been prepared by mixing a PS-b-PVP diblock in acidic solution with a basic solution of a PVP-b-PEO diblock.299
2.15
BLOCK COPOLYMER/SURFACTANT COMPLEXES
The interaction between polymers and surfactants is described by two critical concentrations. The first is the critical aggregation concentration (cac, sometimes denoted CO at which point the binding of surfactant and polymer first occurs. The cac is generally lower than the cmc of the surfactant alone. The second critical concentration (often denoted €2) is associated with the saturation of polymer with surfactant aggregates (Figure 2.34). The cmc of the surfactant (Cm) may also be observed. For some polymer/surfactant systems, Cm and C2 are coincident.371'372 However, in other cases, Cm is less than C2.371'372 The interaction between Pluronic triblocks and the anionic surfactant SDS has been exhaustively investigated.373"380 The interaction between SDS and Pluronic
Neutral Block Copolymers in Dilute Solution
77
Figure 2.34 Critical aggregation concentration €2 resulting from saturated binding of a surfactant to a block copolymer.
F127 has been probed in particular detail, via electromotive force (EMF) measurements with a SDS surfactant selective electrode, isothermal titration calorimetry, light scattering and SANS. The following sequence of events is reported upon addition of SDS to Pluronic F127 at a concentration well above the copolymer cmc:377'379 First, the SDS binds to F127 micelles to form mixed micelles. Further addition of SDS leads to the breakdown of these micelles into smaller ones. At a sufficiently high SDS concentration, SDS micelles can bind directly onto unassociated F127 chains, forming a so-called 'necklace of micelles' (as anticipated theoretically). The SDS binds to the hydrophobic PPO blocks, reducing hydrophobicity and therefore enabling chains to remain unassociated in solution. These coexist with the mixed micelles. Further increase in concentration leads to the binding of more and larger SDS micelles to the F127 chains, until a saturation concentration (~0.1 M SDS) is reached. At saturation conditions, Hecht et al. report binding of 4-6 SDS molecules per polymer chain,375'376 whereas Li et al. report binding of one SDS micelle per chain,377 but this was based on the questionable assumption that the SDS association number (-80) is not affected by binding to the polymer. Figure 2.35 shows the micellization enthalpy determined from DSC as a function of [SDS] for three concentrations of F127. The point at which A//mic starts to decrease signals the binding of SDS onto the polymer, which is completed when A//mic reaches zero.376 Li et al. report that SDS can cause micellization of F127 even below the cmt of the neat polymer.378 This may possibly be due to the disruption of the hydrogen bond network of water by SDS or due to the action of SDS as 'nuclei' for the formation of F127 micelles close to the cmt. Further addition of SDS leads to the breakdown of these micelles as SDS binds to F127 unimers. Almgren et al. studied the micellization of Pluronic L64 and F68 with SDS in the dilute regime using 13C NMR and fluorescence quenching techniques.381 They found that the copolymers form mixed micelles with SDS at concentrations well below the cmc of SDS and that SDS reduces the size of Pluronic micelles. The influence of SDS on the association of Pluronic F127 was investigated by Hecht and coworkers375'376'382 using a variety of experimental methods. They found that SDS binds strongly to F127 and can suppress micelle formation by F127 completely. Addtion of SDS to other Pluronics has also been investigated. The increase in cloud point due to the formation of mixed aggregates with polyelectrolyte character
78
Block Copolymers in Solution: Fundamentals and Applications
Figure 2.35 Micellization enthalpy of F127 from DSC as a function of added SDS concentration at the three SDS concentrations indicated.376 Reproduced by permission of American Chemical Society.
has been reported for L64.383 Fluorescence quenching measurements indicated the formation of mixed micelles upon addition of SDS to L64 and F68.381 Chemical shift measurements (13C) of the methyl carbons in PPO suggested a change in conformation from extended to compact coil upon addition of SDS, going from large micelles to smaller ones (and ultimately unassociated polymer chains).381 The binding of SDS to reverse Pluronics has been studied by isothermal titration calorimetry. This revealed that beyond the cac, complexes are formed in which SDS first binds to the PPO block and then to the PEG block.374 The polymer is incorporated into SDS micelles, which have a lower aggregation number than those of the pure surfactant in water due to dehydration. Increasing length of the PPO block leads to a reduction in the cac. The formation of mixed micelles, micellar clusters and supermicellar aggregates has been observed in solutions of a PS-b-PEO diblock with anionic (SDS) or cationic (CPC1) surfactant.384 The evidence for incorporation of the surfactant into the block copolymer micelles is the increase in micellar size up to a certain concentration. As shown by proton NMR, addition of surfactant leads to a significant increase in mobility in the PS core. The interaction of Pluronic F127 with the nonionic surfactant C^EOs has also been investigated.385 Mixed micelle formation was observed. The synergistic mixing of the two surfactants led to nonideal mixing. The dependence of cmc on mixture composition could however be described using regular solution theory with a nonideal mixing model.
Neutral Block Copolymers in Dilute Solution
2.16
79
COMPLEX MORPHOLOGIES
Numerous complex morphologies have been observed for block copolymers in dilute solution, often these are not in equilibrium as a result of the preparation method. This is especially a problem for micelles containing a glassy core (e.g. PS or PMMA) as discussed below. A variety of morphologies including tubules, vesicles, branched vesicles and large compound vesicles have been observed for PS-b-PAA386 and PS-&-PEO387 diblocks in DMF/water mixtures. Further details are provided in Section 4.1.2. The aggregation of PS-b-PAA diblock micelles into cubic particles (edge length 200600 nm) has been observed (Figure 2.36) upon evaporation of the aqueous/organic solvent mixture.388 The micelles were formed in aqueous solution, to which one of several organic solvents was added. The effect is not fully understood. Hydrogen bonding of the PAA coronas may play a role, and the addition of solvent leads to a ternary system. When the organic solvent evaporates, the phase diagram may pass through a cubic micellar phase.
Figure 2.36 SEM image of cubic microparticles formed after solvent evaporation from a PS-&-PAA micellar solution in water/butanone (97%/3%).388 Reproduced by permission of American Chemical Society.
Compound micelles and onion particles have also been observed for PS-bPCEMA diblocks in solvents slightly selective for the PCEMA.389 Similar to the method of preparation used by Eisenberg et al. the polymers were first dissolved in a nonselective solvent, then a precipitant for PS was added. The structures formed may be nonequilibrium morphologies trapped by PS vitrification. In addition, the PCEMA shells were cross-linked by ultraviolet (UV) radiation. Giant wormlike micelles formed in dilute aqueous solution by a low molecular weight PB-&-PEO block copolymer have been observed by TEM and the structure studied in detail by SANS.345'390 These wormlike rubber micelles can be considered to be giant rubbery (due to the PB block) macromolecules. These elongated micelles successively pack into a nematic and then a columnar phase as the polymer concentration is increased.345
80
Block Copolymers in Solution: Fundamentals and Applications
Several novel morphologies have been observed in dilute aqueous solutions of related PB-b-PEO diblocks.245'391 As noted above, these structures are generally in a nonergodic state, although the local topology is expected to be equilibrated. Branched wormlike micelles containing Y-shaped junctions and a three-dimensional network structure were reported for diblocks with PEO content intermediate between that for diblocks forming bilayer vesicle and cylindrical micelle structures, provided that the PB block is long enough.245'391 Eventually the polymer-rich network phase separates from the solution. Through systematic investigation of the morphology for a range of diblocks, a phase diagram was assembled, as illustrated in Figure 2.37.245'391 Representative examples of the morphologies are also
Figure 2.37 Morphology diagram for PEO-b-PE diblocks in water at a concentration of 1 wt% in terms of the weight fraction of PEO (WPEO) and the PB chain length (WPB).391 The diagram is assembled on the basis of cryo-TEM observations of morphology for two series of diblocks (one with WPB = 46, the other with NPQ = 170) Four basic structural motifs are sketched: bilayer (B), Y-junction branched cylinders (Y), cylindrical micelles (C) and spherical micelles (S). Representative cryo-TEM images showing these structures are also included. A network structure (N) is observed for diblocks with a composition intermediate between that for which bilayer and cylinder morphologies are observed, provided than WPB is sufficiently high. A representative image of the network structure is shown in Figure 2.38. The scale bars in the TEM images indicate 100 nm. Reproduced by permission of Science.
Neutral Block Copolymers in Dilute Solution
81
Figure 2.38 Cryo-TEM image showing a network structure in a 1 wt% aqueous solution of a PEO-b-PB diblock with WPEO = 0.34 and NPB = 170.391 The scale bar indicates 200 nm. Reproduced by permission of Science. included. The network structure is shown in Figure 2.38. The network structure can be broken up by stirring or sonication to produce fragments composed of Y-junctions, spherical caps and cylindrical loops. A structure comprising undulating cylindrical micelles (Figure 2.39) was also reported, close to the boundary between
Figure 2.39 Cryo-TEM images from a solution of a binary mixture of PEO-b-PB diblocks with a composition WPEO — 0.42, close to the boundary between cylinder and sphere structures (NPB = 108.5 for the blend).245 Undulations in the cylindrical micelles, which have pronounced bead-like endcaps can be noted. In (a) short cylinders with an undulation (short arrow) and two undulations (long arrow) are indicated. In (b) and (c) the number of undulations in the cylinder branches is quantized according to the distance between branch and end junction. Reproduced by permission of American Chemical Society.
82
Block Copolymers in Solution: Fundamentals and Applications
sphere and cylinder morphologies for a pre-mixed dispersion of two diblocks.245 Although cylindrical micelles are generally capped with a spherical bulbous end, the formation of cylindrical micelles with periodic thickness variations away from the end-cap appears to be associated with the polydispersity in the binary diblock mixture used and the high molecular weight of the copolymers. It was interpreted as a Rayleigh instability, i.e. an undulational mode propagating from the end cap. Usually the undulations are rapidly damped, but the local variation of curvature in the binary diblock system enables 'local equilibration' in this fascinating structure. An additional octopus structure (Figure 2.40) was noted when using two premixed diblocks to achieve the same overall PEO content as that for the network structure (Figure 2.38).245 Flat octupuses with a variable number of arms were observed, and also hemispherical ones (Figure 2.40). The flat octopus corresponds to a bilayer from which cylindrical arms emanate. The formation of this morphology was ascribed to the distinct interfacial curvature tendencies of the two diblocks in the mixture. Compartmentalized vesicles or micelles are of interest due to the analogies with the portioning of the cell into compartments by lipid membranes, as mentioned in
Figure 2.40 Cryo-TEM showing 'octopus' structures comprised of bilayer cores from which cylindrical arms emanate.245 Flat octopuses and hemispherical octopuses can be noted. This morphology is observed for a solution of two PEO-b-PB diblocks at a composition (WPEO — 0.34) for which the network structure (Figure 2.38) is observed using a single diblock. The competing interfacial curvatures of the two diblocks drive the formation of this complex morphology. Reproduced by permission of American Chemical Society.
Neutral Block Copolymers in Dilute Solution
83
Section 2.12.2. The formation of aggregate structures containing internal arrays of tubes was observed for a PS-£-PAA diblock (S410-^-AA13), initially dissolved in DMF and then dialysed against water.392 Since different projections of the structure were imaged by TEM, a hexagonally packed hollow hoop morphology could be proposed. Giant compound vesicle structures have also been reported for PB-b-PEO diblocks in aqueous solution.393 Polymersomes comprising a bilayer containing a lattice of passages or a network of tubules were observed. The formation of such high genus objects was analysed in terms of the elastic properties of copolymer membranes. Examples of optical micrographs and calculated membrane surfaces are shown in Figure 2.41. Toroidal structures can self-assemble through the collapse of negatively charged cylindrical micelles, driven by interaction with a divalent organic cation (Figure 2.42).432 The micelles were formed by PAA-b-PMA-b-PS triblocks in THF/water mixtures with the divalent cation 2,2'-(ethylenedioxy)diethylamine. The divalent nature of the cation was shown to be essential to this process and analogies were made with the formation of toroidal structures in DNA due to condensation in the presence of multivalent ions. It was also shown to be necessary to control the ratio of divalent ion to acid and to prepare the aggregates from a mixed solvent (THF/water). THF was needed to ensure initial dissolution of the hydrophobic PS core. A range of other intermediate structures, with trifunctional branch points as for the aggregates studied by Jain and Bates, was noted, as illustrated in Figure 2.43. However, interconnected network structures were not observed, in contrast to the observations of Jain and Bates. The formation of figure eight and other complex structures (Figure 2.42) indicates that self-assembly is not simply due to end-to-end association of cylindrical micelles, but that more complex fusion events must occur. In a solution with a low volume ratio of water to THF, compound micelles comprising close-packed arrays of bent cylinders were observed, very similar to those reported by Haluska et al. (Figure 2.41). Giant block copolymer amphiphiles have been prepared by linking an enzyme headgroup to a 40 repeat unit PS chain.394 The 33 kDa enzyme lipase B was used. The aggregation of the copolymer amphiphiles formed in situ at the air-water interface was studied via Langmuir isotherm measurements. Hollow fibrils were observed to self-assemble when the coupling reaction was performed in bulk aqueous solution, aggregating into bundles with a length extending to the urn range.
2.17
VESICLES
Vesicles are spherical shell structures comprising a bilayer of amphiphiles, which may be low molar mass (surfactants, lipids) or may be block copolymers. If there is a slight mismatch in the effective interfacial area per hydrophile compared with hydrophobe, the bilayer membrane can spontaneously close into a vesicle rather than forming a planar lamellar structure. This can be translated into a condition on the so-called surfactant packing parameter which is defined asp = V/al where Vis
84
Block Copolymers in Solution: Fundamentals and Applications
Figure 2.41 Optical micrographs for compound vesicles formed by PE-b-PEO diblocks in aqueous solution (a, c, e) and calculated membrane surfaces (b, d, f, g).393 The scale bars indicate 10 jam. The polymersomes contain (a) small passages, (b) large passages, (e) budded vertices. Reproduced by permission of American Physical Society.
the volume per molecule, a is the effective cross-sectional area per molecule and / is the chain length normal to the interface. Vesicles are observed for l/2< p < 1.103 This can also be related to the development of finite mean curvature.395 A wide variety of chemical approaches have been developed to produce polymeric vesicles using block copolymers, and conditions to1 prepare them have been reviewed.396 Often the vesicle shell is cross-linked to create a hollow sphere.
Neutral Block Copolymers in Dilute Solution
85
Figure 2.42 Toroidal micelles formed by a PAA-b-PMA-b-PS triblock in a water/THF mixed solvent with EDDA divalent cations.432 (a) TEM image of cast film, negatively stained with uranyl acetate, (b) Schematic of toroidal structure showing hydrophobic PS (centre), and PMA (inner shell) with a corona of hydrophilic PAA with closely associated EDDA. Reproduced by permission of Science.
Figure 2.43 TEM images showing examples of intermediate structures formed by quick casting films from THF/water + EDDA solutions of a PAA-b-PMA-b-PS triblock.432 The scale bars indicate 100 nm. (a, b, g) Dumb-bells; (e, f, g, i) interior closed rings; (d, f, h) lariats; (c, j) figure eights; (g) and (i) cylinders with ends connected but not fused toegther. Reproduced by permission of Science.
86
Block Copolymers in Solution: Fundamentals and Applications
Vesicle formation by block copolymers has been the subject of a dedicated review.397 Vesicles formed by low molar mass surfactants and amphiphiles are generally kinetically trapped, nonequilibrium structures.103 In contrast, it has been proposed that thermodynamically stable vesicles can be formed by diblock copolymers due to their intrinsic polydispersity.398'399 The polydispersity leads to selective segregation of short hydrophilic blocks to the inside of vesicles, whereas longer hydrophilic blocks segregate to the outside. The preferred curvature of the bilayer is stabilized in this way. The effect is enhanced for smaller vesicles, as expected since the tendency to segregate will be greater as interfacial curvature is increased.399 Fluorescence probe experiments on PS-&-PAA diblocks with varying hydrophilic block length, labelled with pyrene at the junction point, confirmed this hypothesis.398'399 Earlier data from this group had shown that vesicle size could be changed reversibly by varying solvent composition, pointing towards thermodynamic stability.400'401 Figure 2.44 shows TEM data illustrating reversibility of the vesicle formation and growth process for S3oo-^-AA44 in aqueous solution.399 Bilayer fluidity is also required to reach equilibrium, and this can be achieved using suitable fluidizing solvents which enable the flip-flop exchange of copolymer chains.397
Figure 2.44 TEM images showing reversibility of vesicle formation and growth process upon changing water content for diblock S3,00-b-AA44 in a mixture of water and THF/ dioxane.399 Reproduced by permission of American Chemical Society.
Unilamellar vesicles have been observed for dilute Pluronic L121 in the twophase region between lamellar and micellar liquid phases.402 The high hydrophobe content in this copolymer is believed to favour the formation of vesicles, which were imaged by cryo-TEM. Later, Richtering et al. observed the shear-induced formation of multilamellar vesicles ('onions') in the dilute lamellar phase formed in
Neutral Block Copolymers in Dilute Solution
87
binary solvent (butanol/water) solutions of Pluronic P123 and F127.403'404 The presence of the cosolvent increased the interfacial area per block copolymer molecule and facilitated the mobility of molecules between different layers, both contributing towards an increased tendency for curvature of dilute lamellae into vesicles. Comparison was made to the shear-induced reorientation of a lamellar phase in the same system at higher polymer concentration.403 In this case a shear induced transition from parallel to perpendicular lamellae (Figure 3.24) was observed via SANS, the viscosity decreasing steadily with increasing shear rate. In the more dilute system, increasing shear rate led to an increase in viscosity up to 7 ~ 1, followed by shear thinning at higher shear rates. The increase in viscosity signalled the formation of vesicles, as confirmed by SANS and SALS. The spontaneous formation of multilamellar vesicles in aqueous solutions of PDMS-&-PEO diblocks has been ascribed to the high hydrophobicity of the PDMS block.405 This leads to the formation of bilayers at very low concentration which aggregate into a turbid solution of multilamellar vesicles at higher concentration. The spontaneous formation of multilamellar vesicles has also been observed for low molar mass PEO-b-PBO diblocks in aqueous solution. Their formation was ascribed to the molecular shape asymmetry arising from the large volume to length ratio of the hydrophobic PBO block.406 Wang et a/.407 have prepared vesicles from 'diblocks' in which P4VP is hydrogen bonded to hydroxyl-containing PS random copolymer, denoted PS (OH). The P4VP shell was cross-linked and the PS(OH) polymer in the core was dissolved. Phase-separated nanostructures within block copolymer microspheres have been observed to form via phase separation in a binary blend.408 The droplets were prepared as an oil-in-water emulsion using PCEMA-b-PGMA as surfactant. The P?BA-£-PCEMA within the droplets microphase separated into cylinders or concentric lamellae. The bulk morphology was PrBA cylinders. The PCEMA block could be cross-linked and the PtBA could be hydrolysed to yield internal hydrophilic PAA channels or lamellae. A series of related nanostructures has been predicted on the basis of computer simulation using a self-consistent field approach upon quenching a homogeneous droplet of diblock copolymer in a solvent bath.409 Following the quench, the droplet can absorb or release solvent, depending on the morphology. A series of morphologies is predicted for copolymers with different composition, as shown in Figure 2.45. Particularly interesting is the porous structure formed for a composition where a gyroid phase is stable in bulk. Polymer vesicles or polymersomes based on PEO-b-PEE in aqueous solution (Figure 2.46) show greater toughness than conventional phospholipid vesicles, although the bending and area expansion moduli are comparable.410 The enhanced toughness (defined as the integral of the tension with respect to areal strain) and reduced permeability of the polymersomes could lead to applications in encapsulation. The elastic behaviour was examined via micropipette aspiration of vesicles generated by electroformation (Figure 2.47). In this process, giant vesicles (2050 um in radius) are prepared from a thin film of polymer on adjacent electrodes subjected to an alternating current. Giant vesicles were also prepared by micropipette
88
Block Copolymers in Solution: Fundamentals and Applications
Figure 2.45 Predicted morphologies for nanodroplets containing block copolymers, quenched into a solvent bath.409 The block length ratio is (a) 0.35, (b) 0.30, (c) 0.25, (d) 0.20, (e) 0.15 and (f) 0.10. Reproduced by permission of American Chemical Society.
Figure 2.46 Cryo-TEM image of PEO-b-PEE diblock copolymer vesicles in water.410 The scale bar indicates 20 nm. Reproduced by permission of Science.
Neutral Block Copolymers in Dilute Solution
89
Figure 2.47 Aspiration of a polymer vesicle into a micropipette. The arrow marks the tip of a projection of a vesicle being sucked in by a negative pressure AP.410 Reproduced by permission of Science.
aspiration of PB-b-PEO diblocks.411 The interfacial viscosity and elasticity have been measured.411 The surface shear viscosity is about 500 times larger than that found for common phospholipid vesicles. On the other hand, the bending and stretching elastic constants are similar to those for lipid membranes. By pulling out a tether from the vesicle and monitoring its relaxation, it was possible to study the viscous coupling between the two monolayers comprising the polymeric membrane. The corresponding friction coefficient was about an order of magnitude larger than that found for typical fluid phospholipid membranes. The bending rigidity constant, Kc, was measured via single and dual pipette aspiration techniques for PEO-b-PB diblocks.412 For a diblock with a lengthy hydrophobic block (B125), Kc = (466 ± 157) kBT was more than an order of magnitude larger than values for typical lipids or a PEO-b-PE diblock with a short B46 block, where values range from 13-25 k B T. A quadratic scaling of kc with hydrophobic layer thickness, d, was reported, in agreement with the theoretical expression413,414 kc = ßKAd2. Here KA is the area elastic modulus and ß is a constant, for which values have been calculated depending on whether the bilayer contains coupled or interdigitated monolayers.412 An adaptation of the micropipette aspiration technique involes simultaneous application of voltage pulses to tensioned membranes, with the aim to rupture the membranes in a process termed electroporation.415 Polymersome membranes are able to withstand voltage pulses much higher than those at which lipid membranes rupture. Increasing the mechanical tension reduces the rupture voltage in a parabolic fashion. This can be understood using existing models for interfacial thermodynamics.415 The post-poration dynamics of high molecular weight diblock membranes is however significantly retarded.416 The formation of pores induced by puncturing giant vesicles with a sharp tip,417 or the extrusion of tubes from giant vesicles has been studied for giant unilamellar phospholipid vesicles,418 and extension of these techniques to examine the deformation behaviour and hydrodynamics of block copolymer vesicles should follow.
90
Block Copolymers in Solution: Fundamentals and Applications
The diffusion of probe molecules within PEO-b-PEE and PEO-b-PB polymersome membranes has been studied by FRAP (fluorescence recovery after photobleaching).419 Diffusivity decreases with polymer molar mass, as the membrane rigidity increases. The decrease is particularly marked when the chains become entangled. A transition from Rouse dynamics to reptation dynamics is observed as polymer molar mass increases sufficiently for chains to become entangled. In the Rouse regime the total hydrodynamic friction on a chain is just the accumulated friction £ on each of the N subsegments, DRouse = kBT/NC. The friction factors obtained are consistent with those expected based on composition, and extrapolate to those obtained for the polymer melt. The formation of vesicles by peptide block copolymers420 is considered further in Section 4.8.
2.18
CRYSTALLIZATION IN MICELLES
Crystallization in solution depends sensitively on solvent selectivity. If the solvent is selective for the crystalline block, it can swell the crystalline lamellae (Tm is obviously also reduced). In contrast, if the solvent is selective for the noncrystalline block, the copolymer can precipitate out of solution in a nonequilibrium structure. The crystallization of PEO in a PEO-b-PS-b-PEO triblock, a PEO-b-PPO-b-PEO triblock and a (PPO-&-PEO)4 four-arm starblock in preferential solvents was investigated by Skoulios et a/.421 In dry copolymers and in a poor solvent for PEO they observed crystallization of the PEO blocks. The solvent was found to be located in the PEO layers in aqueous solution, whereas in selective solvents for PS and PPO it was located in the corresponding block structures.421 The degree of crystallinity and chain folding in PS-b-PEO diblocks has been studied as a function of concentration of diethyl phthalate, which is a selective solvent for PS.422 Crystallization of PS-b-PEO diblocks in a selective solvent for the PEO block has been investigated for PS-b-PEO423 and PB-b-PEO.424 For both systems, a lamellar crystal structure is found below about 45 °C for solvent concentrations ranging from zero to a value characteristic of the copolymer. In these materials, PEO crystallizes in two layers separated by solvent but as the solvent concentration increases, the solvent layer gets thicker, separating the PEO layers but without dissolving them. Above a critical PEO layer thickness (50 A), increasing solvent concentration leads to discontinuous decrements in the PEO layer thickness due to step increases in the number of chain folds, whilst the degree of crystallinity decreases. Semicrystalline diblocks in dilute solutions of a solvent selective for the noncrystalline block can form platelet or 'hockey puck' 314 structures. These consist of crystalline chains folded within lamellae between solvated domains of the amorphous block. This constitutes a model system of tethered chains at a flat interface.425 Self-consistent field theory was used to model the density profile of the tethered chains and SANS and SAXS were performed to provide volume fraction
Neutral Block Copolymers in Dilute Solution
91
profiles and crystal domain thicknesses, which were compared with the predictions of theory.425 The core thickness is due to a balance of an entropic contribution from brush stretching and an enthalpic term from crystalline chain folding (and defects due to ethyl branches). Measurements were performed on solutions of PS-b-PEO in cyclopentane or PE-b-PEP in decane (a selective solvent for PEP). Crystalline-amorphous polyolefin diblocks in solution are studied due to their commercial use as 'pour point' depressants in fuels, as viscosity modifiers in lubricating oils and as wax crystal modifiers in middle distallate fuels.426 The aggregate structure of PE-b-PEP diblocks in decane has been investigated using SANS.427 In addition to lamellar plates a superstructure was identified, specifically macroaggregates of lamellae, resulting from van der Waals interactions between lamellar sheets.427 These macroaggregates are needle-like and can be seen by phase contrast optical microscopy. Subsequent work explored the effect of copolymer architecture, via experiments on PEnPEPm mixed arm starblocks.121 With increasing PEP molecular weight, the extension of the PEP chains in the corona and the reduction in core thickness differed from that expected for a diblock. It was also shown that the platelets can be modelled as disks of diameter ~1 mm (and PE core thickness ~4-10 nm). SANS has revealed that syndiotatic polypropylene-b-PEP diblocks in dilute solution in decane form structures on multiple length scales.426 The lamellae resulting from sPP crystallization at sufficiently low temperature form platelets, which associate into rod-shaped aggregates. However, in contrast to structures observed for PE-b-PEP diblocks the rods are not formed from multiply stacked platelets, but probably contain just bilayers. The corresponding length scale is hundreds of Angstroms. The rods associate into bundles that aggregate into ramified structures at a micron length scale. This additional level of order was revealed by optical microscopy and ultra-SANS, which enables structures up to 1 urn to be resolved. The formation of the aggregate structures, which were not observed in decane solutions of sPP homopolymer, was ascribed to the presence of PE sequences in the PEP block that could form small crystalline aggregates. Disk-like micelles formed by crystallization of alkyl chains have also been observed from SAXS/SANS experiments on a PME-b-PHOVE oligomer in water, a selective solvent for hydrophilic ether block.428
REFERENCES 1. Goldraich, M.; Talmon, Y. Direct-imaging cryo-Transmission Electron Microscopy in the Study of Colloids and Polymer Solutions. In Amphiphilic Block Copolymers. Selfassembly and Applications; Alexandridis, P., Lindman, B., Eds. Elsevier: Amsterdam, 2000; pp 253. 2. Danino, D.; Bernheim-Groswasser, A.; Talmon, Y. Coll. Surf. A 2001, 183-185, 113. 3. Brydson, R.; Hammond, C. Generic Methodologies for Nanotechnology: Characterisation. In Nanoscale Science and Technology; Kelsall, R. W., Geoghegan, M., Hamley, I. W., Eds. John Wiley & Sons, Ltd: Chichester, 2005. 4. Bershtein, A. V.; Egorov, V. M. Differential Scanning Calorimetry of Polymers. Ellis Horwood: London, 1994.
92
Block Copolymers in Solution: Fundamentals and Applications
Provencher, S. W. Comput.Phys.Commun. 1982, 27, 229. Berne, B. J.; Pecora, R. Dynamic Light Scattering. Wiley-Interscience: New York, 1976. Vink, H. J Chem. Soc., Faraday Trans. I 1985, 81, 1725. Hamley, I. W. The Physics of Block Copolymers. Oxford University Press: Oxford, 1998. Azzam, R. M. A.; Bashara, N. M. Ellipsometry and Polarized Light. North-Holland: Amsterdam, 1989. 10. Knoll, W. Annu. Rev. Phys. Chem. 1998, 49, 569. 11. Tassin, J. F; Siemens, R. L.; Tang, W. T.; Hadziioannou, G.; Swalen, J. D.; Smith, B. A. J. Phys. Chem. 1989, 93, 2106. 12. Brandani, P.; Stroeve, P. Macromolecules 2003, 36, 9492. 13. Brandani, P.; Stroeve, P. Macromolecules 2003, 36, 9502. 14. Zana, R. Fluorescence Studies of Amphiphilic Block Copolymers in Solution. In Amphiphilic Block Copolymers. Self-assembly and Applications', Alexandridis, P., Lindman, B., Eds. Elsevier: Amsterdam, 2000; p 221. 15. Wanka, G.; Hoffmann, H.; Ulbricht, W. Macromolecules 1994, 27, 4145. 16. Malmsten, M.; Lindman, B. Macromolecules 1992, 25, 5440. 17. Hakansson, B.; Hansson, P.; Regev, O.; Soderman, O. Langmuir 1998, 14, 5730. 18. Wang, R.; Knoll, H.; Rittig, E; Karger, J. Langmuir 2001, 77, 7464. 19. Konak, C; Fleischer, G.; Tuzar, Z.; Bansil, R. J. Polym. Sci. B: Polym. Phys. 2000, 38, 1312. 20. Yu, G.-E.; Zhou, Z.-K.; Attwood, D.; Price, C.; Booth, C.; Griffiths, P. C.; Stilbs, P. J. Chem. Soc., Faraday Trans. 1996, 92, 5021. 21. Macosko, C. W. Rheology. Principles, Measurements and Applications. VCH: New York, 1994. 22. Poole, C. P.; Owens, F. J. Introduction to Nanotechnology. John Wiley & Sons, Ltd: New York, 2003. 23. Kelsall, R. W.; Geoghegan, M.; Hamley, I. W. (Eds) Nanoscale Science and Technology. John Wiley & Sons, Ltd: Chichester, 2005. 24. Hamley, I. W. Curr. Opin. Colloid Interface Sci. 2000, 5, 342. 25. Hamley, I. W. J. Phys., Condens. Matter 2001, 13, R643. 26. Balta-Calleja, F. J.; Vonk, C. G. X-ray Scattering of Synthetic Polymers. Elsevier: London, 1989. 27. Higgins, J. S.; Benoit, H. C. Polymers and Neutron Scattering. Oxford University Press: Oxford, 1994. 28. Guinier, A. X-ray Diffraction. W. H. Freeman: San Francisco, 1963. 29. Kratky, O.; Pilz, I.; Schmitz, P. J. J. Colloid Interface Sci. 1966, 21, 24. 30. Guinier, A.; Fournet, G. Small Angle Scattering of X-rays. John Wiley & Sons, Ltd: New York, 1955. 31. Glatter, O.; Kratky, O. (Eds) Small Angle X-ray Scattering. Academic: London, 1982. 32. Pedersen, J. S. Adv. Colloid Interface Sci. 1997, 70, 171. 33. Mortensen, K. Small-angle Scattering Studies of Block Copolymer Micelles, Micellar Mesophases and Networks. In Amphiphilic Block Copolymers. Self-assembly and Applications; Alexandridis, P., Lindman, B., Eds. Elsevier: Amsterdam, 2000; p 191. 34. Pedersen, J. S.; Svaneborg, C. Curr. Opin. Colloid Interface Sci. 2002, 7, 158. 35. Castelletto, V.; Hamley, I. W. Curr. Opin. Colloid Interface Sci. 2002, 7, 167. 36. Mortensen, K. Polym. Adv. Technol. 2001, 12, 2. 37. Hamley, I. W.; Castelletto, V. Prog. Polym. Sci. 2004, 29, 909. 5. 6. 7. 8. 9.
Neutral Block Copolymers in Dilute Solution 38. 39. 40. 41. 42. 43. 44. 45.
46.
47. 48. 49. 50. 51. 52. 53. 54. 55. 56. 57. 58.
59. 60. 61.
62.
63. 64. 65. 66. 67. 68. 69.
93
Forster, S.; Burger, C. Macromolecules 1998, 31, 879. Pedersen, J. S.; Gerstenberg, M. C. Macromolecules 1996, 29, 1363. Pedersen, J. S. J. Chem. Phys. 2001, 114, 2839. Mortensen, K.; Pedersen, J. S. Macromolecules 1993, 26, 805. Baxter, R. J. J. Chem. Phys. 1968, 49, 2770. Liu, Y.; Chen, S.-H.; Huang, J. S. Macromolecules 1998, 31, 2236. Mortensen, K.; Brown, W.; J0rgensen, E. Macromolecules 1994, 27, 5654. Tuzar, Z.; Kratochvil, P. Scattering from Block Copolymer Micellar Systems. In Light Scattering-Principles and Development; Brown, W., Ed. Oxford University Press: Oxford, 1996; p 327. Booth, C.; Yu, G.-E.; Nace, V. M. Block Copolymers of Ethylene Oxide and 1, 2Butylene Oxide. In Amphiphilic Block Copolymers: Self-assembly and Applications; Alexandridis, P., Lindman, B., Eds. Elsevier: Amsterdam, 2000; p 57. Yu, G.-E.; Yang, Z.; Attwood, D.; Price, C.; Booth, C. Macromolecules 1996, 29, 8479. Vrij, A. J. Chem. Phys. 1978, 69, 1742. Carnahan, N. E; Starling, K. E. J. Chem. Phys. 1969, 57, 635. Evans, D. E; Wennerstrom, H. The Colloidal Domain. Where Physics, Chemistry, Biology and Technology Meet. John Wiley & Sons, Ltd: New York, 1999. Alexandridis, P.; Athanassiou, V.; Fukuda, S.; Hatton, T. A. Langmuir 1994, 10, 2604. Lopes, J. R.; Lori, W. Langmuir 1998, 14, 750. Price, C. Colloidal Properties of Block Copolymers. In Developments in Block Copolymers; Goodman, I., Ed. Applied Science: London, 1982; Vol. 1; p 39. Russell, T. P. Mater. Sci. Rep. 1990, 5, 171. Hamley, I. W.; Pedersen, J. S. J. Appl. Cryst. 1994, 27, 29. Parratt, L. G. Phys. Rev. 1954, 95, 359. Jonsson, B.; Lindman, B.; Holmberg, K.; Kronberg, B. Surfactants and Polymers in Aqueous Solution. John Wiley & Sons, Ltd: Chichester, 1998. Edens, M. W.; Whitmarsh, R. H. Applications of Block Copolymer Surfactants. In Developments in Block Copolymer Science and Technology; Hamley, I. W., Ed. John Wiley & Sons, Ltd: Chichester, 2004; p 325. Alexandridis, P. Curr. Opin. Colloid Interface Sci. 1997, 2, 478. Alexandridis, P.; Holzwarth, J. E; Hatton, T. A. Macromolecules 1994, 27, 2414. Edens, M. W. Applications of Polyoxyethylene Block Copolymer Surfactants. In Nonionic Surfactants. Polyoxyalkylene Block Copolymers; Nace, V. N., Ed. Marcel Dekker: New York, 1996; Vol. 60; p 185. Chu, B.; Zhou, Z. Physical Chemistry of Polyoxyalkylene Block Copolymer Surfactants. In Nonionic Surfactants: Polyoxyalkylene Block Copolymers; Nace, V. M., Ed. Marcel Dekker: New York, 1996. Pandya, K.; Bahadur, P.; Nagar, T. N.; Bahadur, A. Coll. Surf. A 1993, 70, 219. Reddy, N. K.; Fordham, P. J.; Attwood, D.; Booth, C. J.Chem.Soc., Faraday Trans. 1990, 86, 1569. Prasad, K. N.; Luong, T. T.; Florence, A. T.; Paris, J.; Vution, C.; Seiller, M.; Puisieux, F. J. Colloid Interface Sci. 1979, 69, 225. Bahadur, P.; Pandya, K. Langmuir 1992, 8, 2666. Nixon, S. K.; Hvidt, S.; Booth, C. J. Colloid Interface Sci. 2004, 180, 219. Nakashima, K.; Anzai, T.; Fujimoto, Y. Langmuir 1994, 10, 658. Rassing, J.; Attwood, D. Int. J. Pharm. 1983, 13, 47.
94
Block Copolymers in Solution: Fundamentals and Applications
70. Booth, C; Attwood, D. Macromol. Rapid Commun. 2000, 27, 501. 71. Liu, T.; Zhou, Z.; Wu, C.; Nace, V. M.; Chu, B. J. Phys. Chem. B 1998, 102, 2875. 72. Tanodekaew, S.; Deng, N.-J.; Smith, S.; Yang, Y.-W.; Attwood, D.; Booth, C. J. Phys. Chem. 1993, 97, 11847. 73. Crothers, M.; Attwood, D.; Collett, J. H.; Yang, Z.; Booth, C.; Taboada, P.; Mosquera, V.; Ricardo, N. M. P. S.; Martini, L. G. A. Langmuir 2002, 18, 8685. 74. Yang, Z.; Crothers, M.; Ricardo, N. M. P. S.; Chaibundit, C.; Taboada, P.; Mosquera, V.; Kelarakis, A.; Havredaki, V.; Martini, L. G. A.; Valder, C.; Collett, J. H.; Attwood, D.; Heatley, R; Booth, C. Langmuir 2003, 19, 943. 75. Mai, S.-M.; Booth, C.; Nace, V. M. Eur. Polym. J. 1997, 33, 991. 76. Sommer, C.; Pedersen, J. S.; Stein, P. C. /. Phys. Chem. B 2004, 108, 6242. 77. Malcolm, G. N.; Rowlinson, J. S. Trans. Faraday Soc. 1957, 53, 921. 78. Saeki, S.; Kuwahara, N.; Nakata, M.; Kaneko, M. Polymer 1976, 17, 685. 79. Bae, Y. C.; Lambert, S. M.; Soane, D. S.; Prausnitz, J. M. Macromolecules 1991, 24, 4403. 80. Mai, S. M.; Fairclough, J. P. A.; Hamley, I. W.; Denny, R. C.; Liao, B.; Booth, C.; Ryan, A. J. Macromolecules 1996, 29, 6212. 81. Hamley, I. W.; Castelletto, V.; Yang, Z.; Price, C.; Booth, C. Macromolecules 2001, 34, 4079. 82. Dormidontova, E. E. Macromolecules 2002, 35, 987. 83. Dormidontova, E. E. Macromolecules 2004, 37, 7747. 84. Karlstrom, G. J. Phys. Chem. 1985, 89, 4962. 85. Malmsten, M.; Linse, P.; Zhang, K.-W. Macromolecules 1993, 26, 2905. 86. Schillen, K.; Claesson, P. M.; Malmsten, M.; Linse, P.; Booth, C. J. Phys. Chem. B 1997, 101, 4238. 87. Zhang, L.; Eisenberg, A. Science 1995, 268, 1728. 88. Yu, G.-E.; Ameri, M.; Yang, Z.; Attwood, D.; Price, C.; Booth, C. J. Phys. Chem. B 1997, 101, 4394. 89. Altinok, H.; Yu, G.-E.; Nixon, S. K.; Gorry, P. A.; Attwood, D.; Booth, C. Langmuir 1997, 13, 5837. 90. Lodge, T. P.; Bang, J.; Hanley, K. J.; Krocak, J.; Dahlquist, S.; Sujan, B.; Ott, J. Langmuir 2003, 19, 2103. 91. Nace, V. N. Properties of Polyoxyalkylene Block Copolymers. In Nonionic Surfactants. Polyoxyalkylene Block Copolymers; Nace, V. N., Ed. Marcel Dekker: New York, 1996; Vol. 60; p 145. 92. Nace, V. M. J. Am. Oil Chem. Soc. 1996, 73, 1. 93. Elias, H.-G. The Study of Association and Aggregation via Light Scattering. In Light Scattering from Polymer Solutions; Huglin, M. B., Ed. Academic: London, 1972; p 397. 94. Elias, H.-G. J.Macromol.ScL, Chem. 1973, 7, 601. 95. Tuzar, Z.; Kratochvil, P. Adv. Colloid Interface Sci. 1976, 6, 201. 96. Yang, Y.-W.; Deng, N., -J.; Yu, G.-E.; Zhou, Z.-K.; Attwood, D.; Booth, C. Langmuir 1995, 11, 4703. 97. Hall, D. G. Thermodynamics of Micelle Formation. In Nonionic Surfactants. Physical Chemistry; Schick, M. J., Ed. Marcel Dekker: New York, 1987; Vol. 23; p 233. 98. Holtzer, A.; Holtzer, M. F. J. Phys. Chem. 1974, 78, 1443. 99. Attwood, D.; Florence, A. T. Surfactant Systems: Their Chemistry, Pharmacy and Biology. Chapman and Hall: London, 1983.
Neutral Block Copolymers in Dilute Solution
95
100. Yang, Z.; Packard, S.; Deng, N.-J.; Barlow, R. J.; Attwood, D.; Booth, C. Macromolecules 1994, 27, 2371. 101. Kelarakis, A. Phys. Chem., Chem. Phys. 2001, 3, 5550. 102. Tanford, C. The Hydrophobic Effect. Formation of Micelles and Biological Membranes. John Wiley & Sons, Ltd: New York, 1980. 103. Hamley, I. W. Introduction to Soft Matter. John Wiley & Sons, Ltd: Chichester, 2000. 104. Kelarakis, A.; Havredaki, V.; Yu, G.-E.; Derici, L.; Booth, C. Macromolecules 1998, 31, 944. 105. Bedells, A. D.; Arafeh, R. M.; Yang, Z.; Attwood, D.; Heatley, R; Padget, J. C.; Price, C.; Booth, C. / Chem. Soc., Faraday Trans. 1993, 89, 1235. 106. Mortensen, K.; Brown, W. Macromolecules 1993, 26, 4128. 107. Deng, N.-J.; Luo, Y.-Z.; Tanodekaew, S.; Bingham, N.; Attwood, D.; Booth, C. /. Polym. Sci. B: Polym. Phys. 1995, 33, 1085. 108. Li, H.; Yu, G.-E.; Price, C.; Booth, C.; Hecht, E.; Hoffmann, H. Macromolecules 1997, 30, 1347. 109. Luo, Y. Z.; Nicholas, C. V.; Attwood, D.; Collett, J. H.; Price, C.; Booth, C. Colloid Polym. Sci. 1992, 270, 1094. 110. Yu, G.-E.; Yang, Y.-W.; Yang, Z.; Attwood, D.; Booth, C.; Nace, V. M. Langmuir 1996, 72, 3404. 111. Yang, Y.-W.; Yang, Z.; Zhou, Z.-K.; Attwood, D.; Booth, C. Macromolecules 1996, 29, 670. 112. Yang, Z.; Yang, Y.-W.; Zhao, Z.-K.; Attwood, D.; Booth, C. J. Chem. Soc., Faraday Trans. 1996, 92, 257. 113. Zhou, Z.; Chu, B.; Nace, V. M. Langmuir 1996, 12, 5016. 114. Zhou, Z.; Chu, B.; Nace, V. M.; Yang, Y.-W.; Booth, C. Macromolecules 1996, 29, 3663. 115. Zhou, Z.; Yang, Y.-W.; Booth, C.; Chu, B. Macromolecules 1996, 29, 8357. 116. Anderson, B. C.; Cox, S. M.; Bloom, P. D.; Sheares, V. V.; Mallapragada, S. K. Macromolecules 2003, 36, 1670. 117. Robinson, K. L.; de Paz Banez, M. V.; Wang, X. S.; Armes, S. P. Macromolecules 2001, 34, 5799. 118. Minatti, E.; Borsali, R.; Schappacher, M.; Deffieux, A.; Soldi, V.; Narayanan, T. Macromol.Rapid Commun. 2002, 23, 978. 119. Borsali, R.; Minatti, E.; Putaux, J.-L.; Schappacher, M.; Deffieux, A.; Viville, P.; Lazzaroni, R.; Narayanan, T. Langmuir 2003, 19, 6. 120. latrou, H.; Willner, L.; Hadjichristidis, N.; Halperin, A.; Richter, D. Macromolecules 1996, 29, 581. 121. Ramzi, A.; Prager, M.; Richter, D.; Efstratiadis, V.; Hadjichristidis, N.; Young, R. N.; Allgaier, J. B. Macromolecules 1997, 30, 7171. 122. Pispas, S.; Hadjichristidis, N.; Potemkin, L; Khokhlov, A. Macromolecules 2000, 33, 1741. 123. Pispas, S.; Poulos, Y; Hadjichristidis, N. Macromolecules 1998, 31, 4177. 124. Yu, G.-E.; Mistry, D.; Ludhera, S.; Heatley, R; Attwood, D.; Booth, C. J.Chem.Soc., Faraday Trans. 1997, 93, 3383. 125. Sahakaro, K.; Chaibundit, C.; Kaligradaki, Z.; Mai, S.-M.; Heatley, R; Booth, C.; Padget, J. C.; Shirley, I. M. Eur. Polym. J. 2000, 36, 1835. 126. Kelarakis, A.; Mai, S.-M.; Havredaki, V.; Brett, A.; Booth, C. J. Colloid Interface Sci. 2004, 275, 439.
96
Block Copolymers in Solution: Fundamentals and Applications
127. Hodrokoukes, P.; Pispas, S.; Hadjichristidis, N. Macromolecules 2002, 35, 834. 128. Apperloo, J. J.; Janssen, R. A. J.; Malenfant, P. R. L.; Frechet, J. M. J. Macromolecules 2000, 33, 7038. 129. Chang, Y.; Kwon, Y. C.; Lee, S. C; Kim, C. Macromolecules 2000, 33, 4496. 130. Armstrong, J.; B., C.; Mitchell, J.; Beezer, A.; Leharne, S. J. Phys. Chem. 1996, 100, 1738. 131. Pandit, N. K.; Mclntyre, H. J. Pharmaceut. Dev. Technol. 1997, 2, 181. 132. Kwon, K.-S.; Park, M. J.; Hwang, J.; Char, K. Polym. J. 2001, 33, 404. 133. Cheng, Y; Jolicoeur, C. Macromolecules 1995, 28, 2665. 134. Alexandridis, P. A.; Hatton, T. A. Coll. Surf. A 1995, 96, 1. 135. Alexandridis, P.; Holzwarth, J. F. Langmuir 1997, 13, 6074. 136. Bahadur, P.; Li, P.; Almgren, M.; Brown, W. Langmuir 1992, 8, 1903. 137. Bahadur, P.; Pandya, K.; Almgren, M.; Li, P.; Stilbs, P. Colloid Polym. Sci. 1993, 277, 657. 138. Alexandridis, P.; Athanassiou, V.; Hatton, T. A. Langmuir 1995, 11, 2442. 139. Armstrong, J. K.; Chowdbury, B. Z.; Snowden, M. J.; Leharne, S. A. Langmuir 1998, 14, 2004. 140. J0rgensen, E. B.; Hvidt, S.; Brown, W.; Schillen, K. Macromolecules 1997, 30, 2355. 141. de Gennes, P. G. Macromolecules and Liquid Crystals: Reflections on Certain Lines of Research. In Solid State Physics; Liebert, L., Ed. Academic: New York, 1978; Vol. Suppl. 14; p 1. 142. Alexander, S. J. Phys. France 1977, 38, 977. 143. de Gennes, P. G. /. Phys. France 1976, 37, 1443. 144. de Gennes, P. G. Macromolecules 1980, 13, 1069. 145. Daoud, M.; Cotton, J. P. /. Phys. France 1982, 43, 531. 146. Marques, C. M.; Izzo, D.; Charitat, T.; Mendes, E. Eur. Phys. J. B 1998, 3, 353. 147. Zhulina, E. B.; Birshtein, T. M. Vysokomol.Soedin. 1985, 27, 511. 148. Halperin, A. Macromolecules 1987, 20, 2943. 149. Leibler, L.; Orland, H.; Wheeler, J. C. /. Chem. Phys. 1983, 79, 3550. 150. Meier, D. J. J.Polym.Sci. C 1969, 26, 81. 151. Noolandi, J.; Hong, K. M. Macromolecules 1982, 75, 482. 152. Helfand, E.; Tagami, Y. /. Chem. Phys. 1972, 56, 3592. 153. Helfand, E.; Sapse, A. M. J. Chem. Phys. 1975, 62, 1327. 154. Jones, R. A. L. Soft Condensed Matter. Oxford University Press: Oxford, 2002. 155. ten Brinke, G.; Hadziioannou, G. Macromolecules 1987, 20, 486. 156. Balsara, N. P.; Tirrell, M.; Lodge, T. P. Macromolecules 1991, 24, 1975. 157. Mayes, A. M.; Olvera de la Cruz, M. Macromolecules 1988, 21, 2543. 158. Noolandi, J.; Hong, K. M. Macromolecules 1983, 16, 1443. 159. Whitmore, M. D.; Noolandi, J. Macromolecules 1985, 18, 657. 160. Selb, J.; Marie, P.; Rameau, A.; Duplessix, R.; Gallot, Y. Polym. Bull. 1983, 10, 444. 161. Bluhm, T. L.; Whitmore, M. D. Can. J. Chem. 1985, 63, 249. 162. Bluhm, T. L.; Malhotra, S. L. Eur. Polym. J. 1986, 22, 249. 163. Pepin, M. P.; Whitmore, M. D. Macromolecules 2000, 33, 8644. 164. Hong, K. M.; Noolandi, J. Macromolecules 1983, 16, 1083. 165. Scheutjens, J. M. H. M.; Fleer, G. J. J. Phys. Chem. 1979, 83, 1619. 166. Scheutjens, J. M. H. M.; Fleer, G. J. J. Phys. Chem. 1980, 84, 178. 167. van Lent, B.; Scheutjens, J. H. M. Macromolecules 1989, 22, 1931.
Neutral Block Copolymers in Dilute Solution 168. 169. 170. 171. 172. 173. 174. 175. 176. 177. 178. 179. 180. 181. 182. 183. 184. 185. 186. 187. 188. 189. 190. 191. 192. 193. 194. 195. 196. 197. 198. 199. 200. 201. 202. 203. 204. 205. 206. 207. 208.
97
Linse, P.; Malmsten, M. Macromolecules 1992, 25, 5434. Linse, P. Macromolecules 1993, 26, 4437. Linse, P. J. Phys. Chem. 1993, 97, 13896. Linse, P. Macromolecules 1994, 27, 2685. Linse, P. Macromolecules 1994, 27, 6404. Attwood, D.; Collett, J. H.; Tait, C. J. Int. J. Pharm. 1985, 26, 25. Hurter, P. N.; Scheutjens, J. M. H. M.; Hatton, T. A. Macromolecules 1993, 26, 5030. Hurter, P. N.; Scheutjens, J. M. H. M.; Hatton, T. A. Macromolecules 1993, 26, 5592. de Bruijn, V. G.; van den Broeke, L. J. P.; Leermakers, F. A. M.; Keurentjes, J. T. F. Langmuir 2002, 78, 10467. Svensson, M.; Linse, P. Macromolecules 1998, 31, 1427. Nagarajan, R.; Ganesh, K. /. Chem. Phys. 1989, 90, 5843. Rodrigues, K.; Mattice, W. L. Langmuir 1992, 8, 456. Wang, Y.; Mattice, W. L.; Napper, D. H. Langmuir 1993, 9, 66. Adriani, P.; Wang, Y.; Mattice, W. L. J. Chem. Phys. 1994, 100, 7718. Nguyen-Misra, M.; Mattice, W. L. Macromolecules 1995, 28, 1444. Nguyen-Misra, M.; Mattice, W. L. Macromolecules 1995, 28, 6976. Haliloglu, T.; Bahar, I.; Erman, B.; Mattice, W. L. Macromolecules 1996, 29, 4764. Viduna, D.; Milchev, A.; Binder, K. Macromol. Theory Simul. 1998, 7, 649. Milchev, A.; Bhattacharya, A.; Binder, K. Macromolecules 2001, 34, 1881. Nelson, P. H.; Rutledge, G. C.; Hatton, T. A. /. Chem. Phys. 1997, 707, 10777. von Gottberg, F.; Smith, K. A.; Hatton, T. A. J. Chem. Phys. 1997, 706, 9850. Kim, S. H.; Jo, W. H. Macromolecules 2001, 34, 7210. Pedersen, J. S.; Svaneborg, C.; Almdal, K.; Hamley, I. W.; Young, R. N. Macromolecules 2003, 36, 416. Svaneborg, C.; Pedersen, J. S. Phys. Rev. E 2001, 64, 010802(R). Dormidontova, E. E.; Khokhlov, A. R. Macromolecules 1997, 30, 1980. Lodge, T. P.; Hillmyer, M. A.; Zhou, Z.; Talmon, I. Macromolecules 2004, 37, 6680. Li, Z.; Kesselman, E.; Talmon, Y; Hillmyer, M. A.; Lodge, T. P. Science 2004, 506, 98. Forster, S.; Zisenis, M.; Wenz, E.; Antonietti, M. J. Chem. Phys. 1996, 704, 9956. Bahadur, P.; Sastry, N. V.; Marti, S.; Riess, G. Coll. Surf. 1985, 76, 337. Oranli, L.; Bahadur, P.; Riess, G. Can. J. Chem. 1985, 63, 2691. Nguyen, D.; Williams, C. E.; Eisenberg, A. Macromolecules 1994, 27, 5090. Qin, A.; Tian, M.; Ramireddy, C.; Webber, S. E.; Munk, P.; Tuzar, Z. Macromolecules 1994, 27, 120. Tao, J.; Stewart, S.; Liu, G.; Yang, M. Macromolecules 1997, 30, 2738. Xu, R.; Winnik, M. A.; Riess, G.; Chu, B.; Croucher, M. D. Macromolecules 1992, 25, 644. Cogan, K. A.; Gast, A. P.; Capel, M. Macromolecules 1991, 24, 6512. Forster, S.; Wenz, E.; Lindner, P. Phys. Rev. Lett. 1996, 77, 95. Antonietti, M.; Heinz, S.; Schmidt, M.; Rosenauer, C. Macromolecules 1994, 27, 3276. Chaibundit, C.; Mai, S.-M.; Heatley, E; Booth, C. Langmuir 2000, 76, 9645. Kelarakis, A.; Havredaki, V.; Viras, K.; Mingvanish, W.; Heatley, E; Booth, C.; Mai, S.-M. J. Phys. Chem. B 2001, 705, 7384. Forster, S.; Antonietti, M. Adv. Mater. 1998, 70, 195. Israelachvili, J. N.; Mitchell, D. J.; Ninham, B. W. J. Chem. Soc., Faraday Trans. 2 1976, 72, 1525.
98
Block Copolymers in Solution: Fundamentals and Applications
209. Poppe, A.; Willner, L.; Allgaier, J.; Stellbrink, J.; Richter, D. Macromolecules 1997, 30, 7462. 210. Willner, L.; Poppe, A.; Allgaier, J.; Monkenbusch, M.; Lindner, P.; Richter, D. Europhys. Lett. 2000, 57, 628. 211. Witten, T. A.; Pincus, P. A. Macromolecules 1986, 19, 2509. 212. Sebastian, J. M.; Lai, C; Graessley, W. W.; Register, R. A. Macromolecules 2002, 35, 2707. 213. Castelletto, V.; Caillet, C.; Fundin, J.; Hamley, I. W.; Yang, Z.; Kelarakis, A. /. Chem. Phys. 2002, 116, 10947. 214. McConnell, G. A.; Lin, E. K.; Gast, A. P.; Huang, J. S.; Lin, M. Y.; Smith, S. D. Faraday Discuss. 1994, 98, 121. 215. Gast, A. P. Langmuir 1996, 12, 4060. 216. Buitenhuis, J.; Forster, S. /. Chem. Phys. 1997, 107, 262. 217. Brown, G. J.; Richards, R. W.; Heenan, R. K. Polymer 2001, 42, 7663. 218. Liu, Y. C.; Chen, S. H.; Huang, J. S. Phys. Rev. E 1996, 54, 1698. 219. Lobry, L.; Micali, N.; Mallamace, E; Liao, C.; Chen, S.-H. Phys. Rev. E 1999, 60, 7076. 220. Chen, S.-H.; Liao, C.; Fratini, E.; Baglioni, P.; Mallamace, F. Coll. Surf. A 2001, 183-185, 95. 221. Kimura, T.; Kurata, M. Macromolecules 1981, 14, 1104. 222. Honda, C.; Hasegawa, Y; Hirunuma, R.; Nose, T. Macromolecules 1994, 27, 7660. 223. Honda, C.; Abe, Y; Nose, T. Macromolecules 1996, 29, 6778. 224. Dormidontova, E. E. Macromolecules 1999, 32, 7630. 225. Honda, C.; Yamamoto, K.; Nose, T. Polymer 1996, 37, 1975. 226. lyama, K.; Nose, T. Polymer 1998, 39, 651. 227. lyama, K.; Nose, T. Macromolecules 1998, 31, 7356. 228. Goldmints, I.; Holzwarth, J. E; Smith, K. A.; Hatton, T. A. Langmuir 1997, 13, 6130. 229. Aniansson, E. A. G.; Wall, S. N. J. Phys. Chem. 1974, 78, 1024. 230. Aniansson, E. A. G.; Wall, S. N. J. Phys. Chem. 1975, 79, 857. 231. Kositza, M. J.; Bohne, C.; Alexandridis, P.; Hatton, T. A.; Holzwarth, J. F. Macromolecules 1999, 32, 5539. 232. Kositza, M. J.; Bohne, C.; Alexandridis, P.; Hatton, T. A. Langmuir 1999, 75, 322. 233. Kositza, M. J.; Bohne, C.; Hatton, T. A.; Holzwarth, J. F. Prog. Colloid Polym. Sci. 1999, 772, 146. 234. Aniansson, E. A. G.; Wall, S. N.; Almgren, M.; Hoffmann, H.; Kielmann, L; Ulbricht, W.; Zana, R.; Lang, J.; Tondre, C. J. Phys. Chem. 1976, 80, 905. 235. Waton, G.; Michels, B.; Zana, R. Macromolecules 2001, 34, 907. 236. Michels, B.; Waton, G.; Zana, R. Langmuir 1997, 13, 3111. 237. Willner, L.; Poppe, A.; Allgaier, J.; Monkenbusch, M.; Richter, D. Europhys. Lett. 2001, 55, 667. 238. Creutz, S.; van Stam, J.; De Schryver, F. C.; Jerome, R. Macromolecules 1998, 31, 681. 239. Tian, M.; Qin, A.; Ramireddy, C.; Webber, S. E.; Munk, P.; Tuzar, Z.; Prochazka, K. Langmuir 1993, 9, 1741. 240. Esselink, F. J.; Dormidontova, E. E.; Hadziioannou, G. Macromolecules 1998, 31, 2925. 241. Esselink, F. J.; Dormidontova, E. E.; Hadziioannou, G. Macromolecules 1998, 31, 4873. 242. Rager, T.; Meyer, W. H.; Wegner, G. Macromol. Chem. Phys. 2003, 200, 1672. 243. van Stam, J.; Creutz, S.; De Schryver, F. C.; Jerome, R. Macromolecules 2000, 33, 6388. 244. Won, Y.-Y; Davis, H. T.; Bates, F. S. Macromolecules 2003, 36, 953.
Neutral Block Copolymers in Dilute Solution
99
245. Jain, S.; Bates, F. S. Macromolecules 2004, 37, 1511. 246. Teng, Y.; Morrison, M. E.; Munk, P.; Webber, S. E.; Prochazka, K. Macromolecules 1998, 31, 3578. 247. Johnson, B. K.; Prud'homme, R. K. Phys. Rev. Lett. 2003, 91, 118302. 248. Tsunashima, Y.; Hirata, M.; Kawamata, Y. Macromolecules 1990, 23, 1089. 249. Akcasu, A. Z.; Benmouna, M. Macromolecules 1978, 11, 1193. 250. Sigel, R.; Pispas, S.; Hadjichristidis, N.; Vlassopoulos, D.; Fytas, G. Macromolecules 1999, 32, 8447. 251. Tommaseo, G.; Penciu, R. S.; Fytas, G.; Economou, E. N.; Hashimoto, T.; Hadjichristidis, N. Macromolecules 2004, 37, 5006. 252. Candau, F; Heatley, F; Price, C.; Stubbersfield, R. B. Eur. Polym. J. 1984, 20, 685. 253. Spevacek, J. Makromol.Chem., Rapid Commun. 1982, 3, 691. 254. Almgren, M.; Bahadur, P.; Jansson, M.; Li, P.; Brown, W. /. Colloid Interface Sci. 1992, 757, 157. 255. Almgren, M.; Brown, W.; Hvidt, S. Colloid Polym. Sci. 1995, 273, 2. 256. Godward, J.; Heatley, F; Booth, C. J. Chem. Soc., Faraday Trans. 1995, 97, 1491. 257. Gao, Z.; Zhong, X.-F; Eisenberg, A. Macromolecules 1994, 27, 794. 258. Kriz, J.; Plestil, J.; Tuzar, Z.; Pospisil, H.; Brus, J.; Jakes, J.; Masar, B.; Vlcek, P.; Doskocilova, D. Macromolecules 1999, 32, 397. 259. Kriz, J.; Brus, J.; Plestil, J.; Kurkova, D.; Masar, B.; Dybal, J.; Zune, C.; Jerome, R. Macromolecules 2000, 33, 4108. 260. Mochrie, S. G. J.; Mayes, A. M.; Sandy, A. R.; Sutton, M.; Brauer, S.; Stephenson, G. B.; Abernathy, D. L.; Griibel, G. Phys. Rev. Lett. 1997, 78, 1275. 261. Robert, A.; Grubel, G.; Waigh, T. A.; Williams, C. E. submitted. 262. Farago, B.; Monkenbusch, M.; Richter, D.; Huang, J. S.; Fetters, L. J.; Gast, A. P. Phys. Rev. Lett. 1993, 77, 1015. 263. Matsuoka, H.; Yamamoto, Y; Nakano, M.; Endo, H.; Yamaoka, H.; Zorn, R.; Monkenbusch, M.; Richter, D.; Seto, H.; Kawabata, Y; Nagao, M. Langmuir 2000, 16, 9177. 264. Castelletto, V.; Hamley, I. W.; Yang, Z.; Haeussler, W. J. Chem. Phys. 2003, 779, 8158. 265. Dubois-Violette, E.; de Gennes, P.-G. Physics 1967, 3, 181. 266. Annable, T.; Buscall, R.; Ettelaie, R. Coll. Surf. A 1996, 772, 97. 267. Annable, T.; Buscall, R.; Ettelaie, R.; Whittlestone, D. J. Rheol. 1993, 37, 695. 268. Jenkins, R. D.; Silebi, C. A.; El-Aaser, M. S. Polym. Mater. Sci. Eng. 1989, 61, 629. 269. Ma, S. X.; Cooper, S. L. Macromolecules 2001, 34, 3294. 270. Hwang, F. S.; Hogen-Esch, T. E. Macromolecules 1993, 26, 3156. 271. Hwang, F. S.; Hogen-Esch, T. E. Macromolecules 1995, 28, 3328. 272. Xie, X.; Hogen-Esch, T. E. Macromolecules 1996, 29, 1734. 273. Zhang, Y.-X.; Da, A.-H.; Butler, G. B.; Hogen-Esch, T. E. J.Polym.Sci. 1992, 30, 1383. 274. Calvet, D.; Collet, A.; Viguier, M.; Berret, J.-F; Serero, Y. Macromolecules 2003, 36, 449. 275. Cathebras, N.; Collet, A.; Viguier, M. Macromolecules 1998, 37, 1305. 276. Tae, G.; Kornfield, J. A.; Hubbell, J. A.; Lai, J. Macromolecules 2002, 35, 4448. 277. Alami, E.; Almgren, M.; Brown, W.; Frangois, J. Macromolecules 1996, 29, 2229. 278. Francis, J.; Maitre, S.; Rawiso, M.; Sarazin, D.; Beinert, G.; Isel, F. Coll. Surf. A 1996, 772, 251.
100
Block Copolymers in Solution: Fundamentals and Applications
279. Courier, C; Beaudoin, E.; Duval, M.; Sarazin, D.; Maitre, S.; Francois, J. J. Colloid Interface Sci. 2000, 230, 41. 280. Beaudoin, E.; Borisov, O.; Lapp, A.; Billon, L.; Hiorns, R. C.; Francois, J. Macromolecules 2002, 35, 7436. 281. Beaudoin, E.; Hiorns, R. C.; Borisov, O.; Fran§ois, J. Langmuir 2003, 19, 2058. 282. Borisov, O. V.; Halperin, A. Curr. Opin. Colloid Interface Sci. 1998, 3, 415. 283. Winnik, M. A.; Yekta, A. Curr. Opin. Colloid Interface Sci. 1997, 2, 424. 284. Annable, T.; Buscall, R.; Ettelaie, R. Rheology of Transient Networks Formed by the Association of Hydrophobically Modified Water Soluble Polymer. In Amphiphilic Block Copolymers. Self-assembly and Applications; Alexandridis, P., Lindman, B., Eds. Elsevier: Amsterdam, 2000; p 281. 285. Pham, Q. T.; Russel, W. B.; Thibeault, J. C.; Lau, W. Macromolecules 1999, 32, 5139. 286. Pham, Q. T.; Russel, W. B.; Thibeault, J. C.; Lau, W. Macromolecules 1999, 32, 2996. 287. Kelarakis, A.; Havredaki, V.; Yuan, X.-E; Wang, Y.-W.; Booth, C. J. Mater. Chem. 2003, 13, 2779. 288. Kelarakis, A.; Ming, X.-T.; Yuan, X.-E; Booth, C. Langmuir 2004, 20, 2036. 289. Ng, W. K.; Tarn, K. C.; Jenkins, R. D. /. Rheol. 2000, 44, 137. 290. Song, T.; Dai, S.; Tam, K. C.; Lee, S. Y; Goh, S. H. Langmuir 2003, 19, 4798. 291. Hoppenbrouwers, E.; Li, Z.; Liu, G. Macromolecules 2003, 36, 876. 292. Wang, X. S.; Winnik, M. A.; Manners, I. Macromolecules 2002, 35, 9146. 293. Gohy, J. E; Willet, N.; Varshney, S.; Zhang, J. X.; Jerome, R. Angew. Chem., Int. Edn 2001, 40, 3214. 294. Lei, L.; Gohy, J.-E; Willet, N.; Zhang, J.-X.; Varshney, S.; Jerome, R. Macromolecules 2004, 37, 1089. 295. Koutalas, G.; Pispas, S.; Hadjichristidis, N. Euro. Phys. J. E 2004, 15, 1. 296. Yu, G.-E.; Eisenberg, A. Macromolecules 1998, 31, 5546. 297. Patrickios, C. S.; Hertler, W. R.; Abbott, N. L.; Hatton, T. A. Macromolecules 1994, 27, 930. 298. Patrickios, C. S.; Sharma, L. R.; Armes, S. P.; Billingham, N. C. Langmuir 1999, 75, 1613. 299. Plestil, J.; Kriz, J.; Tuzar, Z.; Prochazka, K.; Melnichenko, Y. B.; Wignall, G. D.; Talingting, M. R.; Munk, P.; Webber, S. E. Macromol. Chem. Phys. 2001, 202, 553. 300. Plestil, J.; Pospisil, H.; Kriz, J.; Kadlec, P.; Tuzar, Z.; Cubitt, R. Langmuir 2001, 17, 6699. 301. Plestil, J.; Pospisil, H.; Kadlec, P.; Tuzar, Z.; Kriz, J.; Gordeliy, V. I. Polymer 2001, 42, 2941. 302. Kriz, J.; Plestil, J.; Tuzar, Z.; Pospisil, H.; Doskocilova, D. Macromolecules 1998, 31, 41. 303. Semenov, A. N.; Nyrkova, I. A.; Khokhlov, A. Macromolecules 1995, 28, 7491. 304. Zhou, Z.; Li, Z.; Ren, Y; Hillmyer, M. A.; Lodge, T. P. /. Am. Chem. Soc. 2003, 725, 10182. 305. Zhang, L.; Yu, K.; Eisenberg, A. Science 1996, 272, 1777. 306. Ma, Q.; Remsen, E. E.; Clark, C. J.; Kowalewski, T.; Wooley, K. L. Proc. Natl. Acad. Sci. USA 2002, 99, 5058. 307. Jenekhe, S. A.; Chen, X. L. Science 1998, 279, 1903. 308. Jenekhe, S. A.; Chen, X. L. Science 1999, 283, 372. 309. Zubarev, E. R.; Pralle, M. U.; Sone, E. D.; Stupp, S. I. J. Am. Chem. Soc. 2001, 123, 4105. 310. de Gans, B. J.; Wiegand, S.; Zubarev, E. R.; Stupp, S. I. /. Phys. Chem. B 2002, 706, 9730.
Neutral Block Copolymers in Dilute Solution
101
311. Sommerdijk, N. A. J. M.; Holder, S. J.; Hiorns, R. C.; Jones, R. G.; Nolle, R. J. M. Macromolecules 2000, 33, 8289. 312. Tu, Y.; Wan, X.; Zhang, H.; Fan, X.; Chen, X.; Zhuo, Q.-R; Chau, K. Macromolecules 2003, 36, 6565. 313. Wu, J.; Pearce, E. M.; Kwei, T. K.; Lefebvre, A. A.; Balsara, N. P. Macromolecules 2002, 35, 1791. 314. Williams, D. R. M.; Fredrickson, G. H. Macromolecules 1992, 25, 3561. 315. Wu, C.; Niu, A.; Leung, L. M.; Tarn, T. S. /. Am. Chem. Soc. 1999, 727, 1954. 316. Thurmond, K. B.; Huang, H.; Clark, C. G.; Kowalewski, T.; Wooley, K. L. Coll. Surf. B 1999, 16, 45. 317. Clark, C. G.; Wooley, K. L. Curr. Opin. Colloid Interface Sci. 1999, 4, 122. 318. Wooley, K. L. J. Polym. Sci. A: Polym. Chem. 2000, 38, 1397. 319. Thurmond, K. B.; Kowalewski, T.; Wooley, K. L. /. Am. Chem. Soc. 1997, 779, 6656. 320. Remsen, E. E.; Thurmond, K. B.; Wooley, K. L. Macromolecules 1999, 32, 3685. 321. Thurmond, K. B.; Remsen, E. E.; Kowalewski, T.; Wooley, K. L. Nucleic Acids Res. 1999, 27, 2966. 322. Zhang, Q.; Remsen, E. E.; Wooley, K. L. J. Am. Chem. Soc. 2000, 722, 3642. 323. Ma, Q.; Wooley, K. L. J. Polym. Sci. A: Polym. Chem. 2000, 38, 4805. 324. Zhang, Q.; Clark, C. G.; Wang, M.; Remsen, E. E.; Wooley, K. L. Nano Lett. 2002, 2, 1051. 325. Ma, Q.; Remsen, E. E.; Kowalewski, T.; Schaefer, J.; Wooley, K. L. Nano Lett. 2001, 7, 651. 326. Huang, H.; Remsen, E. E.; Kowalewski, T.; Wooley, K. L. / Am. Chem. Soc. 1999, 727, 3805. 327. Huang, H.; Kowalewski, T.; Remsen, E. E.; Gertzmann, R.; Wooley, K. L. J. Am. Chem. Soc. 1997, 779, 11653. 328. Murthy, K. S.; Ma, Q.; Remsen, E. E.; Kowalewski, T.; Wooley, K. L. J. Mater. Chem. 2003, 13, 2785. 329. Murthy, K. S.; Ma, Q.; Clark, C. G.; Remsen, E. E.; Wooley, K. L. Chem. Commun. 2001, 773. 330. Joralemon, M. J.; Murthy, K. S.; Remsen, E. E.; Becker, M. L.; Wooley, K. L. Biomacromolecules 2004, 5, 903. 331. Pan, D.; Turner, J. L.; Wooley, K. L. Chem. Commun. 2003, 2400. 332. Becker, M. L.; Remsen, E. E.; Wooley, K. L. /. Polym. Sci. A: Polym. Chem. 2001, 39, 4152. 333. Butun, V.; Wang, X.-S.; de Paz-Banez, M. V.; Robinson, K. L.; Billingham, N. C.; Armes, S. P. Macromolecules 2000, 33, 1. 334. Liu, S.; Ma, Y.; Armes, S. P.; Perruchot, C.; Watts, J. F. Langmuir 2002, 18, 7780. 335. Liu, S.; Weaver, J. V. M.; Save, M.; Armes, S. P. Langmuir 2002, 75, 8350. 336. Liu, S.; Armes, S. P. J. Am. Chem. Soc. 2001, 723, 9910. 337. Butun, V.; Billingham, N. C.; Armes, S. P. /. Am. Chem. Soc. 1998, 720, 12135. 338. Butiin, V.; Lowe, A. B.; Billingham, N. C.; Armes, S. P. J. Am. Chem. Soc. 1999, 727, 4288. 339. Ding, J.; Liu, G. Macromolecules 1998, 31, 6554. 340. Ding, J.; Liu, G. J. Phys. Chem. B 1998, 702, 6107. 341. Yan, X.; Liu, G.; Li, H. Langmuir 2004, 20, 4677.
102
Block Copolymers in Solution: Fundamentals and Applications
342. Erhardt, R.; Boker, A.; Zettl, H.; Kaya, H.; Pyckhout-Hintzen, W.; Krausch, G.; Abetz, V.; Muller, A. H. E. Macromolecules 2001, 34, 1069. 343. Erhardt, R.; Zhang, M.; Boker, A.; Zettl, H.; Abetz, C.; Frederik, P.; Krausch, G.; Abetz, V.; Muller, A. H. E. /. Am. Chem. Soc. 2003, 725, 3260. 344. Liu, Y.; Abetz, V.; Muller, A. H. E. Macromolecules 2003, 36, 7894. 345. Won, Y.-Y.; Davis, H. T.; Bates, F. S. Science 1999, 283, 960. 346. Lang, P.; Willner, L.; Pyckhout-Hintzen, W.; Krastev, R. Langmuir 2003, 19, 7597. 347. King, S. M.; Heenan, R. K.; Cloke, V. M.; Washington, C. Macromolecules 1997, 30, 6215. 348. Minatti, E.; Viville, P.; Borsali, R.; Schappacher, M.; Deffieux, A.; Lazzaroni, R. Macromolecules 2003, 36, 4125. 349. Mao, G.; Sukumaran, S.; Beaucage, G.; Saboungi, M. L.; Thiyagarajan, P. Macromolecules 2001, 34, 552. 350. Hamley, I. W.; Pedersen, J. A.; Booth, C.; Nace, V. M. Langmuir 2001, 17, 6386. 351. Kaya, H.; Willner, L.; Allgaier, J.; Stellbrink, J.; Richter, D. Appl. Phys. A 2002, 74, (Part 1 Suppl. S), S499. 352. Schillen, K.; Brown, W.; Johnsen, R. M. Macromolecules 1994, 27, 4825. 353. Glatter, O.; Scherf, G.; Schillen, K.; Brown, W. Macromolecules 1994, 27, 6046. 354. Mortensen, K. /. Phys., Condens. Matter 1996, 8, 103. 355. Pedersen, J. S.; Schurtenberger, P. Macromolecules 1996, 29, 7602. 356. Koyama, R. J. Phys. Soc. Jpn 1973, 34, 1029. 357. LaRue, L; Adam, M.; da Silva, M.; Sheiko, S. S.; Rubinstein, M. Macromolecules 2004, 37, 5002. 358. Lodge, T. P.; Xu, X.; Ryu, C. Y.; Hamley, I. W.; Fairclough, J. P. A.; Ryan, A. J.; Pedersen, J. S. Macromolecules 1996, 29, 5955. 359. Pedersen, J. S.; Hamley, I. W.; Ryu, C. Y; Lodge, T. P. Macromolecules 2000, 33, 542. 360. Nakano, M.; Matsuoka, H.; Yamaoka, H.; Poppe, A.; Richter, D. Macromolecules 1999, 32, 697. 361. Nordskog, A.; Egger, H.; Findenegg, G. H.; Hellweg, T.; Schlaad, H.; von Berlepsch, H.; Bottcher, C. Phys. Rev. E 2003, 68, 011406. 362. Yoshida, E.; Kunugi, S. Macromolecules 2002, 35, 6665. 363. Quintana, J. R.; Janez, M. D.; Hernaez, E.; Garcia, A.; Katime, I. Macromolecules 1998, 31, 6865. 364. Barker, M. C.; Vincent, B. Coll.Surf. 1984, 8, 297. 365. Utiyama, H.; Takenaka, K.; Mizumori, M.; Fukuda, M.; Tsunashima, Y; Kurata, M. Macromolecules 1974, 7, 515. 366. Liu, T.; Nace, V. M.; Chu, B. Langmuir 1999, 15, 3109. 367. Liu, T.; Chu, B. J. Appl. Cryst. 2000, 33, 727. 368. Mingvanish, W.; Chaibundit, C.; Booth, C. Phys. Chem., Chem. Phys. 2002, 4, 778. 369. Borovinskii, A. L.; Khokhlov, A. R. Macromolecules 1998, 31, 7636. 370. Sens, P.; Marques, C. M.; Joanny, J.-F. Macromolecules 1996, 29, 4880. 371. Ghoreishi, S. M.; Li, Y; Bloor, D. M.; Warr, J.; Wyn-Jones, E. Langmuir 1999, 75, 4380. 372. Li, Y; Ghoreishi, S. M.; Warr, J.; Bloor, D. M.; Holzwarth, J. F.; Wyn-Jones, E. Langmuir 1999, 75, 6326. 373. Vadnere, M.; Amidon, G.; Lindenbaum, S.; Haslam, J. L. Int. J. Pharm. 1984, 22, 207. 374. Dai, S.; Tarn, K. C.; Li, L. Macromolecules 2001, 34, 7049. 375. Hecht, E.; Hoffmann, H. Langmuir 1994, 10, 86.
Neutral Block Copolymers in Dilute Solution
103
376. Hecht, E.; Mortensen, K.; Gradzielski, M.; Hoffmann, H. /. Phys. Chem. 1995, 99, 4866. 377. Li, Y.; Xu, R.; Bloor, D. M.; Holzwarth, J. E; Wyn-Jones, E. Langmuir 2000, 76, 10515. 378. Li, Y.; Xu, R.; Couderc, S.; Bloor, D. M.; Wyn-Jones, E.; Holzwarth, J. F. Langmuir 2001, 77, 183. 379. Thurn, T.; Couderc, S.; Sidhu, J.; Bloor, D. M.; Penfold, J.; Holzwarth, J. E; Wyn-Jones, E. Langmuir 2002, 18, 9267. 380. Zhang, K.; Lindman, B.; Coppola, L. Langmuir 1995, 77, 538. 381. Almgren, M.; van Stam, J.; Lindblad, C; Li, P.; Stilbs, P.; Bahadur, P. /. Phys. Chem. 1991, 95, 5677. 382. Hecht, E.; Hoffmann, H. Coll. Surf. A 1995, 96, 181. 383. Pandya, K.; Lad, K.; Bahadur, P. J.Macromol.Sci., Chem. 1993, A30, 1. 384. Bronstein, L. M.; Chernyshov, D. M.; Vorontsov, E.; Timofeeva, G. I.; Dubrovina, L. V.; Valetsky, P. M.; Kazakov, S.; Khokhlov, A. J. Phys. Chem. B 2001, 705, 9077. 385. Couderc, S.; Li, Y; Bloor, D. M.; Holzwarth, J. E; Wyn-Jones, E. Langmuir 2001, 77, 4818. 386. Zhang, L.; Eisenberg, A. Macromolecules 1999, 32, 2239. 387. Yu, K.; Eisenberg, A. Macromolecules 1998, 31, 3509. 388. Zhang, W.; Shi, L.; An, Y; Shen, X.; Guo, Y; Gao, L.; Liu, Z.; He, B. Langmuir 2003, 79, 6026. 389. Ding, J.; Liu, G. Macromolecules 1999, 32, 8413. 390. Won, Y.-Y; Davis, H. T.; Bates, F. S.; Agamalian, M.; Wignall, G. D. / Phys. Chem. B 2000, 704, 7134. 391. Jain, S.; Bates, F. S. Science 2003, 300, 460. 392. Zhang, L.; Bartels, C.; Yu, Y; Shen, H.; Eisenberg, A. Phys. Rev. Lett. 1997, 79, 5034. 393. Haluska, C. K.; Gozdz, W. T.; Dobereiner, H.-G.; Forster, S.; Gompper, G. Phys. Rev. Lett. 2002, 89, 238302. 394. Velonia, K.; Rowan, A. E.; Nolte, R. J. M. /. Am. Chem. Soc. 2002, 724, 4224. 395. Antonietti, M.; Forster, S. Adv. Mater. 2003, 75, 1323. 396. Soo, P. L.; Eisenberg, A. J. Polym. Sci. B: Polym. Phys. 2004, 42, 923. 397. Discher, D. E.; Eisenberg, A. Science 2002, 297, 967. 398. Luo, L.; Eisenberg, A. J. Am. Chem. Soc. 2001, 723, 1012. 399. Luo, L.; Eisenberg, A. Langmuir 2001, 77, 6804. 400. Shen, H.; Eisenberg, A. /. Phys. Chem. B 1999, 703, 9473. 401. Shen, H.; Eisenberg, A. Macromolecules 2000, 33, 2561. 402. Schillen, K.; Bryshke, K.; Mel'nikova, Y. S. Macromolecules 1999, 32, 6885. 403. Zipfel, J.; Lindner, P.; Tsianou, M.; Alexandridis, P.; Richtering, W. Langmuir 1999, 75, 2599. 404. Zipfel, J.; Berghausen, J.; Schmidt, G.; Lindner, P.; Alexandridis, P.; Tsianou, M.; Richtering, W. Phys. Chem., Chem. Phys. 1999, 7, 3905. 405. Kickelback, G.; Bauer, J.; Hiising, N.; Andersson, M.; Palmqvist, A. Langmuir 2003, 79, 3198. 406. Harris, J. K.; Rose, G. D.; Bruening, M. L. Langmuir 2002, 18, 5337. 407. Wang, M.; Jiang, M.; Ning, E; Chen, D.; Liu, S.; Duan, H. Macromolecules 2002, 35, 5980. 408. Lu, Z.; Liu, G.; Liu, F. Macromolecules 2001, 34, 8814. 409. Fraaije, J. G. E. M.; Sevink, G. J. A. Macromolecules 2003, 36, 7891.
104
Block Copolymers in Solution: Fundamentals and Applications
410. Discher, B. M; Won, Y.-Y.; Ege, D. S.; Lee, J. C.-M.; Bates, F. S.; Discher, D. E.; Hammer, D. A. Science 1999, 284, 1143. 411. Dimova, R.; Seifert, U.; Pouligny, B.; Forster, S.; Dobereiner, H.-G. Eur. Phys. J. E 2002, 7, 241. 412. Bermudez, H.; Hammer, D. A.; Discher, D. E. Langmuir 2004, 20, 540. 413. Helfrich, W. Z. Naturforsch., C: Biosci. 1975, 30, 841. 414. Evans, E. Biophys. J. 1975, 14, 923. 415. Aranda-Espinoza, H.; Bermudez, H.; Bates, F. S.; Discher, D. E. Phys. Rev. Lett. 2001, 87, 208301. 416. Dalheimer, P.; Bates, F. S.; Aranda-Espinoza, H.; Discher, D. E. C. R. Physique 2003, 4, 251. 417. Sandre, O.; Moreaux, L.; Brochard-Wyart, F. Proc. Natl. Acad. Sci. USA 1999, 96, 10591. 418. Borghi, N.; Rossier, O.; Brochard-Wyart, F. Europhys. Lett. 2003, 64, 837. 419. Lee, J. C.-M.; Santore, M.; Bates, F. S.; Discher, D. E. Macromolecules 2002, 35, 323. 420. Cornelissen, J. J. L. M.; Fischer, M.; Sommerdijk, N. A. J. M.; Nolte, R. J. M. Science 1998, 280, 1427. 421. Skoulios, A. E.; Tsouladze, G.; Franta, E. J.Polym.Sci. C 1963, 4, 507. 422. Gervais, M.; Gallot, B. Makromol. Chem. 1973, 171, 157. 423. Gervais, M.; Gallot, B. Makromol. Chem. 1973, 174, 193. 424. Gervais, M.; Gallot, B. Makromol. Chem. 1977, 178, 1577. 425. Lin, E. K.; Gast, A. P. Macromolecules 1996, 29, 4432. 426. Radulescu, A.; Mathers, R. T.; Coates, G. W.; Richter, D.; Fetters, L. J. Macromolecules 2004, 37, 6962. 427. Richter, D.; Schneiders, D.; Monkenbusch, M.; Willner, L.; Fetters, L. J.; Huang, J. S.; Lin, M.; Mortensen, K.; Farago, B. Macromolecules 1997, 30, 1053. 428. Nakano, M.; Matsumoto, K.; Matsuoka, H.; Yamaoka, H. Macromolecules 1999, 32, 4023. 429. Hamley, I. W.; Castelletto, V.; Fundin, J.; Crothers, M.; Attwood, D.; Talmon, Y. Colloid Polym. Sci. 2004, 282, 514. 430. Hamley, I. W.; Connell, S. D.; Collins, S. Macromolecules 2004, 37, 5337. 431. Talingting, M. R.; Munk, P.; Webber, S. E.; Tuzar, Z. Macromolecules 1999, 32, 1593. 432. Pochan, D.; Chen, Z.; Cui, H.; Hales, K.; Qi, K.; Wooley, K. L. Science 2004, 306, 94.
3 Concentrated Solutions 3.1 UNDERSTANDING PHASE DIAGRAMS At higher concentrations, block copolymers in solution form a variety of lyotropic mesophases. Lamellar, hexagonal-packed cylinder, cubic-packed micellar and bicontinuous cubic structures have all been observed (Figure 3.1). Due to the fact that such phases possess a finite yield stress and so usually do not flow under their own weight, these are often termed gels. However, as emphasized in Section 3.3.2, the gel properties result from the ordered microstructure rather than any cross-links between polymer chains as in a conventional polymer gel. The symmetry of the ordered phase formed largely depends on the interfacial curvature, as for conventional amphiphiles; however the phase behaviour can also be understood by mapping it onto that for block copolymer melts. To a first approximation, the lyotropic phase behaviour in a highly selective solvent depends primarily on copolymer composition, whereas the thermotropic behaviour depends on the temperature dependence of the solvent quality. The phase diagram for short, not too asymmetric, diblocks resembles that of nonionic surfactants in the richness of lyotropic phase behaviour. For example, micellar liquid, micellar cubic, hexagonal and lamellar phases are all observed for EO18-b-BO10 in water.1 However, as the copolymer compositional asymmetry increases, packing frustration prevents the formation of lamellar and hexagonal phases and interfacial curvature favours formation of spherical micelles. When the effective volume fraction of micelles exceeds that for close packing of hard (or soft) spheres, a cubic micellar phase is formed,2'3 as discussed further in Section 3.3.2. The phase behaviour of a number of PS-b-PI diblocks in solvents of varying selectivity has been investigated by Lodge and coworkers.4-6 They used several of the di-n-alkyl phthalates, all of which are good solvents for PS but the solvent quality for PI varies. The first, DOP is nearly neutral,6 DBP is 'slightly selective', being a near-6 solvent for PI at 90 °C. A further reduction in alkyl chain length in DEP and DMP leads to increasing selectivity for PS. The phase behaviour of symmetric and asymmetric diblocks in these solvents has been investigated via rheometry, static birefringence and light scattering measurements coupled with SAXS. Comparison was also made with phase behaviour in tetradecane, a selective solvent for PL Lai et al. have also investigated the phase behaviour of PS-b-PI
Block Copolymers in Solution: Fundamentals and Applications © 2005 John Wiley & Sons, Ltd.
I. W. Hamley
106
Block Copolymers in Solution: Fundamentals and Applications
Figure 3.1 Lyotropic mesophase structures: (a) micellar cubic; (b) hexagonal; (c) lamellar; (d) bicontinuous cubic.26 Here a portion of the gyroid structure is sketched. The amphiphilic molecules form a bilayer film separating two continuous labyrinths of water. The amphiphilic film is a network with three-fold node points, which defines the gyroid phase. Reproduced by permission of John Wiley & Sons, Ltd.
diblocks in highly selective solvents - tetradecane, squalane and tributylamine, all of which are selective for PI.7 Representative phase diagrams in the four solvents for one asymmetric diblock are presented in Figure 3.2.4 Many aspects of the phase behaviour (in particular the topology of the phase diagram) can be understood via a trajectory map, illustrated in Figure 3.3. This describes a mapping onto the melt phase diagram, illustrated for PS-b-PI diblocks in Figure 3.3. A number of ordered phases are observed depending on copolymer volume fraction, /. Highly asymmetric diblocks adopt cubic-packed sphere structures in which the minority block constitutes the spheres and the majority block forms the matrix. Cylinder and gyroid phases are observed for less asymmetric diblocks. A lamellar phase is observed for symmetric and near symmetric diblocks. In this picture, increasing the concentration of the neutral solvent DOP corresponds to dilution of the melt. The solid line for the gyroidcylinder order-order transition (OOT) follows the 'dilution approximation', whereby the Flory-Huggins interaction parameter scales as XOOT ~ 0"1 [the orderdisorder transition (ODT) scaling is steeper XODT ~ >~ J 4]- The dashed line in
Concentrated Solutions
107
Figure 3.2 Phase diagram for a PS-b-PI diblock with Mn = 3.2 x 10 g mol-1 and fPS = 0.31 in the solvents indicated.4 The volume fraction of polymer is denoted
. The critical micelle temperature in dilute solution is indicated by a filled square. The ordered phases are denoted: L, lamellae; C, hexagonal-packed cylinders; G, gyroid; PL, perforated lamellae; S, cubic-packed spheres. The subscript 1 indicates a normal phase (minority PS component in minority domains) and 2 indicates an inverted phase (PS in majority domains). The smooth curves are guides to the eyes, except for DOP in which the OOT and ODT phase boundaries (solid lines) show the previously determined scaling of the PS-PI interaction parameter. The dashed line corresponds to the 'dilution approximation', XOOT ~ >"'• Biphasic regions were found to be extremely narrow5 and are not indicated. Reproduced by permission of American Chemical Society.
Figure 3.2(a) shows the dilution approximation prediction. The enhanced stability of the disordered phase in comparison to the prediction may be due to thermal composition fluctuations. In the phase diagram trajectory interpretation, dilution corresponds to increasing temperature in the melt reference system.
108
Block Copolymers in Solution: Fundamentals and Applications
Figure 3.3 Phase trajectories of an asymmetric PS-b-PI diblock in DOP, DBP, DEP, and C14.4 The open and closed symbols correspond to OOTs and ODTs, respectively, determined by Khandpur et a/.273 (circles) and Ryu et a/.274 (squares) for PS-b-PI copolymers, with the dashed lines marking the estimated phase boundaries. The trajectories start at the estimated segregation of the neat diblock at 0 °C. Reproduced by permission of American Chemical Society.
In contrast, addition of a selective solvent leads to a renormalization of the copolymer composition, due to selective swelling of one component. This corresponds to a horizontal trajectory across the melt phase diagram (Figure 3.3).4 This type of behaviour is exemplified by the phase diagram in DEP. A rich sequence of successively normal phases (minority PS component) and then inverse phases (majority PS component) is accessed upon increasing solvent concentration. DEP is a much more highly selective solvent than DBP. This leads to a more nearly horizontal trajectory across the phase diagram [and hence more nearly vertical phase boundaries in the (4>, T) plane]. It also leads to an increase in ODT compared with that in DBP. The vertical component of the trajectory corresponds to changes in segregation strength. Addition of the weakly selective solvent DBP corresponds to a reduction in segregation strength. In contrast, DEP is highly PS selective, leading effectively to increased segregation (compared with the melt state at 0 °C) between components. Tetradecane is a selective solvent for PI, and the trajectory is opposite in composition to that for DEP, although again the segregation strength increases as the solvent partitions into PI. The same concept was also used by Lai et al. to collapse phase diagrams for a series of PS-b-PI diblocks in tetradecane into a single phase diagram, expressed in terms of x^ and the overall PS content.7 Returning to the phase diagram trajectory map (Figure 3.3), it is possible to represent increasing temperature by diagonal trajectories. In general, solvent partitioning is less selective at higher temperature, so the effective volume fraction tends to approach that of the neat block copolymer. In other words, /' —»/ with /=/PS = 0.31 for the case of the copolymer considered by Hanley et al.4
Concentrated Solutions
109
Figure 3.4 Phase behaviour in the (>, f) plane for PS-b-PI copolymers in DEP at the temperatures indicated.5 Morphologies are denoted: L, lamellae; C, cylinders; G, gyroid; S, spheres. The subscripts I and S indicate that the minority domain is formed by PI or PS, respectively. The shaded region indicates a glassy structure. Reproduced by permission of American Chemical Society.
110
Block Copolymers in Solution: Fundamentals and Applications
Figure 3.5 Phase behaviour in the (f, T) plane for PS-&-PI copolymers in DEP at the volume fractions indicated.5 Notation for morphologies as in Figure 3.2. Reproduced by permission of American Chemical Society.
Phase diagrams for other PS-&-PI diblocks in several of the di-n-alkyl phthalates were presented in a separate paper.5 In addition, other cuts through the threedimensional 'phase cube' (variables of temperature, copolymer composition and solution concentration) were discussed. Typical phase diagrams in the (>, /) and (T, f) planes are shown in Figures 3.4 and 3.5, respectively. In the (>, /) plane a comparison of the phase diagrams in Figure 3.4 shows that the ordered region increases with decreasing temperature, as expected. In addition, the phase boundaries move to the left because DEP becomes increasingly selective for PS, leading to an increase in its effective volume fraction. This can drive a transition from lamellae to a (normal) structure of PI cylinders (Q) on decreasing > for instance. This is also the origin for the tendency of the OOT phase boundaries to tilt towards the left. The isopleths in Figure 3.5 can be understood similarly. As the amount of solvent increases (from part a to e), the phase diagram deviates more from that for the melt. As 0 increases, the OOT lines move to lower fps and the lamellar and cylinder phases are eliminated, leaving only phases of spherical micelles at 0 = 0.3 and 0 = 0.2. The reduction in ODT follows the decrease in polymer concentration.
Concentrated Solutions
111
A highly selective solvent can even increase the ODT compared with the melt, as for example when squalane is added to PS-rich PS-fr-PI diblocks.7 The explanation is that squalane is such a strongly selective solvent that it is even more incompatible with PS than the PI block.
3.2 PHASE BEHAVIOUR OF PEO-CONTAINING BLOCK COPOLYMERS There has been substantial interest in the phase behaviour of PEO/PPO copolymers in aqueous solution, in particular focused on the Pluronic-type triblocks because of their commercial applications. Their phase behaviour has been studied by several groups as summarized in Table 3.1. It should be recalled when considering these results that the commercial samples have a broad block length distribution which Table 3.1 Studies on the phase behaviour of Pluronic-type copolymers in aqueous solutiion Pluronic
Composition
Solvent
L62 L64
E06P34E06 E013P03oE013
L92 L121 L122 P65 P84
E08P047E08 EO5PO7oEO5 EOnPOvoEOn E020P030E020 EOi9PO43EO19
P85
E026P039E026
Water Water + p-xylene Water Water Water Water Water Water + p-xylene p-xylene Water
P104
E018P058E018
P105
E037P058E037
P123
E020P7oE02o
F68 F127
E080P030E080 EO100P70EO100
25R8
P015E0156P015
Water Water + p-xylene Water Formamide Water + glycerol/propylene glycol/ethanol/glucose Water + butyl acetate/butanol Water + butanol Water Water Water + butyl acetate/butanol Water + butanol Water + p-xylene Water/propylene carbonate; Water/triacetin Water
Ref. 8 9 8 10 11 10, 11 11 12 13 14 (See also refs in Section 2.12 .6) 11 15 8 16 17, 18 11, 19 20 11 11 21, 19 20 21 22 23
112
Block Copolymers in Solution: Fundamentals and Applications
Figure 3.6 Phase diagram in water of EOm-b-POn-b-EOm Pluronics with n — 69 and m — 4 (L121), m = 11 (L122), m = 20 (P123) and m = 99 (F127).11 Reproduced by permission of American Chemical Society.
can vary depending on batch and/or manufacturer. This can clearly influence the observed phase behaviour since it will affect the preferred interfacial curvature. Wanka and coworkers determined binary phase diagrams for a total of twelve PEO-b-PPO-b-PEO copolymers.11 Phase diagrams for EOm-b-POn-b-EOm copolymers with the same PPO block length (n — 69) but with m ranging from 4 to 99 are shown in Figure 3.6. Also for comparison, similar diagrams for copolymers with constant but smaller PPO block lengths and varying PEO content are shown in Figures 3.7 and 3.8 (n = 30 and n = 27, respectively). The essential features of these phase behaviour studies on Pluronic copolymer solutions can be summarized as follows:".24 1. It appears that below a threshold value of molecular weight (~2000 g mol ]) no ordered phase forms. 2. The composition of the copolymer and its total molecular weight has a large influence on its phase behaviour. Wanka et al. concluded that the sequence of mesophases observed in the phase diagram depends largely on the PEO/PPO molar ratio (i.e. the m/n ratio).11 The greater the m/n value, the larger the number of possible mesophases formed. When m/n > 0.5, spherical micelles are formed for ocmc. A disorder-order transition occurs at higher concentrations, leading to the formation of cubic phases. When m/n is reduced to about 0.25, the hexagonal phase becomes the first ordered mesophase, whereas the lamellar phase appears as the first ordered phase for
Concentrated Solutions
113
Figure 3.7 Phase diagram in water of EOm-b-POn-b-EOm Pluronics with n = 30 and m = l l (PE6200), m = 1 3 (PE6400), m = 1 9 (P65) and m = 76 (F68).11 Reproduced by permission of American Chemical Society.
m/n « 0.15. In general, the larger the PPO block and the greater the PEO content, the greater is the gelling ability of the Pluronic copolymer. As the molar mass of the PPO block increases, the minimum copolymer concentration required for forming a gel decreases from 60% to about 20%. 3. The phase sequence is determined by the copolymer composition, which changes the micellar curvature. A geometrical interpretation of phase behaviour is usually employed for low molecular weight surfactants.25 Here the packing of molecules, and the associated interfacial curvature, governs the phase behaviour. However, low molecular weight surfactants are beyond the scope of this chapter and this approach is not detailed here. 4. The phase boundaries in Figures 3.6-3.8 reveal that thermoreversible transitions are possible in Pluronic surfactants at a fixed concentration. Thermally induced gelation of Pluronics at high concentrations is one of the characteristic properties of these systems, in addition to strongly temperature-dependent micellization. The phase behaviour of many Pluronic-type copolymers has been studied by Alexandridis and coworkers. Table 3.1 summarizes the relevant literature. They have studied both binary polymer/water mixtures and ternary polymer/water/ organic solvent mixtures. In general, 2H NMR on solutions in D2O was used to locate one- or two-phase regions in the phase diagram. SAXS was then used to
114
Block Copolymers in Solution: Fundamentals and Applications
Figure 3.8 Phase diagram in water of EOm-b-POn-b-EOm Pluronics with n = 27 and m = 5 (PF20), m = 12 (PF40), m = 73 (PF80).11 Reproduced by permission of American Chemical Society.
determine morphology. Structural parameters were obtained from the SAXS peak positions, together with the volume fraction of polymer. The results were interpreted in terms of the surfactant packing parameter which relates the structure of the self-assembled aggregate to molecular geometry.26 For several systems, additional SANS data provided detailed information on micelle structure from the form 97 98 "7R factor ' and intermicellar interactions from the structure factor. Multiple morphologies have been observed in phase diagrams for Pluronic-type copolymers, including normal and reverse micellar liquid phases, normal and reverse hexagonal-packed cylinder phases, normal and reverse micellar cubic phases and normal and reverse bicontinuous cubic phases and the lamellar phase. All nine types of structure were in fact observed for one ternary system-Pluronic P84 in water/p-xylene mixtures.12 The normal and reverse micellar cubic phases were shown to have space group symmetries Im3m and Fd3m, respectively. The
Concentrated Solutions
115
bicontinuous cubic structures were both of Ia3d symmetry, i.e. consistent with a gyroid structure. Other structures have also been observed - for example, a cubic micellar phase with Pm3n symmetry was observed for P105 in formamide.1 Alexandridis et al. established a number of general trends, as discussed in a good review.29 In binary polymer/water mixtures, the number of lyotropic phases observed increases with PEO content and polymer molar mass. Micellar cubic phases dominate the phase diagram of Pluronics such as F127 and F68 with 70 and 80% PEO, respectively. The phase diagrams of copolymers such as L62 and LI22 containing 20% PEO are dominated by regions of lamellar phase. The lamellar period and the block copolymer interfacial area decrease with increasing polymer concentration. Thermotropic phase transitions are observed - the thermal stability observed is in the order: cubic < hexagonal < lamellar. An increase in temperature causes phase boundaries to shift to lower concentration, i.e. the structures swell with water at high temperature. In ternary systems, an increase in PEO content similarly causes an increase in interfacial curvature which favours the formation of a variety of oil-in-water structures. The polarity of the solvent in ternary mixtures has a large influence on polymorphism - generally systems containing nonpolar organic solvents exhibit a richer polymorphism than that shown by systems containing polar solvents because the interfacial curvature is greater in the presence of nonpolar 'oils'. For several triblocks experimental phase behaviour was compared with that predicted by a mean field lattice model.10'30'31 Linse calculated phase diagrams for P105, P95 and P104.30 These calculated phase diagrams contain regions where a shape transition from spherical to rod-like micelles occurs at high temperature, then at still higher temperature, a two-phase region is found.The features are in qualitative agreement with the phase diagrams of P85 (for which the sphere-rod micelle transition has been studied in great detail - see Section 2.12.6). However, the P85 phase diagram contains a region of cubic micellar phase at high concentration, 14 not considered in the calculations by Linse et al. Semi-quantitative agreement with the phase diagram of LI 22 in water//?-xylene was noted for the case of the EO20-£-PO69-£-EO2o model.10 The theory was used to compute density profiles for the three components, and the dependence of the lamellar domain spacing on chain length. The scaling exponent (approximately 0.5) was in agreement with experiment, although the theory systematically underestimated the magnitude of the domain spacing. Similarly for a EO27-&-PO6i-/?-EO27 model, the phase diagram was qualitatively in agreement with that for the P104/water/ xylene system, at least for the normal phases.31 The model did not predict the location and extent of the reverse phases very accurately. The properties of the lamellar phase were a particular focus, and were in reasonable agreement with experiment. As an example of changes in binary phase diagrams for a series of Pluronics with approximately constant composition (25 wt% PEO), but varying molar mass, Figure 3.9 shows phase diagrams for Pluronics L62, L92 and L122.10 An interesting feature is the presence of both an inverse hexagonal (C2) phase and a normal
116
Block Copolymers in Solution: Fundamentals and Applications
Figure 3.9 Binary Pluronic/water phase diagrams for (a) L62, (b) L92 and (c) L122.10 Notation: isO] and iso2, micellar solutions; C], normal hexagonal phase (block copolymer cylinders); L, lamellar phase; C2, inverse hexagonal phase (water cylinders). Two phase coexistence regions exist between the single phase regions labelled. Reproduced by permission of American Chemical Society.
hexagonal (CO phase for L92 and LI22. The €2 phase is presumably stabilized by the longer hydrophobia PO block. A rich phase behaviour has also been noted for PB-b-PEO diblocks in aqueous solution.32'33 In addition to the classical phases- micellar liquid, bcc-packed spherical micelles, hexagonal-packed cylinders and lamellae- a bicontinuous sponge phase was also observed for a high molecular weight polymer. It was suggested that this might have been due to slow ordering kinetics, or due to the reduced interfacial area per chain which might lead to a decorrelation in the ordering of neighbouring domains. A novel aspect of these papers was the use of cross-linking (of the PB
Concentrated Solutions
117
block using 7-irradiation) to 'fix' lyotropic liquid crystal phase structures prior to TEM. Bates and coworkers have investigated self-assembly in this type of diblock in aqueous solution, as discussed in Section 2.16. Several research groups have also investigated the lyotropic polymorphism of PEO-b-PBO diblock and PEO-b-PBO-b-PEO triblock copolymers. This work is reviewed in detail elsewhere.34'35 Booth and coworkers have also investigated gelation of PEO block copolymers with o, L-lactide,36 and poly(styrene oxide).37'38 Gelation is discussed further in the following section. It has been shown that despite their star architecture, heteroarm star copolymers can form lyotropic liquid crystal structures.39 A hexagonal phase of cylinders was formed in solution in water/xylene mixtures of a heteroarm star comprising PS and PEO chains attached to a small divinylbenzene core. A segregation between the PS arms in the core and the PEO chains in the corona was suggested.
3.3 3.3.1
GELATION RHEOLOGY
The formation of gels by amphiphilic block copolymers in concentrated solution can be detected using tube inversion experiments. The term gel is used to mean a substance with a finite yield stress. It does not indicate an interconnected network formed by physical cross-links, as in conventional polymer gels. The terms 'hard' and 'soft' gels were introduced by Hvidt et al.40 to describe qualitatively the dynamic shear moduli of a concentrated solution of a Pluronic triblock copolymer. A more quantitative definition of 'hard' and 'soft' gels was introduced by Kelarakis et al.41 A solution with a yield stress exceeding cry = 40 Pa was observed not to flow out of an inverted tube, and was classified as a hard gel. The tube inversion method has been shown to give results in excellent agreement with those from rheology and DSC,42 and has been used to prepare gel phase diagrams for a number of systems.35 A hard gel has a storage modulus higher than loss modulus (G1 > G"} and high yield stress. Of the solutions which are mobile in the inverted tube test, those with zero yield stress and G" > G' can be classified as true sols. Between the extremes of hard gel and sol, solutions are found with a small but finite yield stress and G' > G", properties that are characteristic of a gel. Whether or not G' exceeds G" depends upon choice of frequency, and a fixed frequency of 1 Hz has been used by Booth and coworkers to obtain consistent results.35 In keeping with several reports on the rheology of aqueous micellar solutions of block copolymers,40^5 these fluids are termed soft gels. It must be stressed that this convenient division of block copolymer solutions into hard gel, soft gel and sol is based entirely on their rheology, involving arbitrary choices of yield stress and frequency, and is not uniquely related to structure. More rigorously, we consider first the linear viscoelasticity, then nonlinear viscoelastic effects such as yield stress.
118
Block Copolymers in Solution: Fundamentals and Applications
3.3.1.1 Linear Viscoelasticity Gels formed by cubic-packed micelles typically have a mechanical response dominated by elasticity. A representative example of the frequency dependence of the dynamic shear moduli from a micellar cubic phase is shown in Figure 3.10 Characteristically, G' is essentially independent of frequency whereas G" passes through a minimum. The shape of the curves can be described approximately by a Maxwell model at high frequencies and by a Voigt cell near the minimum in G".46 The transition from a mesophase to a micellar liquid leads to dramatic changes in the low frequency response, as the modulus evolves from the terminal response typical of the mesophase to that of a liquid (G' ~ a;2, G" ~ uS). As in block copolymer melts, composition fluctuations may have a pronounced effect on the rheological moduli well above the ODT. This has been shown in measurements on PS-b-PI diblocks in a PS-selective solvent, the fluctuation effect becoming larger and the onset occurring at higher temperature, as the diblock concentration increased.47 At lower concentrations, viscosity is the best rheological technique for probing the state of order in the system. The viscosity of PS-&-PI solutions in a selective solvent was shown to increase on passing from a 'gas-like' phase of
Figure 3.10 Frequency sweep of dynamic shear moduli for a 10 wt% gel of EO^-b-BOig at 20 °C. (o) G', (B)G". The strain amplitude A = 0.3%.77
Concentrated Solutions
119
isolated micelles to a 'liquid-like' phase of interacting micelles.48 On further increasing the temperature, the viscosity decreased due to the breakup of micelles. The structural information was provided by SAXS which showed the development of a maximum in the structure factor in the liquid-like phase, and the disappearance of both structure and form factor features at high temperatures where micellar dissolution occurred.48 A similar correlation between dynamic viscosity and the development of a structure factor peak was reported for a PS-£-PEP diblock in dodecane.49 The formation of a gel in a solution of a PS-b-PEB diblock has been studied Theologically.50 The gel structure was not determined, but based on the frequency dependence of the dynamic shear moduli appears to be a cubic phase. The time dependence of tan<5 = G'/G" measured at different frequencies indicated behaviour similar to that observed for physical gelation, specifically curves of tan<5 intersected each other at a single time.50 Such a frequency-independent tan<5 is a signal of gelation.51'52 The gelation process was also manifested in the frequency dependence of G' and G". At the gel point, G' and G" were both parallel and linear functions of frequency, at low frequency.50 The thermally induced transition from a liquid to a gel in aqueous solutions of Pluronic L64 (EO^-b-PO^Q-b-EO^) has also been analysed via dynamic mechanical shear measurements and SANS experiments. The results suggested that gelation occurs through a percolation transition.53 The theory of percolation applied to complex fluids predicts a universal scaling for structural and rheological parameters at the transition point. At the gel point, the scaling G' w G" « u;A is expected from theory and was observed in the experiments on L64 solutions. The value for the exponent A obtained from the frequency dependence of tan 6 was in excellent agreement with results from computer simulations for percolating systems in three dimensions, although the exponents for G' and G" individually were lower than the theoretical values, one explanation of which is the presence of an excess of effective cross-links.53 The neutron scattering data were analysed using Baxter's sticky hard sphere model for the intermicellar structure factor, although this model is unlikely to provide a unique fit to the data, as discussed by Castelletto et a/.54'55 Earlier, Booth and coworkers had proposed that the transition from micellar solution to a soft gel in aqueous solutions of EO41-b-BO8 occurred via a percolation transition, and that the soft gel was a dynamic network of weakly interacting spherical micelles with a fractal structure.42 It has also been suggested that the solsoft gel transition in other EOm-b-EOn diblock solutions occurs via a percolation transition.41 The sol-soft gel boundary had a very similar shape to that reported for the percolation transition in L64, 53 although the scaling of the dynamic shear moduli at the gel point was not examined. Based on the frequency dependence of the dynamic shear moduli in the linear regime (and dielectric relaxation data), two relaxation processes have been detected for PS-&-PI diblock micelles in a solution of an unentangled PI homopolymer.56 The features of the fast process were qualitatively similar to those for the relaxation of star polymer chains, suggesting that it corresponded to star-like retraction of
120
Block Copolymers in Solution: Fundamentals and Applications
PI blocks tethered to the PS cores of the micelles. This analogy establishes a connection to the tube theory for entangled star polymers. However, there is a difference in the relaxation mechanism for star-arm and corona blocks,56'57 because whereas a star arm can retract along a tube under a constraint on junction location only, the relaxation of the corona block is hindered by the micellar core that acts as an impenetrable barrier. Thus the relaxation is slower for a corona block than for a star arm having the same entanglement density. The relaxation time for the slow process depended on solution concentration. In concentrated solutions where the micelles were entangled through their PI coronas, the slow relaxation time rs was close to the Stokes-Einstein (SE) diffusion time, rSE, evalulated from the viscosity associated with the fast process. Thus, the slow process of the concentrated micelles was attributed to SE diffusion governed by the relaxation of individual PI blocks. On the other hand, for dilute solutions, rs was much shorter than TSE but close to the SE diffusion time evaluated from the PI solvent viscosity, suggesting that the slow process corresponded to SE diffusion in the pure solvent. 3.3.1.2 Nonlinear Viscoelasticity Over usual experimental timescales, block copolymer gels based on cubic micellar structures are characterized by the development of a yield stress, above which plastic flow occurs. Watanabe and coworkers were among the first to study the transition from a micellar solution to a gel with a finite yield stress, in solutions of a PS-b-PB diblock in tetradecane (a selective solvent for PB), via rheology and SAXS.58~61 Using SAXS, a cubic lattice was identified for gels exhibiting a yield stress. At high temperatures, melting of the ordered structure, i.e. an order-disorder transition, was detected from the SAXS data. For diblock concentrations less than about 10%, no structure factor peak was observed in the SAXS patterns and the flow behaviour was Newtonian. The absence of a structure factor peak suggests a 'gas-like' structure of noninteracting micelles (it was stated that the solution was above the cmc). Representative data showing the transition from a Newtonian liquid to a Bingham fluid are presented in Figure 3.11, which shows that the apparent yield stress increases with concentration.58'60'61 Considering transitions as a function of temperature, an intermediate phase was observed between the low temperature gel and high temperature Newtonian liquid for a 20% solution. This phase was characterized by linear viscoelastic properties, but zero yield stress. The corresponding SAXS pattern showed a single broad structure factor peak, indicative of a liquidlike structure of interacting micelles.59-61 In a companion paper to their work on the linear viscoelasticity of PS-b-PI diblock micelles in a low molecular weight PI homopolymer matrix,5 Watanabe and coworkers also investigated the nonlinear stress relaxation behaviour.62 As in the linear regime, a fast and a slow relaxation process were identified. Figure 3.12
Concentrated Solutions
121
Figure 3.11 Steady-state flow behaviour of solutions of a PS-b-PB diblock copolymer (Mn = 52000 g m o - 1 , 29.5 wt% PS) in tetradecane at various concentrations (wt%, indicated).58,60 Reproduced by permission of American Institute of Physics.
shows stress relaxation data obtained for different strain amplitudes. It is apparent that the relaxation times determined in the linear viscoelastic regime agree well with the two relaxation processes that can be seen in the stress relaxation data. Time-strain separability was examined for both processes. At long times the stress relaxation occurs only through the slow process, and G(?, 7) = Gs(t, j). By fitting a series of exponentials, it was possible to subtract the slow term by extrapolation to
122
Block Copolymers in Solution: Fundamentals and Applications
Figure 3.12 Stress relaxation modulus for different strains for a micelle of a PS-b-PI diblock in an unentangled PI homopolymer matrix.62 The arrows indicate the relaxation times Tf and rs for the fast and slow processes determined in the linear viscoelastic regime.56 The solid curve indicates the linear relaxation modulus obtained from the complex modulus G*. Reproduced by permission of American Chemical Society.
t = 0. Thus the stress relaxation of the fast process, Gf(t, j) was obtained. In this way, it was confirmed that time-strain separability holds for both fast and slow processes, i.e. at long time Gx(t, ^} = hx(^}G(t}, where x = f, s. For the fast process, the damping function h^} was found to be very similar to that for linear and/or star PS solutions, supporting the interpretation that the fast process was due to the motion of individual corona chains, as inferred from the linear viscoelastic data. For the slow process, the 7 dependence was much larger than for the fast process, and was also strongly dependent on solution concentration, unlike /Zf(7). The scaling of /zs(7) at large 7 was estimated from a simple argument based on changes in the translational entropy of micelles. This supports the conclusion from the linear viscoelasticity that the slow process for concentrated solutions depends on straininduced changes in the positions of micelles. For dilute micelles, no clear explanation was put forward for the observed weaker dependence of hs on 7. The nonlinear rheology of block copolymer solutions can also be measured under oscillatory shear, by analysis of the shape of the stress for an applied sinusoidal strain, 7 = 70 sin cut. In the linear regime, the stress is also a sinusoidal waveform,
Concentrated Solutions
123
although out of phase with the strain for a viscoelastic material. In the nonlinear regime, the nonsinusoidal stress can be written as a Fourier series:58,61,63
where the terms G'(l) and G"(l) with / 7^ 1 are nonzero in the nonlinear regime and G'(l = 1) and G"(l = 1) are the usual storage and loss moduli, respectively. The nonlinear rheology of solutions of a PS-b-PB diblock in tetradecane has been analysed using Equation (3.1).58 As expected, the magnitude of the higher order coefficients increased with strain amplitude, terms with / = 1, 3, 5 being resolved. For a material which does not suffer wall slip, the even terms in Equation (3.1) are absent.64'65 Watanabe et al. have also used this equation to extract the linear term (/ = 1) in order to study the evolution of its frequency dependence with strain for PS-&-PI-&-PS solutions in a number of Pi-selective solvents.63 This indicated, for example, a transition from plastic to nonlinear viscoelastic to linear viscous flow on increasing the temperature of a PS-fc-PI-&-PS solution in tetradecane. A Fourier analysis has been performed of the nonlinear flow behaviour of an aqueous solution of a PEO-6-PBO diblock forming a face-centred cubic (FCC) micellar phase.66 It was found that the magnitude of the higher order harmonics increased with strain amplitude, and up to the / = 81 harmonic was resolved at 2000% strain. The onset of nonlinear response as defined from the dependence of the isochronal dynamic shear moduli on strain amplitude was found to be in good agreement with that defined by the appearance of a higher harmonic in the stress waveform. The amplitude of the higher order harmonics was also compared with the predictions of two models of differing degrees of sophistication. The first was a simple asymptotic model for the amplitude of the harmonics, using a Carreau-type equation, that is an approximation to the flow in the limit of extreme shear thinning that occurs for very large amplitude strains.65 The second model accounted for the strain dependence of the stress response of a periodic (micellar) lattice modelled as connected MaxwellVoigt elements.46 As a qualitative alternative to analysis of the nonlinear response in terms of a Fourier series expansion of the stress, the flow behaviour can also be can also be represented by plotting stress against strain in a Lissajous figure.61'67 In the linear regime, an elliptical-shaped figure is obtained, nonlinearity being manifested by lozenge or more complicated shaped patterns.61 Careful rheology experiments suggest that in fact cubic micellar structures are neither gels nor yield stress fluids.68 Habas et al. suggest that flow will always occur although the terminal relaxation time can be very long. Nonetheless, the viscosity in the liquid regime can be very high (it shows a strong shear-thinning behaviour as shear rate is increased). They mapped out the dependence of the relaxation time on temperature and stress for a Tetronic (PEO-/?-PPO)4 four-arm star copolymer forming a BCC structure, and were able to probe relaxation times that already reached several hours. A divergence of the relaxation time with decreasing stress
124
Block Copolymers in Solution: Fundamentals and Applications
was indicated. A model based on the stochastic motion of micelles trapped by neighbouring micelles was developed, using the Eyring equation to describe the jump frequency over an activation barrier. This was able to describe the stress dependence of the terminal relaxation time very well. It was established that the motion corresponds to diffusion of caged micelles, not the diffusion of unimers or free micelles. It was also shown that the rheological properties are hardly affected by shear alignment (as noted earlier for a different system69'70) and that the terminal relaxation time is therefore not associated with the alignment of grains of the crystal. 3.3.2
STRUCTURE - PACKING OF MICELLES
The transition from a mobile fluid to a hard cubic gel occurs when the effective volume fraction of spheres in the system, 0, reaches a critical value.14'35'71 This is 4>c = 0.68-0.74 for packed cubic structures of spherical micelles, and >c = 0.52 for the corresponding primitive cubic structure, the latter corresponding approximately to the equilibrium condition for hard sphere crystallization. With an expansion factor, 6, defined as 6 = vs/va, where vs is the effective hard-sphere volume of a micelle in solution and va its anhydrous volume, the critical copolymer concentration for gelation, cgc (in g dm~3), is given by:
where pa is the density of 'dry' copolymer in g cm 3. If the hard sphere volume vs is equated with the thermodynamic volume vt (which is one-eighth of the excluded volume of one micelle for another) then 6 = 6t. Accordingly, for a given structure (given 0C), the determining factor for gelation at a given concentration is 6t, which in turn is largely determined by the exclusion behaviour of the PEO block corona. This depends sensitively on the structure of the micelle (including the association number) and the quality of the solvent. For large spherical micelles, <5t becomes smaller as the solvent is made poorer, e.g. by increasing the temperature or by adding salt. As a consequence, phase boundaries relate only indirectly (through <*>t) to copolymer composition, chain length and architecture. Careful characterization of the exclusion properties of the micelles by scattering techniques (e.g. static light scattering) is required if the hard gel boundary is to be predicted. From Equation (3.2), it is evident that a plot of critical gel concentration versus p&/6t will provide 0C from the slope. Data for aqueous solutions of PEO-b-PBO diblock copolymers plotted in this way (see Figure 3.13) yield (/>c ^ 0.70, i.e. bcc or fee packing of spherical micelles can be inferred on the basis of mobility and light scattering data alone, although SAXS or SANS are required to distinguish between bcc (0C = 0.68) and fee (0C = 0.74) structures. Figure 3.13 shows an interesting effect of block architecture in the PEO/PBO system: while the data points for the diblock PEO-&-PBO copolymers72"74 indicate cubic packing, those for triblock
Concentrated Solutions
125
Figure 3.13 Critical gel concentration for copolymer solutions at 40 °C versus pa/St, where pa is the density of anhydrous liquid copolymer and St is the thermodynamic expansion factor.35 The data points (taken from refs72~75) are for (•) PEO-fc-PBO and (O) PEO-6PBO-b-PEO copolymers. The lines are calculated from Equation (3.2) for bcc, fee and hex structures. Reproduced by permission of Royal Society of Chemistry.
Figure 3.14 Hard gel boundaries (determined by tube inversion experiments) for EO96-&BO18 (•), EOi84-£-BO18 (O), EO3i5-6-BO17 (D) and EO398-6-BOi9 (H).74 Reproduced by permission of American Chemical Society. PEO-b-PBO-b-PEO copolymers75 are consistent with hexagonal packing of cylindrical micelles (
126
Block Copolymers in Solution: Fundamentals and Applications
temperature fluid (so-called 'cold gelation') or on reducing temperature from a high temperature fluid ('hot gelation'). The terminology was introduced by Schmolka.76 The process of hot gelation is driven by the negative coefficient of solubility of PEO in water, i.e. as temperature is reduced the micellar corona expands and the excluded volume of one micelle for another increases, leading to gelation when the critical packing fraction is reached. The lower temperature sol-gel boundary is driven by the enhancement of micellar association with increasing temperature. The sol-gel transition boundary is observed to shift to lower concentration as the PEO block length increases, as also illustrated in Figure 3.14. The gel phase region is also observed to extend to higher temperature as the relative hydrophobe content increases.77
3.3.3 THERMODYNAMICS OF GELATION AND MICELLIZATION IN CONCENTRATED SOLUTION The standard enthalpy of gelation Agel//° may be obtained from the temperature dependence of the cgc, in much the same way as Amic//° is derived from the temperature dependence of the cmc, as discussed in Section 2.6. Expressions for AgeiG° and Agel//° analogous to Equations (2.18) and (2.19) can be found elsewhere.34 Compared with micellization, gelation has been shown to be almost athermal.35 The standard enthalpy of gelation of a 30 wt% solution of copolymer EO41-b-BO8, whether endothermic (cold gelation) or exothermic (hot gelation), is just a few hundred joules per mole of copolymer molecules,42 whereas the corresponding value for micellization in dilute solution is 78 kJ mol"1 of copolymer molecules.78 The large difference arises because micellization involves transfer of PBO blocks from aqueous solution to micelle core, with a large contribution from the hydrophobic effect, while gelation involves small changes in hydration caused by concentration of the PEO blocks in overlapping micelle coronas. The thermodynamics of micellization in concentrated solution differs from that in dilute solution.79 The lower gel boundary is ascribed to a unimer-micelle equilibrium even though in dilute solution it is generally agreed that micellization in dilute solution is complete at concentrations around 10 times the cmc. The thermodynamics of gelation has been explored in some detail based on measurements of the cmt of concentrated solutions using DSC.35'80 The method is more readily applied to copolymers with PPO hydrophobes, as PEO/PBO copolymers may show very small thermal effects at the cmt.81'82 The value of the standard enthalpy of micellization derived from data for EO93-b-PO44-£-EO93 (a fraction obtained from Pluronic F127), for example, is much reduced at moderate concentration: from 300 kJ mol"1 in dilute solution to less than 20 kJ mol"1 at 35 wt% copolymer.80 The explanation is that water-water hydrogen bonding is replaced by water-poly(ether) hydrogen bonding in the more concentrated solution, with consequent reduction in the hydrophobic effect associated (in this case) with the P block.
Concentrated Solutions 3.3.4
127
EFFECT OF ADDED HOMOPOLYMER, SALT OR SURFACTANT
The effect of salt on the micellization and gelation processes of Pluronic copolymers has been studied by a number of groups.43'83"85 Salts such as NaCl, KC1 and KF act as structure makers, leading to an increase in the self-hydration of water through hydrogen bonding and therefore a reduction in polymer solubility.24'43'84 On the other hand, salts such as KI, KSCN (and their sodium analogues), urea and ethyl alcohol, act as structure breakers, reducing self-hydration and increasing the hydration of the polymer. 24-43-84 Adding a 'salting out' agent such as KC1 shifts the gel region and cloud point to lower temperatures, whereas adding a 'salting in' agent moves the gel region and cloud point to higher temperatures.84'85 The effect of addition of K2SO4 on the upper hard gel boundary of a PEO-&-PBO diblock is evident in Figure 3.15.35'86 Addition of the salt has the opposite tendency to reduction in temperature, because the solubility of the PEO blocks is lower in the salt solution, and hence the corona is less swollen. Thus the upper sol-gel boundary is lowered. For cold gelation, the reduction in solubility of the copolymer with increasing temperature enhances micellization. On the other hand, addition of salt reduces the swelling of the micellar corona. These two 'salt effects' are in opposition, and to a greater or lesser extent, cancel out. Consequently, the lower sol-gel boundary is not greatly affected. The effects of salt on gel formation in P85 have been investigated via rheological measurements.85 In this system, a hard gel is formed at low temperatures depending on the concentration. On heating, concentrated solutions (e.g. 28.5% polymer in 1 M KF) melt from a hard gel into a sol, then at higher temperatures the soft gel hexagonal phase is formed. The temperature shifts for the sphere - cylindrical
Figure 3.15 Phase diagram from tube inversion for solutions of copolymer EO9o-£>-BOio in (O) water and (•) 0.2 mol dm~3 aqueous K2SO4.86 Reproduced by permission of Royal Society of Chemistry.
128
Block Copolymers in Solution: Fundamentals and Applications
micelle phase boundary and the soft gel boundary on addition of salt are found to be equal to the temperature shifts in the cloudpoints for the different systems. However, this is not the case for the hard gel. The cloud point and soft gel transitions at high temperature in Pluronic triblocks are usually attributed to changes in the PEO-block solvation only.30'87'88 The onset of micellization is also shifted by the addition of salt in the same direction as the cloud point but by a much smaller amount.89 This shows a weaker effect of salts on the dimensions of the PPO block compared with the PEO block, which explains the reduced influence on the hard gel transition temperature because hard gel formation is controlled by the volume fraction of effective hard spheres, which depends on the swelling of both blocks.85 The effect of addition of PEO and PPO homopolymers on the gel formation of Pluronic F127 in aqueous solution has been studied via rheology.90 The structure of neat F127 solutions in the concentration range 10-20% has been probed by SANS and rheology.91 Addition of PEO can reduce the gel region and/or eliminate it at sufficiently high PEO concentration. The amount of PEO required to 'melt' the gel depends on the copolymer concentration and decreases with increasing PEO molecular weight. Whilst PEO of low molecular weight hardly affects the gelation process and high molecular weight PEO leads to phase separation, an intermediate Mw (e.g. Mw = 6000) homopolymer increases the gelation concentration and decreases the sol-gel transition temperature. In contrast, low molecular weight PPO increases the stability region of the Pluronic F127 gel. However, higher molecular weight (Mw = 4000) PPO is too large to be solubilized.84 Salt-induced gelation of Pluronic L35 upon addition of CdCl2 has been investigated by SAXS.92 Increasing the CdCl2 concentration to 5 wt% led to the formation of an ordered array of nanocrystals within the gel matrix, due to coordination between the metal ions and the O atoms in the PEO units within the corona. As discussed in Section 2.15, the interaction between SDS and Pluronics has been extensively studied, and the relevant literature on self-assembly in dilute solution was discussed there. The ordered hexagonal and lamellar phases formed in the Pluronic L64/water phase at high copolymer concentrations were found to be replaced upon addition of SDS by an isotropic solution.93 At moderate copolymer concentrations, a bicontinuous isotropic solution was transformed into a solution of discrete micelles.93 Thus, addition of conventional ionic surfactants to nonionic polymer surfactants (e.g. Pluronics) can have a dramatic effect on phase behaviour. Because the ionic surfactant often acts as a cosurfactant it can change the packing density at the surfactant-water interface and hence induce transitions to phases with different interfacial curvature, or even lead to the breakup of micellar structures. Hecht et al. also studied the effect of SDS on the cubic gel phase formed at high concentration (c > 20 wt%) of F127.94 The gel phase is destabilized as the amount of added SDS increases, and eventually disappears. This is due to the suppression of F127 micelles.
Concentrated Solutions 3.3.5
129
INFLUENCE OF ARCHITECTURE
Telechelic chains associate into so-called flower micelles in which the midblock is looped so that the endblocks can form the core (Figure 3.16). Bridging increases with polymer concentration, leading to effective attractive interactions between block copolymer micelles. This leads ultimately to jamming of the micelles, i.e. to the formation of an extended network (Figure 3.16). The network has viscoelastic
Figure 3.16 Schematic showing association of telechelic chains into flower micelles (above a critical aggregation concentration) and ultimately an extended network, as concentration is increased.104 Reproduced by permission of American Chemical Society.
properties characteristic of a polymer gel. This has been investigated in particular for copolymers with a hydrophilic PEO midblock and hydrophobic end blocks, including reverse Pluronic 25R823 and PBO-£-PEO-£-PBO triblocks.95'97 A percolation transition has also been observed above a critical concentration for PEO end-capped with hexadecyl alkyl chains, as signalled by a strong increase in viscosity and high-frequency shear modulus.98 The terminal relaxation time is identified as the detachment time of the hydrophobic end groups from the micelles. The influence of end group length (and thus hydrophobicity) on the aggregation number and cac has been examined.99
130
Block Copolymers in Solution: Fundamentals and Applications
Dielectric spectroscopy can be used to probe the fraction of bridged versus looped chains in block copolymers with an appropriate longitudinal dipole moment. Watanabe et al. prepared a PS-b-PI-£>-PI-&-PS block copolymer by head-to-head coupling of two diblocks such that the dipole moments of the PI blocks in the resulting 'triblock' had inverted dipoles.100 Dielectric spectroscopy provided the fraction of looped chains in gels in the Pi-selective solvent n-tetradecane from the low frequency response, using an (unbridged) PS-&-PI diblock as a reference. The fraction of looped chains decreases with increasing copolymer concentration as bridging becomes favourable. The contributions to the modulus of the bridged chains, entangled loops and dangling loops were considered. Increasing the fraction of bridged chains leads to an increase in modulus. The relative contribution from loops and bridges was found to be similar, pointing to the important role of dangling loops which sustain the equilibrium elasticity due to the strong osmotic constraint on the PI block conformation. The influence of bridging chains on Theological behaviour has been investigated in detail. Tan et al. performed steady shear experiments on bcc gels formed by a PSb-PB-b-PS triblock in DBP, which is selective for the midblock.101 A critical shear rate for disruption of the ordered structure was associated with the relaxation rate of concentration fluctuations. They found that the time for recovery of the elastic moduli was relatively insensitive to the shear rate 7, in contrast to the behaviour of a PS-b-PB diblock gel examined previously.102 This was ascribed to the re-formation of bridges from the loops created under shear. The frequency dependence of the shear moduli of PS-£-PI-£-PS triblocks and a PS-&-PI diblock at the gel point in the PI-selective solvent n-tetradecane was found to be described by power laws at high frequency, although an additional slow relaxation mode was observed at low frequency. The absence of a power-law scaling at low a> was ascribed to the disappearance of a self-similar structure on large length scales due to composition fluctuations. A sol-gel transition has been observed for telechelic polymers with strongly hydrophobic endgroups, for example in PEG end-capped with fluoroalkyl segments provided the fluoroalkyl end groups are not too long.103 Copolymers with lengthy PEG midblocks only exhibited single phase behaviour as with other associative thickeners. When the ratio of PEG and fluoroalkyl segments was nearly balanced, sol-gel coexistence was observed. Copolymers with very long fluoroalkyl segments were insoluble. For the copolymers undergoing a sol-gel transition, it was found that the swelling ratio and modulus of the gel phase are determined by the PEG midblock length. On the other hand, the relaxation time is controlled mainly by the length of the hydrophobe. The erosion of gels (in a stream of water) was studied and found to be dominated by the size of the hydrophobe. The viscosity and equilibrium sol concentration depend on the length both of the PEG midblock and of the hydrophobe. The rheology of PEG end-capped with perfluorinated alkyl groups has been shown to depend dramatically on endgroup length - the viscosity was much higher for chains with C8F17 hydrophobes than C6F13 hydrophobes.104 The association of telechelic PEG with partially fluorinated end caps, F(CF2)8(CH2)n,
Concentrated Solutions
131
has been investigated via SANS and rheology, and compared with that of the corresponding diblock with half the molar mass.105 As expected, the viscosity enhancement with increasing concentration is much greater for the triblock system in which bridging occurs. SANS provided information on the association number and dimensions. The temperature dependence of the relaxation time obtained from stress relaxation measurements followed Arrhenius behaviour, with an associated activation energy Ea = 43 - 49k&T, about twice that of telechelics with hydrocarbon endcaps, for which values ranged from Ea = 17 - 28 kBT going from a C12 to a C^ hydrophobe.106 19F NMR has also been used to probe the aggregation of telechelics comprising PEG with perfluorinated hexyl or octyl hydrophobe units.107 Bridging at high concentration leads ultimately to the formation of a network structure. It is now established that the modulus of gels formed by ABA triblocks can be reduced by blending with the corresponding AB diblock,97'108 which obviously reduce the extent of bridging. The effect has been confirmed both for PEO-based97 and styrenic108 block copolymer systems. Grafting of poly(acrylic acid) onto PEO-b-PPO-b-PEO Pluronic triblocks has been shown to lead to thermoreversible gelation in aqueous solution.109 The PA A forms water soluble 'cross-links' between the hydrophobic PPO domains, as shown schematically in Figure 3.17. The rheology of these associating polymers has been studied in detail.109'110 Microgels formed by loose cross-linking of PAA (using a divinyl species) in the presence of Pluronic copolymers have been shown to be thermoresponsive due to reversible aggregation of PPO chains in certain temperature intervals.111 Pluronic L92 was used as a model 'hydrophobic' copolymer and Pluronic F127 as a model 'hydrophilic' copolymer. A fractal structure of crosslinked clusters was revealed by SANS for the former system, whereas micelles were observed in the latter case. The swelling behaviour and kinetics of these materials was examined as a function of pH and temperature.112 The drug absorbing properties of microgels were examined using model weakly basic molecules, and found to be correlated to the PPO content.112'113 An ion exchange mechanism was invoked on the basis of comparisons of the loading of hydrophobic and hydrophilic drugs. Protein loading was also investigated and found to be related to the pore size in the gels.113
Figure 3.17 Highly schematic illustration of amphiphilic polymer network formed by cross-linking linear ABA triblock copolymers.116 Reproduced by permission of Elsevier.
132
Block Copolymers in Solution: Fundamentals and Applications
Patrickios and coworkers have employed the concept of cross-linking of block copolymers to form model networks, using PEGMA as a cross-linker of ABA triblocks in which the hydrophilic blocks are hydrophilic cationic PDMA and the hydrophobic block is PMMA, 114'115 PBMA, 116 or PLMA.116 PDMA-6-PMAA double hydrophilic copolymers were also studied.117 Due to the polylectrolyte character of these networks, the swelling response of the hydrogels is found to be pH dependent, increasing with the degree of ionization of the PDMA block at low pH.115 For the double hydrophilic copolymers, a minimum in the degree of swelling is observed at around the isoelectric point.117 At low pH, the high degree of swelling is due to ionization of PDMA, whilst at high pH, the swelling is due to ionization of the other block, e.g. PMAA in the double hydrophilic copolymers. The concept has been extended to cross-link PDMA-b-PMMA or PEGMA-b-PMAA heteroarm star copolymers, as shown in Figure 3.18.116'118'119 and more complex heteroarm starblock copolymers such as those in which the block sequence in the arms is reversed from one arm to another, or where homopolymer arms are mixed with multiblock arms.118'119 The cross-linking leaves dangling chains of one block, whereas the others are elastically linked into the network. In many of these papers, comparison was made with statistical copolymer and homopolymer networks. Patrickios has also reviewed other methods, not involving cross-linking of block copolymers, to form amphiphilic copolymer networks.116
Figure 3.18 Highly schematic illustration of amphiphilic polymer network formed by cross-linking mixed arm star copolymers.116 Reproduced by permission of Elsevier.
3.4 ORDER-DISORDER PHASE TRANSITION The ODT in block copolymer solutions refers to the transition from an ordered meosphase to a disordered micellar solution (demicellization occurs separately at higher temperatures, although it is not a true phase transition).120 It can be located using a number of methods, of which the most widely used are small-angle
Concentrated Solutions
133
scattering and rheology experiments. Other techniques are employed such as measurement of birefringence. Isotropic phases are characterized by zero birefringence, but ordered lamellar or hexagonal phases have finite birefringence. The scaling of the critical polymer volume fraction for the ordered lamellar disordered phase transition concentration, >ODT> with chain length, was studied via birefringence experiments for a series of nearly symmetric PS-b-PI copolymers in toluene, DOP and in the melt.121 A scaling relationship ^ODT ~ (X^O °626 was obtained, with an exponent close to the theoretical value [Equation (3.11)]. This is illustrated in Figure 3.19 which shows data at a fixed temperature. The scaling was
Figure 3.19 Volume fraction at the ODT versus xN f°r a range of nearly symmetric PS-bPI diblocks in toluene at approximately 36 °C.121 Reproduced by permission of John Wiley & Sons, Inc.
obtained from the experimental data by determination of x from a straight line plot of F/2N(f)1'6 versus l/TODr [Equation (3.3)]. However, in contrast to the theoretical predictions, the scaling ^QDT ~ N~°'62 was observed for copolymer concentrations spanning the concentrated regime (up to the melt limit) as well as the semidilute region. Thus no region of validity of the 'dilution approximation' (which predicts 0oDT~^~') expected for concentrated solutions, was observed. Allowance for composition fluctuations would add a correction term proportional to (7V01 3)~°'33 to Equation (3.11), but this would not modify the scaling of >ODT with TV appreciably.121 For a different system, PEP-b-PEE diblocks in squalene, a scaling 0ooT~ N™ 0 ' 81 was found, intermediate between the scalings for concentrated and semidilute solutions.121 This suggests that the scaling may be nonuniversal.
134
Block Copolymers in Solution: Fundamentals and Applications
The fact that PEP-b-PEE solutions were closer to the dilution approximation predictions may be due to the fact that squalene is a nearly athermal solvent, whereas toluene is not a completely athermal solvent for PS-b-PI. Other factors that could account for the observed differences between the two systems include the difference in x, which is much smaller for PEP-b-PEE than PS-£-PI, thus requiring considerably larger values of N for PEP-fe-PEE to achieve the same degree of segregation. Thus weaker chain stretching and diminished composition fluctuations may be anticipated for the PEP-&-PEE solutions. Finally, excluded volume effects may be important, these decreasing with molar volume of the solvent, squalene being a relatively large diluent.121 The ODT determined for a series of PS-b-PI diblocks in DOP was used to compute the scaling of XODT with polymer volume fraction.6 For diblocks with compositions ranging from/PS = 0.15 - 0.76, the scaling XQDT ~ <j>~a with a= 1.31.6 was observed. This conflicts with the predictions of the dilution approximation (a = 1), as for the scaling of >ODT mentioned above. However, for cylinder-sphere, gyroid-cylinder and lamellar-gyroid order-order phase transitions XOOT scaled as >~ LO , in agreement with the dilution approximation.4'6 The ODT can also be located via small-angle scattering experiments - as for block copolymer melts34 discontinuities in peak width and intensity are observed at the ODT. There is also evidence for a small shift in peak position q* in one system (PS-&-PI in DBP).122 For block copolymer solutions in a nonselective good solvent the inverse structure factor is given by:123'124
Here ¥(q) is a function of radius of gyration and composition of the block copolymer. The effective x parameter in semidilute solution is defined by: Xe&N = XABN
Concentrated Solutions
135
frequency dynamic shear moduli are much higher in an ordered mesophase than in the disordered phase, and there is a discontinuous decrease in modulus on crossing an ODT. The transition from micellar solution to gel (and between hard and soft gel) in solutions of PEO-b-PPO-b-PEO triblocks40'43'53'85 and PEO/PBO copolymers41'42'128 has also been located via the temperature dependence of the isochronal dynamic shear moduli, although the transition between sol and gel is more simply determined from tube inversion tests. Measurements of the shear-rate dependent viscosity are also useful in locating this transition.49'91 The effect of composition fluctuations on the Theological response was investigated in detail for three symmetric PS-&-PI diblocks in the neutral solvent dioctyl phthalate.47 The concentration range extended from disordered, dilute solutions to the (lamellar) ordered state, with the emphasis on the intermediate regime, where large amplitude composition fluctuations were clearly evident. The dynamic shear moduli and apparent relaxation times, all normalized to their respective values in the absence of fluctuations, increased markedly as the ODT was approached by increasing concentration. This effect was most evident when the rheological data were plotted as phase angle versus frequency (reduced with respect to the longest relaxation time in the system, r^. The rate of increase of dynamic shear moduli and relaxation times was found to be consistent with theoretical predictions by Fredrickson and coworkers,129'130 although the magnitude of the effect was larger than anticipated by theory, as for block copolymer melts.34 It was pointed out that the onset of measurable fluctuation effects depends on molecular weight in a manner that suggests that entanglements play an important role in coupling fluctuations to the viscoleastic properties.47 This is consistent with earlier measurements of translational diffusion in a lamellar block copolymer melt.131 Dynamic light scattering has also been used to locate the ODT in a PS-&-PI solution in a neutral solvent (toluene). Two distinct modes were observed in the scattered intensity autocorrelation function.132 The faster one was diffusive and decreased in rate approximately exponentially with increasing concentration. The magnitude of the diffusivity suggested that it reflects the translational diffusion of the block copolymers. The slower mode was attributed to cooperative rearrangements of microdomains. This slower mode was also found to undergo a sharp decrease in rate over a narrow range of concentration, corresponding to the ODT.
3.5 3.5.1
ORDER-ORDER PHASE TRANSITION STRUCTURAL ASPECTS
An intriguing aspect of the lyotropic polymorphism of block copolymers in selective solvents is the observation of distinct cubic packings of spherical micelles. McConnell et al. studied the packing of PS-&-PI micelles in decane via SAXS and SANS.133"135 Their phase diagram is reproduced in Figure 3.20. It is parameterized in terms of the relative size of micellar corona and core, LIRC where L is the corona
136
Block Copolymers in Solution: Fundamentals and Applications
Figure 3.20 Phase diagram obtained for a series of PS-b-PI diblocks with different composition in which 0PS, the volume fraction of the PS micellar core is plotted against the ratio of corona layer thickness, L to micellar core radius, /?c.133'135 Regions of liquidlike disorder (o) and crystalline order (•) are indicated. At higher concentrations a liquidlike state was observed (*) above a melting transition, estimated to occur at the points indicated as (+). At still higher concentrations, a transition to an unidentified anisotropic structure was noted (n)- Reproduced by permission of American Chemical Society.
thickness and Rc is the core radius. Crew-cut micelles with short coronal chains behave as hard spheres, which pack into an fee phase at high concentration. In contrast, hairy micelles with thick coronas behave as soft spheres and pack into a bcc phase. A noteworthy feature of this phase diagram is the fcc-bcc phase transition line, which is independent of PS volume fraction. Self-consistent mean-field theory has been used to analyse SANS data obtained for PS-&-PI diblocks in decane.134 The theory was used to calculate intermicellar pair potentials and combined with liquid-state theory (specifically, the RogersYoung closure to the Ornstein-Zernike equation) this enabled a comparison with the measured structure factor, S(q) (Figure 3.21). Good agreement was obtained between the experimental and theoretical S(q), comparisons being made for a range of core volume fractions for a system with short-ranged interactions (which lead to a fee lattice) and one with longer-range interactions (which lead to a bcc lattice). McConnell and Gast later extended this work by using the homogeneous liquidstate theory as a basis for density functional theory (DFT) for an ordered lattice.136 A modified weighted density approximation was used to estimate the free energy of each solid structure and to predict the liquid-solid phase transition. The DFT theory only predicted a simple liquid-fee transition, although the experimental phase diagram for PS-&-PI diblocks in decane indicates transitions from the liquid state to fee or bcc phases (Figure 3.20). Despite the failure of the theory to predict a bcc
Concentrated Solutions
137
Figure 3.21 Structure factors versus wavevector for dPS-b-Pl diblocks in core-contrast matched decane solutions.134'275 (a) dS393-£-I2o6 at core volume fractions of 0.012 (A), 0.02(+), 0.03(*), 0.04 (A) and 0.05 (o). (b) dS4O2-b-l422 at core volume fractions of 0.006 (o), 0.013 (n) and 0.019 (A). The lines are theoretical fits from the self-consistent field interaction potentials and the Rogers-Young closure to the Ornstein-Zernike equation. Reproduced by permission of Royal Society of Chemistry.
phase, it does provide a reasonable estimate of coexistence curves between the liquid and solid.136 Transitions between fee and bcc structures have been observed for PEO-^-PBO diblocks in the highly selective solvent water as a function of both concentration137 and temperature.70'138 The concentration dependence of the isothermal phase transition boundary contrasts with the phase diagram of McConnell et al. for
138
Block Copolymers in Solution: Fundamentals and Applications
PS-/J-PI diblocks in decane (Figure 3.20) in that a transition from fee to bee can occur simply by changing concentration. The observation of distinct micellar packings for different copolymer concentrations (at constant temperature) was ascribed to increasing interpenetration of micellar coronas with increasing concentration. The transition from bcc to fee observed for PEO40-^-PBO10 upon increasing temperature has been ascribed to the thermally induced change in solvent quality for the E-block corona.138'139 At low temperature, the corona is well solvated. As temperature is increased the 0 temperature of PEO is approached and the corona block contracts, whilst the core radius increases (leading to an increase in association number).71 The contraction of the corona compared with the core leads to more short-range intermicellar interactions and hence a fee phase.139 Lodge and coworkers have shown that changes in the cubic packing of micelles can be induced by varying solvent quality. The phase diagrams in Figure 3.2 show a bcc phase of PI spheres in the phase diagram of PS-&-PI diblocks in the weakly selective solvent DBP, and an fee arrangement in the strongly selective solvent DEP. In addition, a re-entrant ODT is observed for solutions in the latter solvent near (f> = 0.2, where on heating the sequence micellar liquid - ordered cubic phase micellar liquid is observed. The change in cubic packing was rationalized on the basis of differences in the solvent selectivity for the PI core, since the PS chains are well solvated in either solvent. However, the micellar hydrodynamic radius was the same in both solvents, and large differences in the ratio of coronal layer thickness to core radius are unlikely. Therefore, the system appears not to follow the phase diagram of PS-b-PI diblocks in decane established by Gast and coworkers.133'135 A thermoreversible fcc-bcc transition observed for PS-b-PI diblocks in solvents selective for either block was shown to be driven by a decrease in aggregation number as the solvent penetrates the core.140 No discontinuities in q were observed. The phase diagram for the system was in excellent agreement with simulations for star polymer solutions,141 replacing the number of arms with the micellar association number. A thermoreversible transition between fee/hexagonal close-packed (hep) and bcc phases for symmetric PS-b-PI diblocks in PS-selective di-n-alkyl phthalate solvents was examined in detail.142'143 A transition from a low temperature fcc/hcp to a high temperature bcc phase was driven by PI core swelling as solvent selectivity decreased. The mechanism of the transition was studied in detail by preparing shear-aligned samples. Samples were shear aligned in two different cells so that information in two orthogonal planes could be obtained.143 The hep - bcc and fee bcc transitions were found to follow epitaxial transitions: the Burgers mechanism143 and a modified Bain distortion,142'143 respectively, as also observed for atomic metals. In the Bain distortion, a distorted bcc unit cell embedded in an fee unit cell transforms to bcc by compressing along one <100> direction and expanding along two < 110> directions of the fee unit cell. There is three-fold degeneracy associated with this. Bang and Lodge showed that to account for the observed SAXS patterns, the pure Bain distortion had to be modified by allowing infinitessimal axial
Concentrated Solutions
139
distortions and rotations (alternatively viewed as slippage of close-packed {111} planes). The orientational relationships between the fee and bcc lattices followed the Kurdjumov-Sachs and Nishiyama-Wassermann relationships observed for metallic systems.142'143 The Burgers mechanism for the martensitic hep to bcc transformation involves slippage of alternating hep planes, followed by a slight lattice distortion.143 As with the modified Bain distortion, there is an associated three-fold degeneracy. The orientational relationships in this case are those of Burgers and Pitsche-Schraders. Nine orientations of the bcc lattice with respect to the hep lattice arise, as for fee. The observation of the bcc phase at high temperature just below the ODT is significant in view of the Alexander-McTague conjecture144 that a bcc phase is preferred close to the melting line for weakly first order phase transitions, for entropic reasons. The fcc-bcc transition was also observed for Piselective solvents.143 An interesting mechanism for driving transitions between different micellar cubic structures is to prepare distributions of mixed micelles. This has hardly been explored as yet. Replacing Pluronic F127 (EO99-b-EO6g-b-EOgg) with EO45-bBO14-b-EO45 in micellar gels at a fixed total copolymer concentration has been shown to drive a transition from an fee structure to bcc (Pluronic F127 itself forms an fee structure at low concentration in aqueous solutions and a bcc structure at higher concentration).145 Addition of water to an acetone solution of a PHOVE-b-PMOVE diblock was shown to lead to the formation of a bcc gel.146 Both blocks are soluble in acetone but only the latter PMOVE block is soluble in water. The transition was described as a water-induced phase transition. The mechanism was ascribed to a spinodal-type process whereby the ordered phase develops continuously without an initial micelle formation step. Consistent with this, the characteristic spacing of micelles did not vary during lattice formation which was driven by an increase in x- This is in contrast to gelation induced by temperature or change in concentration, where micelles first form and then pack onto a lattice. 3.5.2
ORDERING KINETICS
The transition between hexagonal-packed cylinder and gyroid structures has been probed via SAXS and rheometry for a PS-b-Pl diblock in DBP.147'148 For shallow quenches from the high temperature cylinder phase into the gyroid phase, the transition proceeds via a nucleation and growth processes. For deeper quenches, an intermediate metastable structure was observed, identified as hexagonal perforated lamellae (hpl). The transition was examined for a specimen with a shear-aligned cylinder phase and occurred epitaxially. The pathway and kinetics of the transition was found to be approximately the same whether the sample was shear aligned or not. The results were compared with prior work on this transition in block copolymer melts. Subsequently, the kinetics of grain growth of the gyroid phase in solutions of PS-b-PI diblocks in di-n-alkyl phthalate solutions were investigated
140
Block Copolymers in Solution: Fundamentals and Applications
by polarized optical microscopy.149 It was possible to monitor the growth and subsequent transformation of birefringent domains of a transient hpl phase during a quench from the high temperature disordered phase into the (nonbirefringent) gyroid phase. The transition from hpl to gyroid phase was shown to occur via classical nucleation. Grain growth rates were compared with the theory of Goveas and Milner150 for front propagation at an order-order transition in a weakly segregated block copolymer microphase, and semiquantitative agreement was found. The ordering kinetics of a cylinder phase (from the disordered state) have been analysed similarly.151 The development of both ellipsoidal grains and spherulites was observed. The Goveas-Milner model could account for the growth front velocity. The growth in the fraction of ordered birefringent regions could be fit to an Avrami equation, the exponent n = 3 being consistent with heterogeneous nucleation and three-dimensional growth. The ordering kinetics of a PS-£-PEB-£-PS triblock in mineral oil, a selective solvent for the PEB middle block, have been studied by SAXS.152 A two-stage ordering process of micelles from the disordered phase into a bcc lattice was noted. The first corresponds to temperature equilibration and supercooling of the micellar fluid and the second to the nucleation and growth of the ordered phase. The competition between the increase with quench depth of the thermodynamic driving force for ordering and the decrease in mobility (due to vitrification of PS) lead to a minimum in the induction time for the second ordering stage. The process of electric-field induced alignment of a lamellar phase formed by a PS-&-PI diblock in toluene has been investigated by time-resolved SAXS.153'154 Lamellae were initially oriented parallel to electrode surfaces (due to preferential interactions and/or flow). Electric fields caused reorientation of into a perpendicular configuration. Two dominant mechanisms were identified - grain boundary rotation and rotation of entire grains. The former dominates in weakly segregated systems, and the latter in strongly segregated systems.153'156
3.6 DOMAIN SPACING SCALING, AND SOLVENT DISTRIBUTION PROFILES The scaling of domain spacing is generally described by a power law in volume fraction, d ~ (f)~P. The exponent /3 depends on solvent selectivity (varying with both temperature and the type of solvent). Usually, d increases upon addition of a selective solvent. Addition of a nonselective solvent decreases d as the segregation between blocks is reduced. However, the swelling also depends on the type of ordered phase, for ideal swelling /3 decreasing in the sequence: infinite lamellae /3=l; infinite cylinders (3 = A; spheres /3 = L based on dimensionality arguments. 1 /^ For a selective solvent, a scaling relation for the domain spacing d~(l/T) was 157 obtained from SAXS experiments on a PS-b-PB diblock in tetradecane. At low concentration, d was found to scale as d~^>~ 1/3 , while the micellar radius, R (obtained from the position of form factor oscillations), was found to be approx-
Concentrated Solutions
141
imately independent of temperature. At higher concentrations, d was found to depend less strongly on concentration and R increased slowly.157 A systematic study of the domain spacing scaling in two nearly symmetric PS-b-PI diblocks in neutral solvents was undertaken via SAXS experiments on these polymers dissolved in toluene and dioctyl phthalate (DOP).125 The results were summarized in a scaling relationship for the domain spacing in the ordered phase in the semidilute and concentrated regimes (0.15 < 0 < 0.6):
The scaling with TV is the same as that for strongly segregated block copolymer melts. The scaling with temperature is the same as that observed for PS-&-PB copolymers in a selective solvent,59 while the concentration dependence is reversed, i.e. d decreases with polymer volume fraction $ for selective solvents, but increases for neutral ones. Since the effective x parameter scales as x ~ (0/7)Af~1/2, 158 this equation can be rewritten:
a scaling consistent with data for a series of PS-&-PI diblocks in DOP.15 This scaling behaviour was reproduced by the self-consistent mean field theory calculations of Whitmore and Noolandi, for the specific case of PS-^-PI diblocks in the neutral solvent toluene.159 However, these authors did note a decrease in the scaling exponent on moving from a weakly segregated system to a more strongly segregated one, and also noted an additional scaling with 0 (which arises because X is strictly not inversely proportional to T, since in general in x— AIT + B, with B / 0). Lodge et al. also reported a scaling d ~ 01/3 for a series of PS-&-PI diblocks in DOP, and a relationship d~ %0'25.6 The latter is in agreement within uncertainly with the scaling reported by Mori et a/.158 Hashimoto et al. have also observed that in the disordered phase, the domain spacing is independent of temperature and polymer concentration, i.e. the scaling:
is obtained.125'127'158 This is equivalent to:
This scaling was also predicted by SCMFT calculations by Whitmore and Noolandi.159 The scaling [Equation (3.4)] was obtained for the lamellar phase formed by a symmetric PS-fc-PI diblock in toluene, where the concentration was increased continuously by solvent evaporation.59 However, above 0 = 0.7 it was suggested that microphase separation became kinetically controlled rather than thermodynamically
142
Block Copolymers in Solution: Fundamentals and Applications
limited due to the lower rate of solvent loss, leading to a deviation from the d ~ 01/3 relationship.59'60 Nonequilibrium effects for a symmetric and an asymmetric PS-bPI diblock and an asymmetric PS-b-PB polymer in DOP were investigated in detail using SAXS.160 It was proposed that in the kinetically controlled regime, the interfacial area per block at the A-B interface cannot maintain the equilibrium value but is fixed at an approximately constant value. Then the domain spacing decreases with increasing 0 and decreasing temperature due to deswelling and thermal contraction, respectively. It was found that the crossover to nonequilibrium behaviour occurs at a lower concentration for the spherical micellar phase than the lamellar phase, which was explained on the basis of distinct thermodynamic barriers for growth of different ordered phases.160 These results were later extended, following further SAXS experiments on a series of PS-b-PI diblocks in toluene.161 For spheres, nonequilibrium effects were proposed to result from 'suppressed mutual diffusivity' of the block chains between spheres with increasing 0. For other morphologies, nonequilibrium effects may result from 'grain boundary effects' and/or vitrification of one component. These effects illustrate the problems observed with kinetically trapped nonequilibrium structures in concentrated block copolymer solutions. Using SAXS, the dependence of domain spacing on copolymer volume fraction has been determined for a PS-&-PI diblock and a Pl-b-PS-b-Pl triblock in the slightly selective solvent dibutyl phthalate.48 At high concentrations it was shown that d scales as d~ 0~1/3 showing three-dimensional shrinkage of micelles of finite length (in this case elongated micelles122). This concentration dependence is the same as that obtained earlier for a PS-£-PB diblock in a selective solvent,157 although as mentioned above it differs from the results from the same group for diblocks in neutral solvents.125 The swelling of semidilute solutions of a PS-£>-PMMA or a PMMA-b-dPMMA diblock in the neutral solvent toluene was investigated using SANS.162 The peak position in the disordered phase was found to scale as q ~ 00'05. The data do not agree with the scaling for the structure factor peak position [Equation (3.12)]. The results also differ from those of Hashimoto et al. for PS-&-PI copolymers in toluene, 125 which as mentioned above showed that q* was independent of concentration in the disordered phase. The swelling of PS-b-PI diblocks in selective and nonselective solvents has also been investigated through SAXS measurements of the domain spacing in ordered structures at high polymer volume fraction.163 The nonselective solvent 1, 3, 5-triisopropylbenzene was used, as well as three solvents of varying selectivity for PI. The scaling of domain spacing was described as a power law in volume fraction, with an exponent (3 that increases with solvent selectivity (varying with both temperature and the type of solvent). The distribution of solvent within diblock block copolymer microdomains has been assessed theoretically for block copolymers in neutral solvents. A neutral solvent is expected to be distributed nearly evenly between the two domains, with a slight excess at the interface to screen unfavourable interactions between the
Concentrated Solutions
143
components. For a selective solvent, selective partitioning is expected. Lodge et al. have investigated the partitioning of solvent via analysis of the SANS intensity of reflections from a lamellar structure formed by three PS-b-PI diblocks in toluene.164 Normal and deuterated solvent provided contrast to probe solvent partitioning, which was also analysed using self-consistent field theory. The method was used to investigate the selectivity of three solvents for a PS-b-PI diblock.165 In this way, it was shown that toluene is not a completely neutral solvent for this system - it is slightly selective towards PI. Benzene and THF are better solvents for PS. The segregation of homopolymer to the interface in blends with diblocks has also been thoroughly investigated, as discussed elsewhere.34
3.7
SEMIDILUTE BLOCK COPOLYMER SOLUTION THEORY
The phase behaviour of block copolymers in solution in a neutral solvent can be described in terms of a vertical trajectory through the phase diagram, i.e. a renormalization of xN (Section 3.1) This is the 'dilution approximation', introduced by Helfand and Tagami.166 It is assumed that for a neutral solvent (XAS ~ XBS where these interaction parameters refer to block A and B, respectively, in solvent S) a solution with a volume fraction 0 of copolymer behaves equivalently to a neat diblock melt if x is replaced by an effective value Xeff = <^XAB- The approximation relies on the assumption that the solvent is distributed uniformly throughout the block copolymer microdomains. Naughton and Matsen determined regions of validity for this approximation using self-consistent field theory and gave approximate conditions for the solvent quality and the relative variation of solvent concentration and block copolymer segregation.167 The development of theory for concentrated block copolymer solutions in a neutral good solvent dates back to work done in the iQgQs.123'124'159'168 Hong and Noolandi applied self-consistent field theory to block copolymers in a nonselective solvent.168 They also presented phase diagrams for neutral good solvents. For block copolymer concentrations ~ 0.6, in a poor neutral solvent, they found that the ODT between homogenous and microphase-separated phases can be described using the dilution approximation:166
where, for a symmetric diblock the structure factor function F(f)/2 = 10.5 is predicted using Leibler's mean field theory.169 In the dilution approximation it is assumed that the primary role of added solvent is to reduce the number of unfavourable monomer-monomer contacts in a spatially uniform manner. Fredrickson and Leibler provide a justification for Equation (3.8) for neutral good solvents, whilst pointing out that even as 0 —>• 1, the solvent will not be homogeneously distributed, rather it will segregate in the interfacial regions to
144
Block Copolymers in Solution: Fundamentals and Applications 1 9^
screen monomer-monomer interactions. It was observed that inhomogeneities in solvent concentration exist whenever the solution is microphase separated,123 a result not obtained by Hong and Noolandi.168 This discrepancy was ascribed to an incorrect description of the solvent concentration field which was assumed to have the same period as the A-B composition profile, whereas it is actually half this period, and is out of phase with it.123 The self-consistent field theory method was later developed to consider the lamellar phase of diblocks in a neutral solvent.159 A slight tendency for solvent segregation to the lamellar interphase was noted, with an initial increase in excess solvent (^s) with increasing copolymer concentration, followed by a decrease in the semidilute and concentrated regimes. This decrease necessarily results as > approaches unity and the volume fraction of solvent, 1-0, approaches zero. The resulting maximum in t/>s is enhanced with increasing degree of A-B segregation in the copolymer, ips being approximately proportional to XAB- Competing with this trend is the approximate inverse dependence of ijjs on N in the weak segregation limit.123 Whitmore and Noolandi also obtained approximate scaling relationships for the domain spacing:159
The strongest dependence occurs in the weak segregation limit, with p w 1/3, q w 0.8, r w 0.4, whereas in the strong segregation limit p w 0.2, q « 2/3, r w 0.22. This compares with the values p = 0.l4 and q = 0.64 obtained for strongly segregated block copolymer melts.170'171 Volume fraction profiles were also calculated for PS-&-PI diblocks in the nearly neutral solvent toluene.159 The scaling predictions of the theory were compared with the experimental results of Hashimoto and coworkers, as discussed in the preceding section. Here no distinction was made between the weak and strong segregation limits. Whitmore and Vavasour, also using self-consistent mean field theory, found that phase diagrams for diblocks in a nonselective solvent could be mapped onto essentially the same diagram, when scaled by (f>XARN-172 The mapping onto the melt phase diagram was suggested, although not quantitatively investigated in this limit (0 —> 1). The theories of Fredrickson and Leibler123 and Olvera de la Cruz124 concern microphase separation in semidilute block copolymer solutions in a nonselective good solvent, analysed using the random phase approximation (RPA) and the 'blob' model where monomers are grouped in blobs the size of the correlation length. These authors used Leibler's formulation of the RPA169 for weakly segregated block copolymers. Whereas Olvera de la Cruz only considered the mean field limit, Fredrickson and Leibler also allowed for composition fluctuations, via the method of Brazovskii173 as applied to block copolymer melts.174 Composition fluctuations are important close to the ODT in block copolymer melts, and there is evidence for their existence in semidilute solutions, as discussed in Section 3.4.122 In the semidilute region, excluded volume interactions are important on length scales smaller than the 'blob' size £, but are screened out at distances larger than £. As the
Concentrated Solutions
145
polymer concentration is increased, the screening becomes more important and the correlation length decreases as ^-VKV+^ ~ 0~°-77,175 where ^ = 0.588 is the excluded volume exponent for good solvents. Because of these excluded volume effects at short distances, the probability of contacts between unlike segments is reduced by a factor of c^2"-1)/^3"-1) ~ 0 023 . Renormalization group studies indicate that the contact probability is even further reduced by an additional factor of >XSD, where XSD ~ 0-29 is a crossover exponent given by %SD == Xs/(3^-l), with Xs ~ 0.22.123 (Monte Carlo simulations indicate that xs maY be significantly larger for strongly incompatible blocks.176) In the limit that contributions to the free energy from solvent inhomogeneities are negligible (which is valid for long copolymer chains), mean field theory for semidilute solutions predicts that 0x^V in Equation (3.8) can be replaced by 4>6XN, where 5 = (xs+l )l($v — 1) « 1.6 (the same substitution can be used to relabel the ordinate for phase diagrams within composition fluctuation theory177). This result is valid in the limit that N~°'22
where F differs by a multiplicative constant of order unity from F in Equation (3.8). At the transition, the concentration of copolymer scales as:
Lodge et al. have observed an exponent consistent with this, based on experiments on nearly symmetric PS-b-PI diblocks, as discussed in Section 3.4.121 The position of the structure factor peak, q', for semidilute block copolymer solutions is predicted to scale as:123'124
Mayes et al. tested this prediction using SANS.162 As mentioned in Section 3.6, they did not obtain this scaling, instead a best fit to the data yielded q* ~ 0°'05. This weaker than expected dependence is presently unexplained. However, the expected concentration dependence of the blob size:
was confirmed.162 This scaling was reported earlier for homopolymer solutions in the semidilute regime.178"181 Experimental results for the scaling of q* with 0 yield different exponents. Duplessix et al. found an exponent 0.087 for PS-b-dPS-b-PS triblocks in CS2.182 Daoud et al. obtained the predicted scaling q* ~ <^>a125 for PS solutions in CS2 containing small amounts of dPS.178 Similar experiments by
146
Block Copolymers in Solution: Fundamentals and Applications
King et al. on solutions of dPS in d-toluene containing unlabelled PS produced an exponent 0.078.181 Thus nonuniversal exponents are obtained for homopolymer solutions as well as block copolymers. The Polymer Reference Interaction Site Model (PRISM) theory has also been applied to investigate ordering in block copolymer solutions. The simplest Gaussian thread version of PRISM theory was applied to study the equilibrium properties of diblock copolymer solutions under neutral solvent conditions.183 Analytical predictions were obtained for the influence of local and microdomain scale concentration fluctuations on the relationship between temperature, degree of polymerization and polymer concentration at the ODT. In the semidilute regime, PRISM predictions were found to agree with blob scaling and fluctuation-corrected field-theoretic analyses. However, in concentrated solutions and the melt, strong disagreements were found, and the mean field dilution approximation was found to fail. Apparent scaling laws for the concentration at the ODT with copolymer degree of polymerization were found for concentrated solutions, with effective exponents which depend on solvent quality, melt screening length, chain aspect ratio, and other nonuniversal structural features. The primary reason for failure of the dilution approximation was found to be interchain nonrandom mixing, which is concentration dependent and driven by local entropic packing effects. Composition fluctuations on the microdomain length scale were shown to yield corrections of secondary importance which vanish in the long chain limit. For good solvents, the predicted apparent exponent for the scaling of <^ODT with N is accidentally in close agreement with semidilute blob scaling results, and agrees with the experiments by Lodge et al.121 on PS-b-PI diblocks. Smaller effective exponents were found under 9 solvent conditions. Detailed PRISM calculations for the small-angle scattering intensity and local physical clustering (via contact interchain pair correlation functions) were also presented.183 Distinctive dependences of these quantities on the degree of polymerization, copolymer concentration, solvent quality and polymer screening length were established. In qualitative agreement with TEM experiments on PS-b-PI solutions in neutral solvents,184 significant local clustering of the minority component was anticipated in asymmetric diblocks even very far from the ODT.
3.8 THEORETICAL UNDERSTANDING OF PHASE DIAGRAMS Self-consistent field theory has been used to calulate phase diagrams of ordered micellar solutions. Noolandi et al. compared continuum self-consistent field theory with the lattice version of this theory for triblock copolymers such as the Pluronics in aqueous solution.185 From a different viewpoint, this work represents an extension of the self-consistent field theory employed by Noolandi and coworkers and Matsen and coworkers for the phase behaviour of block copolymer melts34 to block copolymers in solution. The approximations introduced by the adoption of a lattice model are found to lead to some significant differences in the solution phase
Concentrated Solutions
147
Figure 3.22 Phase diagram for aqueous solutions of Pluronic P94 (EO2i-Z>-PO47-£-EO2i). Notation: LI, isotropic (polymer poor) solution; Ij, cubic phase; H ]5 hexagonal phase; La, lamellar phase; H2, inverse hexagonal phase; I2, inverse cubic phase; L2, isotropic (polymer rich) solution.185 The solid and dashed lines are calculated from the continuum and lattice descriptions, respectively. Reproduced by permission of American Chemical Society.
behaviour compared with the continuum theory, as illustrated by Figure 3.22. For example, the continuum theory predicts ordered phases for Pluronic L64, whereas the lattice theory (neglecting polydispersity) predicts none. Linse and coworkers have also used lattice mean field theory to compute phase diagrams for Pluronic copolymers,10'30'31 as discussed in the preceding section. The approach has also been applied to calculate the phase diagram for a specific ABC triblock in aqueous solution.1 The self-assembly of copolymer ME l4-b-PO \2-b-EOu was examined. Phase boundaries were found to be weakly dependent on copolymer concentration, but strongly dependent on temperature. For each of the ordered phases (hexagonal, lamellar, inverse hexagonal and inverse micellar cubic phases) a three layer structure was determined by analysing volume fraction profiles. Self-consistent field theory has been used to investigate the phase behaviour of block copolymers in solvents of varying selectivity.187 For a neutral good solvent, the dilution approximation is followed. For a symmetric diblock in selective solvent, the sequence lamellar - hexagonal-packed cylindrical micelles - bccpacked spherical micelles - micellar solution - homogeneous solution can be anticipated upon dilution, and self-consistent field theory calculations could reproduce this behaviour (the micellar solution phase was not considered however). For asymmetric copolymers, the theory was able to reproduce transitions from normal to inverse cylinders (and ultimately inverse spheres) via lamellae upon reducing the selectively of a solvent for the majority block.
148
Block Copolymers in Solution: Fundamentals and Applications
There have been surprisingly few computer simulations on mesophase structures in block copolymer solutions, although various mesoscopic simulation techniques have been used to study ordering in block copolymer melts. One exception is dynamic mean field density functional theory calculations performed to simulate the phase behaviour of Pluronic copolymers.188 This method is based on timedependent Ginzburg-Landau equation for the free energy of Gaussian chains. A mapping was made from a coarse-grained polymer bead model onto parameters for two specific Pluronics - L64 and the reverse Pluronic 25R4.188 Good agreement was found when predicted structures for a particular polymer concentration were compared with experimental results. Figure 3.23 shows examples of simulated morphologies. A separate paper focused on the shear-induced structure (and
Figure 3.23 Morphologies for Pluronic L64 from mesoscopic computer simulations. Isosurfaces of EO volume fraction are indicated.188 Polymer concentrations and morphologies as follows: (a) 70%, lamellar structure; (b) 60%, lamellar/hexagonal phase boundary; (c) 55%, hexagonal; (d) 50%, micellar. Reproduced by permission of American Chemical Society.
Concentrated Solutions
149
calculated scattering patterns) of the hexagonal phase for L64.189 The effect of confinement in a cylindrical slit on the same system was also probed, in particular the orientation of the cylinders was examined as a function of slit width.190 Molecular based computer simulation techniques cannot yet directly be used to simulate block copolymer structures, since the simulation would of necessity contain too many atoms. However, coarse-grained models have been pioneered. Larson has carried out lattice Monte Carlo simulations of 'diblocks' A4B7, A4B6, A4B4 and A6H4, and observed the formation of lyotropic mesophases.191 Phase diagrams were constructed for binary surfactant/water mixtures and ternary surfactant/water/oil systems. Klein and coworkers have simulated bilayer, cylindrical and spherical micellar morphologies.192 The results were mapped onto the behaviour of PEO-&-PEE diblocks, studied experimentally.193 Computed density profiles for bilayer phases were compared with those of a lipid bilayer, and the membrane thickness was found to scale approximately with the molar mass of hydrophobe as d~Afg , i.e. as for weakly segregated block copolymer melts.
3.9 FLOW ALIGNMENT In this section, we consider flow alignment of lyotropic structures, in particular orientation under shear. Shear fields can be conveniently applied in several geometries, and provided the shear rate and/or amplitude are high enough, macroscopic alignment of the sample is possible. Understanding the processes of alignment under flow may also be relevant to applications, where in many cases block copolymer gels are subjected to strong flow fields (for example in drug delivery or in detergent formulations under agitation). Of course, in these applications, the flow field is much more complex so fundamental studies have focused on well-defined shear flow fields. The effect of shear flow on block copolymer mesophases has been extensively reviewed.35'77'194"197 3.9.1
LAMELLAR PHASE
Under shear, the layers in a lamellar block copolymer phase typically align with their normals either along the shear gradient direction, Vv, in the so-called parallel orientation, or with their normals along the neutral direction, e, in the so-called perpendicular orientation (see Figure 3.24). The transverse orientation (layer normals along the shear direction, v) has also been observed, although usually as a transient nonequilibrium state. The effects of shearing the lamellar phase of a concentrated block copolymer solution in a Couette cell have been investigated using SANS.198'199 A solution of a PS-b-PI diblock in DOP was investigated and the lamellar orientation was monitored below and above the ODT temperature with the neutron beam incident either radially or tangentially to the Couette cell. It was found that below the
150
Block Copolymers in Solution: Fundamentals and Applications
Figure 3.24 Orientations of a lamellar structure with respect to the shear plane (v, e). (a) Parallel; (b) perpendicular; (c) transverse.
quiescent ODT, oscillatory shear produced lamellae parallel to the plane of the shear cell walls. However, steady shear resulted in a reorientation of the lamellae into the perpendicular orientation (i.e. lamellar normals in the neutral direction).199 Above the transient ODT, the alignment induced by steady shear above a critical rate, quantified by the anisotropy of the scattering ring, was found to follow a master curve as a function of reduced shear rate (with respect to the shear rate for the onset of orientation) for all temperatures.198'199 The critical shear rate was found to increase exponentially with temperature. The transition between parallel and perpendicular orientations was probed in detail on a similar diblock solution subjected to oscillatory shear.200 The transition between perpendicular and parallel orientation could be initiated at a fixed temperature by increasing frequency. A systematic difference in the defect density between parallel and perpendicular alignments was also noted. In particular, under shear flow a sample with the parallel alignment was largely disordered due to a large number of defects, and it was pointed out that this shear-induced disordering below the quiescent ODT is not in agreement with existing theories.200 It was also proposed that the observed shearinduced disordering of perpendicular lamellar is driven by increasing lamellar undulations. Zryd and Burghardt have investigated the same system in both oscillatory and steady shear via rheology and flow birefringence, and compared their results with those obtained for lamellar PS-b-PI melts.201 They found similar behaviour for either type of flow, i.e. preferential parallel alignment of the lamellae at high reduced frequencies and perpendicular alignment at low reduced frequency. This behaviour is the same as that observed in the melt. At high temperature, the degree of perpendicular alignment induced by steady flow decreased, in contrast to the high degree of orientation maintained under oscillatory shear. This was ascribed to an increased propensity for defect formation and/or formation of mixed orientation states under steady shear. In aqueous solutions of Pluronic P85 (PEO26-&-PPO41-b-PEO26) forming a lamellar phase, Mortensen reported that following shear, parallel and perpendicular orientations coexist with all other lamellar orientations in which layers are parallel to the flow direction.194 Richtering and coworkers have examined using SANS the shear-induced formation of vesicles in solutions of Pluronic block copolymers in
Concentrated Solutions
151
butanol/water mixtures.202'203 Pluronic P123 and F127 were both studied. Orientation diagrams were constructed showing that a transition from a parallel to a perpendicular alignment of planar lamellae occurred at a certain shear rate for samples in a well defined region of the phase diagram (copolymer concentration and butanol/water ratio).20 Similarly, conditions for shear-induced formation of multilamellar vesicles were identified. This occurred close to the low polymer concentration boundary of the lamellar phase region. The formation of vesicles is signalled by strong shear thickening, whereas the viscosity decreased across the parallel to perpendicular transition in the more concentrated solution.202 SALS data also showed a four-lobe pattern characteristic of vesicles. 3.9.2
HEXAGONAL PHASE
Shear flow leads to alignment of the cylinders in a hexagonal phase parallel to the shear direction. Two possible orientations of the hexagonal lattice in the (Vv, e) plane are possible, and have been observed for block copolymer melts204 although this has not been reported for solutions. A transition from shear-induced order to shear-induced disorder at higher shear rates has been reported above the quiescent ODT of a hexagonal phase in a block copolymer solution.205 The experiments were performed with a solution of a PS-bPI diblock in the neutral solvent DOP, forming cylinders of PS. The samples were investigated by SANS with in situ steady shear. It was pointed out that existing theories for the effects of shear on block copolymers only predict shear-induced ordering due to the suppression of composition fluctuations. However, it was proposed that fluctuations of the cylinders, if characterized by a lifetime larger than the inverse shear rate, lead to the disordering observed at high shear rates. In prior work, Balsara et a/.206 observed a large difference in the ordering transition of this system for shear oriented 'single crystals' and polydomain samples prepared under quiescent conditions. In sheared polydomain samples where the imperfections were partially removed, coexistence of ordered and disordered regions was observed at temperatures between the two limiting ordering transition temperatures. However, these observations have not been satisfactorily accounted for. Shear orientation of the hexagonal-packed cylinder phase formed in aqueous solutions of Pluronic P85 confirmed the long-range nature of the induced alignment.194 A sample was sheared between parallel plates and rotated with respect to the neutron beam. With the neutron beam incident along the shear gradient direction, a SANS pattern with a pair of meridional Bragg reflections is observed, showing orientation of the cylinders along the shear direction. The hexagonal order was confirmed by the hexagonal symmetry observed by SANS when the neutron beam was incident along the shear direction. At lower concentrations of this triblock, shear-orientation of rod-like micelles leads to an aligned nematic phase, as characterized by SANS.194 Richtering and coworkers studied shear orientation in the hexagonal phase of the related Pluronic L64 in aqueous solution via
152
Block Copolymers in Solution: Fundamentals and Applications
birefringence, SALS and SANS.207 The degree of alignment was found to depend on strain, with a common sigmoidal increase in birefringence with strain independent of stress over the range studied. Depolarized light scattering revealed streaks of scattering along the flow direction due to a mesoscopic stripe texture (observed for other surfactant systems208'209 by polarized light microscopy) perpendicular to the flow direction, ascribed to undulations of the director. Zig-zag undulations of cylinders result from a buckling instability that arises from thermal contraction/ expansion. The development of a stripe pattern due to zig-zag undulations of cylinders formed by diblock EO18-&-BO10 in aqueous solution aligned by flow in a capillary has been observed by polarized optical microscopy and analysed through the splitting of SAXS Bragg reflections.210 The effect of large amplitude oscillatory shear on the orientation of a hexagonal phase formed in solution of EO]8-b-BOi0 in 0.2 M K2SO4 has been studied using SAXS on samples subjected to steady shear in a Couette cell.211 The orientation was quantified in terms of order parameters PI, P4 and P6, where Pn denotes an ensemble averaged Legendre polynomial of order n. These order parameters were extracted from the SAXS patterns using a model of scattering from oriented infinitely long cylinders. It was found that the order parameters increased logarithmically with the shear rate. A related study was undertaken for aqueous gels of EO40-^-BO10.212 Here, it was shown that the susceptibility to shear orientation depended on diblock concentration, because a gel containing 25% diblock did not shear orient, whereas gels containing between 30 and 38% diblock did form a shearoriented cylindrical structure. 3.9.3
CUBIC MICELLAR PHASES
The effect of shear on hcp/fcc and bcc phases has been investigated in some detail.77'196 In essence, samples usually align with a close-packed direction along the shear direction, although flow can occur in several possible shear planes depending on shear conditions and geometry. Figure 3.25 presents representative SAXS patterns from shear-aligned PEO-b-PBO diblocks, showing the indexation to hep and bcc structures, respectively.137 One of the first experiments to examine the effect of shear on the orientation of micellar cubic phase was performed on a PS-fr-PEP diblock in dodecane (a selective solvent for the latter block).213 SANS was used to probe the effects of in situ steady shearing, and it was found that long-range order was induced by very low shear rates, whereas shear melting was noted at high shear rates. The data were interpreted on the basis of a distorted fee structure, with an ABC ABC... layer stacking but distorted from a close-packed arrangement due to normal stresses that led to a decreased inter-layer spacing. Distortions from a close-packed structure were also noted for a PEO-b-PSO diblock studied by SAXS under steady shear.214 The distorted lattice was assigned a rhombohedral structure. For the fee phase formed by PS-&-PI micelles in decane, a transition from polycrystallinity to sliding
Concentrated Solutions
153
Figure 3.25 Schematic of SAXS patterns obtained in the (v, e) plane (shear direction horizontal) following large amplitude oscillatory shear for aqueous diblock solutions.l37 (a) 9 wt% gel of EO96-/?-BOi8 forming a hep structure (indexation to hexagonal lattice), (b) 10 wt% gel of EO315-i'-BOi7 (indexation to a bcc structure). The sample is a multiply twinned, directionally oriented crystal. The labels A, B, C and D correspond to different twins obtained by rotation around the [111 ] direction. Reproduced by permission of American Chemical Society.
hep layers was observed on increasing the shear rate.215 The sliding flow of hep planes observed for such gels at relatively low shear rates has been ascribed215 to a slip-stick mechanism of flow, where the micelles hop from registered (A, B or C) sites (see Figure 3.26). A similar observation has also been made for a diblock containing a polyelectrolyte block, PfBS-b-PNaMA.216 Highly aligned SAXS patterns from fee crystals have been obtained simply by hand shearing cubic phases of PS-&-PI micelles between sliding parallel plates.135
Figure 3.26 Arrangement of spheres in planes in a close-packed structure. With respect to a reference layer A, successive layers can be of type B or C.
154
Block Copolymers in Solution: Fundamentals and Applications
Shear-induced orientation effects in fee structures have also been examined using Pluronic copolymers.217"219 The effect of steady shear, applied using a Couette cell, on the orientation of a fee structure in Pluronic F108 was investigated using SAXS, and transitions between different types of shear flows were elucidated.217 A twinned fee structure with a high density of stacking faults due to flow of sliding layers was observed to transform into large homogeneous single crystals of either twin, separated on a millimetre length scale, upon application of large amplitude oscillatory shear (LAOS).218 The same system was subsequently investigated in more detail using SAXS and SANS.219 Different mechanisms of flow were identified, depending on the shear rate. At low shear rates, the fee structure was locally preserved, and the flow was mediated by defects between crystallites. However, at high shear rates, the melting of the structure was observed through the development of a liquid-like structure factor. Intermediate shear rates (7 w 50 s"1) led to layer sliding, where hep planes were aligned parallel to the Couette cell walls. Creep experiments starting from initially polycrystalline or aligned states have provided a wealth of information on flow mechanisms in an fee gel formed by Pluronic F108.220 It was suggested that the stationary flow curve is associated with the onset of layer sliding of hep planes, although shear thinning occurred at much lower shear rates than those for which layer sliding was detected by SAXS, suggesting nucleation of a fraction of sliding planes below the threshold for detection by SAXS. It was also suggested that the incubation time observed in the creep experiments (above the characteristic stress) was related to the time necessary to reorient initially unoriented grains into sliding hep layers. The effect of shear on an aqueous solution of the diblock EO55-b-BO8 forming a fee structure has been investigated in detail, using SAXS on samples sheared in a Couette cell.221 Steady shear was found to orient the mesophase into a polydomain structure with the hep planes both parallel and perpendicular to the shear plane. At low shear rates, a sliding mechanism of flow of hep layers was identified whereas at higher shear rates, partial melting occurred.221 A common flow-induced state of alignment has been reported from several experiments on different block copolymer solutions forming bcc gels, when subjected to oscillatory or steady shear at moderate shear rates. Flow planes are illustrated in Figure 3.27. In the most common type of flow, a twinned bcc crystal is formed with a <111> close-packed direction along the shear direction, with {110} planes in the shear plane, and a (21 l)-type twinning plane. The bcc twins slip along the twinning planes, allowing the crystal to flow at moderate shear rates.215'222'223 This flow mechanism is the same as that observed for charged colloidal suspensions subjected to steady shear,224 and also observed for a diblock copolymer melt subject to reciprocating shear.225 A transition from an initially aligned twinned bcc structure to a distorted hep structure on increasing shear rate due to the distortion of the twinned structure has been reported for a solution of an PS-£-PI diblock in decane.215 However, the twinned structure is again accessed at higher shear rates (in other words it is the stable state, above a critical shear rate, 7 w 50 s~' in these experiments)215 in contrast to the colloidal system studied earlier.224 At still higher
Concentrated Solutions
155
Figure 3.27 Flow planes in a bcc lattice.223 (a) bcc lattice with a {110} plane shaded, (b) Section of a bcc lattice showing a {110} plane, (c) bcc lattice with a {211} plane shaded, (d) Section of a bcc lattice showing a {211} plane. Reproduced by permission of American Institute of Physics.
shear rates, melting was observed for the copolymer solution, a process also observed for sheared aqueous solutions of diblock EOgg-b-BO^.226 A more complex orientation than that with {110} planes parallel to the shear plane has been observed for 40 wt% gels of EO86-£>-BOi0 in 0.2 M K2SO4 subjected to either steady226 or oscillatory227 shear. The detailed elucidation of the mechanism was possible because higher order 200, and especially 211 reflections could be resolved. Indexation of the SANS and SAXS patterns including these higher order reflections indicated that flow in {211} and {321} planes is important as well as flow in {110} planes. Thus, it was proposed that the aligned crystal is a 'directionally oriented' crystal, with a [111] direction along the shear direction, but with flow in {110}, {211} and {321} planes (which intersect in a common [111] zone axis). This mechanism is consistent with flow in the three principal slip systems of a bcc crystal, as observed for example for bcc metals.228 Which of these planes are active under particular flow conditions is described in detail elsewhere.77 An interesting point is that the {211} planes are not the most densely packed. A crystallographic study of a shear oriented gel of EO2io-^-BOi6 has shed further light on the orientation state of an aligned bcc gel.229 Here, a gel was subjected to O'7'Z
156
Block Copolymers in Solution: Fundamentals and Applications
large amplitude oscillatory shear (LAOS), and then 'mesoscopic crystallography' was performed, i.e. the gel was rotated with respect to the neutron beam to map out the location of Bragg reflections on the Ewald sphere. In this case flow was observed in {110} and {211} planes only, and not also in {321} planes. A similar alignment state was reported for a 30 wt% gel of EO86-^-BOi0 in 0.2 M K2SO4 subjected to steady shear, and investigated by SAXS.227 Mortensen has presented SANS patterns for Pluronic copolymers F88194'230'231 and P85194'230 which appear to show the same orientation of a twinned bcc crystal, although it was not indexed in detail by him. The work of Riser et a/.223 has also revealed a rich and complex flow behaviour under steady shear in bcc gels formed in an aqueous solution of Pluronic F68. The flow in a Couette cell was inhomogeneous due to the large stress gradient (a decrease of 20% at the outer wall compared with the inner wall). Two successive orientation transitions were observed to occur at very different shear rates but separated by a narrow difference in stress. The stress as a function of shear rate shows two plateaux (Figure 3.28), within each of which two distinct structural
Figure 3.28 Two stress plateaux observed in the flow curve for Pluronic F68 in water (46 wt% polymer) at 20 °C.223 Reproduced by permission of American Institute of Physics.
organizations coexist in the gap of the Couette cell. This phenomenon has been termed 'shear banding', as observed earlier for sheared solutions of wormlike micelles. The structure of the gel as a function of shear rate was elucidated by SAXS. This revealed that at low shear rates, a viscous polycrystalline state filled the gap. At intermediate rates, a first oriented state was observed with {211} planes parallel to the shear plane. The proportion of this state increased across the first stress plateau. At higher rates, in the second plateau, a second oriented state was observed with {110} planes oriented in the shear plane. This is the so-called layer sliding regime, observed earlier for highly charged colloidal latex suspensions.224 In
Concentrated Solutions
157
both oriented states, the close-packed <111> direction was along the shear direction. Due to the fact that the stress gradient in the Couette cell (20%) was larger than the 2.5% difference in stress separating the two plateaux in the rheology data, the intermediate state with {211} planes parallel to the shear plane was never observed in isolation, but only in combination with the disoriented state (at 7 = 16 s"1, arrowed in Figure 3.28) or with the layer-sliding state (at 7=160 s"1, arrowed). Scanning the X-ray beam across the gap of the Couette cell confirmed that the intermediate orientation state was located close to the outer wall, as expected since the stress was lower there. A question remains as to why stress plateaux were not observed for an fee gel formed by a related Pluronic copolymer (F108) and also studied extensively by Eiser et al.220 despite the observation of strong shear thinning. The difference compared with the bcc gel in Pluronic F68 could be because no purely elastic regime was observed for the fee gel, suggesting a high density of dislocations sufficient to induce flow even for small, applied stresses. The bcc gel may be less defective, alternatively there may be a difference due to the distinct preferred flow planes in bcc compared with fee gels. Hamley and coworkers66 have also compared the flow behaviour of fee and bcc gels subjected to LAOS, and noted the much more elastic response of the bcc gel. In a bcc gel the flow appears to be hindered resulting in a slip-stick motion as planes of micelles shear past one another. In contrast, in a fee gel, sliding of hep layers appears to be unhindered (at least at sufficiently high shear rates). This difference may reflect the difference in the intermicellar potential, which is more long range in the case of a bcc packing of micelles (as discussed in Section 3.5.1). Watanabe et al. have shown that the equilibrium modulus, Ge is hardly affected by shear orientation of a PS-b-PI diblock forming a bcc structure in tetradecane.232 It was noted that this indicates that defects do not have a major influence on Ge. The measured value of Ge is almost an order of magnitude smaller than that predicted on the basis of entropic elasticity theory:
The difference between the measured Ge and Gj] was ascribed to the correlation between the conformation of the PB blocks (which minimizes the osmotic energy) and the influence of this on the effective number density of strands. The nonlinear response of bcc227 and fee66 gels subject to oscillatory strain has been probed, plotting stress against strain in Lissajous figures. The interpretation of the flow mechanism followed that of Doi et al., who employed a lattice model to analyse the viscoelasticity of a (hexagonal) micellar phase formed by a block copolymer.233 The results show that flow under LAOS in a bcc phase occurs through a slip-stick mechanism, and suggest that this is the process that leads to the observed macroscopic orientation of gels, under steady shear. 226>227 This complements other work that has suggested a slip-stick flow mechanism for fee gels under steady shear.215'216 Lissajous figures obtained for an fee gel of a PEO-£»-PBO
158
Block Copolymers in Solution: Fundamentals and Applications
diblock under LAOS indicated strong shear thinning.66 The onset of strong shear thinning is associated with the development of layer sliding flow of hep layers. As mentioned above, the observation of strong shear thinning from creep experiments on an fee gel was also ascribed to this flow mechanism.220 The influence of oscillatory shear on PS-&-PEB-&-PS triblocks in a solvent selective for the midblock has been investigated by Mortensen et al.234 At the concentration examined, a bcc structure was formed. Flow occurred in both {111} and {110}-type planes along <112> and <111> directions, respectively. An orientation diagram in terms of the flow state as a function of frequency and strain was presented. The influence of the duration of shear was also examined. If shear was stopped after only a short time, rapid relaxation of orientation was observed. However, retention of an oriented scattering pattern was observed after prolonged shearing, and this was ascribed to flow involving relatively large grains. The difference in relaxation behaviour suggests a decrease in slip plane density as shearing progresses. The influence of steady shear on a bcc lattice formed by a PS-b-PB diblock in the PS-selective solvent di-n-butyl phthalate has been examined by shear rheometry.102 The flow behaviour is unsurprisingly non-Newtonian. At low shear rates, steady state was achieved after a significant stress overshoot followed by a thixotropic (time-dependent shear thinning) stress decay. This plastic behaviour reflected mild disruption of the lattice. At high shear rates, the lattice was greatly disrupted, leading to a much weaker dependence of viscosity on shear rate. The pre-sheared system exhibited recovery to an elastic response following cessation of shear. The slowest recovery was observed at a frequency close to that associated with concentration fluctuations. It follows that shear at this (intermediate) frequency was most effective in disrupting the lattice. Thus shear at the highest shear rate does not always lead to the strongest shear-induced disordering. It was noted that the soft nature of the PB core in these micelles allows fluctuations in the number of chains per micelle that varies as defects are created (during shear) or annihilated (during recovery). The shear melting of a bcc lattice has been examined for melts and concentrated solutions of PS-&-PI and PS-&-PEP diblock, triblock and starblock copolymers.235 Steady shear destroys the bcc lattice at a critical shear stress, ac — 0.038G°CC, where G®cc is the low frequency modulus of the bcc lattice. When comparing data at a common reduced temperature, this quantity was found to scale approximately as G£CC ~ RT/d3, where d is the measured domain spacing of the {110} planes. This scaling is consistent with a potential of the form t/(r)~&B71n(l/r), which is similar to that proposed by Witten and Pincus 236 for spheres with interacting grafted polymer layers (see Section 2.9), although the data obtained by Register and coworkers did not scale with / as predicted. A universal scaling was found when G°ccd3/RT was plotted against reduced temperature (this quantity being employed in the absence of quantitative data on x) and found to vary exponentially. In a companion paper, the flow curve (viscosity as a function of shear stress) of bcc block copolymer melts was discussed.237 The melting of the bcc lattice at a critical
Concentrated Solutions
159
shear stress led to a dramatic reduction in viscosity. The kinetics of the ordering process were also studied.237 The low frequency deformation behaviour was examined via creep and dynamic viscoelasticity measurements.238 Using a model for high temperature creep in metals and ceramics, data for zero shear viscosity as a function of temperature could be rescaled onto a master curve. The resulting Arrhenius plots provided an activation energy that is related to the temperature dependence of the monomeric friction coefficients, with an additional contribution from block segregation. The effect of grain size (controlled through thermal history) on the viscosity was also investigated. The effect of extensional flow applied to orient a bcc gel formed by a PS-&-PEBb-PS triblock in a midblock selective oil has been examined by SAXS.239'240 The deformation was found (by measuring lattice dimensions by SAXS) to be affine for low strains, the onset deformation for nonaffine behaviour decreasing with decreasing polymer concentration.239 At large strains a 'butterfly' pattern was observed, similar to that previously obtained for stretched polymer networks and ascribed either to physical inhomogeneities or anisotropic concentration fluctuations.240 Its observation for the triblock gel was attributed to cluster formation - stretching brings endblock domains closer together in the direction perpendicular to the deformation. The effect of extension rate was also examined. It also found that the deformation was nonaffine at low extension rates, but became affine at higher rates.241
3.10 DYNAMICS 3.10.1
DYNAMIC MODES
DLS has been used to probe the dynamic modes for dilute, semidilute and concentrated solutions of PS-&-PI diblocks in the neutral good solvent toluene.242"244 Three modes have been observed244 - cooperative diffusion (with diffusion coefficient Dc), corresponding to relaxations of fluctuations in polymer concentration, a nondiffusive (q independent) internal mode (with decay rate 1^) that reflects the relative motion of the centres of mass of the blocks on a single chain243 and a heterogeneity mode (DH), due to fluctuations in composition from chain to chain. The observation of these modes is in agreement with theoretical predictions by Benmouna et a/.245'246 and Semenov and coworkers.243 Benmouna et al. predicted cooperative and internal modes, whilst Semenov and coworkers considered a heterogeneity mode. The observation of a fourth mode due to clusters or long-range density fluctuations243 is controversial, and has been ascribed to nonequilibrium or metastable states.244 Lodge and coworkers found that in dilute solution of their PS-b-PI diblocks, a single diffusive mode is present, and is attributed to a superposition of Dc and DH modes. ^ In semidilute and concentrated solutions these two modes are clearly resolved. The decay rate for the cooperative and heterogeneity modes exhibit the predicted dependence on concentration, and molecular weight. Field gradient NMR
160
Block Copolymers in Solution: Fundamentals and Applications
studies were also performed to determine the translational diffusivities (Ds) of the same copolymers. Values of Ds were found to be very similar to DH. The internal mode was hard to resolve, but was observed for the highest molecular weight sample examined. The concentration dependences of DH and Ds were found to be insensitive to the ODT. The relationship of these results to earlier DLS studies of block copolymers in neutral good solvents was discussed in detail (this is not reiterated here). Apparent inconsistencies could be resolved by appropriate assignment of the heterogeneity mode. In a companion paper, index matching of the copolymer was achieved in zero average contrast conditions by the use of appropriate solvents for the PS-&-PI diblocks studied.247 The cooperative diffusion mode could be eliminated, enhancing the relative amplitude of the internal mode, which was observed together with the heterogeneity mode. The decay rate of the internal mode was found to decrease with concentration. Values of DH were again in agreement with those obtained from pulsed field gradient NMR. As the concentration increased towards the ODT, the total scattered intensity increased markedly due to the onset of large amplitude concentration fluctuations. Forced Rayleigh scattering revealed a decrease in diffusivity with increasing concentration for matched asymmetric PS-b-PI and PS-£-PI-b-PS copolymers.248 This was ascribed to the retardation of chain diffusion due to concentration fluctuations. The diffusivity decreased rapidly and discontinuously across the ODT. Since the retardation of mobility due to ordering or concentration fluctuations was larger for the diblock, for which the ordering concentration was greater, the effect was ascribed to the increased monomeric friction in the PS-rich cylindrical domains. A comparison was made to the melt behaviour of the same polymers, and it was noted that the onset of fluctuation effects in the diffusivity occurs at the same relative distance from the ODT in both melts and solutions. The dynamics of giant micelles formed by high molecular weight PS-b-PI diblocks (Mw > 1 x 106 g mol"1) in an unidentified solvent were compared with those of colloidal particles. In contrast to the latter systems, no slow-down in the collective diffusional dynamics is observed in the vicinity of q*, the peak of the static structure factor.249 This was ascribed to nonlocal mobility due to interactions between the tethered corona chains. Since the chains are long enough to be entangled, reptation contributes substantially to the dynamics.250 The same group have also investigated the dynamics of model colloidal particles prepared by crosslinking the PCEMA core of PS-&-PCEMA micelles.251 Two relaxation processes were observed - a fast cooperative diffusion due to interactions among coronal chains and a slow one due to self-diffusion of the core. Analogies with the dynamics of star polymers252 were made, although this is outside the scope of the present text.
3.10.2
DYNAMICS OF GELATION
Gelation is manifest by changes in the intermediate scattering function S(q, i) from the simple exponential relaxation dynamics of a simple liquid. In fact, the system
Concentrated Solutions
161
Figure 3.29 DLS field correlation function data for a hard gel (o), compared with the single exponential relaxation observed for the dilute liquid phase (n) formed by EO92-^-BOi8 in aqueous solution.262 Reproduced by permission of American Institute of Physics.
becomes nonergodic (i.e. the dynamics vary from point to point in the sample). Nevertheless, by averaging measurements over the sample volume, reliable DLS data can be obtained. A plateau is commonly observed as the dynamics slow down over a certain relaxation time interval (see, for example, Figure 3.29). The dynamics of Pluronic copolymers in structured mesophases provide an early example of the observation of this phenomenon.23'253 In particular, DLS has been used to probe the dynamics of reverse Pluronic 25R8 in aqueous solution.23 At low temperatures and copolymer concentrations the DLS correlation functions were characterized by two dynamic modes. These modes were associated with the presence of free molecules and networks of copolymer strands interconnected through the PPO blocks. At higher copolymer concentrations the correlation functions showed a fast decay followed by a slower relaxation with a much longer decay time. Indeed, a pronounced plateau in the correlation function indicated a nonergodic system in which dynamic fluctuations were suppressed for an intermediate range of timescales. The system was then described as a network of micelles, interconnected by PPO strands. An increase in the polymer concentration of this network at low temperatures, resulted in an ordered solid-like mesophase.23 The dynamics of Pluronic P85 in aqueous solution were also characterized by DLS.253 SANS indicated a single lamellar mesophase at high temperatures. At low temperatures two phases were present, one of which resembled the high temperature phase while the other was similar to the lamellar melt phase.253 The DLS correlation function exhibited a single exponential decay in the concentrated lamellar phase at high temperature. Upon dilution, a fast mode appeared also at
162
Block Copolymers in Solution: Fundamentals and Applications
high temperature as a second lyotropic lamellar phase forms. Here again, a pronounced plateau was observed in the decay of the correlation function. The angular dependence indicated large anisotropic grains in this nonergodic 9^ system. Analogies exist between the dynamics of block copolymer gels and glassy states observed for colloidal systems. A system of hard spheres is dominated by the repulsive interactions between the particles. Particles are increasingly caged by their neighbours (simultaneously each particle makes part of a cage) as the volume fraction of hard spheres increases. At a critical volume fraction, this caging becomes permanent and stops all long-range motion. The system can then be considered nonergodic or glassy.254 However, a hard-sphere glass can be perturbed by the introduction of a short-range particle attraction.255'256 Short-range attractive forces lead to clustering of the particles forming the cage and opens holes, leading to the 'softening' of the hard sphere glasses. This is a qualitatively different glassy state dominated by attractive interactions, where the particles tend to form polydisperse aggregates, some of which can span the whole sample (percolating clusters). Colloidal glassy states dominated by repulsive and/or attractive interactions are well described in terms of the mode-coupling theory (MCT) of supercooled liquids in the vicinity of a glass transition.256"258 Recently DLS has also been used to investigate the application of MCT to the sol-gel transition of block copolymer micellar systems259"261 through the determination of the field correlation function gw(t) via DLS. Dense micellar solutions of Pluronic L64 were studied by DLS.259'261 The field correlation function showed a logarithmic decay, attributed to a higher-order glass transition singularity predicted by MCT, which points to the coexistence of both percolation and structural arrest in the system.259'261 The dynamics of a PEO-bPBO diblock in aqueous solution were studied in micellar, soft gel and hard gel phases.262 Both the liquid and the soft gel phases are micellar phases, although the structural order is higher in the soft gel phase than in the liquid phase. The hard gel phase corresponds to an fee arrangement of micelles. DLS results revealed that the relaxation time in the dilute liquid phase exhibits a single characteristic time associated with the diffusion of the micelles. In addition, a second characteristic time associated with the presence of micellar clusters in the system is identified in the concentrated liquid and in the soft gel phases. It was thus suggested that the structure of the soft gel phase comprises micellar clusters coexisting with micellar fluid, in good agreement with hypotheses from previous work.54'55 The dynamics of the system slows down as the hard gel phase is approached and a plateau is observed in the DLS correlation function (Figure 3.29). The structure of the hard gel is 'softened' upon increasing temperature and/or decreasing concentration. As for the Pluronic copolymers studied by Chen and coworkers,259'261 gelation in these diblock systems is not related to bridging of copolymer chains between micelles, but is due to packing of micelles into cubic structures. The dynamics of micellar solutions of PS-b-PAA block copolymers were investigated by rheology in the vicinity of the gel point263 and in the gel at high
Concentrated Solutions
163
concentration via rheology and DLS.260 The PAA was not completely hydrolysed, and contained a fraction of hydrophobic ethyl aery late 'stickers'. At the gel point, the expected power law dependence of moduli on frequency, G' ~ G" ~ u;A (Section 3.3.1) was observed. The exponent A was consistent with that predicted by percolation theory.53 A re-entrant liquid-to-glass transition was observed for the PS-b-PAA system. This was in agreement with the MCT phase diagram for a system of hard spheres with short range attractive interactions, which presents a reentrant repulsive glass-to-liquid-to attractive glass transition for a particular range of hydrophobic stickers concentration. In addition, the field correlation function showed an extended logarithmic decay with increasing concentration of sticky soft spheres above the re-entrant transition branch. This, as already mentioned, indicates the coexistence of attractive and repulsive interactions in the system according to MCT. It was therefore concluded that the PS-b-PAA system obeyed all the theoretical predictions of MCT. However not all the DLS results on polymeric micelle or cluster dynamics found in the literature have been explained in terms of MCT. In some cases the DLS data may be reminiscent of supercooled liquids and colloids close to the glass transition, but the final analysis of the results was other than that based on the MCT.264~267 Semidilute micellar solutions of PS-b-PI-b-PS triblock copolymers in a midblock selective solvent were studied by DLS.267 The system could be described as a micellar network where the micelles, well ordered in a cubic structure, were linked by nodes formed by the terminal PS blocks. The correlation function measured by DLS was similar to that expected for supercooled liquid and colloids near the glass transition, but it was interpreted in terms of PS node cooperative diffusion. The dynamics of PS-6-PEB-£-PS micelles in a midblock selective solvent were similarly investigated.265 The dilute solution correlation function only showed the translational diffusion of flower-like micelles. At intermediate concentrations, the correlation functions were characterized by a slow mode and a fast mode. The slow mode was ascribed to the diffusion of polydisperse molecular aggregates, formed by random bridging of triblock copolymer molecules, while the fast mode was associated with the diffusion of flower-like micelles. The semidilute region was identified as a physical gel, characterized by the presence of physical PS nodes. Three dynamic processes were extracted from the correlation function in the semidilute region: a fast mode, corresponding to the collective diffusion of interconnected PS nodes in the physical gel; an intermediate mode, due to the local dynamics of PS nodes trapped in the network of the ordered gel; and a slow mode, associated with the presence of aggregates of insoluble homopolymer impurities (PS) in the system.265 The dynamics of aqueous solutions of a hydrophobically modified poly(oxyethylene) was studied by DLS.266 In this case the time correlation function was described by a single exponential followed by a fractional exponential relaxation function. At short times, the correlation function indicated a diffusive mode whereas at long times, the analysis provided a mean relaxation time, associated with the release of coupled clusters.
164
Block Copolymers in Solution: Fundamentals and Applications
The influence of thermal gelation on the dynamics of a giant PS-b-PI diblock (M = 2 x 106 g moF1) in decane was compared with the gelation of star polymers with /= 128 arms (the association number of the diblock micelles was p = 1500).268 The gelation led to a slow down in dynamics due to the overlap of 'arms' (corona chains for the micelles), and a plateau developed in S(q, i). Further details on thermal gelation in the star polymer systems can be found elsewhere.269'270 NMR has also been used to investigate gelation. For example, changes in the proton relaxation times for the the PEO block (in particular the transverse relaxation time r2) were noted at the lower and upper gel transitions for several PEO-b-PEO diblock and PEO-b-PBO-b-PEO triblock copolymers.271 A two-step relaxation model indicated that the slow mode representing large scale motions was affected much more significantly by gelation than the fast mode ascribed to local segmental motions. Transient nuclear Overhauser effect measurements revealed interpenetration of PEO and PBO blocks at the core-corona interface, as for a PEO-b-PPO-bPEO triblock studied earlier.272 For the Pluronic-type copolymer, changes in the relaxation times of both PEO and PPO blocks were observed at the gel transitions (although the fast mode for the PEO block was not sensitive to gelation).272
REFERENCES 1. Alexandridis, P.; Olsson, U.; Lindman, B. Langmuir 1997, 13, 23. 2. Mortensen, K.; Brown, W.; Norden, B. Phys. Rev. Lett. 1992, 68, 2340. 3. Luo, Y. Z.; Nicholas, C. V.; Attwood, D.; Collett, J. H.; Price, C; Booth, C. Colloid Polym. Sci. 1992, 270, 1094. 4. Hanley, K. J.; Lodge, T. P.; Huang, C.-I. Macromolecules 2000, 33, 5918. 5. Lodge, T. P.; Pudil, B.; Hanley, K. J. Macromolecules 2002, 35, 4707. 6. Lodge, T. P.; Hanley, K. J.; Pudil, B.; Alahapperuma, V. Macromolecules 2003, 36, 816. 7. Lai, C.; Russel, W. B.; Register, R. A. Macromolecules 2002, 35, 841. 8. Alexandridis, P.; Zhuo, D.; Khan, A. Langmuir 1996, 12, 2690. 9. Alexandridis, P.; Olsson, U.; Lindman, B. Macromolecules 1995, 38, 7700. 10. Svensson, M.; Alexandridis, P.; Linse, P. Macromolecules 1999, 32, 637. 11. Wanka, G.; Hoffmann, H.; Ulbricht, W. Macromolecules 1994, 27, 4145. 12. Alexandridis, P.; Olsson, U.; Lindman, B. Langmuir 1998, 14, 2627. 13. Svensson, B.; Olsson, U. Macromolecules 2000, 33, 7413. 14. Mortensen, K.; Pedersen, J. S. Macromolecules 1993, 26, 805. 15. Svensson, B.; Alexandridis, P.; Olsson, U. J. Phys. Chem. B 1998, 102, 7541. 16. Alexandridis, P. Macromolecules 1998, 31, 6935. 17. Ivanova, R.; Lindman, B.; Alexandridis, P. Langmuir 2000, 16, 3660. 18. Alexandridis, P.; Ivanova, R.; Lindman, B. Langmuir 2000, 16, 3676. 19. Holmqvist, P.; Alexandridis, P.; Lindman, B. / Phys. Chem. B 1998, 702, 1149. 20. Zipfel, J.; Berghausen, J.; Schmidt, G.; Lindner, P.; Alexandridis, P.; Richtering, W. Macromolecules 2002, 35, 4064. 21. Holmqvist, P.; Alexandridis, P.; Lindman, B. Macromolecules 1997, 30, 6788.
Concentrated Solutions
165
22. Ivanova, R.; Lindman, B.; Alexandridis, P. Langmuir 2000, 16, 9058. 23. Mortensen, K.; Brown, W.; J0rgensen, E. Macromolecules 1994, 27, 5654. 24. Chu, B.; Zhou, Z. Physical Chemistry of Polyoxyalkylene Block Copolymer Surfactants. In Nonionic Surfactants: Polyoxyalkylene Block Copolymers; Nace, V. M., Ed. Marcel Dekker: New York, 1996. 25. Israelachvili, J. N.; Mitchell, D. J.; Ninham, B. W. /. Chem. Soc. Faraday Trans. 2 1976, 72, 1525. 26. Hamley, I. W. Introduction to Soft Matter. John Wiley & Sons, Ltd: Chichester, 2000. 27. Yang, L.; Alexandridis, P.; Steytler, D. C.; Kositza, M. J.; Holzwarth, J. F. Langmuir 2000, 16, 8555. 28. Yang, L.; Alexandridis, P. Langmuir 2000, 16, 4819. 29. Alexandridis, P. Curr. Opin. Colloid Interface Sci. 1997, 2, 478. 30. Linse, P. J. Phys. Chem. 1993, 97, 13896. 31. Svensson, M.; Alexandridis, P.; Linse, P. Macromolecules 1999, 32, 5435. 32. Hentze, H.-P; Kramer, E.; Berton, B.; Forster, S.; Antonietti, M.; Dreja, M. Macromolecules 1999, 32, 5803. 33. Forster, S.; Berton, B.; Hentze, H.-P.; Kramer, E.; Antonietti, M.; Lindner, P. Macromolecules 2001, 34, 4610. 34. Hamley, I. W. The Physics of Block Copolymers. Oxford University Press: Oxford, 1998. 35. Hamley, I. W.; Mai, S.-M.; Ryan, A. J.; Fairclough, J. P. A.; Booth, C. Phys. Chem., Chem. Phys. 2001, 3, 2972. 36. Tanodekaew, S.; Godward, J.; Heatley, F.; Booth, C. Macromol. Chem. Phys. 1997, 198, 3385. 37. Crothers, M.; Attwood, D.; Collett, J. H.; Yang, Z.; Booth, C.; Taboada, P.; Mosquera, V.; Ricardo, N. M. P. S.; Martini, L. G. A. Langmuir 2002, 18, 8685. 38. Yang, Z.; Crothers, M.; Ricardo, N. M. P. S.; Chaibundit, C.; Taboada, P.; Mosquera, V.; Kelarakis, A.; Havredaki, V.; Martini, L. G. A.; Valder, C.; Collett, J. H.; Attwood, D.; Heatley, E; Booth, C. Langmuir 2003, 19, 943. 39. Tsitsilianis, C.; Alexandridis, P.; Lindman, B. Macromolecules 2001, 34, 5979. 40. Hvidt, S.; J0rgensen, E. B.; Schillen, K.; Brown, W. J. Phys. Chem. 1994, 98, 12320. 41. Kelarakis, A.; Mingvanish, W.; Daniel, C.; Li, H.; Havredaki, V.; Booth, C.; Hamley, I. W.; Ryan, A. J. Phys. Chem., Chem. Phys. 2000, 2, 2755. 42. Li, H.; Yu, G.-E.; Price, C.; Booth, C.; Hecht, E.; Hoffmann, H. Macromolecules 1997, 30, 1347. 43. Almgren, M.; Brown, W.; Hvidt, S. Colloid Polym. Sci. 1995, 273, 2. 44. Mingvanish, W.; Kelarakis, A.; Mai, S.-M.; Daniel, C.; Yang, Z.; Havredaki, V.; Hamley, I. W.; Ryan, A. J.; Booth, C. J. Phys. Chem. B 2000, 104, 9788. 45. Kelarakis, A.; Havredaki, V.; Rekatas, C.; Mai, S.-M.; Attwood, D.; Booth, C.; Ryan, A. J.; Hamley, I. W.; Martini, L. G. A. Macromol. Chem. Phys. 2001, 202, 1345. 46. Jones, J. L.; McLeish, T. C. B. Langmuir 1995, //, 785. 47. Jin, X.; Lodge, T. P. Rheol Acta 1997, 36, 229. 48. Hamley, I. W.; Fairclough, J. P. A.; Ryan, A. J.; Ryu, C. Y; Lodge, T. P.; Gleeson, A. J.; Pedersen, J. S. Macromolecules 1998, 31, 1188. 49. Higgins, J. S.; Blake, S.; Tomlins, P. E.; Ross-Murphy, S. B.; Staples, E.; Penfold, J.; Dawkins, J. V. Polymer 1988, 29, 1968. 50. Soenen, H.; Berghmans, H.; Winter, H. H.; Overbergh, N. Polymer 1997, 38, 5653. 51. Chambon, E; Winter, H. H. /. Rheol. 1987, 31, 683.
166
Block Copolymers in Solution: Fundamentals and Applications
52. Winter, H. H.; Mours, M. Adv. Polym. Sci. 1997, 134, 165. 53. Lobry, L.; Micali, N.; Mallamace, K; Liao, C; Chen, S.-H. Phys. Rev. E 1999, 60, 7076. 54. Castelletto, V.; Caillet, C.; Fundin, J.; Hamley, I. W.; Yang, Z.; Kelarakis, A. J. Chem. Phys. 2002, 116, 10947. 55. Castelletto, V.; Caillet, C.; Hamley, I. W.; Yang, Z. Phys. Rev. E 2002, 65, 050601(R). 56. Sato, T.; Watanabe, H.; Osaki, K.; Yao, M.-L. Macromolecules 1996, 29, 3881. 57. Watanabe, H. Acta Polym. 1997, 48, 215. 58. Watanabe, H.; Kotaka, T.; Hashimoto, T.; Shibayama, M.; Kawai, H. J. Rheol. 1982, 26, 153. 59. Shibayama, M.; Hashimoto, T.; Kawai, H. Macromolecules 1983, 16, 16. 60. Hashimoto, T.; Shibayama, M.; Kawai, H.; Watanabe, H.; Kotaka, H. Macromolecules 1983, 16, 361. 61. Watanabe, H.; Kotaka, T. Polym. Eng. Rev. 1984, 4, 73. 62. Watanabe, H.; Sato, T.; Osaki, K.; Yao, M.-L. Macromolecules 1996, 29, 3890. 63. Watanabe, H.; Sato, T.; Osaki, K.; Yao, M.-L.; Yamagishi, A. Macromolecules 1997, 30, 5877. 64. Reimers, M. J.; Dealy, J. M. /. Rheol. 1996, 40, 167. 65. Wilhelm, M.; Maring, D.; Spiess, H.-W. Rheol. Acta 1998, 27, 399. 66. Daniel, C.; Hamley, I. W.; Wilhelm, M.; Mingvanish, W. Rheol. Acta 2001, 40, 39. 67. Watanabe, H. Nihon Reorogi Gakkaishi (J. Soc. Rheol. Jpn) 1980, 8, 26. 68. Habas, J.-R; Pavie, E.; Lapp, A.; Peyrelasse, J. /. Rheol. 2004, 48, 1. 69. Hamley, I. W.; Pople, J. A.; Fairclough, J. P. A.; Ryan, A. J.; Booth, C.; Yang, Y.-W. Macromolecules 1998, 31, 3906. 70. Pople, J. A.; Hamley, I. W.; Fairclough, J. P. A.; Ryan, A. J.; Komanschek, B. U.; Gleeson, A. J.; Yu, G.-E.; Booth, C. Macromolecules 1997, 30, 5721. 71. Deng, N.-J.; Luo, Y.-Z.; Tanodekaew, S.; Bingham, N.; Attwood, D.; Booth, C. J. Polym. Sci. B: Polym. Phys. 1995, 33, 1085. 72. Yang, Y.-W.; Ali-Adib, Z.; McKeown, N. B.; Ryan, A. J.; Attwood, D.; Booth, C. Langmuir 1997, 13, 1860. 73. Kelarakis, A.; Havredaki, V.; Derici, L.; Yu, G.-E.; Booth, C.; Hamley, I. W. /. Chem. Soc., Faraday Trans. 1998, 94, 3639. 74. Mingvanish, W.; Mai, S.-M.; Heatley, F.; Booth, C.; Attwood, D. J. Phys. Chem. B 1999, 103, 11269. 75. Chaibundit, C.; Mai, S.-M.; Heatley, F; Booth, C. Langmuir 2000, 16, 9645. 76. Schmolka, I. R. /. Biomed. Mater. Res 1972, 6, 571. 77. Hamley, I. W. Phil. Trans. R. Soc. Lond. 2001, 559, 1017. 78. Yu, G.-E.; Ameri, M.; Yang, Z.; Attwood, D.; Price, C.; Booth, C. J. Phys. Chem. B 1997, 101, 4394. 79. Bedells, A. D.; Arafeh, R. M.; Yang, Z.; Attwood, D.; Padget, J. C.; Price, C.; Booth, C. J. Chem. Soc., Faraday Trans. 1993, 89, 1243. 80. Yu, G.-E.; Deng, Y.-L.; Dalton, S.; Wang, Q.-G.; Attwood, D.; Price, C.; Booth, C. J. Chem. Soc., Faraday Trans. 1992, 88, 2537. 81. Booth, C.; Attwood, D. Macromol. Rapid Commun. 2000, 27, 501. 82. Kelarakis, A.; Havredaki, V.; Yu, G.-E.; Derici, L.; Booth, C. Macromolecules 1998, 31, 944. 83. Vadnere, M.; Amidon, G.; Lindenbaum, S.; Haslam, J. L. Int. J. Pharm. 1984, 22, 207. 84. Malmsten, M.; Lindman, B. Macromolecules 1992, 25, 5440.
Concentrated Solutions
167
85. J0rgensen, E. B.; Hvidt, S.; Brown, W.; Schillen, K. Macromolecules 1997, 30, 2355. 86. Derici, L.; Ledger, S.; Mai, S.-M.; Booth, C.; Hamley, I. W.; Pedersen, J. S. Phys. Chem., Chem. Phys. 1999, 7, 2773. 87. Kjellander, R.; Florin, E. J. Chem. Soc. Faraday Trans. 1 1981, 77, 2053. 88. Linse, P.; Bjorling, M. Macromolecules 1991, 24, 6700. 89. Bahadur, P.; Pandya, K.; Almgren, M.; Li, P.; Stilbs, P. Colloid Polym. Sci. 1993, 272, 657. 90. Malmsten, M.; Lindman, B. Macromolecules 1993, 26, 1282. 91. Prud'homme, R. K.; Wu, G.; Schneider, D. K. Langmuir 1996, 72, 4651. 92. Liu, L.-Z.; Wan, Q.; Liu, T.; Hsiao, B. S.; Chu, B. Langmuir 2002, 28, 10402. 93. Zhang, K.; Lindman, B.; Coppola, L. Langmuir 1995, 77, 538. 94. Hecht, E.; Mortensen, K.; Gradzielski, M.; Hoffmann, H. / Phys. Chem. 1995, 99, 4866. 95. Kelarakis, A.; Yuan, X.-E; Mai, S.-M.; Yang, Y.-W.; Booth, C. Phys. Chem., Chem. Phys. 2003, 5, 2628. 96. Kelarakis, A.; Havredaki, V.; Yuan, X.-E; Wang, Y.-W.; Booth, C. J. Mater. Chem. 2003, 23, 2779. 97. Kelarakis, A.; Ming, X.-T.; Yuan, X.-E; Booth, C. Langmuir 2004, 20, 2036. 98. Lafleche, E; Durand, D.; Nicolai, T. Macromolecules 2003, 36, 1331. 99. Lafleche, E; Nicolai, T.; Durand, D.; Gnanou, Y; Taton, D. Macromolecules 2003, 36, 1341. 100. Watanabe, H.; Sato, T.; Osaki, K. Macromolecules 2000, 33, 2545. 101. Tan, H.; Watanabe, H.; Matsumiya, Y; Kanaya, T.; Takahashi, K. Macromolecules 2003, 36, 2886. 102. Watanabe, H.; Matsumiya, Y; Kanaya, T.; Takahashi, Y. Macromolecules 2001, 34, 6742. 103. Tae, G.; Kornfield, J. A.; Hubbell, J. A.; Johannsmann, D.; Hogen-Esch, T. E. Macromolecules 2001, 34, 6409. 104. Xu, B.; Li, L.; Yekta, A.; Masoumi, A.; Kanagalingam, S.; Winnik, M. A.; Zhang, K.; Macdonald, P. M.; Menchen, S. Langmuir 1997, 13, 2447. 105. Serero, Y; Aznar, R.; Porte, G.; Berret, J.-E; Calvet, D.; Collet, A.; Viguier, M. Phys. Rev. Lett. 1998, 82, 5584. 106. Annable, T.; Buscall, R.; Ettelaie, R.; Whittlestone, D. J. Rheol. 1993, 37, 695. 107. Preuschen, J.; Menchen, S.; Winnik, M. A.; Heuer, A.; Spiess, H. W. Macromolecules 1999, 32, 2690. 108. Spontak, R. J.; Wilder, E. A.; Smith, S. D. Langmuir 2001, 77, 2294. 109. Bromberg, L. Macromolecules 1998, 31, 6148. 110. Bromberg, L. J. Phys. Chem. B 1998, 702, 10736. 111. Bromberg, L.; Temchenko, M.; Moeser, G. D.; Hatton, T. A. Langmuir 2004, 20, 5683. 112. Bromberg, L.; Temchenko, M.; Hatton, T. A. Langmuir 2002, 28, 4944. 113. Bromberg, L.; Temchenko, M.; Hatton, T. A. Langmuir 2003, 79, 8675. 114. Simmons, M. R.; Yamasaki, E. N.; Patrickios, C. S. Macromolecules 2000, 33, 3176. 115. Triftaridou, A. I.; Hadjiyannakou, S. C.; Vamvakaki, M.; Patrickios, C. S. Macromolecules 2002, 35, 2506. 116. Patrickios, C. S.; Georgiou, T. K. Curr. Opin. Colloid Interface Sci. 2003, 8, 76. 117. Demosthenous, E.; Hadjiyannakou, S. C.; Vamvakaki, M.; Patrickios, C. S. Macromolecules 2002, 35, 2252. 118. Vamvakaki, M.; Patrickios, C. S. Chem. Mater. 2002, 14, 1630.
168
Block Copolymers in Solution: Fundamentals and Applications
119. Georgiades, S. N.; Vamvakaki, M.; Patrickios, C. S. Macromolecules 2002, 35, 4903. 120. Dormidontova, E. E.; Lodge, T. P. Macromolecules 2001, 34, 9143. 121. Lodge, T. P.; Pan, C.; Jin, X.; Liu, Z.; Zhao, J.; Maurer, W. W.; Bates, F. S. /. Polym. ScL, Polym. Phys. Ed. 1995, 33, 2289. 122. Lodge, T. P.; Xu, X.; Ryu, C. Y.; Hamley, I. W.; Fairclough, J. P. A.; Ryan, A. J.; Pedersen, J. S. Macromolecules 1996, 29, 5955. 123. Fredrickson, G. H.; Leibler, L. Macromolecules 1989, 22, 1238. 124. Olvera de la Cruz, M. J. Chem. Phys. 1989, 90, 1995. 125. Hashimoto, T.; Shibayama, M.; Kawai, H. Macromolecules 1983, 16, 1093. 126. Bates, F. S.; Rosedale, J. H.; Fredrickson, G. H.; Glinka, C. J. Phys. Rev. Lett. 1988, 61, 2229. 127. Mori, K.; Okawara, A.; Hashimoto, T. J. Chem. Phys. 1996, 104, 7765. 128. Yu, G.-E.; Li, H.; Fairclough, J. P. A.; Ryan, A. J.; McKeown, N.; Ali-Adib, Z.; Price, C.; Booth, C. Langmuir 1998, 14, 5782. 129. Fredrickson, G. H.; Larson, R. G. J. Chem. Phys. 1987, 86, 1553. 130. Fredrickson, G. H.; Helfand, E. /. Chem. Phys. 1988, 89, 5890. 131. Lodge, T. P.; Dalvi, M. C. Phys. Rev. Lett. 1995, 75, 657. 132. Balsara, N. P.; Stepanek, P.; Lodge, T. P.; Tirrell, M. Macromolecules 1991, 24, 6227. 133. McConnell, G. A.; Gast, A. P.; Huang, J. S.; Smith, S. D. Phys. Rev. Lett. 1993, 71, 2102. 134. McConnell, G. A.; Lin, E. K.; Gast, A. P.; Huang, J. S.; Lin, M. Y; Smith, S. D. Faraday Discuss. 1994, 98, 121. 135. McConnell, G. A.; Gast, A. P. Macromolecules 1997, 30, 435. 136. McConnell, G. A.; Gast, A. P. Phys. Rev. E 1996, 54, 5447. 137. Hamley, I. W.; Daniel, C.; Mingvanish, W.; Mai, S.-M.; Booth, C.; Messe, L.; Ryan, A. J. Langmuir 2000, 16, 2508. 138. Hamley, I. W.; Pople, J. A.; Fairclough, J. P. A.; Terrill, N. J.; Ryan, A. J.; Booth, C.; Yu, G.-E.; Diat, O.; Almdal, K.; Mortensen, K.; Vigild, M. /. Chem. Phys. 1998, 108, 6929. 139. Hamley, I. W.; Pople, J. A.; Diat, O. Colloid Polym. Sci. 1998, 276, 446. 140. Lodge, T. P.; Bang, J.; Park, M. J.; Char, K. Phys. Rev. Lett. 2004, 92, 145501. 141. Watzlawek, M.; Likos, C. N.; Lowen, H. Phys. Rev. Lett. 1999, 82, 5289. 142. Bang, J.; Lodge, T. P.; Wang, X.; Brinker, K. L.; Burghardt, W. R. Phys. Rev. Lett. 2002, 89, 215505. 143. Bang, J.; Lodge, T. P. J. Phys. Chem. B 2004, 108, 12071. 144. Alexander, S.; McTague, J. Phys. Rev. Lett. 1978, 41, 702. 145. Liu, T.; Chu, B. J. Appl. Cryst. 2000, 33, 727. 146. Fuse, C.; Okabe, S.; Sugihara, S.; Aoshima, S.; Shibayama, M. Macromolecules 2004, 37, 7791. 147. Wang, C.-Y; Lodge, T. P. Macromolecules 2002, 35, 6997. 148. Wang, C.-Y; Lodge, T. P. Macromol. Rapid Commun. 2002, 23, 49. 149. Chastek, T. Q.; Lodge, T. P. Macromolecules 2003, 36, 7672. 150. Goveas, J. L.; Milner, S. T. Macromolecules 1997, 30, 2605. 151. Chastek, T. Q.; Lodge, T. P. Macromolecules 2004, 37, 4891. 152. Nie, H.; Bansil, R.; Ludwig, K. F; Steinhart, M.; Konak, C.; Bang, J. Macromolecules 2003, 36, 8097. 153. Boker, A.; Knoll, A.; Elbs, H.; Abetz, V.; Miiller, A. H. E.; Krausch, G. Macromolecules 2002, 35, 1319.
Concentrated Solutions
169
154. Boker, A.; Elbs, H.; Hansel, H.; Knoll, A.; Ludwigs, S.; Zettl, H.; Urban, V.; Abetz, V.; Miiller, A. H. E.; Krausch, G. Phys. Rev. Lett. 2002, 89, 135502. 155. Zvelindovsky, A. V.; Sevink, G. J. A. Phys. Rev. Lett. 2003, 90, 049601. 156. Boker, A.; Abetz, V.; Krausch, G. Phys. Rev. Lett. 2003, 90, 049602. 157. Shibayama, M.; Hashimoto, T.; Kawai, H. Macromolecules 1983, 16, 1434. 158. Mori, K.; Hasegawa, H.; Hashimoto, T. Polymer 2001, 42, 3009. 159. Whitmore, M. D.; Noolandi, J. J. Chem. Phys. 1990, 93, 2946. 160. Shibayama, M.; Hashimoto, T.; Hasegawa, H.; Kawai, H. Macromolecules 1983, 16, 1427. 161. Mori, K.; Hasegawa, H.; Hashimoto, T. Polymer 1990, 31, 2368. 162. Mayes, A.; Barker, J. G.; Russell, T. P. J. Chem. Phys. 1994, 101, 5213. 163. Lai, C.; Russel, W. B.; Register, R. A. Macromolecules 2002, 35, 4044. 164. Lodge, T. P.; Hamersky, M. W.; Hanley, K. J.; Huang, C.-I. Macromolecules 1997, 30, 6139. 165. Huang, C.-I.; Chapman, B. R.; Lodge, T. P.; Balsara, N. P. Macromolecules 1998, 31, 9384. 166. Helfand, E.; Tagami, Y. J. Chem. Phys. 1972, 56, 3592. 167. Naughton, J. R.; Matsen, M. W. Macromolecules 2002, 35, 5688. 168. Hong, K. M.; Noolandi, J. Macromolecules 1983, 16, 1083. 169. Leibler, L. Macromolecules 1980, 13, 1602. 170. Helfand, E.; Wasserman, Z. R. Macromolecules 1976, 9, 879. 171. Helfand, E.; Wasserman, Z. R. Microdomain Structure and the Interface in Block Copolymers. In Developments in Block Copolymers 1; Goodman, L, Ed. Applied Science: London, 1982; pp 99. 172. Whitmore, M. D.; Vavasour, J. D. Macromolecules 1992, 25, 2041. 173. Brazovskii, S. A. Sov. Phys. JETP 1975, 41, 85. 174. Fredrickson, G. H.; Helfand, E. /. Chem. Phys. 1987, 87, 697. 175. de Gennes, P. G. Scaling Concepts in Polymer Physics. Cornell University Press: Ithaca, 1979. 176. Bassler, K. E.; Olvera de la Cruz, M. J.Phys.France 11993, 3, 2387. 177. Fredrickson, G. H. J. Chem. Phys. 1986, 85, 5306. 178. Daoud, M.; Cotton, J. P.; Farnoux, B.; Jannink, B.; Benoit, H.; Duplessix, R.; Picot, C.; de Gennes, P. G. Macromolecules 1975, 8, 804. 179. Brown, W.; Mortensen, K.; Floudas, G. Macromolecules 1992, 25, 6904. 180. Hamada, R; Kinugasa, S.; Hayashi, H.; Nakajima, A. Macromolecules 1985, 18, 2290. 181. King, J. S.; Boyer, W.; Wignall, G. D.; Ullman, R. Macromolecules 1985, 18, 709. 182. Duplessix, R.; Cotton, J. P.; Benoit, H.; Picot, C. Polymer 1979, 20, 1181. 183. Guenza, M.; Schweizer, K. S. Macromolecules 1997, 30, 4205. 184. Liu, Z.; Kobayashi, K.; Lodge, T. P. J. Polym. Sci. B: Polym. Phys. 1998, 36, 1831. 185. Noolandi, J.; Shi, A.-C; Linse, P. Macromolecules 1996, 29, 5907. 186. Shusharina, N. P.; Alexandridis, P.; Linse, P.; Balijepalli, S.; Gruenbauer, H. J. M. Eur. Phys. J. E 2003, 10, 45. 187. Huang, C.-I.; Lodge, T. P. Macromolecules 1998, 31, 3556. 188. van Vlimmeren, B. A. C.; Maurits, N. M.; Zvelinodvsky, A. V.; Sevink, G. J. A.; Fraaije, J. G. E. M. Macromolecules 1999, 32, 646. 189. Zvelindovsky, A. V. M.; van Vlimmeren, B. A. C.; Sevink, G. J. A.; Maurits, N. M.; Fraaije, J. G. E. M. J. Chem. Phys. 1998, 709, 8751.
170
Block Copolymers in Solution: Fundamentals and Applications
190. Sevink, G. J. A.; Fraaije, J. G. E. M.; Huinink, H. P. Macromolecules 2002, 35, 1848. 191. Larson, R. G. J. Phys. France II1996, 6, 1441. 192. Srinivas, S.; Discher, D. E.; Klein, M. L. Nature Mater. 2004, 3, 638. 193. Hillmyer, M. A.; Bates, F. S.; Almdal, K.; Mortensen, K.; Ryan, A. J.; Fairclough, J. P. A. Science 1996, 277, 976. 194. Mortensen, K. /. Phys., Condens. Matt. 1996, 8, 103. 195. Mortensen, K. Coll. Surf. A 2001, 183-185, 211. 196. Hamley, I. W. Curr. Opin. Colloid Interface Sci. 2000, 5, 342. 197. Mortensen, K. Polym. Adv. Technol. 2001, 12, 2. 198. Balsara, N. P.; Hammouda, B. Phys. Rev. Lett. 1994, 72, 360. 199. Balsara, N. P.; Hammouda, B.; Kesani, P. K.; Jonnalagadda, S. V; Straty, G. C. Macromolecules 1994, 27, 2566. 200. Wang, H.; Kesani, P. K.; Balsara, N. P.; Hammouda, B. Macromolecules 1997, 30, 982. 201. Zryd, J. L.; Burghardt, W. R. Macromolecules 1998, 31, 3656. 202. Zipfel, J.; Lindner, P.; Tsianou, M.; Alexandridis, P.; Richtering, W. Langmuir 1999, 75, 2599. 203. Zipfel, J.; Berghausen, J.; Schmidt, G.; Lindner, P.; Alexandridis, P.; Tsianou, M.; Richtering, W. Phys. Chem., Chem. Phys. 1999, 1, 3905. 204. Tepe, T.; Schulz, M. E; Zhao, J.; Tirrell, M.; Bates, F. S.; Mortensen, K.; Almdal, K. Macromolecules 1995, 28, 3008. 205. Balsara, N. P.; Dai, H. J. J. Chem. Phys. 1996, 705, 2942. 206. Balsara, N. P.; Dai, H. J.; Kesani, P. K.; Garetz, B. A.; Hammouda, B. Macromolecules 1994, 27, 7406. 207. Schmidt, G.; Richtering, W.; Lindner, P.; Alexandridis, P. Macromolecules 1998, 37, 2293. 208. Oswald, P.; Geminard, J. C.; Lejcek, L.; Sallen, L. J. Phys. France II1996, 6, 281. 209. Schmidt, G.; Miiller, S.; Lindner, P.; Schmidt, C.; Richtering, W. J. Phys. Chem. B 1998, 702, 507. 210. Imperor-Clerc, M.; Hamley, I. W.; Davidson, P. Macromolecules 2001, 34, 3503. 211. Pople, J. A.; Hamley, I. W.; Fairclough, J. P. A.; Ryan, A. J.; Booth, C. Macromolecules 1998, 37, 2952. 212. Pople, J. A.; Hamley, I. W.; Terrill, N. J.; Fairclough, J. P. A.; Ryan, A. J.; Yu, G.-E.; Booth, C. Polymer 1998, 39, 4891. 213. Phoon, C. L.; Higgins, J. S.; Allegra, G.; van Leeuwen, P.; Staples, E. Proc. R. Soc. Land. A 1993, 442, 221. 214. Castelletto, V.; Hamley, I. W.; Holmqvist, P.; Rekatas, C.; Booth, C.; Grossmann, J. G. Colloid Polym. Sci. 2001, 279, 621. 215. McConnell, G. A.; Lin, M. Y.; Gast, A. P. Macromolecules 1995, 28, 6754. 216. Leyh, B.; Creutz, S.; Gaspard, J.-P.; Bourgaux, C.; Jerome, R. Macromolecules 1998, 37, 9258. 217. Berret, J.-E; Molino, E; Porte, G.; Diat, O.; Lindner, P. J. Phys., Condens. Matt. 1996, 8, 9513. 218. Diat, O.; Porte, G.; Berret, J.-F. Phys. Rev. B 1996, 54, 14869. 219. Molino, F. R.; Berret, J.-E; Porte, G.; Diat, O.; Lindner, P. Eur. Phys. J. B 1998, 3, 59. 220. Eiser, E.; Molino, E; Porte, G. Eur. Phys. J. E 2000, 2, 39. 221. Daniel, C.; Hamley, I. W.; Mingvanish, W.; Booth, C. Macromolecules 2000, 33, 2163.
Concentrated Solutions
171
222. Kleppinger, R.; Mischenko, N.; Theunissen, E.; Reynaers, H. L.; Koch, M. H. J.; Almdal, K.; Mortensen, K. Macromolecules 1997, 30, 7012. 223. Eiser, E.; Molino, R; Porte, G.; Diat, O. Phys. Rev. E 2000, 61, 6759. 224. Ackerson, B. J.; Clark, N. A. Phys. Rev. A 1984, 30, 906. 225. Almdal, K.; Koppi, K.; Bates, F. S. Macromolecules 1993, 26, 4058. 226. Hamley, I. W.; Pople, J. A.; Booth, C.; Yang, Y.-W.; King, S. M. Langmuir 1998, 14, 3182. 227. Hamley, I. W.; Pople, J. A.; Booth, C.; Derici, L.; Imperor-Clerc, M.; Davidson, P. Phys. Rev. E 1998, 58, 7620. 228. Hull, D.; Bacon, D. J. Introduction to Dislocations. Pergamon: Oxford, 1984. 229. Hamley, I. W.; Mortensen, K.; Yu, G.-E.; Booth, C. Macromolecules 1998, 31, 6958. 230. Mortensen, K. Europhys. Lett. 1992, 19, 599. 231. Mortensen, K. Progr. Colloid Polym. Sci. 1993, 91, 69. 232. Watanabe, H.; Kanaya, T.; Takahashi, Y. Macromolecules 2001, 34, 662. 233. Doi, M.; Harden, J. L.; Ohta, T. Macromolecules 1993, 26, 4935. 234. Mortensen, K.; Theunissen, E.; Kleppinger, R.; Almdal, K.; Reynaers, H. Macromolecules 2002, 35, 7773. 235. Sebastian, J. M.; Lai, C.; Graessley, W. W.; Register, R. A. Macromolecules 2002, 35, 2707. 236. Witten, T. A.; Pincus, P. A. Macromolecules 1986, 19, 2509. 237. Sebastian, J. M.; Lai, C.; Graessley, W. W.; Register, R. A.; Marchand, G. R. Macromolecules 2002, 35, 2700. 238. Sebastian, J. M.; Graessley, W. W.; Register, R. A. J. Rheol. 2002, 46, 863. 239. Reynders, K.; Mischenko, N.; Mortensen, K.; Overbergh, N.; Reynaers, H. Macromolecules 1995, 28, 8699. 240. Mischenko, N.; Reynders, K.; Mortensen, K.; Overberg, N.; Reynaers, H. J. Polym. Sci. B: Polym. Phys. 1996, 34, 2739. 241. Kleppinger, R.; van Es, M.; Mischenko, N.; Koch, M. H. J.; Reynaers, H. Macromolecules 1998, 31, 5805. 242. Jian, T.; Anastasiadis, S. H.; Fytas, G.; Adachi, M.; Kotaka, T. Macromolecules 1993, 26, 4706. 243. Jian, T.; Anastasiadis, S. H.; Semenov, A. N.; Fytas, G.; Adachi, K.; Kotaka, T. Macromolecules 1994, 27, 4762. 244. Pan, C.; Maurer, W.; Liu, Z.; Lodge, T. P.; Stepanek, P.; von Meerwall, E. D.; Watanabe, H. Macromolecules 1995, 28, 1643. 245. Benmouna, M.; Duval, M.; Borsali, R. J. Polym. Sci. B: Polym. Phys. 1987, 25, 1839. 246. Benmouna, M.; Benoit, H.; Borsali, R.; Duval, M. Macromolecules 1987, 20, 2620. 247. Liu, Z.; Pan, C.; Lodge, T. P.; Stepanek, P. Macromolecules 1995, 28, 3221. 248. Lodge, T. P.; Blazey, M. A.; Liu, Z.; Hamley, I. W. Macromol Chem. Phys. 1994, 198, 983. 249. Sigel, R.; Pispas, S.; Vlassopoulos, D.; Hadjichristidis, N.; Fytas, G. Phys. Rev. Lett. 1999, 83, 4666. 250. Sigel, R.; Pispas, S.; Hadjichristidis, N.; Vlassopoulos, D.; Fytas, G. Macromolecules 1999, 32, 8447. 251. Loppinet, B.; Sigel, R.; Larsen, A.; Fytas, G.; Vlassopoulos, D.; Liu, G. Langmuir 2000, 16, 6480. 252. Vlassopoulos, D. J. Phys.: Condens. Matter 2001, 13, R855.
172
Block Copolymers in Solution: Fundamentals and Applications
253. Mortensen, K.; Brown, W.; J0rgensen, E. Macromolecules 1995, 28, 1458. 254. Pusey, P. N.; van Megen, W. Phys. Rev. Lett. 1987, 59, 2083. 255. Pham, K. N.; Puertas, A. M.; Bergenholtz, J.; Egelhaaf, S. U.; Mousaaid, A.; Pusey, P. N.; Schofield, A. B.; Gates, M. E.; Fuchs, M.; Poon, W. C. K. Science 2002, 296, 104. 256. Poon, W. C. K. MRS Bull. 2004, 29, 96. 257. Bergenholtz, J.; Fuchs, M. Phys. Rev. E 1999, 59, 5706. 258. Fabbian, L.; Gotze, W.; Scortino, R; Tartaglia, P.; Thiery, F. Phys. Rev. E 1999, 59, R1347. 259. Mallamace, R; Gambadauro, P.; Micali, N.; Tartaglia, P.; Liao, C.; Chen, S.-H. Phys. Rev. Lett. 2000, 84, 5431. 260. Grandjean, J.; Mourchid, A. Europhys. Lett. 2004, 65, 712. 261. Chen, S.-H.; Chen, W.-R.; Mallamace, R Science 2003, 300, 619. 262. Castelletto, V.; Hamley, I. W.; Waigh, T. A. /. Chem. Phys. 2004, 727, 11474. 263. Bhatia, S. R.; Mourchid, A. Langmuir 2002, 18, 6469. 264. Konak, C.; Helmstedt, M.; Bansil, R. Macromolecules 1997, 30, 4342. 265. Konak, C.; Fleischer, G.; Tuzar, Z.; Bansil, R. J. Polym. Sci. B: Polym. Phys. 2000, 38, 1312. 266. Nystrom, B.; Walderhaug, H.; Hansen, F. K. J. Phys. Chem. 1993, 97, 7743. 267. Raspaud, E.; Lairez, D.; Adam, M.; Carton, J.-P. Macromolecules 1996, 29, 1269. 268. Kapnistos, M.; Vlassopoulos, D.; Fytas, G.; Mortensen, K.; Fleischer, G.; Roovers, J. Phys. Rev. Lett. 2000, 85, 4072. 269. Loppinet, B.; Stiakakis, E.; Vlassopoulos, D.; Fytas, G.; Roovers, J. Macromolecules 2001, 34, 8216. 270. Stiakakis, E.; Vlassopoulos, D.; Loppinet, B.; Roovers, J.; Meier, G. Phys. Rev. E 2002, 66, 051804. 271. Godward, J.; Heatley, R; Smith, S.; Tanodekaew, S.; Yang, Y.-W.; Booth, C. J.Chem.Soc., Faraday Trans. 1995, 97, 3461. 272. Godward, J.; Heatley, R; Booth, C. J. Chem. Soc., Faraday Trans. 1995, 97, 1491. 273. Khandpur, A. K.; Porster, S.; Bates, R S.; Hamley, I. W.; Ryan, A. J.; Bras, W.; Almdal, K.; Mortensen, K. Macromolecules 1995, 28, 8796. 274. Ryu, C. Y; Lodge, T. P. Macromolecules 1999, 32, 7190. 275. Cast, A. P. Langmuir 1996, 72, 4060.
4 Polyelectrolyte Block Copolymers 4.1 4.1.1
MICELLIZATION GENERAL REMARKS
As for nonionic block copolymers, micellization occurs in a solvent that is selective for one of the blocks. However, the ionic character introduces new parameters governing the structure and properties of the micellar solution. To date, there have been relatively few reviews focused on the subject.1'2 Early work by Selb and Gallot, largely on PS-b-qP4VP diblocks, was reviewed by them3 and also by Hamley.' The properties of polyelectrolyte block copolymers in solution depend greatly on salt concentration. Regimes of 'salted brush' and 'osmotic brush' behaviour have been identified, depending on whether the added salt concentration is greater or less than the counterion concentration in the corona of the micelle (Figure 4.1).4 A further 'Pincus regime'5 where the Coulombic interaction among coronal chains is unscreened because counterions are located outside the corona has also been identified, as mentioned in Section 4.3 below. The strength of the polyelectrolyte plays an important role in micellization in polyelectrolyte-containing block copolymers.6"8 In strongly charged polyelectrolytes, such as poly(styrene sulfonate), the ionic units are essentially fully dissociated, irrespective of ionic strength or pH. Polyelectrolytes in which the number and arrangement of charges is fixed are termed 'quenched' poly electrolytes. In 'annealed' polyelectrolytes the degree of dissocation depends on pH, as the total number and position of charges is not fixed. For example, for polyacrylic acid, at low pH ~ pKa the degree of dissociation is suppressed due to trapped counterions. Addition of salt can lead to substitution of counterions (e.g. Na+ for H+) and an increase in ionization. Due to the difference in pH within the micelle and in bulk solution, large changes in aggregation number and shape in response to changes in ionic strength are possible for this type of polyelectrolyte. For polyelectrolytes the ionic strength plays an important role in the chain conformation, and the presence of a high charge density leads to some specific properties unique to ionic block copolymers. Many of the studies on ionic block
Block Copolymers in Solution: Fundamentals and Applications © 2005 John Wiley & Sons, Ltd.
I. W. Hamley
174
Figure 4.1
Block Copolymers in Solution: Fundamentals and Applications
(a) Salted and (b) osmotic brushes in block copolymer micelle coronas.
copolymers have been undertaken with solvents selective for the ionic polyelectrolyte block, generally water or related solvents, such as water/methanol mixtures. However, it has been observed that it is often difficult to dissolve hydrophilic-hydrophobic block copolymers in water. These dissolution problems are far more pronounced than for block copolymers in nonaqueous selective solvents, although they do not always reflect real insolubility. In many cases, dissolution can be achieved if a better solvent is used first, as discussed in the next section. This process however can influence the charge and hence conformation of the polyelectrolyte block, possibly leading to trapped nonequilibrium structures. Other important parameters governing the self-assembly of ionic block copolymers include the nature of the backbone, in particular whether it is hydrophilic (e.g. PAA) or hydrophobic (e.g. PSS). For the former, charge effects govern chain conformation whereas in the latter case, chain configuration depends both on charge and on the effect of hydrophobicity. For polyelectrolytes with a hydrophobic backbone, electrostatic interactions tend to favour an extended conformation, whereas a collapsed configuration reduces hydrophobic interactions. This leads to a pearl necklace conformation.9'10 This consists of regions of charged globules ('pearls') separated by narrow elongated strings. The formation of this conformation has been explained theoretically by analogy with the Rayleigh instability of a charged droplet, whereby addition of charge eventually exceeds the interfacial tension. 11-13 The flexibility of the chain is also important-the influence of charge on flexible chains will differ from that of semi-flexible diblocks or rod-coil systems. The latter type of behaviour is exemplified by block copolymers containing charged peptide 'rod' units, as discussed below. Aggregation into micelles of polyelectrolyte block copolymers modifies the electrostatic component of the interaction potential compared with that for unagreggated polyelectrolytes. In the former case the electrostatic potential is spherically symmetric (and decays inversely with distance), whereas in the latter case it is locally cylindrically symmetric. This has important consequences for counterion condensation.
Polyelectrolyte Block Copolymers
175
4.1.2 MICELLIZATION IN BLOCK COPOLYMERS CONTAINING ANIONIC BLOCKS The main class of anionic polyelectrolytes are polyacrylates and polymethacrylates, usually present as sodium (or other alkali) salts but also in acidic form at sufficiently high pH. Eisenberg and coworkers have investigated micellization of PS-b-PNaA diblock copolyelectrolytes in water over a number of years.14 The PS block length ranged from 6 to 100, while that of the polyelectrolyte varied from ca. 300 to ca. 1400.15 At a constant polyelectrolyte block length, increasing the hydrophobic block length lowered the cmc, as for nonionic block copolymers. Changing the soluble block length from 300 to 1400 typically changed the cmc values by less than a factor of 2, also in agreement with trends for neutral block copolymers. For very short PS blocks, the cmc decreased very rapidly with increasing length of the insoluble block, whereas for more than 12 repeat units, the drop in the cmc was more gradual.15 Micelles formed by PS-b-PC&A and PS-b-PCsMA in toluene, a selective solvent for the PS block, have been characterized using SAXS.16 Comparisons of the micelle core dimensions were made with the Zhulina-Birshtein and Halperin models for type IV polymeric micelles, as discussed in Section 2.7.1. The cmc of a family of PS-£-PNaA diblocks with fixed PS block length (660 repeat units) and 2.6 to 14 ionic block units in THF has been investigated using light scattering.17 Micelles formed by PS-£-PNaA and PS-b-PAA dissolved in this and other organic solvents have been characterized by the same group using a variety of techniques including size exclusion chromatography and static and dynamic light scattering.18 The cmcs and micellar dimensions of the same type of PS-&-PNaA copolymer in aqueous and NaCl salt solutions have been the subject of studies using fluorescence measurements19 and static light scattering experiments.20 A complex dependence of the cmc on block lengths was observed in the fluorescence experiments, with a maximum in the cmc as a function of the soluble PNaA block length.19 The effect of salt concentration on the association numbers, radii of gyration and second virial coefficients of micelles was investigated using static light scattering.20 It was found that the association number p increased as a function of salt concentration at low salt contents, but remained constant above ca. 0.10 M NaCl. The data were compared with the predictions of the neutral star polymer scaling theories, 21-23 and several of the various mean field models for block copolymer micelles (as discussed in Section 2.7).24-26 Good agreement was found with several of the models, although it was not possible to descriminate between the predictions of the theories with the data available.20 'Crew-cut' micelles with a short coronal block attached to a long core block have been the focus of much work by the Eisenberg group.27-32 They first reported the formation of such micelles in aqueous solutions of PS-b-P4VP diblocks quaternized with methyl iodide.32 In subsequent work, crew-cut micelles of highly asymmetric PS-b-PAA28 and nonionic PS-b-PEO31 diblocks in water were studied using TEM. Multiple micellar morphologies for the former were observed28'30 for copolymers
176
Block Copolymers in Solution: Fundamentals and Applications
with 80-98 % PS. Micellization occured in the range 3-6 wt% polymer, depending on its composition.30 For a series of diblocks with a fixed number of PS repeats (200) transitions from spheres to rods to vesicles to compound micelles was observed on decreasing the PAA block length, as illustrated in Figure 4.2. The micrometre-size compound micelles were observed to have an internal structure made up of a core containing reverse micelles and PS in the corona [Figure 4.2(e) and (f)]. Such compound structures have been predicted theoretically for ABC
Figure 4.2 Morphologies formed by PS-b-PAA diblocks in aqueous solution: (a) S200-bAA21; (b) S200-b-AA15; (c) S200-b-AA8; (d) S20o-^-AA4.28 (e) Enlargement of one of the large complex micelles in (d) showing the internal structure. The elongation in one direction was probably caused by the strong shear forces during microtoming. (f) A schematic of the structure of the large complex micelle filled with bulk reverse micelles. Reproduced by permission of Science.
Polyelectrolyte Block Copolymers
177
triblock copolymer melts in the 'superstrong segregation' limit.33 In this limit, segregation between, say, B and C blocks can occur within micelles in the A matrix. A similar effect is likely to underly the observation of compound micelles in block copolymer solutions. Zhang and Eisenberg occasionally observed lamellar structures for compositions where vesicles were otherwise seen (i.e. an alternative bilayer structure), and simple reverse micelle-like aggregates were also noted.28'30 The addition of homopolystyrene was observed to drive transitions from the bilayer or cylindrical morphologies into the spherical micelle structure.30 A transition between morphologies of crew-cut micelles was also observed for nonionic PS-bPEO diblocks, in particular a transition from spherical to rod-like or vesicular structures was induced by increasing the length of the hydrophilic PEO blocks.31 The effect of ionic strength in inducing transitions between micellar structures was also explored for PS-b-PAA and PS-b-PEO diblocks in water.29 It was found that addition of ions in micromolar (CaCl2 or HC1) or millimolar (NaCl) quantities can change the morphology of PS-b-PAA micelles in dilute aqueous solutions. In particular, on increasing the acid or salt concentration, the morphology changed from spheres to rods to vesicles and then to large compound vesicles. Gelation could be also be induced with the addition of ions to the solution.29 Later, morphology diagrams for PS-b-PAA diblocks in dioxane/water were presented.34-36 In general, increasing water content favoured the sequence of morphologies: single chains-spheres - rods-bilayers-inverted structures, with coexistence regions between the various phases. A variety of complex morphologies was also observed for PS-b-PAA31 and PS-b-PEO38 diblocks by Eisenberg et al., as mentioned in Section 2.16. The polymers were dissolved in DMF/water mixtures. DMF is a good solvent for PS, PAA, and PEO whereas water is a nonsolvent for PS. The resulting structures are not in equilibrium when the water content in the mixed solvent is high (the range 7.5-9.5% water was studied), and the aggregate structure depends on the method of preparation.37 The aggregate structure also depends on the addition of electrolytes (which changes the conformation of the water-soluble block) and the preparation temperature.36-38 Structures with combinations of the morphological features mentioned above were also commonly observed.38 The influence of the initial common solvent on the final morphology in aqueous solution was studied by comparing the aggregate structures observed for PS-b-PAA diblocks dissolved in DMF, THF and dioxane and mixtures of these.39 The selectivity of the solvent for either block influences the association number and chain packing. A weakly selective solvent leads to a swollen core. Less polar solvents reduce the PAA-solvent interaction and hence reduce repulsive interactions between corona chains, leading to an increase in association number and the stretching of the PS core chains. The reversibility of morphological transitions upon addition/removal of water to a dioxane solution of diblock PS310-b-PAA52 was studied using cryo-TEM to determine morphology.34 The binary phase diagram for this copolymer is shown in Figure 4.3. The transitions follow the sequence expected based on interfacial curvature/molecular packing considerations. The reversibility of vesicle formation and growth is discussed further in Section 2.17.
178
Block Copolymers in Solution: Fundamentals and Applications
Figure 4.3 Binary morphology diagram for diblock S310-b-AA52 in THF/water mixtures as a function of polymer concentration and water content in the solvent mixture.34,182 Reproduced by permission of Science.
Another interesting morphology observed by Eisenberg and coworkers, again a nonequilibrium structure resulting from the solvent exchange during preparation, is bowl-shaped micelles.40 These were observed for crew-cut micelles formed by PAI-b-PS-b-PAI triblocks initially dissolved in a common solvent and then dialysed against water to induce association of the PS chains. It was proposed that the bowlshaped structure forms from initially spherical compound micelles due to the formation and coalescence (followed by break-through to the surface) of waterfilled bubbles as water is added, and solvent is extracted from the core leading to PS vitrification, as illustrated in Figure 4.4.
Figure 4.4 (a) Proposed mechanism of formation of bowl-shaped structure, (b) Micrographs illustrating the steps (selected from different micrographs) for a PAI-b PS-bPAI triblock.40 Reproduced by permission of American Chemical Society.
Micellization in di- and triblock copolymers with a PMAA polyelectrolyte block and one or two hydrophobic blocks bearing a perfluorinated side chain has been investigated.41 Despite the expected high hydrophobicity, the copolymers were found not to be very surface active. In exceptional cases, surface tension measurements can be insensitive to micellization. Matsuoka and coworkers have observed micelle formation in
Polyelectrolyte Block Copolymers
179
PS-b-PSS diblocks via dye solubilization and dynamic light scattering.42 However, no decrease in surface tension with increasing concentration over the appropriate range was observed, and no adsorption at the air/water interface was detected by X-ray reflectivity. When salt was added, surface activity was noted and foam formation was detected. For the unscreened solutions, it was proposed that the highly charged nature of PSS was due to an image charge effect. The image charges are significant because of the dielectric contrast between air and water. A strong electrostatic repulsion of PSS from the interface results. Image charge repulsion was considered theoretically by Wittmer and Joanny43 who showed that the corresponding electrostatic energy can exceed thermal energy in the early stage of adsorption when there are lateral variations in charge density. 4.1.3 MICELLIZATION IN BLOCK COPOLYMERS CONTAINING CATIONIC BLOCKS Armes and coworkers have published extensively on the solution properties of tertiary amine methacrylate block copolymers (Figure 4.5). These are prepared by atom transfer radical polymerization or group transfer polymerization. For these ionic block copolymers, the size and association number of the micelles can be systematically controlled by changing the temperature, pH or ionic strength of the solution. The properties are exemplified by the PDMA-b-PDEA system. Diblocks of this type dissolve molecularly at low pH, but as pH is increased deprotonation occurs and the PDEA block becomes hydrophobic, leading to micellization.44'45 At low pH, the copolymers can be considered to be double hydrophilic.44 The micellization of block copolymers containing the permanently hydrophobic PMMA block and PDMA has also been investigated, via static and dynamic light scattering, ultracentrifugation and surface tension measurements.46 For these copolymers, the cmc was found to increase with the length of the hydrophobic block if the hydrophilic block length was fixed, in contrast to the trend for most neutral and ionic block copolymers. Increasing the overall molar mass of the copolymer also increased the cmc for a given composition. Increasing the overall molar mass of the copolymer for a given composition produced larger micelles with a lower association number. Both micelle size and association number decreased with the length of the hydrophobic block if the hydrophilic block length was fixed. Another key concept introduced by Armes and coworkers is that of so-called 'schizophrenic' self-assembly, whereby the core and corona blocks can be interchanged according to solution pH and electrolyte concentration. An example is the PMEMA-b-PDEA system, as illustrated schematically in Figure 4.6, although several other systems exhibit this behaviour, as described in a recent review.47 When the micellar core is charged, ionomer effects become important.48 This is because the concentration of solvent in the core is minimal so that counterions and polyelectrolyte chains remain bound, forming dipoles, as the interaction between charges is not reduced by the dielectric constant of the solvent.
180
Block Copolymers in Solution: Fundamentals and Applications
Figure 4.5 Repeat units in tertiary amine block copolymers studied by Armes and coworkers.183 Reproduced by permission of American Chemical Society.
Work by the Armes group on micellization in tertiary amine methacrylate block copolymers is summarized in Table 4.1. The list covers most aspects of their work, but is not exhaustive. A review from 2001 discusses the development of polymeric amphiphiles by Armes' and other groups up to that date.49
Figure 4.6 Formation of micelles and reverse micelles of a PDEA-b-PMEMA diblock according to solution conditions.59 Reproduced by permission of American Chemical Society.
Poly electrolyte Block Copolymers
181
Table 4.1 Studies on micellization in tertiary amine methacrylate-based block copolymers by Armes and coworkers Copolymer type
Study
PMMA-b-PDMA
Dependence of cmc and association number on block composition and molar mass Micellization/micelle dimensions and structure pH-dependent micellization. Added salt favours micellization due to charge screening. Detailed SANS study of micelle structure Reversible micellization upon adjustment of pH, temperature or electrolyte concentration
PDMA-b-PDEA PDMA-b-PDEA
PDMA-b-PDEA PDMA-b-PDPA PDMA-b-PMEMA PMDPS-b-PMMA PMDPS-b-poly(alkyl methacrylates) qPDMA-b-PDEA PNaSS-fc-PNaSCOO
PDMA-b-PMAA PDMA-b-PEO PDEA-b-PEO PDMS-b-PDMA
PMEMA-b-PDMA PMEMA-b-PDEA PPO-b-PDEA
PVBA-ZP-PDEA
PVBA-b-PMEMA
Former block obtained by betainization of PDMA. Micellar characteristics obtained Reduced surface activity upon betaine formation. Polydisperse micelle formation, unless dialysed against nonselective solvent Quaternization of PDMA suppresses LCST behaviour and enhances pH range of micelle stability pH responsive micellization of strong/weak acid double hydrophilic diblocks. Polymerization using NaPSS as macro-initiator with other polymers also examinedunsuccessful Zwitterionic system. Thermoreversible micellization Micellization at high temperature Micellization at high pH Micellization in organic solvents as well as water. High surface activity, particularly in alkaline solution. Commercial relevance of silicone surfactants Schizophrenic micellization Schizophrenic micellization Schizophrenic micellization Schizophrenic micellization of a zwitterionic diblock Schizophrenic micellization of a zwitterionic diblock
Reference 46 44, 45 45
51
52 53
54 55
56 57 57 58
59 60 61 62 63
Micellization in related copolymers has been studied by Gohy et a/.50 who prepared P2VP-b-PDMA diblocks. At low pH polydisperse loose aggregates were formed, and micelles with a protonated PDMA corona were observed at intermediate pH. At higher pH, the PDMA corona becomes uncharged, leading to micellar aggregation. Armes and coworkers have also investigated micellization in vinyl ether-based block copolymers prepared by living cationic polymerization. Micellization in aqueous solutions of diblocks of PMTEGVE (hydrophilic) and PIBVE (hydrophobic) has been characterized using dynamic light scattering, aqueous GPC and
182
Block Copolymers in Solution: Fundamentals and Applications
dye solubilization.64 Dihydrophilic block copolymers of PMVE and PMTEGVE65 or PVA66 have been shown to form micelles at elevated temperatures, due to the hydrophobic nature of PVME above the cloud point. Proton NMR spectroscopy confirmed that the PVME block formed the micellar core, and that micellization was reversible. Reversible schizophrenic micellization in neutral-fr-anionic OEGMA-b-PMAA diblocks has also been studied.67
4.1.4
MICELLIZATION OF POLYAMPHOLYTE BLOCK COPOLYMERS
Polyampholytes contain monomers with opposite charges.48 Block copolyampholytes contain blocks with opposite charges. There have been few studies to date on this type of system. The association properties of PDMA-b-PNaMA polyampholyte diblock copolymers has been investigated as a function of pH and the copolymer composition.68 At the isoelectric point aggregation led to phase separation. Moving away from the isoelectric point (balanced charge), charged micelles formed. The size of the aggregates increased with concentration but was almost independent of pH. This has also been noted for nonblock polyampholytes. Addition of salt also did not change the dimensions significantly. Aggregates were also still observed for pH > 10 or pH < 4 where one block was neutral. Schizophrenic micellization has been observed for a PVBA-b-PDEA polyampholyte diblock.62 Gohy et al. investigated the association of PDMA-b-PMAA diblocks.69 Depending on pH, the DMA units can be positively charged and the MAA unit neutral, the MAA units can be negatively charged and the DMA units neutral or both blocks can be charged. At the isoelectric point, strong electrostatic interactions lead to the formation of insoluble complexes. The electrostatic interactions can be screened by the addition of salt. Conditions for the formation of spherical micelles were discussed.
4.1.5 MICELLIZATION OF POLYELECTROLYTE-CONTAINING ABC TRIBLOCKS The solution properties of ABC triblocks containing a hydrophobic midblock and oppositely charged end blocks have been investigated.70 The PAI-b-PS-b-PMAA copolymers behaved as polyampholytes, with two inflection points observed during potentiometric titration (PMAA being deprotonated prior to PAI). Large vesicular structures (Rh = 150-250 nm) were observed, the internal vesicular structure depending on pH. The association properties of an ABC triblock with a polyelectrolyte midblock and short hydrophobic end blocks has been investigated for PS-fr-PNaA-b-PBMA in water.71 The high viscosity in dilute solution was ascribed to the formation of a transient network structure, due to association between the end blocks.
Polyelectrolyte Block Copolymers
183
The pH-dependent changes in micellar structure and intermicellar interactions of P2VP-b-PMMA-b-PAA triblocks in aqueous solution have been investigated in bulk by static and dynamic light scattering and for adsorbed micelles by AFM.72 P2VP is protonated at low pH, and this leads to polyampholytic heteroarm micelles for pH < 1. At higher pH, close to the isoelectric point, electrostatic interactions increase to the extent that phase separation was observed. In alkaline solution, nonequilibrium aggregates of heteroarm star micelles were observed.
4.1.6 MICELLIZATION OF BLOCK COPOLYMERS CONTAINING GRAFTED POLYELECTROLYTES Micellization of Pluronic block copolymers grafted with PAA which confers polyelectrolyte properties has been studied by surface tensiometry,73 viscometry73 and probe solubilization,74 the latter as a function of pH. Gel formation by this type of copolymer is discussed in Section 3.3.5.
4.1.7 MICELLIZATION IN BLOCK COPOLYMERS CONTAINING SULFONATED POLYISOPRENE The micellization of a nonionic PS-b-PI diblock in a PS-selective solvent has been compared with that of a PS-b-PI diblock modified by the addition of a single sulfonate group at the end of the PI block.75 The effect of added salt was also examined. The dimensions of both the micellar core and corona decreased upon addition of the charged endgroup, although addition of salt led to screening and an increase in micellar radius. This was interpreted using a polymer brush model. The micellization of ionically end-capped diblocks prepared from the same parent PS-b-PI diblock was later investigated by SANS.76 The ionic groups were located either on the soluble PS end (quaternary ammonium group) or insoluble PI block (sulfonate group) or both (zwitterionic case). Figure 4.7 schematically presents the findings concerning chain dimensions. The coronal chains in sample Q where the soluble PS chain is end functionalized are extended. In sample S where the insoluble block is end functionalized, the end groups were found to lie on the surface of the micellar core, leading to folding of the chains. The same was found for the zwitterionic end-capped diblock (Z), although here the coronal chain is also folded due to attractive interactions between the ionic end groups. The micellization of block copolymers containing the strong polyelectrolye, sulfonated polyisoprene (sPI) have been reported. Micellization of both PS-6-sPI77 diblock and PS-&-sPI-b-PS78 triblock has been examined. The effect of ionic strength (added NaCl concentration) and polymer concentration on micelle dimensions was studied via light scattering. Fluorescence experiments using pyrene probes were used to locate cmc values.
184
Block Copolymers in Solution: Fundamentals and Applications
Figure 4.7 Schematic of chain configuration deduced from SANS studies on micelles formed by ionically end-capped PS-fe-PI diblocks in a PS-selective solvent.76 Reproduced by permission of American Chemical Society.
4.2
CHAIN CONFORMATION
The structure of polyelectrolyte block copolymer micelles has been examined via SANS, static and dynamic light scattering and cryo-TEM.4 It was reported that at low added salt concentration, the PSSA polyelectrolyte shell in the PEE-b-PSSA diblock examined has typical features of a polymer brush. However, if the added salt concentration exceeds the intrinsic ionic strength of the polyelectrolyte shell, the micellar aggregation number increases due to the screening of the interaction between chains (salted brush behaviour) for this particular copolymer. Guenoun and coworkers suggest that changes in aggregation number depend on copolymer asymmetry79-an increase in association number p has been observed for the symmetric diblock studied by Forster et al., whereas p decreases for a highly asymmetric diblock. In the case of an intermediate asymmetry, p was found to be independent of added salt concentration.79,80 The corona contracts when the added salt concentration exceeds a threshold equal to the inner salt concentration in the corona in the absence of salt.79,81,82 The scaling of corona chain dimensions has been compared with previous results and the predictions of scaling theories. Forster et al. reported L ~ c~°'13, where cs is salt concentration. The exponent reported by Guenoun et al. was -0.14 for ?BS26-£-NaSS404, -0.11 for rBS27-£-NaSS75781 and —0.18 for EP52-b-NaSS25179 (an exponent —0.17 was noted for dense adsorbed layers of this copolymer on polystyrene82). Scaling theory predicts L ~ cs ' in the salted brush regime where salt concentration exceeds the counterion concentra-
Polyelectrolyte Block Copolymers
185
tion.83 A scaling L ~ c$ is predicted for polyelectrolyte star micelles with glassy cores, and also for equilibrium micelles in an appropriate regime of salt concentration.84 This scaling has been observed for polyelectrolyte brushes tethered to a flat PS interface in a PS-b-PAA diblock, in the salted brush regime.85 The transition from neutral to osmotic brush on increasing ionization was examined, and the increase in brush thickness probed. The same scaling has also been observed for a PEBS-b-PNaSS diblock adsorbed onto hydrophobized mica,86 as discussed in Section 5.2.2. The scaling behaviour of the association number of micelles ofn neutral-Z?polyelectrolyte diblocks as a function of core block length, p ~ ßB has been investigated by several groups. An exponent (3 = 1.62 or 1.89 was obtained for micelles of PS/PMAA diblock and triblock copolymers in a dioxane/water mixture containing 20% water.87 The dilute solutions were characterized using static and quasi-elastic light scattering. However, it was also found that the association number (and hydrodynamic radius) depended on corona block length NA with an exponent yu = —0.86 or —0.41 for di- and tri-blocks respectively. Since the core radius was not much smaller than the micellar radius, the micelles can be classified in regime III of Zhulina and Birshtein22 (Section 2.7), in which case a dependence of p and micellar radius on NA is expected. However, the exponents (3 and m were not in agreement with the scaling theory predictions (Table 2.3). Other work on PS-b-PAA diblocks in aqueous solution also provided scaling relationships for core radius RB not in agreement with theoretical predictions for nonionic block copolymer micelles.88 In addition to the ionic nature of the solutions studied, thermodynamic equilibrium may not have been reached in the solutions, as shown by the slow exchange kinetics of PS-&-PAA micelles in water such that they remained intact after elution from a GPC column, and as discussed further in Section 4.1.2. Micelles formed by PS-£-PCsA and PS-b-PCsMA in toluene, a selective solvent for the PS block, have been characterized using SAXS.16 Since the scattering was dominated by the cores, which were much smaller than the distance between them, the structure factor and form factor were well separated in the SAXS pattern, enabling detailed information to be extracted on the core dimensions. It was found that the core radius scales as N3/5 B, in agreement with the prediction of the ZhulinaBirshtein and Halperin models for type IV polymeric micelles. Also in agreement with these theories, the association number scaled as NJ and the surface area per 2/5 chain as NE' . Thus, these ionic block copolymer micelles obeyed the same scaling laws with core block length as micelles formed by neutral block copolymers. The conformation of the polyelectrolyte block in PtBS-b-PNaSS diblocks has been probed by SANS with contrast variation using H2O/D2O mixtures.79,89 The PNaSS block was found to adopt a rod-like conformation, as characterized by an intensity variation with wavenumber / ~ q~1 89 This behaviour was observed above a critical wavenumber which increased with salt concentration. The corresponding persistence length was surprisingly long, and proportional to the Debye screening length.79 A high persistence length of worm-like corona chains was observed even
186
Block Copolymers in Solution: Fundamentals and Applications
Figure 4.8 SANS intensity profiles obtained for a 1 wt% solution in D2O of a PtBS-&PNaSS diblock for salinities 5= 8.5 x 10~4 M (•), 10~2 M (A), 0.16 M (x) and 1 M (•).89 The critical salt concentration is S — 10~2 M. Deviations from the q~l behaviour can be observed in a Kratky plot (see Guenoun et al.).89 Reproduced by permission of American Physical Society.
for high salt concentrations (up to 1 M NaCl), i.e. despite screening of the charge. This is in striking contrast to the compact coil conformation adopted by uncharged polymer brushes. The extended conformation was adopted, up to a threshold salt concentration corresponding to the free counterion concentration. The free counterions contribute to an osmotic pressure which is reduced by screening when salt is added. The concentration of free counterions was estimated from the Manning model for counterion condensation. Figure 4.8 shows SANS data with the intensity varying as I(q) ~q~l, as expected for rod-like objects. Deviations from this behaviour set in above a salt concentration S « 10~2 M. These can be seen more clearly in a Kratky plot of q2I(q) versus q (not shown). Dan and Tirrell predicted p to be independent of salt concentration,90 whereas the experiments by Guenoun et al. suggested that this was only the case at low ionic strengths, since p decreased with salt concentration for higher ionic strengths.81 Later, the same group showed that the hydrodynamic radius obtained from pulsed field gradient NMR and fluorescence recovery after fringe pattern photobleaching (FRAP) was consistent with that obtained from DLS.91 Cryo-TEM also provided images of the micelles, confirming the existence of both spherical and elongated objects. The distribution of counterions formed by a PrBS-&-PNaSS diblock (with a lengthy polyelectrolyte block and short hydrophobic block) was probed by SAXS, using dialysis to change the counterion and hence the contrast.92 The counterions were found not to be homogeneously distributed in the extended rod-like brush corona. Instead, the
Poly electrolyte Block Copolymers
187
radial distribution function extracted from the partial structure factor suggested a Poisson-Boltzmann distribution of counterions around a coronal chain. The conformation of the coronal chains in spherical micelles formed by a PS-b-PCsA diblock in toluene (selective for PS) has been analysed by SANS using contrast variation.93 The conformation of PS chains as a function of distance from the ionic core was probed by selective deuteration of segments of the PS block, varying the position of the labelled block. The chains next to the core were rod-like as revealed by the / ~ q~1 variation of small-angle scattering intensity. The power law exponent decreased with distance of dPS blocks from the core. Semidilute behaviour of swollen coils, close to the / ~ g~ 5 / 3 scaling anticipated by the Daoud-Cotton model94 in the intermediate q range, was observed for segments well separated from the core. The contribution from inter-chain scattering also decreased with distance from the core. This scaling behaviour should not be influenced too much by the presence of a pronounced structure factor peak, since this occurred at significantly lower q. The presence of such a feature indicates interactions among the micelles. The chain conformation and distribution of counterions in the corona of a PS-b-PAA diblock have been studied by detailed SANS experiments.95 The polymer contained 50% ionized PAA units. Under these conditions the corona chains were found to be almost fully stretched. The counterions were found to be sequestered within the corona. There was no evidence of charge annealing (recombination and dissociation of the weak polyacid units) and/ or migration of counterions towards the outer coronal layer. The change in PAA coronal layer thickness as a function of degree of ionization was also studied by SANS.96 The core radius and aggregation number were found not to change significantly. In these studies, partial structure factors (PS-PS, PAA-PAA and those for the counterion) were obtained via contrast matching experiments, enabling detailed information about the location of particular species to be obtained. The behaviour of neutral-b-ionic block copolymers has been investigated in semidilute conditions (up to 20 wt% polymer).97 SANS, SAXS and light scattering were used to characterize PEP-b-PNaSS diblocks in aqueous solution. Liquid-like ordering of micelles was observed above a critical concentration, as confirmed by the scaling of SANS peak position with concentration, q* ~ c1/3. The structure of PS-b-PAA diblocks in water has also been investigated via SAXS and SANS up to concentrations where coronal chain overlap occurs.98 The same scaling behaviour of SAXS peak position was noted. The micelle dimensions decrease with increased packing fraction, irrespective of ionic strength and micellar charge. This is due to increased counterion adsorption and/or Donnan salt partitioning between the coronal layer and surrounding medium. At high salt concentration, where charge interactions are screened, the density profile decreases as r~ 4 / 3 , in agreement with scaling theory for star polymers and uncharged block copolymer micelles [Equation (2.23)]. At high charge and in the absence of salt, the polyelectrolyte chains remain almost fully stretched and they interdigitate once the volume fraction of micelles exceeds that for hard sphere crystallization. The interpenetration of the chains leads to the formation of a physical gel, as characterized by measurements of the dynamic
188
Block Copolymers in Solution: Fundamentals and Applications
shear moduli. The countenon distribution was determined from SAXS and found to follow closely the radial density profile of the corona segments.98 The distribution of counterions in the corona of PS-b-PAA micelles has been probed by anomalous SAXS." The technique relies on contrast created at the absorption edge of metal atoms/ions-in this case rubidium ions. The ions were found to be closely correlated to the corona chains, much more so than in linear polyelectrolytes due to interactions among the polyelectrolyte brushes 'grafted' to spherical surfaces. The results were compared with recent models for interacting polyelectrolyte brushes and star polyelectrolytes100,101 (this theory is discussed in the following section). The conformation of corona chains in PS-b-PAA micelles has been studied by contrast variation SANS.102 The interaction between the micelles was described using the hard sphere structure factor, for salt-free solutions. In the presence of excess salt, intermicellar interactions were suppressed. The PAA chains were found to contract with increasing polymer concentration, independent of ionic strength. Interpenetration of coronal chains was observed when the packing fraction exceeded a critical value
4.3
THEORY
We focus here on theory developed specifically for block copolymers containing a polyelectrolyte block. When considering adsorption for example, models are often based on polyelectrolyte brushes. The large literature on this subject is not covered here, although suitable reviews can be found elsewhere.103'104 Dan and Tirrell introduced a polymer brush scaling model for diblocks comprising a strongly charged block and a hydrophobic block in aqueous salt solution.90 The system was studied in the regime of high salt concentrations where the electrostatic persistence length is smaller than the range of excluded volume interactions. The formation of micelles in bulk and adsorption on a planar substrate were both considered. The aggregation number of micelles and surface density of the adsorbed layer was predicted to increase with salt concentration, whilst the charged block layer thickness decreased. Micellar properties were predicted to be similar to those of neutral copolymers in nonpolar solvents. Micellization in neutral-b-polyelectrolyte diblocks in a solvent selective for the polyelectrolyte block has been analysed using brush scaling theory for both weakly and strongly charged diblocks.105 Analytical expressions for the cmc and micelle dimensions at the cmc were obtained. For strongly charged diblocks, micellization was suppressed unless the charged block was very short compared with the neutral block. Strong condensation of counterions onto the charged block occurs when the chain is long. The stretching penalty associated with this is high and micelle formation is prevented. In the weakly charged limit, micellization is not hindered in this way. Addition of salt was found not to significantly affect the micelles unless the
Polyelectrolyte Block Copolymers
189
screening length was smaller than the micelle size. Wittmer and Joanny used the scaling approach to investigate conditions under which micellization is possible. This occurs provided the charged monomer content is not too high. The model was mainly used to examine adsorption on a solid substrate or at the oil-water interface, as discussed in Section 5.1.2. The scaling approach leads to the following results. The micellar core radius is predicted to scale as given by Equations (2.21) and (2.22):106
The corresponding free energy scales as [in the following, prefactors O(l) are omitted, and expressions are written as equalities]:
where rB is the core-corona interfacial tension. For the polyelectrolyte corona, the free energy comprises a Gaussian elastic energy and an electrostatic energy:43
where Rm is the micellar radius, a is a monomer size, T is the fraction of charged monomers and lB is the Bjerrum length. The thermodynamic potential that governs micellization is the grand canonical free energy and is given by:105
where u is the chemical potential of the solution. By minimizing n, the equilibrium structure of the micelle can be determined. The cmc occurs at a chemical potential:43
The association number at the cmc is:
The size of the micelle is approximately equal to the size of the corona. Micelles only form if the association number is greater than unity, which imposes a condition on the charge fraction43
190
Block Copolymers in Solution: Fundamentals and Applications
If the charge fraction is smaller than Tm, the cmc is roughly given by:43
The foregoing neglects the condensation of counterions within the polyelectrolyte brush (osmotic brush). Allowance for this in the so-called strong charge limit is discussed by Marko and Rabin.105 More general expressions allowing for the density of coronal chains as well as copolymer composition were obtained by Shusharina et al. as summarized in Table 4.2. Table 4.2 Scaling relationships for association number p, corona thickness RA and core radius RB for micelles formed by neutral-b-polyelectrolyte diblocks5 Regime
Labela
p
RA/a
RB/a
z
0mic > e~
Individual quasineutral chains Planar quasineutral brush (1) Planar quasineutral brush (2) Spherical quasineutral brush Individual Pincus chains Planar Pincus brush (1) Planar osmotic brush (1) Planar osmotic brush (2) Spherical osmotic brush 0mic < e~z
Planar Pincus brush (2) Spherical Pincus brush
I(i)
NB
NA'
NB
I(pl)
NB
NA/NB
NB
I(p2)
NB
NB4/5
I(s) II(i) II(pl) III(pl) III(p2) III(s)
2/NA18/11
NA9/11
NB/Na6/11
NA3/5 /NB4//25
NB3/5
NAT4/7 NA7/3T4/3/NB2/9 NAT2/5 NB2/NA3T3 NAr2/5 AB2/NA3T3 N AT 2 / 5
NB NB NB
NB NB2/3 NB2/3 NB/NAr N B /N A T
II(p2) NB2/NA39/8 T15/4 NA5/4T1/2 7 /11 21/22T1511 8/11 2 II(s) N B /N NA N B'11T2/11
NB/NA13/8T5/4 NB8/33/NA7/22T5/11
aI, II and III denote, respectively, regimes with insignificant electrostatic interactions, Pincus regimes and osmotic regimes; i, p and s denote individually grafted polyions, planar brush and spherical brush, respectively (some of which are subdivided into regimes with different scaling properties).
Huang and coworkers have developed the model of Marko and Rabin to consider constraints on counterion condensation in the corona of micelles formed by diblocks with a long charged block and a short hydrophobic chain.107 Huang et al. argue that corona chains will not be fully stretched due to counterion pressure. The association number and the number of condensed counterions per micelle were computed, considering the chemical potentials of copolymer chains and of counterions. The cmc and polydispersity were found to be large in neutralb-polyelectrolyte micelles compared with the analogous uncharged diblock micelles. The number of free chains in the charged diblock decreases above the cmc.
Polyelectrolyte Block Copolymers
191
Figure 4.9 Spherical lattice model used in lattice mean-field theory by Shusharina and coworkers.108 The micelle contains six copolymer chains with solvophobic cores and charged coronas. Counterions are also shown. Solvent is indicated by unfilled cells. Reproduced by permission of American Chemical Society.
Shusharina et al. have performed lattice mean field simulations of micellization in neutral-b-polyelectrolyte diblocks containing a short solvophobic block and a weakly charged poly electrolyte.108 Figure 4.9 illustrates a micelle formed by six chains on the spherical lattice used. The cmc was investigated as a function of charge and length of the hydrophilic block and also as a function of solvent quality and salt concentration. Volume fraction profiles reveal a sharp interface between micellar core and corona. The corona contains very dilute hydrophilic segments. The scaling of micelle aggregation number, core radius and corona thicknesss are in qualitative agreement with scaling predictions. In a subsequent paper, a more detailed comparison with scaling theory was made and a very comprehensive summary of scaling behaviour was provided.5 Three scaling regimes were identified: a quasi-neutral regime at low fractional charge; a Tincus regime'I09 where the Coulomb interaction in the corona is unscreened since the counterions are predominantly outside the charged corona; and an 'osmotic regime' where the Coulomb interactions are screened by counterions which are located within the charged corona. Analytical scaling relations for the micellar aggregation number, the radius of the core, and the thickness of the corona were presented as a function of the length of both blocks, and the fractional charge of the hydrophilic block. These are summarized in Table 4.2. Figure 4.10 summarizes 'phase diagrams' illustrating conditions under which various micellar structures are observed. Diagrams are parameterized in terms of the chain length of the polyelectrolyte block, NA and the fractional charge on each monomer, T. Two limiting cases were considered, depending on the parameter z, the electrostatic
192
Block Copolymers in Solution: Fundamentals and Applications
Polyelectrolyte Block Copolymers
193
interaction energy between the charged micelle and a counterion located on the periphery of the corona. For 0mic > e- z , most counterions are within the spherical corona (osmotic brush), whereas for 0mic < e-z most of the counterions are outside the corona. The composition and grafting density (i.e. interfacial area per block in the case of micelles) leads to a further subdivision of scaling regimes. For very short charged blocks and for low grafting densities, the chains are isolated, and termed 'individual grafted polyions'. If the charged blocks are still short but the grafting density is higher, planar brushes are formed. When the charged block is of significant length compared with the neutral block it forms the corona of micelles, i.e. a spherical brush. The mean-field lattice model provided volume fraction profiles of counterions, as well as coronal layer thickness. Changes in aggregate morphology in neutral-b-polyelectrolyte diblocks as a function of salt concentration have been analysed using a scaling model by Netz.110 The neutral block was hydrophobic. For low salt concentrations, electrostatic interactions are dominant, and a transition from spherical to cylindrical micelles occurs as salt concentration is increased. At high salt concentration, the elastic energy of the hydrophobic block dominates and bilayers are predicted. It was suggested that the presence of coexisting structures at intermediate salt concentrations could indicate the formation of complex interconnected structures such as those observed by Eisenberg and coworkers as discussed in Section 2.16. The endcap energy of cylinders was found to be high, pointing towards the possibility of the formation of closed rings (toroids), as observed experimentally.111 Scaling theory has also been developed by Borisov and Zhulina to describe micellization in diblocks containing a weak polyelectrolyte block and a hydrophobic block.6,84At low pH and/or salt concentration, quasi-neutral micelles in which the corona chains were undissociated were predicted.6 An increase in salt concentration leads to the ionization of the corona chains and hence a decrease in association number. This of course is in contrast to the behaviour of strong or 'quenched' polyelectrolytes. For crew-cut micelles, the corresponding transition occurs continuously, whereas for star-like micelles it occurs discontinuously at a critical salt concentration. The decrease in association number leads to an increase in osmotic repulsion in the corona that cannot be balanced by an increase in free energy of the core interface. The starlike micelles then become unstable and dissociate into smaller ones. This can also occur for crew-cut micelles with short insoluble blocks. Similar jump-like transitions were predicted on changing pH. Phase diagrams were presented in terms of salt concentration and the bulk degree of dissociation, and scaling laws in the various regimes were derived.6 This model was
Figure 4.10 Diagrams as a function of charged block length NA and fraction of singly charged monomers, T, illustrating different scaling regimes for neutral-b-polyelectrolyte diblocks and the location of counterions (dots).5 (a) 0mic > e- z , (b) 0mic < e-z The key to the labels is provided in Table 4.2. Reproduced by permission of American Chemical Society.
194
Block Copolymers in Solution: Fundamentals and Applications
later developed to consider changes in morphology for crew-cut micelles, driven by changes in ionic strength.7 Increasing salt concentration (or decreasing degree of ionization of corona blocks) was predicted to lead to transitions from spherical micelles to cylindrical micelles to lamellar aggregates. Morphology diagrams were presented in terms of block length (for either block) and the second virial coefficient which describes short-range interactions between charged monomers (which depends quadratically on the degree of ionization). The influence of salt on the association number was also investigated for diblocks containing strong polyelectrolyte blocks.84 An increase in p with salt concentration was observed, together with a small decrease in micelle radius due to screening of charges on coronal chains. Scaling laws for micelle dimensions and association numbers for different regimes of salt concentration were presented. The association properties of a diblock polyampholyte have been analysed using random phase approximation theory.112 When the charges on the polycationic block and polyanionic block are balanced, the solution phase separates into a polyelectrolyte network and almost pure solvent. However, above a certain charge asymmetry ratio, spherical micelles are formed. For weakly charged diblocks, the excess charge varies with pH. Changes in the dimensions of the micelles were found to be in qualitative agreement with experiments on adsorbed weakly charged diblocks.113 The association behaviour of telechelic polyelectrolytes has been examined using an analytical model based on an extended Flory network model. Telechelic polyelectrolytes are ABA triblocks with a lengthy charged solvophilic midblock and short hydrophobic endblock 'stickers'. The swelling and collapse of gels formed by telechelic polyelectrolytes was investigated, and phase diagrams constructed.114 The formation of a gel leads to phase separation since concentrated gel phase coexists with a dilute supernatant phase. Increasing the degree of ionization leads to destabilization of gels. Added salt screens the interactions between the charged end groups and facilitates the association of telechelic molecules. The formation of physical clusters instead of a macroscopic physical gel was also considered.115 Depending on the attraction energy between stickers, clusters are found to be stabilized if the Coulomb interaction and translational entropy of counterions are sufficiently large. The association of multiblock ionomers, i.e. chains comprising a small fraction of charged units inserted into neutral chains led to the concept of superstrong segregation.116 In this regime the ionic units and counterions form an ion pair with a dipole moment. Interaction between these dipoles causes attraction between them to form 'multiplets'. The limiting size and spacing of of multiplets was analysed in the melt limit. The analysis will also apply to solutions of ionomers, and other systems of copolymers with strongly interacting groups.117 This work predicted the formation of disk-like and ultimately lamellar structures as the multiplet interaction parameter increases. Such structures were considered in the analysis of superstrong segregation in ABC triblocks (Section 2.7.6) and have been observed by Lodge et al. in micelles formed by ABC block copolymers.118,119
Polyelectrolyte Block Copolymers
195
Little theoretical or simulation effort has been devoted to the behaviour of concentrated solutions of polyelectrolyte block copolymers. Microphase separation of concentrated solutions of associating polyelectrolytes has been investigated using strong segregation limit theory.120 Spherical, cylindrical and lamellar structures were all predicted for a concentration range lower than that at which physical gelation occurred. Dynamic mean field density functional theory has been used to compute the phase behaviour of concentrated solutions of a weakly charged triblock polyelectrolyte in the case that a Donnan equilibrium is established, i.e. long-range electrostatic interactions are screened.121 The lyotropic phase diagram was found to be modified depending on the charge on the solvophilic end-blocks. Likos and coworkers have recently used the interaction potential they derived for polyelectrolyte solutions to investigate the structure and phase behaviour of polyelectrolyte star solutions.100'101 As for the comparison between neutral star polymers and block copolymer micelles,122 analogies may be expected with the structure of concentrated solutions of polyelectrolyte stars and micelles. The theory led to phase diagrams containing multiple ordered phases, including simple cubic, fcc and bcc phases as well as interesting body-centred orthogonal and diamond packings of stars.101 Micellization in dilute solutions of a neutral-b-polyelectrolyte diblock with oppositely charged linear homopolymers was examined using a brush model to calculate contributions to the free energy of micelles.123 The system was studied in the limit that the polyelectrolyte block was large relative to the neutral block, whilst the homopolymer chains were shorter than the charged block in the copolymer. Polyelectrolyte complexes with different compositions were found to coexist with well-defined micelles. The micelles comprise the polyelectrolyte block and the oppositely charged homopolymer polyelectrolyte surrounded by a corona of the hydrophilic block.
4.4
POLYION COMPLEXES
Electrostatic interactions between a pair of neutral-b-polyelectrolyte block copolymers containing blocks with equal but opposite charges can lead to the formation of a polyion complex (PIC). When the charge is not balanced, overcharging and charge reversal of micellar aggregates are possible. To form PIC micelles, it was shown that the charge on oppositely charged segments has to be matched.124,125 This is illustrated in Figure 4.11 which shows complex formation by PEG-b-PAsp and PEG-b-PLys diblocks, with oppositely charged polyelectrolyte blocks.125 Chains with matched polyelectrolyte block length can form micelles, whereas unmatched chains form charge neutralized bimolecular complexes. Kataoka and coworkers have investigated PIC micellization in this system in a series of papers.124-126 The core radius and volume fraction of PEG at the core/corona interface were obtained from light scattering experiments, and the importance of these two parameters in determining micellar size was emphasized.126 The volume
196
Block Copolymers in Solution: Fundamentals and Applications
Figure 4.11 Polyion complex formation by pairs of block copolymers with oppositely charged polyelectrolyte blocks.125 When the polyelectrolyte chain length is matched, micelles can self-assemble, whereas for unmatched chains, bimolecular complexation is observed. Reproduced by permission of Science.
fraction of PEG at the interface was shown to be proportional to the length of the PAsp block. Polyion complexes have also been prepared by the Kataoka group in which the micelles have been functionalized by the incorporation of aldehyde groups at the end of the PEG corona block.127 These ligands could assist in targeted delivery. The core-forming block was PDMA, to which plasmid DNA is complexed thus entrapping the DNA for use in gene delivery. It is established that the resistance of plasmid DNA against nuclease digestion is increased remarkably when it was incorporated into the core of PIC micelles.128 The incorporation of the model protein, chicken egg white lysozyme (positively charged over a wide pH range) into PIC micelles formed with PEG-b-PAsp diblocks was also studied via light scattering.129 The stoichiometry between aspartic acid residues and the total number of lysine and arginine residues in lysozyme in the micelle core was quantified. Further details on drug delivery applications of PIC micelles can be found in Section 6.3. The co-preciptation of P2VP blocks in PS-b-P2VP and P2VP-b-PEO blocks leads to the formation of onion-type micelles with a PS core, a P2VP inner shell and a PEO outer shell (see Figure 2.27). 13° The PS-b-P2VP polymer was dissolved in a methanol/dioxane/water mixture and then dialysed against 0.1 M HC1, leading to micelles with a protonated P2VP corona. Addition of P2VP-b-PEO at pH > 10 leads to the formation of onion micelles. These were investigated later by SANS 131 (see Section 2.12.2). Kabanov and coworkers have investigated micelles formed in several types of block ionomer complex. For example, they have investigated micellization in mixtures of diblock cations with anionic surfactant. The diblocks contained a
Polyelectrolyte Block Copolymers
197
cationic block such as PE4VP) and a PS132 or PAMS133 block. The anionic surfactant was bis(2-ethylhexyl)sulfosuccinate (AOT). The micelle characteristics (dimensions and aggregation numbers) were found to be similar to those of neutral block copolymers in a selective solvent. They also observed polyion complexes in mixtures of a diblock containing an anionic block, PEO-b-PNaMA, and the cationic homopolymer PE4VP.134 They showed that micelles containing a core consisting of stoichiometric neutralized polyion complex and a PEO corona form in water, in contrast to the interpolyelectrolyte complexes formed by the homopolymers which lose solubility when the charge is neutralized. This points to the important role of the PEO corona in enhancing solubility. It was also shown that the PIC micelles are highly salt sensitive, since dissociation was observed at high salt concentration. Similar conclusions were reached from a study on complexes of the same diblock system (although a different composition) but with cationic ammonium and pyridinium bromide salts.135-137 The influence of temperature on the dehydration of the PEO chain was also examined.137 At room temperature, dispersions were stable, but on increasing temperature aggregation occurred due to the dehydration of the PEO 'steric stabilization' layers. The onset temperature for aggregation decreased with increasing salt concentration, depending also on the nature of the salt anion, following the Hofmeister series (cf. Section 2.6.3). Diblocks containing an insoluble block such as PS-b-PNaA formed complexes with N-cetylpyridinium cations-the nonstoichiometric complexes being swollen with ionized PAA chains.138 Complexes between PEO-b-PNaMA diblock copolymers and single, double or triple cationic surfactants have been found to lead to stable dispersions with particle size in the range 100-200 nm, the stability depending on the length of the blocks.139 Complexation between block copolymers containing a cationic block and DNA, which is anionic, to yield so-called polyplexes is relevant to drug delivery applications.140,141 The cationic block copolymer should charge compensate the DNA, inducing a coil-globule transition so that the DNA condenses and has enhanced stability under physiological conditions. It is also necessary that the polymer has sufficiently low molar mass (<30 000 g mol"1)141 so that it can be eliminated from the body. Addition of a hydrophobic block is believed to enhance cell interactions and tissue permeability. Hydrophilic polyethylene glycol on the other hand enhances solubility of the polyplex. Further details are provided in Section 6.3. Polyelectrolyte complexation between a neutral-b-cationic PEO-b-PDEA diblock and an anionic PMAA homopolymer leads to the formation of three types of aggregate structure, depending only on solution pH.142 Above pH 8.5, simple micelles are formed with hydrophobic PDEA cores and PEO corona chains, the PMAA not participating in micelle formation. However, at pH 6-8.5, polyion complex micelles with charge-compensated PDEA/PMAA cores and PEO coronas are observed. Below pH 3, a third micelle structure is formed with a core comprising hydrogen bonded PEO and PMAA with a PDEA corona. Complexation between an ABC triblock with a cationic inner shell B block and a neutral-anionic
198
Block Copolymers in Solution: Fundamentals and Applications
diblock (or an anionic homopolymer) has been used to prepare shell cross-linked micelles.143 Onion-type micelles were prepared from PEO-b-qPDMA-b-DEA triblocks in alkaline solution. Sodium polystyrene sulfonate homopolymer or PEO-b-PNaSS diblock were added to effect ionic cross-linking. The diblock copolymer cross-linker proved to be much more effective than the homopolymer.
4.5
COPOLYMER-SURFACTANT COMPLEXES
The interaction between polymers and surfactants is described by two critical concentrations. The first is the critical aggregation concentration (cac, sometimes denoted C1) at which point the binding of surfactant and polymer first occurs. The cac is generally lower than the cmc of the surfactant alone. The second critical concentration (often denoted C2) is associated with the saturation of polymer with surfactant aggregates (Figure 2.34). The cmc of the surfactant (Cm) may also be observed. For some polymer/surfactant systems, Cm and C2 are coincident.144,145 However, in other cases, Cm is less than C2.144'145 Complex formation between P2VP-b-PEO diblocks and fluorinated anionic surfactants has been investigated in aqueous solution.146 The two types of fluorinated surfactant and P2VP have opposite charges at pH 3, leading to strong electrostatic interactions. TEM revealed the presence of vesicles as well as other complex polymicellar aggregates. Berret and coworkers have performed small-angle scattering (SANS and SAXS) and DLS studies of colloidal complexes resulting from the self-assembly of neutralb-polyelectrolyte diblocks and oppositely charged surfactants.147-150 An aggregate with a core-shell structure forms, as sketched in Figure 4.12. The core is described as a coacervated microphase comprising surfactant micelles linked by the polyelectrolyte blocks. The corona formed by the lengthy neutral chains forms a stabilizing coating. The complexes form when the stoichiometric charge ratio Z (equal to the ratio of surfactant and polyelectrolyte repeat unit concentrations) is in the range Z = O.l-l.150 Copolymers with both type of charged block were studied. Cationic polyelectrolyte blocks of PTMEMS formed complexes with the surfactant SDS, and anionic PNaA or PSS blocks formed complexes with DTAB.150 The neutral block was polyaerylamide. The complexes have hydrodynamic radii in the range 13-58 nm. The cores contain typically a few hundred micelles.148 There is no evidence of crystalline order in the core, in contrast to very well ordered cubic or hexagonal lattices observed for the precipitate formed from the phase-separated polyelectrolyte homopolymer/surfactant mixtures when Z = 1.150 Indeed, Monte Carlo simulations of aggregation within the core indicate hard sphere interactions among the micelles in the core,149 which lead to the observed structure factor peak at large q. The phase separation observed for the homopolymer/surfactant complex is prevented in the block copolymer/surfactant complex by the neutral micellar corona.148 The scaling properties of the colloidal complexes differ from those of conventional block copolymer micelles.150 The core radius decreases more slowly
Polyelectrolyte Block Copolymers
199
Figure 4.12 Schematic of core-shell structure formed by complexation between a diblock comprising an anionic poly(sodium acrylate) block and neutral poly(acrylamide) and the cationic surfactant DTAB.147 Reproduced by permission of EDP.
with core block length than predicted and the neutral block does not significantly influence the core radius. The nonequilibrium nature of the self-assembly process was emphasized by the observed dependence of the size of complexes on the initial mixing conditions.148 Cryo-TEM on the complexes provided direct evidence for internal structure within the micelles, as well as providing dimensions for comparison with the indirect scattering techniques.149 It is possible that the systems studied by Kabanov and coworkers discussed above also form structures with microphaseseparated core structures. Examination of this requires the use of a probe sensitive to internal micellar structure, such as SANS.
4.6 COMPLEXATION WITH OTHER MOLECULES The complexation of PEl-b-PPG-b-PEl and PEl-b-PEG-b-PEl triblock copolymers with cyclodextrins leads to fascinating pseudo-polyrotaxanes, in which the cyclodextrins thread over the hydrophobic block domains, since the cavity in the cyclodextrin molecules is itself hydrophobic. In the PEI-b-PPG-b-PEI system, B-cyclodextrins moved away from fluorescein-4-isothiocyanate end caps towards the PPG block as temperature increased, due to increased hydrophobicity of the midblock.151,152 In the PEl-b-PEG-b-PEl system, pH responsiveness was observed.153 At pH < 8, a-cyclodextrin (a-CD) molecules selectively thread over
200
Block Copolymers in Solution: Fundamentals and Applications
Figure 4.13 Formation of pseudo-polyrotaxanes by threading of a-cyclodextrin molecules around a PEl-b-PEG-b-PEl triblock.153 The self-assembly is pH responsive due to protonation of PEI below pKa = 8.9. Reproduced by permission of American Chemical Society.
the PEG midblock due to repulsive electrostatic interactions (Figure 4.13), since PEI is protonated under these conditions. At pH 11, all blocks threaded a-CD molecules (Figure 4.13). Using these two systems, thermoresponsive and pH-responsive formation of pseudo-polyrotaxanes is possible via a self-assembly process.
4.7
GELATION
As in neutral block copolymers (Section 3.3), block copolymers containing polyelectrolyte chains can undergo gelation at sufficiently high concentration. At low pH, electrostatic interactions dominate the self-assembly of PAA-b-P2VP-b-PAA triblocks in aqueous solution.154 The P2VP units are protonated, and the PAA units are negatively charged. As sketched in Figure 4.14, an increase in concentration leads to interactions between the oppositely charged units and hence gelation. The resulting gel exhibits interesting rheological behaviour, including strain hardening and shear thickening. These shear-induced changes in network structure were ascribed to an increase in the number of elastically active chains due to an increase in the extent of intermolecular bridging.
4.8 HIERARCHICAL ORDER IN PEPTIDE BLOCK COPOLYELECTROLYTE SOLUTIONS Block copolymers composed of a synthetic block and a charged polypeptide block are a particularly interesting class of polyelectrolyte block copolymer. Hierarchical order and additional functionality can be programmed into the self-assembly of
Polyelectrolyte Block Copolymers
201
Figure 4.14 Schematic of self-assembly in a PAA-b-P2VP-b-PAA triblock polyelectrolyte as a function of concentration.154 Here cg denotes the critical concentration for gelation and c' is a concentration at which percolation is observed. Reproduced by permission of American Chemical Society.
202
Block Copolymers in Solution: Fundamentals and Applications
Figure 4.15 Peptide sequence of samples of PEG-peptide block copolymers containing switch peptide sequences shown in B-strand conformation (a) and as a-helical wheel representation (b).160 Rectangles and ovals mark hydrophobic and hydrophilic amino acids, respectively. Reproduced by permission of Royal Society of Chemistry.
block copolymers by incorporating one or more polypeptide chains. These can form secondary structures-a helices due to intramolecular hydrogen bonding or (3 sheets due to intermolecular hydrogen bonding. Furthermore, by appropriate positioning of amino acid residues it is possible to design structures with localized hydrophobicity or hydrophilicity, for example a helices with hydrophobic and hydrophilic faces (see Figure 4.15). The solution self-assembly of block copolymers containing peptide units has been reviewed.155,156 4.8.1
a HELIX STRUCTURES
A range of aggregate structures of PS-b-polypeptide diblocks in aqueous solution have been revealed by TEM and AFM on dried films.157 The synthetic poly (isocyanide) peptides studied form a-helices. Poly(isocyano-L-alanine-L-alanine)
Polyelectrolyte Block Copolymers
203
forms a right-handed helix, whereas poly(isocyano-L-alanine-L-histidine) forms a left-handed helix. The aggregate structure was found to depend on the length of the poly(isocyanide) block, the pH and the anion-headgroup interactions. Micelles, vesicles and bilayer aggregates were all observed. In addition, superhelices were imaged, and ascribed to coiled rods. The superhelices had an opposite twist to that of the constituent blocks, a puzzling observation that was not explained. The influence of PEG chains on the aggregation of peptides containing heptad sequences designed to form coiled coil superstructures has been investigated by two groups.158-160 Klok and coworkers employed the coiled coil sequence LAEIEAK designed de novo by Hodges and coworkers.161 Incorporation of PEG chains in hybrid diblocks significantly enhanced the stability of coiled coil superstructures towards denaturing, an effect ascribed to the formation of a hydrophilic, protective shell of PEG around the hydrophobic interior. The ability of the peptide to form an a helix was not disturbed. Pechar et al. reached a similar conclusion for their diblocks comprising PEG(2000) and (VSSLESK)n, n = 3-6.158 Based on evidence from CD and analytical ultracentrifugation, the self-organization in the PEGpeptides containing the LAEIEAK sequence was described as an equilibrium between copolymer unimers and dimeric and tetrameric coiled coil aggregates, as illustrated in Figure 4.16. The relative amounts of these species are temperature and concentration dependent. The helix content decreases with temperature, suggesting cooperative unfolding. The process is reversible. Increasing concentration favours the formation of dimeric and tetrameric aggregates. In the peptides containing the (VSSLESK)n sequence with n = 5 or 6, dimers were observed. The helix content increased with n. Electron paramagnetic resonance on spin-labelled samples of the LAEIEAK-containing polymers confirmed that the PEG chains are folded closely around the peptide core, as sketched in Figure 4.16.162 The self-assembly of
Figure 4.16 Model for the self-assembly of PEG-b-peptide diblocks containing the heptad sequence LAEIEAK.160-162 Reproduced by permission of American Chemical Society.
204
Block Copolymers in Solution: Fundamentals and Applications
homopolypeptides and PEG-peptide diblocks containing one or two heptad repeat sequences was later compared.160 The helix content was similar whether the peptide contained one or two heptad repeats. On the other hand, for the peptides containing the (VSSLESK)n sequence, the helix content increased with n. For the LAEIEAKcontaining polymers, the helix content was enhanced in the polypeptide compared with the diblocks (and decreased with PEG content) due to steric hindrance caused by the conjugated PEG which probably hampers association and folding of the peptide.160,163 The helix content increased with concentration, and decreased with temperature. Nanostructure formation in aqueous solutions of PEG-peptide diblocks and PEG-b-peptide-£-PEG triblocks containing 'switch' peptide sequences has also been explored.160 Switch peptides are patterned with hydrophobic and hydrophilic substituents in such a way that they can form either amphiphilic a-helices or amphiphilic B-strands, depending on pH,164 as illustrated in Figure 4.15. TEM and AFM suggested the formation of fibrillar structures, which is particularly interesting since circular dichroism (CD) indicated that in the low pH conditions, the a-helical structure of the peptide was retained upon conjugation to PEG. A model for the formation of fibrils based on side-to-side stacking of the 'Janus' helices [Figure 4.15(b)] was presented. The solution self-assembly of PS-b-PLys diblocks was investigated by SANS and light scattering.165 Extended micellar structures were observed, whilst the a helix conformation of the PLys was retained. Changing solvent conditions can induce a transition in the conformation of PBLG from a random coil to a rod-like helical conformation. This has been studied via SANS on PBLG-b-PS diblocks for which a transition from a coil-coil to a rod-coil conformation has been detected.166 4.8.2
(3 SHEET STRUCTURES
It has been shown that conjugation to PEG can be used to control the lateral aggregation of B-sheet forming peptides into fibrils. This is relevant to fibril formation in amyloidosis, which is responsible for diseases such as BSE and Alzheimer's. Lynn and coworkers observed fibril formation in aqueous solutions of PEG-peptide diblocks where the peptide block was based on the central domain of the B-amyloid peptide. 167-169 They found from SANS and TEM that the PEG forms a coating around the fibril-thus acting as a 'steric stabilization' layer. Stupp and coworkers have studied fibril formation in diblocks with a peptide block linked to fatty acid 'tail groups'.170 The peptides were designed to contain terminal cysteine residues to facilitate intramolecular cross-linking via disulfide bonds. This enabled the reversible polymerization of nanofibres. By controlling the pH, gelation was also observed due to intermolecular cross-linking. In a subsequent paper,171 it was shown that it was possible to mineralize a nanostructured fibrous scaffold prepared by the self-assembly of a fatty acid-b-peptide diblock, to produce artificial bone.
Polyelectrolyte Block Copolymers
205
Collier and Messersmith have investigated the effect of conjugation to PEG on peptide fibril width for PEG-b-peptide and peptide-b-PEG-b-peptide Copolymers containing ß-sheet forming sequences.172 The self-assembly in aqueous solution of a number of hybrid block Copolymers containing amphiphilic ß strand sequences flanked by one or two PEG terminal chains was investigated via CD spectroscopy, SAXS and TEM.173 CD revealed primarily ß-strand secondary structures. In comparison with the native peptide sequence, it was found that the secondary structure was stabilized against pH and temperature variation in the di- and triblock Copolymers with PEG. SAXS indicated the presence of fibrillar structures, and the dimensions of these are comparable with the estimated width of a (3 strand (with terminal PEG chains in the case of the Copolymers). Transmission electron microscopy on selectively stained and dried specimens shows directly the presence of fibrils (Figure 4.17). It was proposed that
Figure 4.17 TEM image of dried films of a PEG/peptide block eopolymer in which the peptide forms a ß-sheet structure.173 Reproduced by permission of American Chemical Society.
206
Block Copolymers in Solution: Fundamentals and Applications
these fibrils result from the hierarchical aggregation of B strands into helical tapes which then stack into fibrils. Details on nanofibre formation, due in part to B-sheet hydrogen bonding, in PEG-peptide amphiphiles,170-172,174 is not discussed in detail here, since the chains are too short for them to be considered as block copolymers. 4.8.3
HYDROGELS
Hybrid hydrogels have been prepared via the self-assembly of a water soluble synthetic polymer and coiled-coil peptides, which were genetically engineered, and expressed using Escherichia coli.115,176 The peptides were linked via a pendant metal-chelating ligand to a hydrophilic copolymer of HPMA and the metalchelating monomer DAMA (Figure 4.18). Two coiled coil peptides were employed-
Figure 4.18 Hybrid hydrogels in which a water soluble copolymer is linked via coiled coil peptide domains.176 The structure of the cross-link is shown enlarged. Reproduced by permission of Wiley-VCH.
one based on a natural sequence from the motor protein kinesin (containing a six histidine tag), and the other a block copolymer HHHHHH-b-(noncoiled 30 amino acid sequence)-b-(VSSLESK)6. The heptad sequence favours coiled coil formation. The histidine block acts as a tag for Ni2+ chelation. Hydrogels containing the wild type peptide cross-linker underwent a volume phase transition driven by thermal unfolding. Those containing the block polypeptide did not show a volume phase transition. However, for hydrogels containing either type of cross-linker, gel
Polyelectrolyte Block Copolymers
207
swelling was observed by replacement of the histidine tag by a metal-chelating ligand.177 Later, peptides containing block sequences were studied, comprising noncoiled regions (A) alternating with sequences from the kinesin stalk protein with a tendency to form coiled coils.178 Histidine-tagged AB diblock, ABAB tetrablock and ABABAB hexablock polymers were prepared via recombinant DNA techniques. Size exclusion chromatography revealed the presence of multimer aggregate structures (mainly dimers and tetramers). The degree of swelling decreased with temperature as unfolding occurred. The use of synthetic polypeptide block Copolymers to create hydrogels has been explored by Deming and coworkers.179,180 They investigated gelation for a range of diblocks containing a hydrophilic PLys [or poly(L-glumatic acid)] Polyelectrolyte block and a hydrophobic block based on an a-helix for poly(L-leucine) or a ß-sheet in the case of poly(L-valine).179 It was found that a high degree of conformational order of the hydrophobic chain was required to observe gelation, and that gelation occurred slightly better with a-helical domains. The importance of chain packing was highlighted by the fact that gelation was not observed in a mixture containing racemic L-leucine at a concentration at which gelation occurred for a copolymer containing the enantiomeric form. Hierarchical ordering was observed with a nanoscale interpenetrating network structure and a porous bicontinuous morphol1 80 ogy comprising gel matrix and water channels at the microscale. Laser scanning confocal microscopy and Cryo-TEM were used to probe the cellular structure for a PLys-&-poly(L-leucine-,stat-L-valme) diblock (Figure 4.19). SANS confirmed a nanoscale ordered structure.
Figure 4.19 Cryo-TEM image showing cellular structure in a gel formed by a poly(Llysine)-b-poly(L-leucine-stat-L-valine) diblock.180 Reproduced by permission of American Chemical Society.
4.8.4
POLYPEPTIDE BLOCK COPOLYMER-BASED COMPLEXES
Complexes of a PEO-b-PLys diblock with retinoic acid (vitamin A) have been investigated via a number of techniques.181 In aqueous solution, micelles comprised
208
Block Copolymers in Solution: Fundamentals and Applications
a compact core of the cationic PLys block complexed with retinoic acid in a lamellar structure surrounded by a PEO corona. The PLys block formed an a-helix at high pH, the coil-helix transition spanning the pH range from 3.7 to 9.0. The high stability of the a-helix was ascribed to the stabilization of the retinoate. The binding of retinoic acid within a sterically stabilized micelle is important to ensure protection, solubility and transport by body fluids. REFERENCES 1. Hamley, I. W. The Physics of Block Copolymers. Oxford University Press: Oxford, 1998. 2. Forster, S.; Abetz, V.; Miiller, A. H. E. Adv. Polym. Sci. 2004, 166, 173. 3. Selb, J.; Gallot, Y. Ionic Block Copolymers. In Developments in Block Copolymers; Goodman, I., Ed.; Elsevier: London, 1985; Vol. 2; p 27. 4. Forster, S.; Hermsdorf, N.; Bottcher, C; Lindner, P. Macromolecules 2002, 35, 4096. 5. Shusharina, N. P.; Linse, P.; Khokhlov, A. R. Macromolecules 2000, 33, 3892. 6. Zhulina, E. B.; Borisov, O. V. Macromolecules 2002, 35, 9191. 7. Borisov, O.; Zhulina, E. B. Macromolecules 2003, 36, 10029. 8. Barrat, J. L.; Joanny, J.-F. Adv. Chem. Phys. 1996, 94, 1. 9. Baigl, D.; Ober, R.; Qu, D.; Fery, A.; Williams, C. E. Europhys. Lett. 2003, 62, 588. 10. Baigl, D.; Sferrazza, M.; Williams, C. E. Europhys. Lett. 2003, 62, 110. 11. Kantor, Y.; Kardar, M. Europhys. Lett. 1994, 27, 643. 12. Dobrynin, A. V.; Rubinstein, M.; Obukhov, S. P. Macromolecules 1996, 29, 2974. 13. Dobrynin, A. V.; Rubinstein, M. Macromolecules 1999, 32, 915. 14. Zhang, L.; Khougaz, K.; Moffitt, M.; Eisenberg, A. Self-assembly of Block Polyelectrolytes. In Amphiphilic Block Copolymers: Self-assembly and Applications; Alexandridis, P., Lindman, B., Eds. Elsevier: Amsterdam, 2000; p. 87. 15. Astafieva, I; Zhong, X. F.; Eisenberg, A. Macromolecules 1993, 26, 7339. 16. Nguyen, D.; Williams, C. E.; Eisenberg, A. Macromolecules 1994, 27, 5090. 17. Khougaz, K.; Gao, Z.; Eisenberg, A. Macromolecules 1994, 27, 6341. 18. Khougaz, K.; Zhong, X. F; Eisenberg, A. Macromolecules 1996, 29, 3937. 19. Astafieva, L; Khogaz, K.; Eisenberg, A. Macromolecules 1995, 28, 7127. 20. Khougaz, K.; Astafieva, I; Eisenberg, A. Macromolecules 1995, 28, 7135. 21. Halperin, A. Macromolecules 1987, 20, 2943. 22. Zhulina, E. B.; Birshtein, T. M. Vysokomol.Soedin. 1985, 27, 511. 23. Zhulina, Y. B.; Birshtein, T. M. Polym. Sci. USSR 1986, 27, 570. 24. Leibler, L.; Orland, H.; Wheeler, J. C. J. Chem. Phys. 1983, 79, 3550. 25. Nagarajan, R.; Ganesh, K. J. Chem. Phys. 1989, 90, 5843. 26. Whitmore, M. D.; Noolandi, J. Macromolecules 1985, 18, 657. 27. Gao, Z.; Varshney, S. K.; Wong, S.; Eisenberg, A. Macromolecules 1994, 27, 7923. 28. Zhang, L.; Eisenberg, A. Science 1995, 268, 1728. 29. Zhang, L.; Yu, K.; Eisenberg, A. Science 1996, 272, 1777. 30. Zhang, L.; Eisenberg, A. J. Am. Chem. Soc. 1996, 118, 3168. 31. Yu, K.; Eisenberg, A. Macromolecules 1996, 29, 6359. 32. Gao, Z.; Zhong, X.-F; Eisenberg, A. Macromolecules 1994, 27, 794. 33. Dormidontova, E. E.; Khokhlov, A. R. Macromolecules 1997, 30, 1980. 34. Shen, H.; Eisenberg, A. J. Phys. Chem. B 1999, 103, 9473.
Polyelectrolyte Block Copolymers
209
Shen, H.; Eisenberg, A. Macromolecules 2000, 33, 2561. Choucair, A.; Eisenberg, A. Eur. Phys. J. E 2003, 10, 37. Zhang, L.; Eisenberg, A. Macromolecules 1999, 32, 2239. Yu, K.; Eisenberg, A. Macromolecules 1998, 31, 3509. Yu, Y; Zhang, L.; Eisenberg, A. Macromolecules 1998, 31, 1144. Riegel, I. C.; Eisenberg, A.; Petzhold, C. L.; Samios, D. Langmuir 2002, 18, 3358. Busse, K.; Kressler, J.; van Eck, D.; Horing, S. Macromolecules 2002, 2002, 178. Matsuoka, H.; Maeda, H.; Kaewsaiha, P.; Matsumoto, K. Langmuir 2004, 20, 7412. Wittmer, J.; Joanny, J. F. Macromolecules 1993, 26, 2691. Butun, V.; Billingham, N. C.; Armes, S. P. Chem. Commun. 1997, 671. Lee, A. S.; Gast, A. P.; Butun, V.; Armes, S. P. Macromolecules 1999, 32, 4302. Baines, F. L.; Armes, S. P.; Billingham, N. C.; Tuzar, Z. Macromolecules 1996, 29, 8151. Butun, V.; Liu, S.; Weaver, J. V. M.; Bories-Azeau, X.; Cai, Y; Armes, S. P. in press. Dobrynin, A. V; Colby, R. H.; Rubinstein, M. J. Polym. Sci. B: Polym. Phys. 2004, 42, 3513. 49. Liu, S.; Armes, S. P. Curr. Opin. Colloid Interface Sci. 2001, 6, 249. 50. Gohy, J.-E; Antoun, S.; Jerome, R. Macromolecules 2001, 34, 7435. 51. Butun, V.; Armes, S. P.; Billingham, N. C. Polymer 2000, 42, 5993. 52. Tuzar, Z.; Pospisil, H.; Plestil, J.; Lowe, A. B.; Baines, F. L.; Billingham, N. C.; Armes, S. P. Macromolecules 1997, 30, 2509. 53. Lowe, A. B.; Billingham, N. C.; Armes, S. P. Macromolecules 1999, 32, 2141. 54. Vamvakaki, M.; Unali, G.-E; Butun, V.; Boucher, S.; Robinson, K. L.; Billingham, N. C.; Armes, S. P. Macromolecules 2001, 34, 6839. 55. Gabaston, L. I.; Furlong, S. A.; Jackson, R. A.; Armes, S. P. Polymer 1999, 40, 4505. 56. Lowe, A. B.; Billingham, N. C.; Armes, S. P. Macromolecules 1998, 31, 5991. 57. Vamvakaki, M.; Billingham, N. C.; Armes, S. P. Macromolecules 1999, 32, 2088. 58. de Arcangelis, L.; Del Gado, E.; Coniglio, A. Eur. Phys. J. E 2002, 9, 277. 59. Boutun, V.; Billingham, N. C.; Armes, S. P. J. Am. Chem. Soc. 1998, 720, 11818. 60. de Paz Banez, M. V.; Robinson, K. L.; Armes, S. P. Macromolecules 2000, 33, 451. 61. Liu, S.; Billingham, N. C.; Armes, S. P. Angew. Chem., Int. Ed. 2001, 40, 2328. 62. Liu, S.; Armes, S. P. Angew. Chem., Int. Ed. 2002, 41, 1413. 63. Liu, S.; Armes, S. P. Langmuir 2003, 19, 4432. 64. Patrickios, C. S.; Forder, C.; Armes, S. P.; Billingham, N. C. J.Polym.Sci.A: Polym.Chem. 1996, 34, 1529. 65. Forder, C.; Patrickios, C. S.; Armes, S. P.; Billingham, N. C. Macromolecules 1996, 29, 8160. 66. Forder, C.; Patrickios, C. S.; Billingham, N. C.; Armes, S. P. Chem. Commun. 1996, 883. 67. Butun, V.; Vamvakaki, M.; Billingham, N. C.; Armes, S. P. Polymer 2000, 41, 3173. 68. Goloub, T.; de Keizer, A.; Cohen Stuart, M. A. Macromolecules 1999, 32, 8441. 69. Gohy, J.-F.; Creutz, S.; Garcia, M.; Mahltig, B.; Stamm, M.; Jerome, R. Macromolecules 2000, 33, 6378. 70. Bieringer, R.; Abetz, V; Miiller, A. H. E. Eur. Phys. J. E 2001, 5, 5. 71. Tsitsilianis, C.; Katsampas, I.; Sfika, V. Macromolecules 2000, 33, 9054. 72. Sfika, V; Tsitsilianis, C.; Kiriy, A.; Gorodyska, G.; Stamm, M. Macromolecules 2004, 37, 9551. 73. Bromberg, L. Langmuir 1998, 14, 5806. 74. Bromberg, L.; Barr, D. P. Macromolecules 1999, 32, 3649. 35. 36. 37. 38. 39. 40. 41. 42. 43. 44. 45. 46. 47. 48.
210
Block Copolymers in Solution: Fundamentals and Applications
75. Mendes, E.; Schadler, V.; Marques, C. M; Lindner, P.; Wiesner, U. Europhys. Lett. 1997, 40, 521. 76. Schadler, V.; Lindner, P.; Wiesner, U.; Mendes, E. J Phys. Chem. B 1998, 702, 7316. 77. Szczubialka, K.; Ishikawa, K.; Morishima, Y. Langmuir 1999, 75, 454. 78. Szczubialka, K.; Ishikawa, K.; Morishima, Y. Langmuir 2000, 16, 2083. 79. Muller, R; Guenoun, P.; Delsanti, M.; Deme, B.; Auvray, L.; Yang, J.; Mays, J. W. Eur. Phys. J. E 2004, 75, 1. 80. van der Maarel, J. R. C.; Groenewegen, W.; Egelhaaf, S. U.; Lapp, A. Langmuir 2000,16, 7510. 81. Guenoun, P.; Davis, H. T.; Tirrell, M.; Mays, J. W. Macromolecules 1996, 29, 3965. 82. Hariharan, R.; Biver, C.; Mays, J.; Russel, W. B. Macromolecules 1998, 31, 7506. 83. Zhulina, E. B.; Borisov, O.; Birshtein, T. M. J. Phys. France II1992, 2, 63. 84. Borisov, O. V.; Zhulina, E. B. Macromolecules 2002, 35, 4472. 85. Bendejacq, D.; Ponsinet, V.; Joanicot, M.; Loo, Y.-L.; Register, R. A. Macromolecules 2002, 35, 6645. 86. Balastre, M.; Li, E; Schorr, P.; Yang, J.; Mays, J. W.; Tirrell, M. V. Macromolecules 2002, 35, 9480. 87. Qin, A.; Tian, M.; Ramireddy, C.; Webber, S. E.; Munk, P.; Tuzar, Z. Macromolecules 1994, 27, 120. 88. Zhang, L.; Barlow, R. J.; Eisenberg, A. Macromolecules 1995, 28, 6055. 89. Guenoun, P.; Muller, E; Delsanti, M.; Auvray, L.; Chen, Y. J.; Mays, J. W.; Tirrell, M. Phys. Rev. Lett. 1998, 81, 3872. 90. Dan, N.; Tirrell, M. Macromolecules 1993, 26, 4310. 91. Guenoun, P.; Davis, H. T.; Doumaux, H. A.; Maldonado, A.; Mays, J. W.; Talmon, Y; Taulier, N.; Tirrell, M.; Urbach, W.; Zheng, Y. Langmuir 2000, 16, 4436. 92. Muller, E; Fontaine, P.; Delsanti, M.; Belloni, L.; Yang, J.; Chen, Y. J.; Mays, J. W.; Lesieur, P.; Tirrell, M.; Guenoun, P. Eur. Phys. J. E 2001, 6, 109. 93. Moffitt, M.; Yu, Y; Nguyen, D.; Graziano, V.; Schneider, D. K.; Eisenberg, A. Macromolecules 1998, 31, 2190. 94. Daoud, M.; Cotton, J. P. J. Phys. France 1982, 43, 531. 95. Groenewegen, W.; Lapp, A.; Egelhaaf, S. U.; van der Maarel, J. R. C. Macromolecules 2000, 33, 4080. 96. Groenewegen, W.; Egelhaaf, S. U.; Lapp, A.; van der Maarel, J. R. C. Macromolecules 2000, 33, 3283. 97. Guenoun, P.; Delsanti, M.; Gazeau, D.; Mays, J. W.; Cook, D. C.; Tirrell, M.; Auvray, L. Eur. Phys. J. B 1998, 7, 77. 98. Korobko, A. V.; Jesse, W.; Lapp, A.; Egelhaaf, S. U.; van der Maarel, J. R. C. J. Chem. Phys. 2005, 722, 024902. 99. Dingenouts, N.; Patel, M.; Rosenfeldt, S.; Pontoni, D.; Narayanan, T.; Ballauff, M. Macromolecules 2004, 37, 8152. 100. Jusufi, A.; Likos, C. N.; Ballauff, M. Colloid Polym. Sci. 2004, 282, 910. 101. Hoffmann, N.; Likos, C. N.; Lowen, H. J. Chem. Phys. 2004, 727, 7009. 102. Korobko, A. V.; Jesse, W.; Egelhaaf, S. U.; Lapp, A.; van der Maarel, J. R. C. Phys. Rev. Lett. 2004, 93, 177801. 103. Holm, C.; Joanny, J.-E; Kremer, E; Netz, R. R.; Reineker, P.; Seidel, C.; Vllgis, T. A.; Winkler, R. G. Adv. Polym. Sci. 2004, 766, 67. 104. Ruhe, J.; et al. Adv. Polym. Sci. 2004, 765, 79.
Polyelectrolyte Block Copolymers
211
105. Marko, J. K; Rabin, Y. Macromolecules 1992, 25, 1503. 106. de Gennes, P. G. Macromolecules and Liquid Crystals: Reflections on Certain Lines of Research. In Solid State Physics; Liebert, L., Ed. Academic: New York, 1978; Vol. suppl. 14; pp 1. 107. Huang, C.; Olvera de la Cruz, M.; Delsanti, M.; Guenoun, P. Macromolecules 1997, 30, 8019. 108. Shusharina, N. P.; Linse, P.; Khokhlov, A. R. Macromolecules 2000, 33, 8488. 109. Pincus, P. Macromolecules 1991, 24, 2912. 110. Netz, R. R. Europhys. Lett. 1999, 47, 391. 111. Pochan, D.; Chen, Z.; Cui, H.; Hales, K.; Qi, K.; Wooley, K. L. Science 2004, 306, 94. 112. Castelnovo, M.; Joanny, J.-F. Macromolecules 2002, 35, 4531. 113. Walter, H.; Harrats, C.; Muller-Buschbaum, P.; Jerome, R.; Stamm, M. Langmuir 1999, 15, 1260. 114. Vasilevskaya, V. V.; Potemkin, I. I.; Khokhlov, A. R. Langmuir 1999, 75, 7918. 115. Potemkin, I. I.; Vasilevskaya, V. V; Khokhlov, A. R. J. Chem. Phys. 1999, 111, 2809. 116. Nyrkova, I. A.; Khokhlov, A. R.; Doi, M. Macromolecules 1993, 26, 3601. 117. Semenov, A. N.; Nyrkova, I. A.; Khokhlov, A. Macromolecules 1995, 28, 7491. 118. Lodge, T. P.; Hillmyer, M. A.; Zhou, Z.; Talmon, I. Macromolecules 2004, 37, 6680. 119. Li, Z.; Kesselman, E.; Talmon, Y; Hillmyer, M. A.; Lodge, T. P. Science 2004, 306, 98. 120. Limberger, R. E.; Potemkin, I. L; Khokhlov, A. R. J. Chem. Phys. 2003, 119, 12023. 121. Kyrylyuk, A. V; Fraaije, J. G. E. M. J. Chem. Phys. 2004, 727, 2806. 122. Lodge, T. P.; Bang, J.; Park, M. J.; Char, K. Phys. Rev. Lett. 2004, 92, 145501. 123. Kramerenko, E. Y; Khokhlov, A. R.; Reineker, P. J. Chem. Phys. 2003, 779, 4945. 124. Harada, A.; Kataoka, K. Macromolecules 1995, 28, 5294. 125. Harada, A.; Kataoka, K. Science 1999, 283, 65. 126. Harada, A.; Kataoka, K. Macromolecules 2003, 36, 4995. 127. Kataoka, K.; Harada, A.; Wakebayashi, D.; Nagasaki, Y. Macromolecules 1999, 32, 6892. 128. Katayose, S.; Kataoka, K. J. Pharm. Sci. 1998, 87, 160. 129. Harada, A.; Kataoka, K. Macromolecules 1998, 31, 288. 130. Talingting, M. R.; Munk, P.; Webber, S. E.; Tuzar, Z. Macromolecules 1999, 32, 1593. 131. Plestil, J.; Kriz, J.; Tuzar, Z.; Prochazka, K.; Melnichenko, Y. B.; Wignall, G. D.; Talingting, M. R.; Munk, P.; Webber, S. E. Macromol. Chem. Phys. 2001, 202, 553. 132. Lysenko, E. A.; Bronich, T. K.; Slonkina, E. V; Eisenberg, A.; Kabanov, V. A.; Kabanov, A. V. Macromolecules 2002, 35, 6344. 133. Lysenko, E. A.; Bronich, T. K.; Eisenberg, A.; Kabanov, V. A.; Kabanov, A. V. Macromolecules 1998, 37, 4516. 134. Kabanov, A. V; Bronich, T. K.; Kabanov, V. A.; Yu, K.; Eisenberg, A. Macromolecules 1996, 29, 6797. 135. Bronich, T. K.; Kabanov, A. V; Kabanov, V. A.; Yu, K.; Eisenberg, A. Macromolecules 1997, 30, 3519. 136. Solomatin, S. V; Bronich, T. K.; Bargar, T. W.; Eisenberg, A.; Kabanov, V. A.; Kabanov, A. V. Langmuir 2003, 79, 8069. 137. Solomatin, S. V; Bronich, T. K.; Eisenberg, A.; Kabanov, V. A.; Kabanov, A. V. Langmuir 2004, 20, 2066. 138. Lysenko, E. A.; Bronich, T. K.; Eisenberg, A.; Kabanov, V. A.; Kabanov, A. V. Macromolecules 1998, 37, 4511.
212
Block Copolymers in Solution: Fundamentals and Applications
139. Bronich, T. K.; Popov, A. M.; Eisenberg, A.; Kabanov, V. A.; Kabanov, A. V. Langmuir 2000, 16, 481. 140. Kabanov, A. V; Kabanov, V. A. Adv Drug Deliv Rev 1998, 30, 49. 141. Shuai, X.; Merdan, T.; Unger, R; Wittmar, M.; Kissel, T. Macromolecules 2003,36, 5751. 142. Weaver, J. V. M.; Armes, S. P.; Liu, S. Macromolecules 2003, 36, 9994. 143. Weaver, J. V. M.; Tang, Y.; Liu, S.; Iddon, P. D.; Grigg, R.; Billingham, N. C; Armes, S. P.; Hunter, R.; Rannard, S. P. Angew. Chem., Int. Ed. 2004, 43, 1389. 144. Ghoreishi, S. M.; Li, Y.; Bloor, D. M.; Warr, J.; Wyn-Jones, E. Langmuir 1999,15, 4380. 145. Li, Y; Ghoreishi, S. M.; Warr, J.; Bloor, D. M.; Holzwarth, J. E; Wyn-Jones, E. Langmuir 1999, 75, 6326. 146. Gohy, J.-E; Mores, S.; Varshney, S. K.; Jerome, R. Macromolecules 2003, 36, 2579. 147. Herve, P.; Destarac, M.; Berret, J.-E; Lal, J.; Oberdisse, J.; Grillo, I. Eur. Lett. 2002, 58, 912. 148. Berret, J.-E; Cristobal, G.; Herve, P.; Oberdisse, J.; Grillo, I. Eur. Phys. J. E 2002, 9, 301. 149. Berret, J.-E; Herve, P.; Aguerre-Chariol, O.; Oberdisse, J. J Phys. Chem. B 2003, 107, 8111. 150. Berret, J.-E; Vigolo, B.; Eng, R.; Herve, P.; Grillo, I.; Yang, L. Macromolecules 2004,37, 4922. 151. Fujita, H.; Ooya, T.; Yui, N. Macromolecules 1999, 32, 2534. 152. Ikeda, T.; Watabe, N.; Ooya, T.; Yui, N. Macromol. Chem. Phys. 2001, 202, 1338. 153. Lee, S. C.; Choi, H. S.; Ooya, T.; Yui, N. Macromolecules 2004, 37, 7464. 154. Bossard, E; Sfika, V.; Tsitsilianis, C. Macromolecules 2004, 37, 3899. 155. Schlaad, H.; Antonietti, M. Eur. Phys. J. E 2003, 10, 17. 156. Vandermeulen, G. W. M.; Klok, H.-A. Macromol. Biosci. 2004, 4, 383. 157. Cornelissen, J. J. L. M.; Fischer, M.; Sommerdijk, N. A. J. M.; Nolte, R. J. M. Science 1998, 280, 1427. 158. Pechar, M.; Kopecekova, P.; Joss, L.; Kopecek, J. Macromol. Biosci. 2002, 2, 199. 159. Vendermeulen, G. W. M.; Tziakos, C.; Klok, H.-A. Macromolecules 2003, 36, 4107. 160. Klok, H.-A.; Vandermeulen, G. W. M.; Nuhn, H.; Rosier, A.; Hamley, I. W.; Castelletto, V.; Xu, H.; Sheiko, S. Faraday Discuss. 2005, 128, 29. 161. Su, J. Y; Hodges, R. S.; Kay, C. M. Biochemistry 1994, 33, 15501. 162. Vandermeulen, G. W. M.; Hinderberger, D.; Xu, H.; Sheiko, S.; Jeschke, G.; Klok, H.-A. Chem. Phys. Chem. 2004, 5, 488. 163. Vandermeulen, G. W. M.; Tziatzios, C.; Klok, H.-A. Macromolecules 2003, 36, 4107. 164. Mutter, M.; Gassmann, R.; Buttkus, U.; Altmann, K.-H. Angew. Chem., Int. Ed. 1991, 30, 1514. 165. Lubbert, A.; Castelletto, V.; Hamley, I. W.; Klok, H.-A. Langmuir submitted. 166. Crespo, J. S.; Lecommandoux, S.; Borsali, R.; Klok, H.-A.; Soldi, V. Macromolecules 2003, 36, 1253. 167. Burkoth, T. S.; Benzinger, T. L. S.; Jones, D. N. M.; Hallenga, K.; Meredith, S. C.; Lynn, D. G. J. Am. Chem. Soc. 1998, 120, 7655. 168. Burkoth, T. S.; Benzinger, T. L. S.; Urban, V; Lynn, D. G.; Meredith, S. C.; Thiyagarajan, P. J. Am. Chem. Soc. 1999, 727, 7429. 169. Thiyagarajan, P.; Burkoth, T. S.; Urban, V; Seifert, S.; Benzinger, T. L. S.; Morgan, D. M.; Gordon, D.; Meredith, S. C.; Lynn, D. G. J. Appl. Cryst. 2000, 33, 535. 170. Hartgerink, J. D.; Beniash, E.; Stupp, S. I. Proc. Natl Acad. Sci. USA 2002, 99, 5133. 171. Hartgerink, J. D.; Beniash, E.; Stupp, S. I. Science 2001, 294, 1684.
Polyelectrolyte Block Copolymers 172. 173. 174. 175. 176. 177. 178. 179. 180. 181. 182. 183.
213
Collier, J. H.; Messersmith, P. B. Adv. Mater. 2004, 16, 907. Hamley, I. W.; Ansari, I. A.; Castelletto, V.; Klok, H.-A.; Nuhn, H.; Rosier, A. 2004. Santoso, S. S.; Vauthey, S.; Zhang, S. Curr. Opin. Colloid Interface Sci. 2002, 7, 262. Wang, C.; Stewart, R. J.; Kopecek, J. Nature (London) 1999, 397, 417. Kopecek, J.; Tang, A.; Wang, C.; Stewart, R. J. Macromol. Symp. 2001, 774, 31. Tang, A.; Wang, C.; Stewart, R. J.; Kopecek, J. J. Controlled Release 2001, 72, 57. Wang, C.; Kopecek, J.; Stewart, R. J. Biomacromolecules 2001, 2, 912. Nowak, A. P.; Breedveld, V.; Pakstis, L.; Ozbas, B.; Pine, D. J.; Pochan, D.; Deming, T. J. Nature 2002, 417, 424. Pochan, D. J.; Pakstis, L.; Ozbas, B.; Nowak, A. P.; Deming, T. J. Macromolecules 2002, 35, 5358. Thunemann, A. R; Beyermann, J.; Kukula, H. Macromolecules 2000, 33, 5906. Discher, D. E.; Eisenberg, A. Science 2002, 297, 967. Ma, Y.; Tang, Y.; Billingham, N. C.; Armes, S. P.; Lewis, A. L.; Lloyd, A. W.; Salvage, J. Macromolecules 2003, 36, 3475.
5 Adsorption 5.1
INTRODUCTION
Adsorption of block copolymers is relevant to their interfacial activity and hence applications for example in lubrication, coatings technology, emulsification or as templates for patterning inorganic materials. Adsorption of block copolymers has been discussed in a review of polymers at interfaces.1 The subject is also briefly covered in several texts.2,3
5.2 ADSORPTION AT THE AIR-WATER INTERFACE 5.2.1
ADSORPTION OF NEUTRAL BLOCK COPOLYMERS
Adsorption is typically studied as a function of surface area, often correlated to surface pressure-area (n-A) isotherms measured using a Langmuir trough. Two types of conformational transition are observed upon compression of adsorbed block copolymer layers, depending on the interaction of the soluble block with the surface. In the case of no preferable interaction, a gradual transition between noninteracting polymer mushrooms and stretched polymer brushes is observed upon compression. These systems are studied as models for 'tethered chains'. Representative data obtained for a PEO-b-PFMA diblock are shown in Figure 5.1.4 In contrast, when the soluble block has a tendency to adsorb to the interface then at low surface density an adsorbed pancake conformation is observed. Upon compression, this block is repelled from the surface and a brush is formed in which the insoluble block acts as a 'buoy' (Figure 5.2). The continuous change of conformation is exemplified by PDMS-b-PS diblocks at the surface of ethyl benzoate5 or DOP.6 The surface-active PDMS block acts as a buoy for the PS block (Figure 5.2), which is consequently depleted from the surface (the range of the depletion layer increases with chain length6). The scaling behaviour of tethered layer thickness with surface coverage and copolymer chain length was probed, and compared with theory. Good agreement was found with numerical self-consistent field (SCF) calculations,7,8 although not with analytical
Block Copolymers in Solution: Fundamentals and Applications © 2005 John Wiley & Sons, Ltd.
I. W. Hamley
216
Block Copolymers in Solution: Fundamentals and Applications
Figure 5.1 Representative surface pressure-area isotherm for a PEO-b-PFMA diblock film at the air-water interface, with schematic of the monolayer structures. Figure adapted from Hussain et al. Reproduced by permission of American Chemical Society.
Figure 5.2 Adsorption of block copolymers at the air-water interface in the case that the hydrophobic block forms a 'buoy' to which the hydrophilic block is tethered.
results in the strong segregation limit.9 Figure 5.3 shows fits to neutron reflectivity data for a PS-b-PDMS diblock on ethyl benzoate, along with model density profiles.5 These are based on parabolic functions, as anticipated for tethered polymer brushes. The quality of fit can be improved by allowing for an exponential tail and a depletion layer near the air-liquid interface. The adsorption behaviour of amphiphilic block copolymers at the air-water interface has mainly been studied for copolymers containing PEO. In particular, experiments have largely focused on two main classes: PS-b-PEO diblocks10-16 and PMMA-b-PEO17-20 diblocks. Phase transitions in monolayers of these block
Adsorption
217
Figure 5.3 Neutron reflectivity data and model fits for a PDMS-b-PS diblock on ethyl benzoate.5 (a) Neutron reflectivity, shown normalized by the reflectivity from the solvent, (b) Corresponding PS density profiles. The long dashed line corresponds to a parabolic function with exponential tail, the solid line to a parabolic function with a depletion layer and the short dashed line to a parabolic function with both an exponential tail and a depletion layer. Reproduced by permission of American Institute of Physics.
copolymers at the air-water interface have been investigated via surface pressurearea isotherm measurements, neutron reflectivity and Brewster angle microscopy. In the PS-b-PEO system, at low coverage, the favourable interaction between PEO and water leads to adsorption of PEO at the interface.11,14-16 As the layer is compressed, the repulsive interactions between PS and PEO lead to desorption of PEO from the interface. Upon further compression, the PEO is pushed out of the surface layer into the water subphase to form a brush, and the surface is covered with PS blocks.12,15 This surface phase transition has a first order character,11,15 although SCF theory predicts it to be second order.10 Due to the effective repulsive interaction between the PEO and the interface, a depletion region develops near the
218
Block Copolymers in Solution: Fundamentals and Applications
Figure 5.4 Surface pressure-area isotherms for a series of PS-b-PEO diblocks with constant PS chain length (NPS = 31) and increasing PEO chain length, as indicated.14 Reproduced by permission of Springer Verlag.
interface. The depletion layer is particularly evident for longer PEO chains. Indeed, for shorter chains the plateau in the n-A isotherm is much less pronounced,4,14,15 as shown in Figure 5.4. It was shown that such n-A curves collapse into a master curve when the area per monomer is normalized by the PEO chain length.14,16 Brewster angle microscopy confirmed a first-order transition from an adsorbed PEO layer to a PEO brush in the plateau region of the n-A isotherm for copolymers with long PEO blocks.15 The concentration profiles obtained from neutron reflectivity have been compared with those computed using various 'brush' models, especially SCF theory.10,14 Similarly, the adsorbed amount and the hydrodynamic layer thickness (determined via optical microscopy and dynamic light scattering, respectively) have been compared with SCF calculations.21 The organization at the air-water interface of a PS-b-PEO diblock as a function of bulk copolymer concentration has been probed via neutron reflectivity, using selectively labelled chains to obtain the full set of partial structure factors.20 Surface micelles were observed well below the bulk cmc. The PS cores form buoys (lenses) at the surface [Figure 5.5(a)]. As the copolymer concentration increases towards and over the bulk cmc, the adsorbed amount increases and the thickness of the PS
Adsorption
219
Figure 5.5 (a) Schematic of structure of adsorbed PS-b-PEO diblocks at the air-water interface on increasing copolymer concentration (A-C).20 (b) Schematic of structure of diblock EE 21 2-b-EOn2 as a function of increasing surface pressure.23 Reproduced by permission of American Chemical Society.
top layer decreases (and its density increases) as the PS forms less curved lenses. The PEO chains are consequently forced to adopt conformations that extend more into the aqueous subphase. At these concentrations, the reflectivity data on selectively labelled chains indicates structuring of the PS layer, which may result from a phase transition to a structure with in-plane order. The concentration at
220
Block Copolymers in Solution: Fundamentals and Applications
which an onset of layering in the PS is observed corresponds to a point at which an apparent kink in the surface tension data is observed. Surface quasi-elastic light scattering (SQELS) has been used to probe the viscoelasticity of adsorbed layers, in particular to determine the shear moduli and associated relaxation times.22 This technique was also used to obtain the dynamic elasticity and dilational viscosity of a series of PS-b-PEO and PEO-b-PS-b-PEO copolymers.16 In the PMMA-b-PEO system, detailed contrast variation neutron reflectivity revealed that the adsorbed layer contains both components at low surface coverage whereas upon compression the PEO extends into the aqueous subphase and the PMMA is expelled into the air phase.18,22 In the case of an asymmetric PEE-b-PEQ diblock, compression leads initially to the formation of finite domains.23 When the area fraction of domains reaches about 95%, corresponding to a surface area of around 6 A2 per EO monomer, PEO desorption is observed. This surface area is lower than values typically observed for other (symmetric) diblocks. The resulting layer structure is indicated in Figure 5.5(b). Ultimately a homogeneous PEE surface layer is formed, the thickness of which increases linealy with grafting density. Adsorption of amphiphilic silicon-containing diblock copolymers at the air/water interface has been investigated by Matsuoka and coworkers.24-26 Using n-A measurements, and X-ray reflectivity, they confirmed that a monolayer can be formed at an appropriate surface pressure.25 An interesting observation has been made for a diblock containing a highly ionic hydrophilic block, PDESCB attached to a hydrophobic PMMA block.26 It was found by SAXS, DLS and dye solubilization that micelles were formed in bulk, and yet no decrease of surface tension and no evidence of adsorption at the air-water interface was observed. This puzzling observation has yet to be fully explained, although it was suggested that destabilization of the amphiphilic monolayers due to electrostatic repulsion might play a role (however, in this case, micelles should also not be stable). The structure of an adsorbed monolayer of a PhI-b-PSSA at the air-water interface has been studied by X-ray reflectivity by Matsuoka and coworkers.27 The PhI forms a buoy to which the PSSA brush is tethered. Surface pressure-area isotherms for monolayers of telechelic PEO-based block copolymers (with C12 and C16 end groups) at the air-water interface reveal two plateaux.28 The first is ascribed to loops in the PEO midblock (and is observed for the homopolymer alone). Following this plateau the pressure rises rapidly as the end groups anchor onto the substrate. The second plateau at high density is due to dissolution of the polymer chains into bulk water. Brush theory was used to calculate isotherms which were in reasonable agreement with the experiments. In addition, polymer adsorption and desorption kinetics were calculated. The structure of aggregates formed at the air-water interface by copolymers with more complex architectures has also been investigated for mixed arm PEO-b-PS3 and PEO-b-PS2 star copolymers.29 Langmuir-Blodgett deposition of the monolayers onto silicon substrates enabled the surface morphology, comprising circular domains, to be imaged. As the surface pressure increased, the packing density of circular domains increased.
Adsorption 5.2.2
221
ADSORPTION OF POLYELECTROLYTE BLOCK COPOLYMERS
Adsorption of diblock copolymers containing a polyelectrolyte at the air-water interface has been studied by neutron reflectivity. Thomas and coworkers have studied the adsorption of a number of PDMA-b-PMMA diblocks as a function of concentration and pH.30-32 Reducing pH increases the positive charge on the PDMA block, which has a pKa of about 7.3. Considering the effect of concentration at a fixed high pH, at low concentration the adsorbed layer was found to consist of a mix of the two blocks, with about half of the layer immersed in the underlying water. At higher concentration, a multilayer structure was formed with a central layer rich in the more hydrophobic PMMA and the outer layers comprising PDMA. This structure was ascribed to the formation of surface micelles with a PMMA core and PDMA extending up into the air-water interface and down into the aqueous subphase. A transition to this structure, characterized by a large increase in adsorbed layer thickness, is induced by increasing pH or by increasing concentration at a fixed high pH. Increasing the fractional charge on the polyelectrolyte however disfavours the formation of micelles.31 A study of the structure of adsorbed layer of a PDMA-b-PnBMA was undertaken using neutron and X-ray reflectivity suggested a similar structure with a submerged PDMA block attached to a PnBMA 'buoy'.33 Ellipsometry was used to probe the pH-controlled adsorption of diblocks comprising hydrophobic PDEA or PDPA and methyl- or benzyl-quaternized PDMA or sulfopropyl betainized PDMA.34 At low solution pH, where the copolymers are molecularly dissolved, low adsorbed amounts were observed. At higher solution pH, a pronounced increase in adsorbed amount accompanied the formation of micelles. Changes in the conformation of polyelectrolyte brushes in PtBS-b-PSS diblocks upon varying the interfacial area/molecule have been investigated by neutron reflectivity.35 For salted brushes, i.e. those for which charge is screened by the added NaCl, a parabolic density profile was observed (cf. Figure 5.3), as for uncharged polymer brushes. An additional excess density observed near the interface is due to adsorption of the hydrophobic backbone of the polyelectrolyte. The density profile was also measured for so-called 'osmotic brushes' in which the brush is swollen by its own trapped counterions. The profiles could be fitted using Gaussian functions, in agreement with the predictions of SCF calculations.36 However, the thickness of the brush was dependent on grafting density, in contrast to the SCF calculations but in agreement with models that account for the lateral inhomogeneity of counterion distribution37,38 and the reduction in entropy of counterions due to the space occupied by the polymer backbone (which is compensated by an extension of chains with increasing grafting density38). The latter is supposed to be the main factor. Forster and coworkers investigated the adsorption of a PEE-b-PSSA diblock at the air-water interface using X-ray reflectivity.39 The polyelectrolyte block was shown to form an osmotically swollen brush with thickness independent of polymer concentration (the hydrophobic block stretched in proportion to grafting density).
222
Block Copolymers in Solution: Fundamentals and Applications
At high salt concentration the contraction of the brush was observed, the thickness scaling as L ~ cs-1/3, as discussed further in Sections 4.2 and 5.3.2.
5.3 ADSORPTION ON SOLID SUBSTRATES The adsorption of surfactants and copolymers onto solid substrates is relevant to their use in applications including detergency, oil recovery and lubrication.40,41 The adsorption isotherm has been investigated for neutral and polyelectrolyte block copolymers on different substrates. In addition, the morphology of surface micelles has been examined, and the use of these in patterning inorganic materials (metals, metal oxides) explored. 5.3.1
ADSORPTION OF NEUTRAL BLOCK COPOLYMERS
The adsorption of PEO-containing block copolymers on solid substrates has been extensively investigated. Non-fouling surfaces coated with PEO are of interest since protein, plasma adsorption, etc., is reduced. There has been controversy as to whether surface micellization is correlated to the bulk cmc.42 Ellipsometry on Pluronic copolymers adsorbed onto (hydrophilic) silica provided evidence for an increase in adsorbed amount just below the bulk cmc.43 Surface association was discussed, especially since the increase in adsorbed amount occurred cooperatively but since the adsorbed layer was substantially less thick than the hydrodynamic radius, adsorption of intact micelles did not occur. It was concluded that PEO rather than PPO has the stronger tendency to adsorb,43 although the PPO block will have a stronger tendency to adsorb onto hydrophobic substrates.44,45 For PEO-b-PTHF-b-PEO and PEO-b-PPO-b-PEO copolymers, Tiberg and coworkers found that the amount of adsorbed copolymer increases two orders of magnitude below the cmc for the former system, and close to the cmc for the latter one.46 This increase was related to the cooperative formation of surface micellar structures on hydrophilic surfaces. In contrast, at hydrophobic surfaces, monolayers were observed. AFM on related alcohol ethoxylate surfactants provided direct evidence for the formation of surface micelles on a hydrophilic substrate and a monolayer on a hydrophobic substrate.47 Measurements using the surface forces apparatus on a PEO-£-PTHF-/?-PEO triblock also provided evidence for short-range attractions between surface aggregates due to bridging of the gap between hydrophilic quartz surfaces.48 In contrast to these findings, Muller and coworkers reported no correlation between the adsorption of PEO-b-PDLL diblocks onto hydrophilic silica and micellization in bulk.49 By comparing the adsorption of the diblock and that of PEO homopolymer, it was found that the hydrophobic block drives adsorption (in contrast to the conclusions of Malmsten et al. on PEO-b-PPOb-PEO triblocks43). On hydrophobized silica, the same diblocks formed a monolayer, anchored by the adsorbed PDLL blocks.50
Adsorption
223
Adsorption of PEQ-b-PPO-b-PEO block copolymers onto hydrophobic surfaces has been studied by surface plasmon resonance (SPR). SPR exploits the coupling between incident photons from a laser and surface electrons (plasmons) at the metal/solution interface. The angle at which there is a minimum in the intensity of internally reflected light is sensitive to the dielectric properties of the substrate. Using this technique, Green et al. found a dense layer of PPO adsorbed at the hydrophobic substrate.51 Adsorption of a series of Pluronics onto model hydrophobic self-assembled monolayer surfaces has been studied by AFM and SPR.45 Adsorption isotherms were obtained from SPR, and the adsorbed amount was found to go through a maximum near the cmc, contrary to the Langmuir isotherm. The adsorption process was found to be partly irreversible. Surface micelle structures were only observed (by AFM) for the more hydrophobic copolymers, the more hydrophilic ones adsorbing as unimers to form a uniform layer. Force-distance curves suggested a brush-like structure of adsorbed micelles. The surface structure was shown to depend on the tip force applied. It was suggested that higher force scans revealed the micellar core structure, in contrast to low force scans which probe the polymer brush coronas. In a sister paper,52 the kinetics of adsorption was analysed by SPR. Enhanced adsorption rates for more hydrophobic triblocks was observed above the cmc, due to adsorption of micelles. Adsorption of unimers was inferred for more hydrophilic copolymers. Adsorption of Pluronics P105, F108, F88 and F68 onto PS latex particles has been investigated by sedimentation field-flow fractionation and dynamic light scattering.53 The thickness and density of the adsorbed layer was studied as a function of particle size. The mobility of the adsorbed chains was studied by electron spin resonance (ESR) using spin-probe labelled polymers. For a given particle size, the layer thickness was found to increase with PEO endblock chain length, and the dynamics to speed up, i.e. the rotational correlation rate from ESR increased. Later, Hariharan et al. analysed the adsorption data and showed it was consistent with the star-like polymer model (Section 2.7.1), i.e. the layer thickness L and particle radius R satisfied the relationship:54
Here a is the surface chain density, k is a constant (close to unity), v is a binary cluster integral characterizing the polymer-solvent interaction, lc is the contour length of the chains and aK is the statistical segment or Kuhn length. A plot of (LIR + 1)5/3 against vo 1 / 3 /R provided a straight line with a slope that yielded v/aK.54 It has been shown using neutron reflectivity that layering of block copolymer micelles occurs at a hard wall.55 Adsorption of Pluronic P85 onto quartz from solution in D2O was studied. Monte Carlo simulations of interacting hard spheres were able to account for the observed density profiles.55,56 Later, the model was refined to account for the tethered chains.56
224
Block Copolymers in Solution: Fundamentals and Applications
Emoto et al. studied adsorption of PEG-b-PLA micelles onto SAM-coated Si, glass and plasma-treated PDMS.57 They found that the micelle structure was disrupted upon deposition. However, cross-linking the PLA core via terminal methacryloyl moieties led to stable micelles that adsorbed intact. This group have reviewed the applications of PEG-based block copolymers at surfaces for biomedical applications.58 Their own work involves the used of PEG-b-PLA in which the PEG chain is end-terminated with an acetal or aldehyde group (Figure 5.6),59 which can be used to immobilize ligands including proteins, peptides and sugars. The PLA can be cross-linked (via end-functionalization with methacryloyl units) and confers additional biodegradability. Similarly, surfaces with biorecognition properties have been developed using PEG-b-PLA diblocks with biotinylated PEG chains.60,61
Figure 5.6 Functionalized PEG layer on a biodegradable PLA substrate, prepared from the self-assembly of PEG-b-PLA diblocks.59 Reproduced by permission of American Chemical Society.
The adsorption of copolymers with more complex architectures has also been investigated. For example, AFM has been used to probe the adsorption of PS7P2VP7 heteroarm polymers on silicon and mica.62 In acidic aqueous solutions, adsorption of single molecules and of micelles was noted, the rigid PS micellar core forming pronounced 'humps'. Addition of toluene, a good solvent for PS, lead to a conformational transition that caused flattening of the micelles as the PS chains extended from the micellar core. Surface micellar structures at the air-water
Adsorption
225
interface have also been imaged (by deposition of the film onto silicon wafers) for a PEO-b-PS3 starblock copolymer.29 Responsive polymer surfaces sensitive to humidity have been prepared from PS-b-PI diblocks with hydrophilic dimethylamine or sulfobetaine end groups.63 The PS block adsorbed onto the silicon substrate, leaving dangling end groups. The reduction in the contact angle was monitored upon exposure to water vapour. The stratification resulting from adsorption of bidisperse mixtures of PS-b-P2VP diblocks onto silicon in good solvent conditions has been investigated by neutron reflectivity.64 The system was considered as a model bidisperse 'polymer brush'. The volume fraction profiles were compared with analytical self-consistent mean field theory. Experiments in poorer, near-6, conditions in cyclohexane showed much reduced layering of long and short chains. Tao et al. determined adsorption isotherms for a series of near symmetric PS-bPCEMA diblocks, adsorbing as brushes onto silica from a PS-selective solvent.65 The brush areal densities were compared with the predictions of the model due to Marques et al. (Section 5.5). The influence of a surface on the isotropic-nematic transition in a solution of cylindrical micelles formed by PB-b-PEO diblocks in aqueous solution has been compared with the bulk behaviour using neutron reflectivity to probe the surfaceinduced ordering and SANS to probe the bulk phase transition.66 The phase transition was shifted to lower polymer volume fractions in the layer close to the silicon-solution interface. 5.3.2
ADSORPTION OF POLYELECTROLYTE BLOCK COPOLYMERS
The adsorption of polyelectrolyte block copolymers has also been investigated, for polyampholyte diblocks and also for neutral-b-polyelectrolyte copolymers. Obviously for this type of copolymer, adsorption will be strongly influenced by the charge on the copolymer relative to the surface. Surface micellization by these types of copolymer has also been observed, as discussed in the following section. The adsorption of a PtBS-b-PSS diblock onto PS latices from water has been studied via gravimetric determination of the adsorbed amount, following centrifugation.54 The adsorbed amount increases with ionic strength (added NaCl) to give a dense polyelectrolyte brush. The thickness of the adsorbed layer was compared with scaling theories, as discussed in the previous section. A simple blob model for polyelectrolyte brushes on spherical surfaces was developed, leading to scaling predictions for brush layer thickness L ~ cs-m with m = —1/6 for brushes adsorbed on a sphere with infinite radius (m = —0.17 was observed for dense adsorbed layers) and m — —1/10 for adsorption on a sphere with zero radius (m = —0.11 was observed for bulk micelles). This behaviour depends much less weakly on salt concentration than predicted by earlier scaling models. The adsorption of this type of diblock (PtBS-b-PNaSS) on hydrophobized mica in the salted brush regime was reported to be consistent with L ~ cs-1/3.67 The same behaviour was reported earlier for a PEE-b-PSSA diblock adsorbed at the air-water interface.39
226
Block Copolymers in Solution: Fundamentals and Applications
Stamm and coworkers have investigated the adsorption of PDMA-b-PMAA polyampholyte diblocks from solution onto silicon substrates.68,69 The adsorbed amount determined via ellipsometry using the Langmuir isotherm was measured as a function of pH, polymer concentration and salt concentration. The adsorption was maximal at the isoelectric point (Figure 5.7). At a fixed pH, increasing salt
Figure 5.7 Amount of PDMA-b-PMAA diblock adsorbed as a function of pH at a fixed polymer concentration.69 The arrows indicate where the silicon surface and the polyampholyte carry a positive or negative charge. Reproduced by permission of Springer Verlag.
concentration increased the adsorbed amount.68 Adsorption kinetics were also analysed, allowing the diffusion coefficient for the early stage of adsorption to be measured.68 Later, the structure of dried adsorbed layers was imaged by AFM.69 This confirmed the presence of surface micelles, the size of which was weakly dependent on pH, although the adsorbed amount was strongly pH dependent. An and coworkers have used neutron reflectivity to investigate the conformation of the PDMA block in aqueous salt solutions of a PDMA-b-PMMA diblock, where the PMMA is adsorbed onto a SAM layer (of octadecyl trichlorosilane) on silica.70 The adsorption of the same diblock at the air-water interface was also studied (see Section 5.2.2). It was found that the PDMA chain was not ionized in the vicinity of the PMMA anchor, in agreement with a SCF theory model for weakly grafted poly electrolytes.1}J2 5.3.3
SURFACE MICELLES
Under appropriate conditions, it is possible for adsorption of micelles to occur, rather than single chains. This provides a route, for example, for the controlled
Adsorption
227
Figure 5.8 AFM topographic image of a PS-P2VP diblock micellar film loaded with HAuCl4 deposited onto a Si wafer.74 Reproduced by permission of American Chemical Society.
deposition of inorganic nanoparticles, prepared from precursors (such as metal salts) sequestered within the micellar core.73,74 The structure of block copolymer micelles on solid substrates has been probed on dried films by AFM.74-76 Figure 5.8 shows a representative topography image. In addition, TEM can be applied for specimens that have been suitably cryo-cooled (sometimes the specimens are also stained to enhance contrast).75,77-82 Eisenberg and coworkers have imaged surface micelles formed by amphiphilic diblocks deposited by the Langmuir-Blodgett technique at the air-water interface.83-85 The films were picked up onto supported TEM grids, then dried and shadowed with metal vapour and examined by TEM. The same group has also performed AFM on dried micelles on glass.86 Images were first obtained for PS-bqP4VP diblocks, quaternized with decyl iodide, spread as a monolayer at the airwater interface. TEM images revealed self-assembled circular surface micelles at low surface pressures (<2mNm~ 1 ). For a S26o-b-q4VP240 diblock, the surface micelles were quite regular and had an aggregation number of 131 ± 35.87 The distance between micelles (at low surface pressures) was found to be consistent with fully extended qP4VP chains extending from a central core of PS coils. An apparent first-order phase transition was detected at high pressures from a plateau in n-A isotherms. This was correlated to TEM images which suggested that the polyelectrolyte block changed from a surface adsorbed two-dimensional (2D) state 83 to a submerged aqueous state with less order (quasi-2D) at this transition. This transition was more graphically termed as being between 'starfish' micelles and 'jellyfish' micelles (Figure 5.9).86,88 In subsequent work, Langmuir films of more asymmetric copolymers of the same type at the air-water interface were investigated.84,88 Circular or starfish micelles were observed for copolymers with less than 86% PS, rod or ribbon-shaped surface micelles were observed for copolymers with between 86% and 94% PS, and planar micelles were observed for PS contents between 94% and 97% PS. These structures are shown in Figure 5.10. This sequence of phase transitions is distinct from phase
228
Block Copolymers in Solution: Fundamentals and Applications
Figure 5.9 Schematic of 'starfish' and 'jellyfish' micelles. Typical association numbers are actually p ~ 100.129 Reproduced by permission of Oxford University Press
Figure 5.10 Transmission electron micrographs of thin films of Sm-b-4VPn/C10H21I. (a) Circular surface micelles; m = 480, n = 200.89 (b) Ribbon surface micelles; m = 480, n = 34; (c) Planar surface micelles; m = 480, n = 13. Reproduced by permission of Royal Society of Chemistry.
transitions in the bulk, where the sequence lamellae - cylinders - spheres is observed on increasing the asymmetry of strongly segregated block copolymers. The rod micelles were found to be made up of two layers of the diblocks, whilst the planar aggregates appeared to be one PS block thick, although they extend up to several microns in width.86 In parallel with bulk observations, addition of PS homopolymer was found to induce changes in the morphology of the surface micelles (circle, ribbon or plane). For example, addition of homopolymer PS was observed to induce a transition from ribbon surface micelles to planar micelles in a
Adsorption
229
S260-b-qVP24o diblock. A cellular or foam morphology (i.e. the inverse of the circular surface micelle phase) was observed in some areas (up to 10%) of Langmuir-Blodgett films formed by PS-b-PNaA and PS-b-PNaMA block copolymers. The cell areas of these nanofoams ranged from 100 to 5000 nm2.89 Circular surface micelles have also been observed for nonionic block copolymers. Li et al. reported observations via TEM and AFM of starfish micelles formed by PS-b-PnBMA, PS-b-PtBMA, PS-b-PtBA and PS-b-PDMS diblock copolymers.85 Plateaux observed in n-A isotherms of the Langmuir-Blodgett films at the air-water interface were consistent with the formation of surface micelles. Meiners et al.90 and Spatz et al.14,15 have observed surface micelles of PS-bP2VP on mica or carbon. Samples were prepared by drying dip coated films of diblocks (dissolved in toluene), vitrification on solvent evaporation leading to trapping of micellar structures. Regenbrecht et al.91 probed the structure of micelles of the neutral-b-polyelectrolyte diblock copolymer PEE-b-PSSA adsorbed on solid substrates in aqueous solution by AFM on dried samples. A comparison was made between adsorption onto a neutral hydrophobic graphite substrate, and onto mica modified to be positively charged by a coating of PEI, or by ion exchange. A range of structures from spherical micelles to network-like structures, was observed upon varying ionic strength. Adsorption of micelles and vesicles was observed onto graphite. Adsorption onto strongly charged substrates led to the formation of structures not found in the bulk.76,91 The ordering of micelles of a PS-b-P2VP diblock from pH = 1 aqueous solution (where the P2VP block is protonated) has been studied for films dip coated onto silicon.92 Even in these highly acidic conditions, the deposited micelles were strongly adsorbed, and could not be removed by washing. The highest density and ordering of micelles was achieved by controlling the solvent evaporation rate, the most important experimental variable. The rate of substrate withdrawal also influenced the ordering. Surface micellar structures (stripe and spherical domain patterns) have been imaged for dried PB-b-PS diblocks using TEM.81 Webber et al.93 investigated the adsorption of PDMA-b-PMMA diblocks onto mica as a function of pH. They observed surface micelles at low pH, where PDMA is protonated, and regions of close-packed structure at higher pH for this polyelectrolyte diblock. Subsequently, they have observed hcp structures in an adsorbed layer of a related diblock, qPDMA-b-PDEA.94 The structure of the polymer layer was influenced by the electrostatic interaction between the PDMA block and the negatively charged mica. Surface plasmon resonance experiments revealed a correlation between the rate of adsorption of PS-b-P2VP diblocks from toluene and the bulk cmc - the initial rate of adsorption increasing above this concentration.95 According to Tassin et al., the adsorbed surface micelles act as a barrier that slow down the subsequent adsorption of unimers, observed as an additional slow process. A study of the adsorption kinetics of two PS-b-PEO copolymers onto dielectric substrates (glass and sapphire) using optical interferometry provided evidence for the adsorption of micelles above the cmc.96 The adsorbed amount increased steeply to the cmc,
230
Block Copolymers in Solution: Fundamentals and Applications
however it then decreased sharply and then a slower increase was observed (Figure 5.11). The rate of adsorption also showed a discontinuity at the cmc (Figure 5.11). Adsorption from very dilute solution (well below the cmc) has been found to occur via a two-step process. Ellipsometry has been used to probe the
Figure 5.11 Data from adsorption experiments (using optical interferometry to measure the surface concentration) of a PS-b-PEO diblock from cyclopentane onto dielectric substances.96 (a) Adsorbed concentration as a function of bulk copolymer concentration for adsorption onto glass (A) and sapphire (o, •), (b) Initial rate of adsorption as a function of bulk copolymer concentration. Symbols as for (a). Reproduced by permission of American Chemical Society.
Adsorption
231
adsorption of PS-b-PEO diblocks from toluene - the initial rapid adsorption was ascribed to the diffusive transport of individual chains.97 The later slower growth in thickness of adsorbed layers corresponded to strongly overlapped chains, and brush-like behaviour was then observed. A similar study on PS-b-P2VP diblock adsorption from toluene reached similar conclusions.98 The adsorption of a PPO-b-PEO diblock from aqueous solution onto hydrophilic mica and hydrophobic silica substrates has been studied using in situ AFM.42'44 Adsorption was studied as a function of concentration - close to and below the bulk cmc the copolymers adsorb as single chains or 'pre-micelle aggregates'. As concentration was increased, micelles formed densely packed structures (Figure 2.2). At a given concentration, the ordering was higher for the strongly adsorbed micelles on silica. Adsorption onto the hydrophilic silicon nitride AFM tip was observed to compete with adsorption onto either substrate (although this was a particularly strong effect when imaging weakly adsorbed layers on mica). The substrate coverage with micelles on mica increased roughly linearly with scan time at low micelle concentration. Weak adsorption on mica was analysed to provide a 2D diffusion coefficient. The effect of tip oscillation frequency on the AFM images was investigated - higher frequencies were required to image more weakly adsorbed polymer aggregates on mica. Finally, force-distance curves provided information on the nanomechanical properties of the adsorbed polymer layers. It was suggested that micelles and aggregates adsorb to silica via the hydrophobic PPO block, whilst they adsorb to mica via the hydrophilic PEO block. A model for the adsorbed structure is shown in Figure 5.12. Adsorption of shell cross-linked knedel (SCK) nanoparticles onto mica has been investigated by Wooley and coworkers. Dried films have been imaged by tappingmode AFM. The role of the charge on adsorption was explored by modification of the negatively charged mica surface with a self-assembled monolayer to make it positively charged, and also variation of the charge on the SCK nanoparticles (varying as a function of the extent of cross-linking).99 In situ measurements in a liquid cell of pH-induced changes in the adsorbed 2D array structure have also been performed.100 Further details on related structures can be found in Section 2.12.4.
5.4 SURFACE FORCES EXPERIMENTS Experiments using the surface forces apparatus (SEA) provide information on the interaction between polymer brushes tethered to locally flat surfaces (the actual geometry comprises curved half cylinders, oriented to be mutually orthogonal). The first report on the use of the surface force apparatus for measuring forces betwen adsorbed layers of block copolymers was by Hadziioannou et al.101 They attached symmetric PS-b-PVP diblock copolymers via adsorption of the PVP block, the PS block being nonadsorbing, to mica surfaces in the good solvent toluene. It was demonstrated that the PS chains were strongly stretched normal to the adsorbing surface forming a polymer brush at the liquid-solid interface. A force-distance
232
Block Copolymers in Solution: Fundamentals and Applications
Figure 5.12 Models for adsorption of diblock PO94-b-EO316 onto silica and mica at the concentrations indicated.42 The gray triangle represents the AFM tip. The grey corona chains are PEO and PPO chains (black) from the micelle cores. Reproduced by permission of American Chemical Society.
profile for a PS-b-P2VP diblock adsorbed on mica from a cylcohexane solution is illustrated in Figure 5.13. This shows that well below the 6 temperature for PS in cyclohexane (70 — 34°C) there was pronounced attraction at large plate separations, whereas at higher temperatures the force was purely repulsive.101 To mimic a layer of end-grafted polymer chains more closely, Taunton et al. investigated highly asymmetric diblock copolymers and demonstrated that PS-b-PEO diblock copolymer chains with a short PEO block can be terminally anchored on mica through the adsorption of ethylene oxide segments on the mica surface with the PS blocks dangling freely into the solvent (e.g. toluene).102,103 At high surface coverage, the PS chains were found to be strongly stretched, and the monotonically repulsive interaction profiles of two such polymer brushes were in good agreement with theoretical predictions.102,103 The interaction of end-adsorbed diblock and triblock copolymers in toluene against a bare mica surface was investigated by surface force measurements.104
Adsorption
233
Figure 5.13 Force-distance profile obtained from surface forces experiments on a symmetric P2VP-b-PS diblock adsorbed on mica in cyclohexane at 21 °C.101 Reproduced by permission of American Chemical Society.
Only monotonically repulsive forces were observed when a bare mica sheet interacted with a single layer of PS-b-PEO diblock, end adsorbed at a toluene mica interface. In contrast, clearly detectable attractive forces were found for PEOb-PS-b-PEO triblock copolymers under similar conditions. The attraction was attributed to polymer bridges in which the two PEO end blocks of a single triblock chain simultaneously adsorbed onto the two mica surfaces. The experimentally detected evolution of the force profiles from attractive to repulsive with successive compression-decompression cycles was proposed to result from the dynamics of conformational rearrangements and the subsequent migration of chains between the surfaces. The results further indicated that the relative populations of chains adsorbed in a loop or tail conformation depend on the adsorption energy of the anchoring blocks, as does the degree of stretching of the bridging chains before rupture upon separation of the mica surfaces.104 Surface force experiments were
234
Block Copolymers in Solution: Fundamentals and Applications
combined with simultaneous neutron reflectivity experiments on PS-b-PEO diblocks adsorbed on glass plates.105 Adsorption from toluene led to preferential adsorption of PEO, however due to the very high asymmetry of the copolymers studied, and the favourable adsorption energy, PS was also adsorbed. On the approach of a second coated plate, the PS volume fraction increased near the interface, indicative of strong interlayer repulsion.105 Surface force measurements on a PEO-b-PBO diblock in aqueous solution have revealed purely monotonic repulsive forces between the BO8-b-EO41 layers adsorbed onto hydrophobically modified mica (to which PBO is adsorbed).106 Ellipsometry was used to determine adsorption isotherms (adsorbed amount as a function of concentration), and the adsorbed amount as a function of time. The range of the steric repulsion was found to increase with increasing bulk copolymer concentration, whereas the concentration of an inert salt (up to 0.1 M KBr) did not influence the surface forces profiles. The adsorption kinetics were found to be described by two adsorption regimes. Initially, diffusion of polymer to the interface to the solid surface dominated, thus this was a diffusion-controlled regime. When the surface was fully covered, a crossover into a second regime occurred and the increase in surface coverage was less rapid due to interactions between block copolymer molecules at the surface. Measurements using the surface forces apparatus on a PEO-b-PTHF-b-PEO triblock also provided evidence for shortrange attractions between surface aggregates due to bridging of the gap between hydrophilic quartz surfaces.48
5.5 MODELLING ADSORPTION The adsorption of block copolymers to solid substrates has been investigated extensively by lattice self-consistent mean field calculations.7,106-110 These models consider adsorption of polymer chains, rather than micelles, and properties such as surface density, segment density profiles and adsorbed layer thickness can be calculated. Calculations for specific Pluronic copolymers (P85, P103, P104 and P105) provided concentration- and temperature-dependent data on surface tension in good agreement with measured values.111 In addition, information on the distribution of PPO and PEO segments was obtained that indicated a top layer of PPO, with little water pentration and a thicker PEO layer underneath, extending into the aqueous subphase. Allowing for polydispersity of the block copolymer led to a less clear discontinuity in surface tension data at the cmc, in agreement with experimental observations. Schillen and coworkers examined the adsorption of BO8-b-EO41 using experiment (as mentioned above) and theory.106 The theory was able to reproduce the observed adsorption isotherm, especially when allowance was made for polydispersity. In a polydisperse polymer, preferential adsorption of the chains with longer hydrophobes was noted. Lattice Monte Carlo simulations have been employed to investigate block copolymer adsorption onto solid substrates. Balazs et al. investigated the adsorption of ABA triblocks with solvophobic endblocks onto flat substrates, and compared it
Adsorption
235
with that of AB diblocks.112 They explored the optimal conditions for adsorption of a 'steric stabilization' layer, and found that this could be achieved by adsorption from dilute solution of relatively short polymers, since these adsorb in a loop conformation. The presence of two hydrophobes per chain for the triblock leads to stronger adsorption than for the case of diblocks. Later, they investigated adsorption onto a corrugated surface and found selective adsorption of the end blocks of an ABA triblock onto the ridges, whilst the midblock spanned the intervening wells.113 Consequently, the adsorption of triblocks acts to smooth a rough surface. Zhan and Mattice performed lattice Monte Carlo simulations of adsorption of model A10B10 diblocks in a selective solvent for A.114,115 They found that below and close to the cmc, adsorption of free chains only occurs. In contrast, at concentrations much higher than the cmc, adsorption of micelles predominates.115 They were able to reproduce the observations of Munch and Cast concerning the adsorbed mass per unit area, F, as a function of concentration - in particular a linear increase in F below the cmc, followed by a drop to a lower adsorbed amount above the cmc (only weakly dependent on concentration) (Figure 5.11). At a concentration about ten times the cmc, the adsorption of micelles was marked by an increase in the concentration dependence of adsorbed polymer.115 This regime was not encountered by Munch and Gast, who suggest adsorption of micelles commences at the cmc. Zhan and Mattice suggest that the increased adsorption rate above the cmc observed by Munch and Gast can be interpreted as arising from collisions of micelles with the surface, that do not subsequently adsorb. Wang and coworkers have developed the lattice Monte Carlo simulation approach to model adsorption of di- and triblocks from aqueous solution by considering the interaction between the blocks as well as that between each block and the substrate.116 This is in contrast to the work by Balazs et al. who only considered the interaction between the adsorbing block and the substrate (Zhan and Mattice did allow for nonzero interaction parameters in their study of adsorption from a selective solvent,115 although not for a nonselective solvent117,118). Wang et al. found that a strongly adsorbed diblock behaves like an adsorbed homopolymer. The formation of a collapsed 'globule' structure in a poor solvent was also examined, and found to be inhibited somewhat by the presence of the substrate. Binder and coworkers performed off-lattice Monte Carlo simulations (using a bead-spring model for the polymer chains) of asymmetric diblocks (/= 0.25) with short-range attraction to the surface for the minority block.119 Depending on surface coverage and surface interaction energy, the formation of mushrooms, brushes or surface micelles were observed. The cmc was found to be similar to that in bulk. Large surface micelles were found to contain very dense micellar cores containing crystalline layers. On the basis of a scaling argument (also the Daoud-Cotton model), Johner and Joanny have suggested that even from a micellar solution, only free chains adsorb and never complete micelles, due to the potential barrier for micellar adsorption.120 On the other hand, the formation of surface micelles by adsorption of diblocks at a solid surface from a selective solvent has been predicted on the basis of a modified 'brush' theory.121 The result of Johner and Joanny is also contradicted by experiment. Earlier work by this group discussed the adsorption of brush layers
236
Block Copolymers in Solution: Fundamentals and Applications
from micellar and lamellar solutions.122 Again adsorption of micelles was not considered. Adsorption was strongly influenced by the balance of van der Waals attraction between the adsorbed block and the stretching energy of the 'buoy' brush, which depends on copolymer asymmetry. The adsorption of neutral-b-polyelectrolyte diblocks was analysed in a subsequent paper.123 The formation of a dense brush layer was found to be possible only if the charge on the polyelectrolyte block was not too high. The formation of micelles in bulk was also constrained by the fraction of charged monomers. Adsorption at the oil-water interface was also discussed, and a range of parameters identified for which the interfacial tension was reduced to zero in the presence of added copolymer. Monte Carlo simulations have also been performed of the adsorption of triblock copolymers from a nonselective solvent onto a flat surface from a nonselective solvent.124 The results were compared with Langmuir adsorption studies using ellipsometry to probe adsorption via the end blocks of a PEO-b-PS-b-PEO triblock copolymer from a nonselective solvent (toluene).125 Haliloglu et al. additionally calculated Langmuir adsorption isotherms and analysed adsorption kinetics.124 Comparison was also made with the simulations for diblocks, and it was found that the adsorbed amount is lower for the diblock than for the triblock at low concentration. However, the surface coverage is the same for both copolymers when weakly adsorbed to the surface. Surface density profiles were also compared. Finally, scaling relationships for triblock copolymer adsorption under weak adsorption conditions were derived.124 In a related paper,126 adsorption and bridging of BAB copolymers in an athermal solvent and confined between two parallel flat surfaces was studied, and the dynamic response of the system to sinusoidal and step shear was examined. The dimensions of surface micelles have been analysed using a brush model, modelling adsorbed blocks as a 2D melt. The surface micelles were approximated by part of a sphere. The association number, radius and height of the micelles were 1 /2 all shown to scale with the length of the nonadsorbed block as NA 1/2(i.e. a melt-like 127 scaling). Adsorption of diblocks comprising a highly charged block and a hydrophobic block onto planar substrates has been analysed using a brush scaling model, in the case of high salt concentration where interactions between charged blocks are screened.128 Layers adsorbed from dilute solution below the bulk cmc were found to be much denser than layers in equilibrium with micellar solutions. The adsorbed layer thickness was found to decrease with increasing salt concentration, whereas the surface density increased.
REFERENCES 1. Netz, R. R.; Andelman, D. Phys. Rep. 2003, 380, 1. 2. Cosgrove, T.; Fleer, G. J.; Cohen Stuart, M. A.; Scheutjens, J. M. H. M.; Vincent, B. Polymers at Interfaces. Kluwer: Amsterdam, 1993.
Adsorption
237
3. Jones, R. A. L.; Richards, R. W. Polymers at Surfaces and Interfaces. Cambridge University Press: Cambridge, 1999. 4. Hussain, H.; Kerth, A.; Blume, A.; Kressler, J. J. Phys. Chem. B 2004, 108, 9962. 5. Kent, M. S.; Lee, L. T.; Factor, B. J.; Rondelez, R; Smith, G. S. J. Chem. Phys. 1995,103, 2320. 6. Kent, M. S.; Majewski, J.; Smith, G.; Lee, L. T.; Satija, S. K. J. Chem. Phys. 1998,108, 5635. 7. Whitmore, M. D.; Noolandi, J. Macromolecules 1990, 23, 3321. 8. Carignano, M. A.; Szleifer, I. J. Chem. Phys. 1994, 100, 3210. 9. Zhulina, E. B.; Borisov, O. V.; Pryamitsyn, V.; Birshtein, T. M. Macromolecules 1991,24, 140. 10. Bijsterbosch, H. D.; de Haan, V. O.; de Graaf, A. W.; Mellema, M.; Leermakers, F. A. M.; Cohen Stuart, M. A.; van Well, A. A. Langmuir 1995, 11, 4467. 11. Goncalves da Silva, A. M.; Filipe, E. J. M.; d'Oliveira, J. M. R.; Martinho, J. M. G. Langmuir 1996, 12, 6547. 12. Rother, G.; Findenegg, G. H. Colloid Polym. Sci. 1998, 276, 496. 13. Gongalves da Silva, A. M.; Simoes Gamboa, A. L.; Martinho, J. M. G. Langmuir 1998, 14, 5327. 14. Faure, M. C.; Bassereau, P.; Carignano, M. A.; Szleifer, I.; Gallot, Y.; Andelman, D. Eur. Phys. J. B 1998, 3, 365. 15. Faure, M. C.; Bassereau, P.; Lee, L. T.; Menelle, A.; Lheveder, C. Macromolecules 1999, 32, 8538. 16. Rivillon, S.; Munoz, M. G.; Monroy, F; Ortega, F; Rubio, R. G. Macromolecules 2003, 36, 4068. 17. Gissing, S. K.; Richards, R. W.; Rochford, B. R. Coll. Surf. A 1994, 86, 171. 18. Richards, R. W.; Rochford, B. R.; Webster, J. R. P. Faraday Discuss. 1994, 98, 263. 19. Richards, R. W.; Rochford, B. R.; Webster, J. R. P. Polymer 1997, 38, 1169. 20. Dewhurst, P. R; Lovell, M. R.; Jones, J. L.; Richards, R. W.; Webster, J. R. P. Macromolecules 1998, 31, 7851. 21. Bijsterbosch, H. D.; Cohen Stuart, M. A.; Fleer, G. J.; van Caeter, P.; Goethals, E. J. Macromolecules 1998, 31, 7436. 22. Richards, R. W.; Rochford, B. R.; Taylor, M. R. Macromolecules 1996, 29, 1980. 23. Wesemann, A.; Ahrens, H.; Forster, S.; Helm, C. A. Langmuir 2004, 20, 11528. 24. Nakano, M.; Deguchi, M.; Endo, H.; Matsumoto, K.; Matsuoka, H.; Yamaoka, H. Macromolecules 1999, 32, 6088. 25. Matsumoto, K.; Mizuno, U.; Matsuoka, H.; Yamaoka, H. Macromolecules 2002,35, 555. 26. Matsuoka, H.; Matsutani, M.; Mouri, E.; Matsumoto, K. Macromolecules 2003,36, 5321. 27. Kaewsaiha, P.; Matsumoto, K.; Matsuoka, H. Langmuir 2004, 20, 6754. 28. Barentin, C.; Muller, P.; Joanny, J.-F Macromolecules 1998, 31, 2198. 29. Peleshanko, S.; Jeong, J.-B.; Gunawidjaja, R.; Tsukruk, V. V. Macromolecules 2004, 37, 6511. 30. An, S. W.; Su, T. J.; Thomas, R. K.; Baines, R L.; Billingham, N. C.; Armes, S. P.; Penfold, J. J. Phys. Chem. B 1998, 102, 387. 31. An, S. W.; Thomas, R. K.; Baines, F. L.; Billingham, N. C.; Armes, S. P.; Penfold, J. J. Phys. Chem. 1998, 102, 5120. 32. An, S. W.; Thomas, R. K.; Baines, F. L.; Billingham, N. C.; Armes, S. P.; Penfold, J. Macromolecules 1998, 31, 7877.
238
Block Copolymers in Solution: Fundamentals and Applications
33. Su, T. J.; Styrkas, D. A.; Thomas, R. K.; Baines, F. L.; Billingham, N. C.; Armes, S. P. Macromolecules 1996, 29, 6892. 34. Styrkas, D. A.; Butun, V.; Lu, J. R.; Keddie, J. L.; Armes, S. P. Langmuir 2000,16, 5980. 35. Romet-Lemonne, G.; Daillant, J.; Guenoun, P.; Yang, J.; Mays, J. W. Phys. Rev. Lett. 2004, 93, 148301. 36. Zhulina, E. B.; Borisov, O.; Birshtein, T. M. J. Phys. France II 1992, 2, 63. 37. Seidel, C. Macromolecules 2003, 36, 2536. 38. Naji, A.; Netz, R. R.; Seidel, C. Eur. Phys. J. E 2003, 12, 223. 39. Ahrens, H.; Forster, S.; Helm, C. A. Phys. Rev. Lett. 1998, 81, 4172. 40. Sharma, R. (Ed.) Surfactant Adsorption and Surface Solubilization. American Chemical Society: Washington, 1995; Vol. 615. 41. Edens, M. W.; Whitmarsh, R. H. Applications of Block Copolymer Surfactants. In Developments in Block Copolymer Science and Technology, Hamley, I. W., Ed. John Wiley & Sons, Ltd: Chichester, 2004; p 325. 42. Hamley, I. W.; Connell, S. D.; Collins, S. Macromolecules 2004, 37, 5337. 43. Malmsten, M.; Linse, P.; Cosgrove, T. Macromolecules 1992, 25, 2474. 44. Connell, S. D.; Collins, S.; Yang, Z.; Hamley, I. W. Langmuir 2003, 19, 10449. 45. Brandani, P.; Stroeve, P. Macromolecules 2003, 36, 9492. 46. Eskilsson, K.; Tiberg, F. Macromolecules 1998, 31, 5075. 47. Grant, L. M.; Tiberg, F.; Ducker, W. A. J. Phys. Chem. B 1998, 102, 4288. 48. Eskilsson, K.; Ninham, B. W.; Tiberg, F.; Yaminsky, V. V. Langmuir 1998, 14, 7287. 49. Muller, D.; Malmsten, M.; Tanodekaew, S.; Booth, C. J. Colloid Interface Sci. 2000,228, 317. 50. Muller, D.; Malmsten, M.; Tanodekaew, S.; Booth, C. J. Colloid Interface Sci. 2000,228, 326. 51. Green, R. J.; Tasker, S.; Davies, J.; Davies, M. C.; Roberts, C. J.; Tendler, S. J. B. Langmuir 1997, 13, 6510. 52. Brandani, P.; Stroeve, P. Macromolecules 2003, 36, 9502. 53. Li, J.-T.; Caldwell, K. D.; Rapoport, N. Langmuir 1994, 10, 4475. 54. Hariharan, R.; Biver, C.; Mays, J.; Russel, W. B. Macromolecules 1998, 31, 7506. 55. Gerstenberg, M. C.; Pedersen, J. S.; Smith, G. S. Phys. Rev. E 1998, 58, 8028. 56. Gerstenberg, M. C.; Pedersen, J. S. Langmuir 2001, 17, 7040. 57. Emoto, K.; Nagasaki, Y; Kataoka, K. Langmuir 1999, 15, 5212. 58. Otsuka, H.; Nagasaki, Y; Kataoka, K. Curr. Opin. Colloid Interface Sci. 2001, 6, 3. 59. Otsuka, H.; Nagasaki, Y; Kataoka, K. Biomacromolecules 2000, 1, 39. 60. Cannizzaro, S. M.; Padera, R. F.; Langer, R.; Rogers, R. A.; Black, F. E.; Davies, M. C.; Tendler, S. J. B.; Shakesheff, K. M. Biotechnol Bioeng. 1998, 58, 529. 61. Black, F. E.; Hartshorne, M.; Davies, M. C.; Roberts, C. J.; Tendler, S. J. B.; Williams, P. M.; Shakesheff, K. M.; Cannizzaro, S. M.; Kim, L; Langer, R. Langmuir 1999,15, 3157. 62. Kiriy, A.; Gorodyska, G.; Minko, S.; Stamm, M.; Tsitsilianis, C. Macromolecules 2003, 36, 8704. 63. Anastasiadis, S. H.; Retsos, H.; Pispas, S.; Hadjichristidis, N.; Neophytides, S. Macromolecules 2003, 36, 1994. 64. Levicky, R.; Koneripalli, N.; Tirrell, M.; Satija, S. Macromolecules 1998, 31, 2616. 65. Tao, J.; Guo, A.; Stewart, S.; Birss, V. L; Liu, G. Macromolecules 1998, 31, 172. 66. Lang, P.; Willner, L.; Pyckhout-Hintzen, W.; Krastev, R. Langmuir 2003, 19, 7597.
Adsorption
239
67. Balastre, M.; Li, R; Schorr, P.; Yang, J.; Mays, J. W.; Tirrell, M. V. Macromolecules 2002, 35, 9480. 68. Walter, H.; Harrats, C; Muller-Buschbaum, P.; Jerome, R.; Stamm, M. Langmuir 1999, 15, 1260. 69. Mahltig, B.; Gohy, J.-R; Jerome, R.; Bellmann, C.; Stamm, M. ColloidPolym. Sci. 2000, 278, 502. 70. An, S. W.; Thirtle, P. N.; Thomas, R. K.; Baines, R L.; Billingham, N. C.; Armes, S. P.; Penfold, J. Macromolecules 1999, 32, 2731. 71. Israels, R.; Leermakers, R A. M.; Fleer, G. J. Macromolecules 1994, 27, 3087. 72. Israels, R.; Leermakers, R A. M.; Fleer, G. J.; Zhulina, E. B. Macromolecules 1994, 27, 3249. 73. Spatz, J. P.; Roescher, A.; Moller, M. Adv. Mater. 1996, 8, 337. 74. Spatz, J. P.; Mossmer, S.; Hartmann, C.; Moller, M.; Herzog, T.; Krieger, M.; Boyen, H.-G.; Ziemann, P.; Kabius, B. Langmuir 2000, 16, 407. 75. Spatz, J. P.; Sheiko, S.; Moller, M. Macromolecules 1996, 29, 3220. 76. Forster, S.; Hermsdorf, N.; Leube, W.; Schnablegger, H.; Lindner, P.; Bottcher, C. J. Phys. Chem. B 1999, 103, 6657. 77. Price, C.; Woods, D. Eur. Polym. J. 1973, 9, 827. 78. Booth, C.; Naylor, T. D.; Price, C.; Rajab, N. S.; Stubbersfield, R. B. J. Chem. Soc., Faraday Trans. 11978, 74, 2352. 79. Zhang, L.; Eisenberg, A. Science 1995, 268, 1728. 80. Mortensen, K.; Talmon, Y. Macromolecules 1995, 28, 8829. 81. Breulmann, M.; Forster, S.; Antonietti, M. Macromol. Chem. Phys. 2000, 201, 204. 82. Gohy, J.-E; Antoun, S.; Jerome, R. Macromolecules 2001, 34, 7435. 83. Zhu, J.; Eisenberg, A.; Lennox, R. B. J. Am. Chem. Soc. 1991, 113, 5583. 84. Zhu, J.; Lennox, R. B.; Eisenberg, A. J. Phys. Chem. 1992, 96, 4727. 85. Li, S.; Hanley, S.; Khan, I.; Varshney, S. K.; Eisenberg, A.; Lennox, R. B. Langmuir 1993, 9, 2243. 86. Zhu, J.; Hanley, S.; Eisenberg, A.; Lennox, R. B. Makromol.Chem., Macromol.Symp 1992,53,211. 87. Zhu, J.; Lennox, R. B.; Eisenberg, A. Langmuir 1991, 7, 1579. 88. Zhu, J.; Eisenberg, A.; Lennox, R. B. Macromolecules 1992, 25, 6547. 89. Meszaros, M.; Eisenberg, A.; Lennox, R. B. Faraday Discuss. 1994, 98, 283. 90. Meiners, J. C.; Ritzi, A.; Rafailovich, M. H.; Sokolov, J.; Mlynek, J.; Krausch, G. Appl. Phys. A 1995, 61, 519. 91. Regenbrecht, M.; Akari, S.; Forster, S.; Mohwald, H. J. Phys. Chem. B 1999, 703, 6669. 92. Hahn, J.; Webber, S. E. Langmuir 2004, 20, 4211. 93. Webber, G. B.; Wanless, E. J.; Armes, S. P.; Baines, R L.; Biggs, S. Langmuir 2001,17, 5551. 94. Webber, G. B.; Wanless, E. J.; Butun, V.; Armes, S. P.; Biggs, S.Nano Letters 2002,2,1307. 95. Tassin, J. R; Siemens, R. L.; Tang, W. T.; Hadziioannou, G.; Swalen, J. D.; Smith, B. A. J. Phys. Chem. 1989, 93, 2106. 96. Munch, M. R.; Gast, A. P. Macromolecules 1990, 23, 2313. 97. Motschmann, H.; Stamm, M.; Toprakcioglu, C. Macromolecules 1991, 24, 3681. 98. Toomey, R.; Mays, J.; Tirrell, M. Macromolecules 2004, 37, 905. 99. Ma, Q.; Remsen, E. E.; Kowalewski, T.; Wooley, K. L. J. Am. Chem. Soc. 2001, 723, 4627.
240
Block Copolymers in Solution: Fundamentals and Applications
100. Ma, Q.; Remsen, E. E.; Kowalewski, T.; Schaefer, J.; Wooley, K. L. Nano Letters 2001,1, 651. 101. Hadziioannou, G.; Patel, S.; Granick, S.; Tirrell, M. J. Am. Chem. Soc. 1986,108, 2869. 102. Taunton, H. J.; Toprakcioglu, C.; Fetters, L. J.; Klein, J. Nature 1988, 332, 712. 103. Taunton, H. J.; Toprakcioglu, C.; Fetters, L. J.; Klein, J. Macromolecules 1990, 23, 571. 104. Dai, L.; Toprakcioglu, C. Macromolecules 1992, 25, 6000. 105. Cosgrove, T.; Zarbakhsh, A.; Luckham, P. F.; Hair, M. L.; Webster, J. R. P. W. Faraday Discuss. 1994, 98, 189. 106. Schillen, K.; Claesson, P. M.; Malmsten, M.; Linse, P.; Booth, C. J. Phys. Chem. B 1997, 101, 4238. 107. van Lent, B.; Scheutjens, J. H. M. Macromolecules 1989, 22, 1931. 108. Evers, O. A.; Scheutjens, J. M. H. M.; Fleer, G. J. Macromolecules 1990, 23, 5221. 109. Currie, E. P. K.; Leermakers, F. A. M.; Cohen Stuart, M. A. Macromolecules 1999, 32, 487. 110. Currie, E. P. K.; Norde, W.; Stuart, M. A. C. Adv. Colloid Interface Sci. 2003, 700, 205. 111. Linse, P.; Hatton, T. A. Langmuir 1997, 13, 4066. 112. Balazs, A. C.; Lewandowski, S. Macromolecules 1990, 23, 839. 113. Balazs, A. C.; Huang, K.; Lantman, C. W. Macromolecules 1990, 23, 4641. 114. Zhan, Y.; Mattice, W. L. Macromolecules 1994, 27, 677. 115. Zhan, Y.; Mattice, W. L. Macromolecules 1994, 27, 683. 116. Wang, W.; Lin, Z.-Z.; Ebner, C.; Nikas, Y. Phys. Rev. E 1993, 48, 4622. 117. Zhan, Y; Mattice, W. L.; Napper, D. H. J. Chem. Phys. 1993, 98, 7502. 118. Zhan, Y; Mattice, W. L.; Napper, D. H. J. Chem. Phys. 1993, 98, 7508. 119. Milchev, A.; Binder, K. Langmuir 1999, 75, 3232. 120. Johner, A.; Joanny, J. F. Macromolecules 1990, 23, 5299. 121. Ligoure, C. Macromolecules 1991, 24, 2968. 122. Marques, C.; Joanny, J. F.; Leibler, L. Macromolecules 1988, 27, 1051. 123. Wittmer, J.; Joanny, J. F. Macromolecules 1993, 26, 2691. 124. Haliloglu, T.; Stevenson, D. C.; Mattice, W. L. J. Chem. Phys. 1997, 706, 3365. 125. Dorgan, J. R.; Stamm, M.; Toprakcioglu, C.; Jerome, R.; Fetters, L. J. Macromolecules 1993, 26, 5321. 126. Nguyen-Misra, M.; Misra, S.; Mattice, W. L. Macromolecules 1996, 29, 1407. 127. Kramerenko, E. Y; Potemkin, I. L; Khokhlov, A. R.; Winkler, R. G.; Reineker, P. Macromolecules 1999, 32, 3495. 128. Dan, N.; Tirrell, M. Macromolecules 1993, 26, 4310. 129. Hamley, I. W. The Physics of Block Copolymers. Oxford University Press: Oxford, 1998.
6
Applications
As mentioned in Chapter 1, industrial applications of block copolymer surfactants have been discussed by industrial researchers from the United States.1-3 The texts edited by Nace4 and by Alexandridis and Lindman5 cover many aspects of the behaviour and properties of PEO-based amphiphilic block copolymers, with several chapters devoted to applications. Biomedical applications of block copolymers have been reviewed extensively, as discussed in Section 6.3. From an academic perspective, a number of reviews6-9 have covered the selfassembly of amphiphilic block copolymers, and the numerous potential applications of the resulting nanostructures. This is the main focus of the present chapter, which is concerned with the science underpinning several key applications. It is not intended to provide an overview of industrial products or processes.
6.1 SURFACTANCY/DETERGENCY The major application of amphiphilic block copolymers by value and volume is of course detergency. The PEO-b-PPO-b-PEO copolymers are widely used as nonionic polymeric surfactants in aqueous solution. As the properties of these systems have been extensively discussed in Chapters 2 and 3, further details are not provided here.
6.2 SOLUBILIZATION, EMUSIFICATION AND STABILIZATION 6.2.1 6.2.1.1
SOLUBILIZATION Experiments
There have been numerous studies on the use of amphiphilic block copolymers to solubilize molecules, in particular the use of water soluble block copolymers to solubilize organic compounds ('oils') which are immiscible in water. In fact, industrial applications of block copolymers, such as agrochemical dispersions or pharmaceutical formulations,3 relies on this solubilization capacity. Solubilization for applications in drug delivery is discussed in detail in Section 6.3. Block Copolymers in Solution: Fundamentals and Applications © 2005 John Wiley & Sons, Ltd.
I. W. Hamley
242
Block Copolymers in Solution: Fundamentals and Applications
Solubilization of benzene in solutions of PS-b-PMAA diblocks in aqueous solution was investigated via fluorescence spectroscopy (using tagged chains) and quasi-elastic light scattering.10 Solubilization of organic substances in micelles formed by PMMA-b-PAA copolymers neutralized to various degrees with Li+, Na+, or K+ counterions, has been studied by SANS and NMR. 11 Chloroform or chlorobenzene were solubilized in the PMMA core, to an extent that depends on the charge on the polyelectrolyte block. Solubilization of various organic solutes in PS-b-PEO diblocks has also been quantitatively studied.12 The Solubilization of perfluorinated compounds within block copolymer micelles in aqueous solution has been shown to be more efficient when the micelles are formed from a block copolymer with a partly fluorinated hydrophobic block.13 Furthermore, when a mixture of perfluorinated and hydrocarbon compounds was added, selective Solubilization of each type into the corresponding micelle was observed. Pluronic copolymers have been extensively used to solubilize aromatic hydrocarbons - see for example papers by Hurter and Hatton14 and others.15 Gadelle et al. investigated the Solubilization of aromatic solvents in micelles formed by PEO/PPO (Pluronic type) triblocks in water using head-space gas chromatography and determined Solubilization isotherms and solute partition coefficients.16 They found that the addition of apolar solubilizates promotes association of the copolymer, and that Solubilization is initially a replacement process in which water molecules in the core are displaced from the micellar core by the solubilizate. The partitioning of hydrophobic solutes into micelles of Pluronic block copolymers has been studied using fluorescence probe techniques.17 The partition coefficient, P, defined as the ratio of solute concentration in the micelle microphase and in the aqueous phase were determined for a series of copolymers. A simple inverse relationship between P and the cmc was found using pyrene as a probe (Figure 6.1). An increase in partition coefficient with hydrophobic block content was noted for a number of Pluronic copolymers.14,15 The Solubilization capacity also increases with polymer concentration.14,15 By varying the length of the alkyl chain in alkyl fluorescein probes, the incremental contribution of methylene groups to the free energy (of transfer from aqueous phase into micelle) could be determined - values ranged from —1.3 to —3.0 kJ mol -1 , increasing with the length of the hydrophobic PPO block.17 The Solubilization of water within reverse micelles formed by Pluronics (and related PEO-b-PBO-b-PEO triblocks) in p-xylene has been probed by gravimetry.18 The Solubilization capacity increased considerably when the PEO content in the copolymers decreased. Solubilization of alkyl cyanobiphenyl liquid crystals in the core of micelles formed by PEO-b-PPO diblock copolymers has been examined via light scattering and SAXS.19-22 The Solubilization capacity of hydrocarbon solutes in PEO-b-PPO diblocks and PEO-b-PPO-b-PEO triblocks was also compared.23 The Solubilization was measured by mixing aqueous micellar solutions with the hydrocarbons, separating the aqueous phase from the unsolubilized material and analysing the amount of
Applications
243
Figure 6.1 Linear relationship between logP and log (cmc) where P is the partition coefficient of pyrene into micelles of a series of Pluronic PEO-b-PPO-b-PEO triblocks (sample codes indicated).17 Reproduced by permission of American Chemical Society.
solubilized hydrocarbon by direct gas chromatography.24 The presence of solubilizate was shown to lower the cmc. The polymerization of monomer within block copolymer micelles has been studied. Kriz et al. showed25 using !H NMR that methyl methacrylate segregates to the core-shell interface of PS-b-PMA diblocks in aqueous solution, where it can undergo polymerization to produce an 'onion' structure with a core surrounded by two shells. Solubilization of large molecules such as fullerenes is possible in large block copolymer assemblies (hollow micelles with a monolayer shell,26 rather than a bilayer as in vesicles).26,27 Hollow spheres formed by two poly(phenylquinoline)-b-PS diblocks in binary organic solvent mixtures were able to solubilize C60 and C70 fullerenes. The aggregate dimensions increased from 1-5 (um initially to over 30 um upon solubilization. Block copolymers have been used as dispersants for pigments. For example, P4VP-b-PNaMA diblocks and PNaMA-b-P4VP-£-PNaMA triblocks have been used to disperse alumina-coated titanium dioxide.28 Gemini-type block copolymers, with two hydrophilic PEO blocks linked at a single junction point to a hydrophobic PS block have been prepared.29 They are expected to have superior wetting, foaming and dispersing properties than linear copolymers. Their use as surfactant stabilizers in emulsion polymerization was examined.29 6.2.1.2 Theory and Computer Simulation A scaling analysis based on the star polymer model of Daoud and Cotton for the solubilization of a low molecular weight solute in spherical micelles formed by an
244
Block Copolymers in Solution: Fundamentals and Applications
AB diblock in selective solvent was presented by Nagarajan and Ganesh.30 Scaling relations for the association number, the core radius, the coronal layer thickness L and the volume fraction of copolymer in the core, 0B were derived for a number of situations, including the case where the solvent S is a good or 0 solvent for the shell block A, and the solute J being a good or G solvent for the core block. The solubilization of low molecular weight solute J in micelles of AB diblocks formed in a selective solvent S was considered theoretically by Nagarajan and Ganesh.31 Predictions were made for the cmc, the size and composition distribution of the micelles containing solute J, the association number of the micelles, the maximum extent of solubilization, and the core radius and shell thickness of the micelles. Calculations indicated that micelles containing solubilizate tend to be essentially monodisperse in size and in the extent of solubilization. The solubilization behaviour of the micelles and their geometrical characteristics were found to be significantly influenced by the interactions between the solute and the solventincompatible B block of the copolymer as well as the J-S interfacial tension. The micellar structure was also found to be affected by the interactions between S and the coronal A block, though to a somewhat lesser degree when compared with the corresponding solute-free systems. The predictions of the theory were compared with experimental data for water-soluble PEO-b-PPO diblocks.24 The solubilization of hydrocarbons in PEO-b-PPO-b-PEO has also been examined theoretically, focusing on aggregates with different shape (spherical, cylindrical, lamellar) and transitions between different morphologies induced by solubilization.32 Theory and experiments on the solubilization of hydrophobic substances by block copolymer solvents have been reviewed by Nagarajan.33 Block copolymer micellization in a mixture with two immiscible solvents was considered by Cogan et al.,34 also employing modified Scheutjens-Fleer theory. They considered the case where each of the solvents is selective for one of the blocks and found that for highly incompatible solvents, one segregates in the centre of the micelle whilst the other selectively solvates the corona. The capacity of the micelle to disperse a second solvent was found to be controlled by the size of the core block. It was reported that the number of dispersed molecules and copolymer chains assembled in a micelle change most rapidly as the background solvent concentration approaches saturation levels. The theory was applied to interpret earlier experimental results35,36 on PEO-b-PS diblock copolymers in cyclopentane with trace amounts of water, which is a selective solvent for the PEO block that forms the micellar core. A simple thermodynamic model to describe solubilization in block copolymer micelles was developed by Munk and coworkers.37 The theory assumed that the association number of copolymer molecules within the micelle is not affected by the process of solubilization. The extent of solubilization represents an equilibrium between the favourable interaction of the core component with the solubilizate (with both enthalpic and entropic contributions) and the surface tension at the core boundary that acts against solubilization. The theoretical predictions were found to agree reasonably well with experimental results for neutral-b-polyelectrolyte PS-bPMAA diblocks in solution in 80:20 dioxane/water mixtures (selective for PMAA),
Applications
245
with a variety of organic solubilizates. In the experiments, static light scattering was used to determine the extent of solubilization and dynamic light scattering was used to obtain the micellar hydrodynamic radius.37 Dan and Tirrell presented a model for the phase diagram of a diblock copolymer in a mixture of incompatible and highly selective solvents using a mean field scaling theory approach.38 The properties of the diblock copolymer as an emulsifier were investigated, and the microdomain geometry and size were determined by an analysis of the interfacial layer. They found that solvent selectivity leads to copolymer association at the liquid-liquid interface and the formation of stable microdomains. Symmetric copolymers were found to form lamellae at all copolymer concentrations, whereas asymmetric copolymers were shown to associate in spherical micelles in equilibrium with the two bulk solvent phases. The interfacial curvature and surface density was found to increase, and the volume of the emulsion phase to decrease, with the copolymer degree of asymmetry. The cac, above which chains assemble at the liquid-liquid interface and microdomains form, was found to be significantly lower than the cmc. The emulsion phase volume was found to increase linearly with copolymer concentration until, when one of the solvents is exhausted, cylindrical domains appear. At high copolymer concentrations lamellae replace the cylindrical domains. As noted by Dan and Tirrell,38 this sequence of phase transitions is similar to that predicted for block copolymer/homopolymer blends.39,40 Earlier work by Cantor addressed diblock copolymers as emulsifiers in mixtures of immiscible solvents using mean field Flory-type theory for semidilute solutions.41 However, only a lamellar phase was considered. Emulsification in polymer blends containing block copolymers is discussed elsewhere.42 Pepin and Whitmore have investigated homopolymer solubilization limits in diblock copolymer micelles via Monte Carlo simulations.43 The distribution of low molecular weight homopolymer solubilized in the core of the micelles (formed by near symmetric diblocks) was found to be relatively uniform. This conclusion is in agreement with prior experimental and theoretical results on homopolymer solubilization in block copolymers.44 A threshold volume fraction of homopolymer beyond which it was no longer solubilized, and instead macrophase separated was identified and found to decrease with increasing chain length of homopolymer relative to the corresponding core block in the micelle. Monte Carlo simulations of solubilization in triblock copolymers revealed that solubilization enhances micellization - the fraction of chains associated into micelles and the association number are both increased.45 Structure formation in block copolymer + solvent + solubilizate systems has also been simulated, considering both solubilization in micelles and also microemulsion formation.46 6.2.2
EMULSIFICATION AND STABILIZATION
Amphiphilic block copolymers are widely used to stabilize emulsions and microemulsions. Due to their interfacial activity, they segregate to the oil-water interface, reducing the interfacial tension to facilitate mixing.
246
Block Copolymers in Solution: Fundamentals and Applications
The structure of the adsorbed layer of PEO-b-PBO diblocks at the interface in a perfluorodecalin-in-water microemulsion has been studied by SANS. Volume fraction profiles were obtained and found to resemble those for polymer brushes calculated using SCF theory.47 Block copolymers have been used as stabilizers in microemulsion polymerization to prepare hollow PS nanospheres.48 Methyl methacrylate was solubilized in the core of Pluronic micelles, and then polymerized. Styrene was then added with divinylbenzene as cross-linker. The PMMA core was removed by methylene chloride etching to leave PS nanospheres, as illustrated in Figure 6.2. The particle diameter was 15-30 nm, with a shell thickness 2-5 nm. Similarly, near monodisperse PS microspheres have been prepared by dispersion polymerization using PDMA-bpoly(alkyl methacyrylate) diblocks as stabilizers.49
Figure 6.2 Fabrication of hollow polystyrene nanospheres using block copolymer micelles as nanoreactors for sequential polymerization of PMMA, followed by PS, and subsequent removal of PMMA.48 Reproduced by permission of American Chemical Society.
Addition of PEP-b-PEO diblocks has been shown to dramatically improve the solubilization capacity of nonionic surfactants such as CioEO4 in water/ decane amphiphile microemulsions.50,51 The structure of PEP-b-PEO is of course chemically similar to that of the alcohol ethoxylate surfactants (although with branching) - for example, PEP10-b-PEO10 corresponds roughly to C715EO210.51 The interfacial tension is reduced dramatically upon addition of the block copolymer and the microemulsion can swell to a greater extent. The surfactant volume fraction required to form a balanced one-phase microemulsion is reduced by a factor of about four by using the surfactant/block copolymer mixture. The formation of lamellar mesophases, usually observed when the hydrophobic content of the amphiphiles is increased, is not observed using the block copolymer additive. The effect was ascribed to the ability of the polymer chains to extend further into the adjacent subphases. Contrast variation SANS was used to probe the curvature of the amphiphilic
Applications
247
membrane and the dependence of membrane volume fraction on block copolymer number density and chain dimensions was probed.51 PEO-based block copolymers have been used as stabilizers to prevent flocculation of PMMA microparticles sprayed into liquid CO2.52 Reverse Pluronics PPO-b-PEO-b-PPO (and also Tetronic stars) and PEO-b-PEO diblocks were used, the PEO acting as steric stabilizer, the hydrophobic block anchoring on the PMMA. The stabilization of foams by block copolymers has been investigated.53 For example, measurements of the disjoining pressure in films containing PEO-b-PEO diblock or PEO-b-PEO-b-PEO triblock copolymer have been combined with timeresolved ellipsometric determination of the adsorption rate and time-resolved surface tensiometry. Measurements of the interactions between the two air-liquid interfaces in the foam were performed via the thin film balance technique, in which the film thickness and disjoining pressure isotherm are monitored. It was found that film stability at high pressure was higher for the films containing diblock copolymers due to more efficient packing at the interface. Micellization of block copolymers in supercritical CO2 has been studied by several groups.54-58 The work has been motivated by the fact that block copolymers containing CO2-philic blocks have been shown to be very effective surfactants in emulsion polymerization.59,60 The micellar corona is usually formed by CO2-philic perfluorinated55-58 or partially fluorinated blocks. The use of block copolymers as stabilizers in supercritical fluids has been reviewed.61 Micellization of fluorinecontaining block copolymers to form particles with highly hydrophobic cores has been investigated.13,62
6.3 DRUG DELIVERY Applications of block copolymers in the pharmaceutical industry have been extensively reviewed, in particular in several special issues of Advanced Drug Delivery Reviews, 63-71 but also elsewhere.72-77 It is not intended to reiterate all of this information here, but salient features will be highlighted. The main types of polymer investigated have been PEO-based block copolymers, either PEO-b-PPO-b-PEO Pluronics, PEO-b-polyesters or PEO-£-poly(L-amino acid)s.76 Pluronics are readily available commercially, PEO-b-polyesters benefit from the susceptibility of the polyester block to hydrolytic degradation and PEO-b-poly(L-amino acid)s contain functional groups that may be used for targeted delivery. All of these copolymers can be used as drug delivery systems, sustained release agents, coatings, adjuvants (materials that increases the immune response to an antigen when combined with it), emulsifiers and dispersants and anti-hemolytic agents among other applications. Polymeric micelles containing PEO offer advantages compared with other systems (microspheres, liposomes, etc.) of greater stability, decreased uptake by the reticulo-endothelial system (RES, a system of cells responsible for removal of foreign substances and pathogens from the bloodstream) and thus longer circulation
248
Block Copolymers in Solution: Fundamentals and Applications
times and enhanced passive accumulation in tumours via the enhanced permeability and retention (EPR) mechanism.78 Several reviews cover specific applications such as the use of PEG-based micelles as carriers for contrast agents in medical imaging (MRI or computed tomography)68 or for delivery of photosensitizers for photodynamic therapy.71 Another example is the use of polyether-b-polyester diblocks in formulations of paclitaxel, due to problems of hypersensitivity with the current formulation of taxol, the anti-cancer drug.69 The development of Pluronic copolymers for applications in drug delivery has largely been pioneered by Kabanov and coworkers.79 These materials are attractive due to the ready commercial availability of a wide range of copolymers (with different composition and molar mass), low toxicity, and well characterized physicochemical properties. They investigated the use of Pluronic micelles as delivery agents for drug targeting across the blood-brain barrier.80,81 They have also shown that unimers inhibit immune response. They have recently studied block ionomer complexes as carriers for DNA, as reviewed in detail elsewhere63,82 (and mentioned in Section 4.4). Kataoka et al. have studied the biomedical and pharmaceutical applications of PEG-b-PLA diblocks with functional group terminated PEG corona chains. The relevant papers are cited in a comprehensive review.70 A schematic summarizing various applications exploiting micelles in bulk and also for surface modification is shown in Figure 6.3. PLA is biodegradable and nontoxic and is widely used in implant materials. The PEG corona block bears an acetal or aldehyde end group to
Figure 6.3 Schematic showing applications of PEG-b-PLA block copolymers for biomedical applications.70 Reproduced by permission of Elsevier.
Applications
249
enable functionalization. Nanospheres have been prepared by core cross-linking of heterobifunctional diblocks.83 The polylactide block was terminated by a methacryloyl moiety for cross-linking of the core. The micelles with polymerized core could be adsorbed intact onto a coated surface (in contrast to micelles without polymerized cores).84 Applications of PEO-b-poly(L-amino acid) micelles for drug delivery have been reviewed by Kataoka et al.65 Kwon and coworkers also discuss such systems.67 The core-forming poly(L-amino acid) is nonpolar and can solubilize hydrophobic drugs. Conjugates of the drug molecules with the block copolymer have also been prepared. This type of copolymer has also been used for DNA delivery applications. A PEG-b-PAsp diblock has been used as steric stabilizer in the coprecipitation of DNA with calcium phosphate, to control the growth of Ca3(PO4)2 crystals containing DNA.85,86 Controlling the size distribution is essential since transfection efficiency is strongly influenced. Calcium phosphate/DNA coprecipitates have attracted interest in transfection of plasmid DNA for applications in gene therapy. The copolymer solution containing phosphate ions was mixed with DNA in the presence of calcium ions and crystals with a relatively narrow size distribution were formed. Kataoka and coworkers have studied the properties of PEO-b-PAsp block copolymers, and potential applications which arise from the fact that both blocks are biocompatible and the steric stabilization conferred by PEG.87-89 Conjugation with adriamycin (ADR) to give PEO-b-poly(Asp(ADR)) diblocks enables the preparation of an anticancer drug delivery system. Conjugation inhibits side reactions which can damage ADR - unconjugated material can be removed from the micelles formed in aqueous solution by dialysis.89 The antitumour action in vivo, long blood circulation times (conferred by the PEG micellar coating) and cytotoxicity have been extensively characterized.87-89 The same group earlier investigated the effect of modification of secondary structure through side group substitution in PEG-PLys diblocks.90 The substitution (to a varying extent) of hydrocinnamoyl groups in the N£ position of the amino acid afforded control of the hydrophilic/hydrophobic balance, and of TT-TT interactions. Increasing the degree of substitution, or pH, led to a transition from a random coil to a (3 sheet structure (which is stabilized by side-chain interaction between aromatic groups). Micellization of polymers with a ß-strand structure was also observed. Applications of block copolymer gelation in the development of thermally controlled delivery systems for slow drug release have also been explored. A micellar solution containing solubilized drug molecules is injected into the body at a suitable temperature (Figure 6.4). Gelation then occurs at body temperature (37 °C), providing a matrix for slow release. The use of PEO-based copolymers in this context has been widely reported, and several reviews are available.2,3,79 In particular, PEO-bPPO-b-PEO triblocks are commercially available materials suitable for such applications. The concept has also been demonstrated with block copolymer micelles containing biodegradable polylactide blocks, which further facilitates elimination of the polymeric drug delivery system following release. Jeong et al. have shown that by adjusting the length of the biodegradable PLLA block in PLLA-b-PEO
250
Block Copolymers in Solution: Fundamentals and Applications
Figure 6.4 Principle for a slow release drug delivery system using a block copolymer gel.91 A sol of block copolymer micelles containing solubilized drug is injected at room temperature. It then forms a gel in vivo (gel temperature Tt), which is slowly degraded to release the drug. Reproduced by permission of Nature.
diblocks, and the copolymer concentration, the transition from a high temperature solution to a gel can be manipulated to achieve this.91,92 They further demonstrated the slow release in vitro of a model drug. By use of a different copolymer, PEG-bPLGA-b-PEG [PLGA = poly(D,L-lactic acid-co-glycolic acid], a system exhibiting a sol at room temperature and a gel at body temperature was demonstrated.93,94 Control of the sol-gel phase transition boundary through the addition of salts was also explored.93 The integrity of the gels was examined in vivo using a rat model.94 The same group has also examined the use of graft copolymers, for example PEG grafted with PLGA.95 The use of (PEO-b-polylactide)n (n = 4 or 8) star-block copolymers has also been suggested,96 because it has been shown that degradation is slower for this architecture than for linear copolymer, due presumably to steric effects. The hydrophobic polylactide was either poly(L-lactide) or poly(L-lactide-coglycolide). Micellization of PEO-b-poly(ethyl acrylate) diblocks has also been studied in the context of drug delivery systems.97 Poly(ethyl acrylate) is moderately hydrophobic and has a low glass transition temperature, facilitating dynamic micellization. Micellization of PEG-b-poly(lactide) block copolymers has also been investigated by other groups, who have highlighted the enhanced blood circulation times and reduced uptake by the liver of these PEG-coated micelles, compared with poly(lactide) nanoparticles.98'99 The biodegradability of the poly(lactide) block is also important. Booth and coworkers have also studied block copolymer systems in which the micellar solution can be injected at room temperature, gelation occurring at high temperature.100,101 In particular, they have demonstrated sustained release of salicylic acid from a gel formed by a PEO-b-PBO-b-PEO triblock in aqueous
Applications
251
solution102 and have investigated the solubilization of another model hydrophobic drug, griseofulvin, in micelles formed by diblocks and triblocks of EO with PO or SO.103 They have also investigated the gelation of PEO-£-poly(D,L-lactic acid) diblocks - only a high temperature gel - sol transition was observed, i.e. cold gelation necessary for applications was not observed.104 Drug release properties were not examined. Armes and coworkers have recently developed biocompatible, stimulusresponsive gelators based on triblock copolymers containing 2-methacryloyloxyethyl phosphorylcholine (MPC).105,106 The phosphorylcholine motif is an important component of cell membranes and surface coatings containing it are remarkably resistant to protein adsorption and bacterial/cellular adhesion. The triblocks with symmetric DEA or DPA end blocks are molecularly dissolved at low pH, but form flower micelles or gels at pH 8, depending on polymer concentration (Figure 6.5).106 The microstructure and rheology of the gels and solutions were later studied in detail.107 It was noted that the dynamic elastic moduli were considerably
Figure 6.5 Schematic showing flower micelle formation and gelation in PDPA-b-PMPC-bPDPA triblocks.106 Reproduced by permission of American Chemical Society.
lower than those of Pluronic triblock gels in aqueous solution at similar concentrations, an effect ascribed to reduced bridging of micelles in the MPC-based block copolymers. Release studies using a model hydrophobic drug indicated that PDPAb-PMPC-b-PDPA gels retain the drug longer than PDEA-b-PMPC-b-PDEA gels, probably due to the greater hydrophobicity of PDPA than PDEA.106 Triggered release is possible because gel dissolution occurs at low pH. Polyion complex micelles have also been exploited to encapsulate charged molecules, such as DNA or certain enzymes due to electrostatic interactions with an oppositely charged core-forming block.63,65,66 The PEG block forms a steric stabilization layer, also termed 'palisade'. PEG-b-cationic block copolymers employed to form PIC micelles with DNA include PEG-b-PLys, PEG-b-PEI and PEG-b-PDMA.66 Further information is provided in Section 4.4. Shuai et al. examined DNA condensation using PEG-b-PLys diblocks.108 The transfection using complexes containing monomethoxyl PEG-b-PCL has also been examined, also for copolymers in which these blocks are grafted onto hyperbranched PEL109
252
Block Copolymers in Solution: Fundamentals and Applications
The development of neutral-b-cationic block copolymers for gene therapy applications has been explored by Kabanov and coworkers. Several systems with PEO as the neutral block and different cationic polymer blocks have been investigated.63,79 Compared with cationic homopolymers used for gene delivery, incorporation of PEO in a block copolymer confers improved solubility (even if all electrostatic charges are neutralized63) and steric stabilization. A proposed mechanism for transfection of plasmid DNA via endocytosis is illustrated in Figure 6.6.
Figure 6.6 Schematic showing transfection of plasmid DNA into a cell via endocytosis using a block polyelectrolyte.258 Reproduced by permission of Royal Society of Chemistry.
The incorporation of sugar units onto PEG chain ends that form the coronas of micelles has been reported for PEG-b-PDLL diblocks.110 The sugar groups (e.g. glucose or galactose) should enable cellular-specific targeting via glycol receptors on the cell membrane. The potential use of such sugar-coated micelles in drug delivery is evident. The Kataoka group has shown that specific recognition of lectin proteins (i.e. those that specifically bind carbohydrates) is possible with this type of micelle.111 Surface plasmon resonance has been used to investigate the interaction between lectins immobilized on a gold surface and lactose-coated PEG-b-PDLL diblock micelles, adsorbing from solution.112,113 Micelles from diblock copolymers containing glucose and galactose groups end groups have been characterized, and
Applications
253
their interaction with lectin immobilized on a GPC column was examined.114 Diblocks were prepared using protected glucose and galactose containing initiators, used for ATRP of PEGMA-b-PBzMA. The preparation of sugar-coated shell crosslinked micelles is discussed in Section 2.12.4. Vesicles may also be used in drug delivery applications, as discussed in Section 6.7 and reviewed elsewhere.74 Other applications of block copolymers in medicine are discussed elsewhere.3,79 Examples include the use of Pluronic copolymers in artificial blood formulations, anti-tumour and anti-inflammatory agents and surgical dressings.
6.4 BIODEGRADABLE BLOCK COPOLYMER MICELLES Applications of biodegradable block copolymers have recently been reviewed.115 The main class of biodegradable copolymer contains polyesters or polyester sequences. Applications include surface modification, hydrogel formation and micelles for drug delivery. The latter subject is discussed in detail above. Surface modification generally involves coating with PEG chains by adsorption of a block copolymer. PEG can prevent biofouling. As mentioned in the preceding section, Kataoka and coworkers have investigated end-functionalization of the PEG as a route to conjugate proteins to the surface, for example.70,116 Biodegradable hydrogels based on polylactide-containing copolymers have been extensively studied,91,117,120 due to potential applications in drug delivery. The copolymers comprise a water soluble PEG midblock with hydrophobic polylactide end blocks. The gelation transition depends on copolymer composition and concentration. Further details are provided in the preceding section. Gelation of triblock copolymers containing a PLGA midblock and short PEO endblocks has been investigated by Char and coworkers.121,122 A particular focus was rheological and structural characterization of the turbid gels formed at high temperature, as well as transparent gels formed at low temperature. The turbidity results from aggregation of micelles, which leads to complex dynamics, as manifested for example by a plateau in the decay of the intensity autocorrelation function obtained from DLS122 (see also Section 3.10.2). This was ascribed to a phase separated structure containing heterogeneities resulting from the percolation of micelle clusters. These copolymers were contrasted to Pluronic copolymers in that the PLGA midblock is more hydrophobic than PPO, leading to enhanced attractive interactions between micellar cores. Tew and coworkers have examined the gelation of PLLA-b-PEO-b-PLLA triblocks via rheological experiments.123 Biodegradable micelles formed from block copolymers containing PCL have been investigated by Nie et al.124 The enzymatic degradation kinetics were investigated and it was noted that simply monitoring pH provided a good measure of degradation rate. The enzymatic biodegradation of the PCL core in PCL-b-PEO micelles has been probed by light scattering.125
254
Block Copolymers in Solution: Fundamentals and Applications
6.5 THERMORESPONSIVE MICELLAR SYSTEMS Thermoresponsive block copolymers have been prepared using PNIPAM as the thermosensitive hydrophobic block. PNIPAM exhibits a lower critical solution temperature (LCST) at Tc= 32 °C. Cross-linked gels pass through a volume phase transition at this temperature, shrinking dramatically as water is expelled.126'127 Although there is an extensive literature on random copolymers with NIPAM (for recent examples see references127-129 there have been fewer studies on block copolymers. The micellization of PNIPAM-b-PEO, poly(NIPAM-co-HPMA)-b-PEO and poly(NIPAM-co-HPMA-lactate)-fr-PEO diblocks has been investigated.130 PHPMA is interesting since it was originally developed as a biocompatible polymer for drug delivery applications. Increasing the PHPMA content enables the cloud point to be reduced, indicating an increase in the hydrophobicity of the corresponding block. On the other hand, the hydrolysed block containing PHPMA lactate becomes more hydrophilic. By varying the composition of the copolymer it was possible to engineer a system forming micelles at body temperature that break up during hydrolysis. Furthermore, the hydrolysis kinetics can be controlled through the number and length of oligolactate grafts in the copolymer. This is attractive for drug delivery applications. In a similar vein, the thermoresponsive behaviour of PHPMA-b-PNIPAM diblocks has been studied.131 Large increases in hydrodynamic radius and aggregate molar mass were observed above the LCST of PNIPAM. This was ascribed to enhanced hydrogen bonding in the collapsed state, leading to the formation of aggregates. Several groups have investigated micellization in PNIPAM-b-PEO diblocks.130'132'133 Topp et al. also prepared PNIPAM-b-PEO-bPNIPAM triblocks.132 These studies report that the PEO block increases the LCST and stabilizes micelles. Okano and coworkers have investigated the micellization of diblocks comprising PNIPAM with hydrophobic blocks - either PS 134 or biodegradable PDLL.135 Micelles were formed below the LCST. Above the LCST clouding was observed due to micellar aggregation as the PNIPAM chains collapsed. The aggregation process was found to be reversible. It was noted that the low cmc values observed would favour long circulation times of the micelles in blood and that the hydrated outer shell (below the LCST) would hinder uptake by the RES.134 Since the micelles can also solubilize hydrophobic molecules including an anti-cancer drug, they have potential as drug carriers. Since tumour sites may have a locally higher temperature, the thermoresponsivity of the PNIPAM shell can be exploited to target delivery above the LCST where binding interactions with cell membrane surfaces are favoured.134 In a further development of PNIPAM-based copolymers, double responsive PNIPAM-b-PAA diblocks have been prepared.136 Since PAA is a pH-sensitive polyelectrolyte, the system exhibits thermal and pHresponsiveness. Micelles with a PNIPAM core and PAA corona are formed at high pH and T > Tc where the PAA is deprotonated and PNIPAM is collapsed. For pH < 4 and T < Tc, micelles with a protonated PAA core and PNIPAM corona are formed. At pH < 4, T > Tc aggregation is observed.
Applications
255
Cationic polymerization has been used to prepare another type of thermoresponsive diblock copolymer based on poly(vinyl ethers.137 The hydrophilic block is PHOVE, and the thermoresponsive hydrophobic block was one of three types of poly(vinyl ether) with a pendant terminal alkyl unit. A transition from a micellar solution to a gel (ordered cubic micellar structure) was observed on heating. A practical disadvantage of the system may be the closeness of the critical micelle temperature and critical gel temperature, these being separated by less than 0.5 °C. The sol-gel transition in thermoresponsive lactide-based linear91'94 and star96'138 block copolymers has also been mapped out, as discussed in the previous section.
6.6 METAL-CONTAINING COPOLYMER MICELLES AND NANOREACTORS Block copolymer domains can be used as 'nanoreactors' for the synthesis of inorganic nanoparticles. Reviews of the subject are available.8'139'140 Two basic approaches have been developed. The first involves the binding of inorganic species to the monomer prior to polymerization or to one of the blocks of a copolymer prior to micellization (which may be induced by the ion binding ). The most important approach, however, involves the loading of preformed micelles, whether in solution or in bulk. Cohen et al. have systematically examined the preparation of diblock copolymer nanoreactors, which are subsequently loaded with salt that is then reduced to form metal nanodomains. They have demonstrated the patterning of complexes of silver,141-143 gold141 and zinc.144 The metal salt is added after formation of the diblock, and coordinates to a norbornene-based block, whilst the other block is built from PMTD. The copolymer/metal salt solution is then dried and the thin film morphology observed. For the silver and gold systems, spherical, cylindrical and lamellar metallic domains were observed, after heat treatment.141-143-145 Using zinc-containing complexes, zinc fluoride and zinc sulfide clusters were obtained, within spherical domains 144 and also lamellae.146 This concept was later generalized by creating carboxyl-functionalized nanoreactors within which a variety of metal salts can be sequestered (Figure 6.7).147'148 The carboxyl moiety was again a norbornene derivative. Metal oxide nanoparticles such as CoFe2O4 (of interest for high-density magnetic recording applications) have been prepared in a similar way.149 NaOH was used to form the oxide from the chloride salt precursor. Metal sulfides have also been produced by reaction of the precursor complexes with H2S.147'150'151 Particularly relevant to quantum dot applications of semiconductor nanoparticles, Moffitt et al. described the preparation of CdS nanoparticles within the cores of PS-b-poly(cadmium acrylate) diblocks from organic solvents.152'153 The particle size could be controlled in the range 2.9-5 nm by changing the length of the ionic poly(cadmium acrylate) block. The preparation of PbS nanoparticles was also described.152 An alternative approach to the preparation of CdS
256
Block Copolymers in Solution: Fundamentals and Applications
Figure 6.7 Schematic of block copolymer nanoreactors, whereby a carboxy functionalized domain is loaded with metal ions or complexes.148 Reproduced by permission of Elsevier.
nanoparticles exploits complexation with the P2VP core of PS-b-P2VP micelles.154 Here, the nanoparticles are aggregated into 'raspberry' morphology clusters. A disadvantage of the methods introduced by Cohen et al. is the need to synthesize polynorbornene-based block copolymers, which furthermore are not available commercially. Other complexation chemistries may be preferred. A useful compilation of functional blocks for the complexation of inorganic materials is available.8 The same authors also attempt to rationalize the ability of polymers to solubilize inorganic molecules by reference to hard-soft acid-base interactions. Spatz et al. have produced titania nanoparticles within block copolymer micelle nanoreactors.155 Titania nanoparticles are interesting for applications such as catalysis, water purification and UV blocking. They prepared micelles from a PS-b-PEO diblock in a nonpolar solvent with a PS corona and a PEO core. HC1 was mixed with the micelle solution to create reservoirs within the micelle cores. Titanium alkoxides were then added, the reaction with the acid and subsequent
Applications
257
heating leading to the formation of TiO2 particles. Individual particles and chains were both obtained. The same group have also used aqueous solutions of PS-bP2VP diblocks to form gold nanoparticles by reduction of HAuCl4 solubilized in the micellar core (as discussed also in Section 5.3.3). Below the cmc, nonspherical agglomerated particles were observed, showing that it is necessary for the reduction to occur in micellar 'nanoreactors' to get well defined spherical nanoparticles. Figure 6.8 illustrates the low polydispersity of nanoparticles that can be obtained in
Figure 6.8 TEM image showing well-defined gold nanoparticles in a dried film of goldcontaining PS-b-P2VP micelles.259 Reproduced by permission of American Chemical Society.
this way. Sakai and Alexandridis have shown that gold nanoparticles can be fabricated from HAuCl4.3H2O using Pluronic block copolymers, without additional reducing agents.156 The AuCl-4 ions are reduced by the PEO (the mechanism for this is discussed), and the block copolymer stabilizes the colloidal dipersion of the resulting nanoparticles. The effect of copolymer composition and molar mass on the nanoparticle formation was also investigated. Micellar nanoreactors for metallic nanoparticles based around several different block copolymers have been investigated by Bronstein and coworkers. In many cases, the ability of poly(vinylpyridine), P2VP or P4VP, to form complexes with metal salts due to the presence of an N atom is exploited.157-159 For example, Pd
258
Block Copolymers in Solution: Fundamentals and Applications
clusters have been produced by reduction of Pd(CH3COO)2 that coordinates to the P4VP micellar core formed by PS-b-P4VP diblocks in toluene.157,160 Other metal nanoclusters including cobalt,161 gold,157,159,160 rhodium157 and platinum157 have been prepared in a similar way. The catalytic hydrogenation properties of the nanoclusters have been compared.157,160 Bimetallic colloids have also been prepared.157 The methods may be extended to thin films where surface interactions may be exploited to prepare highly aligned nanostructures. Antonietti and coworkers have extended the range of amphiphilic diblocks that act as nanoreactors by demonstrating the production of metal nanoparticles within micelles of diblocks derived from PS-b-PB.162 The PB was first epoxidized, then the oxirane ring was opened to introduce a variety of functional side groups (nucleophiles, acids, acid chlorides), designed to complex with metal salts. Loaded micelles were formed by dissolution in toluene and mixing with the metal salts. Colloidal gold, silver, palladium and rhodium were all prepared by reduction of the metal salts in the block copolymer micelles. The formation of metal complexes with the double bonds in PB in block copolymers with PS has been examined via 1H NMR and IR spectroscopy.163 It has been shown that a core consisting of PB complexed to metal can be prepared by dissolution in a solvent selective for PS.164 In a variant of the method for nanoparticle templating introduced by Bronstein et al., the patterning of palladium particles at the domain interface in a PI-b-P2VP diblock has been demonstrated, and ascribed to 'polarization' of the nanoparticle at the interface.165 Using PS-b-PEO12 or P4VP-b-PEO166 diblocks, it is possible to prepare nanoparticles in aqueous media, although the micellization is influenced by pH and by the presence of the metal salts. The same group have also developed mixed micelles containing a nonionic PS-b-PEO diblock and a cationic surfactant (cetylpyridinium chloride).167 The aim was to prepare nanoreactors based on charged micellar cores (similar to the concept introduced by Clay and Cohen; (Figure 6.7). The counterions in the micellar corona are exchanged by anions such as PtCl2-6 and PdCl26-. The concept can be extended to mixed micelles containing anionic surfactants with cationic counterion exchange, as exemplified by the system PS-b-PEO/SDS.168 Exchange of the Na+ cation with a rhodium species enabled the loading of the corona with the latter. Reduction using NaBH4 led to clusters of rhodium nanoparticles within micellar aggregates. Nanoreactors containing transition metal binding sites in the core have been prepared from amphiphilic poly(2-oxazoline) block copolymers in aqueous solution.169'170 The sequestered triphenylphosphine moieties were expected to lead to catalytic activity. These hydrophobic groups were introduced into the block copolymer by a polymer analogous Pd-catalyzed phosphorus–carbon coupling reaction, leading to a statistical copolymer with pendant functional groups. Double hydrophilic block copolymers have also been used to template the production of metal hydrous oxide nanoparticles.171'172 The block copolymers contained a PAA anionic metal-binding block and a neutral block of PAM or PHEMA that conferred colloidal stability. Elongated and spherical lanthanum and aluminium hydrous oxide particles were observed.
Applications
259
Rod-like block copolymer micelles can be used as scaffolds to prepare metal nanowires. This has been demonstrated using diblock copolymers consisting of a carbosilane dendrimer and a polyisocyanopeptide, silver ions preferentially forming complexes with the peptide units that form the micellar core.134 The concept can presumably be extended to simpler block copolymers containing a polyelectrolyte block. Polymer/palladium nanowire hybrids have been prepared using nanotubes fabricated from core-shell tetrablock copolymer cylinders, as illustrated in Figure 6.9.174 The procedure is illustrated in Figure 6.10.174 Tetrablock
Figure 6.9 TEM image showing Pd-loaded nanofibres prepared from nanotubes prepared by cross-linking of the P(CEMA-ran-HEMA) domain and removal of the PI cores formed in core-shell cylinders of PI-b-PfBA-b-P(CEMA-ran-HEMA)-6-PSMA tetrablock copolymers.174 Reproduced by permission of American Chemical Society.
Figure 6.10 Procedure for preparation of Pd-loaded nanotubes from PI-b-PtBA-b-P(CEMAra«-HEMA)-b-PSMA tetrablock copolymers (A). (B) Micelle formation. (C) Cross-linking of the P(CEMA-rarc-HEMA) domain. (D) Removal of the PI cores. (E) Loading of nanotubes with PdCl2. (F) Reduction to form Pd-plated nanotubes.174 Reproduced by permission of American Chemical Society.
260
Block Copolymers in Solution: Fundamentals and Applications
PI-b-P?BA-b-poly(CEMA-ran-HEMA)-b-PGMA copolymers self-assembled into cylindrical aggregates in aqueous solution. The core comprised core-shell cylinders of PI then PtEA then P(CEMA-ran-HEMA). The PGMA formed the water soluble corona. The P(CEMA-ran-HEMA) layer was cross-linked followed by ozone etching to remove the PI core to form nanotubes, which may then be coated with Pd using an electroless method. An alternative approach has been used by this group in which bulk nanostructures formed by an ABC triblock are fixed by shell cross-linking.175 The nanocylinders formed are then dispersed in solvent. PBMA-bPCEMA-b-PriBA triblocks formed cylinders with PtBA cores and PCEMA shells in a PBMA matrix. The PCEMA was photochemically cross-linked, prior to dissolving the polymer film in the THF to produce nanofibres. The PtBA units were finally hydrolysed to PAA to produce PAA-lined nanotubes. It was shown that the nanotubes could be loaded with Fe2O3 due to ion exchange of Fe2+ with the protons of the acrylic groups. The concept of fixing bulk structures formed by ABC triblocks and then dissolving in a selective solvent was used to prepare Janus micelles, as discussed in Section 2.12.5. An alternative approach has been developed by Stamm and coworkers using the concept of dissolution of a melt structure (see Section 2.12.5) to prepare isolated PS nanoparticles.176 In this case the melt structure comprising shear-aligned cylinders of PS in a matrix of P4VP (to which the amphiphile pentadecylphenol was conjugated) were dissolved in a P4VP-selective solvent. This led to the formation of PS cylinders coated with P4VP. Pd nanoclusters could be deposited directly onto the nanorod templates by electroless deposition due to a direct redox reduction with the P4VP. In a similar fashion, CdSe nanowires could be fabricated. The harnessing of supramolecular chemistry to prepare diblocks containing a metal complex has recently been used to create metal-containing micelles.177,178 Diblocks were formed by complexation between PS and PEO homopolymers, each of which was end-capped with a terpyridine ligand that forms a tri-coordination complex with ruthenium. The result is a diblock, denoted PS-[Ru]-PEO that contains a Ru complex linking group (Figure 6.11). In aqueous solution, micelles are formed with a PS core surrounded by a Ru complex shell, to which PEO corona chains are tethered. The micellization was compared with that of a covalently bonded PS-b-PEO diblock.178 The presence of the charged metal-ligand complex at the interface between the blocks was found to lead to a large increase in hydrodynamic radius. Addition of salt screens the charge on the Ru complex and leads to a reduction in radius towards that of the covalently bound copolymer. This was ascribed to reduced stretching of the PEO block which appeared to have certain characteristics of polyelectrolyte chains. The concept has been extended to ABC triblocks - spherical core-shell-corona micelles were observed for PS-b-P2VP[Ru]-PEO in aqueous solution in which the P2VP shell is pH sensitive.179 The micellization of block copolymers containing the organometallic block PFS has been investigated by Manners and coworkers. 180-184 Block copolymers containing PFS are of interest because PFS has intriguing charge transport properties, and has been investigated as a precursor material to magnetic ceramics. Spherical
Applications
261
Figure 6.11 Schematic of a PS-[Ru]-PEO diblock that contains a Ru complex linking group.178 Reproduced by permission of American Chemical Society.
and rod-like micelles were observed for a PFS-b-PDMS-b-PFS triblock 180,183 and several PFS-b-PDMS diblocks 181,184 in n-alkanes, which are selective solvents for PDMS. By varying the temperature of micelle formation with respect to the melting point of the crystalline PFS block, it was found that crystallization has a major influence on micelle shape - spherical micelles were observed above Tm, and rodlike ones below. Flower-like aggregates of rod-like micelles were also noted,180,183 again being ascribed to the effect of PFS crystallization. Cylinders of these diblocks deposited onto GaAs substrates by aerosol spraying have been used as templates for patterning of lines into ceramic materials.185 Hydrogen plasma etching removed the organic material leaving a ridge of ceramic containing Si, Fe, O and C. Spherical micelles were also observed for a PFS-b-PEO diblock in water.186 Spherical micelles were observed for several PFS-b-PI diblocks in THF, a selective solvent for PI.182 In addition, the PI corona of the micelles was cross-linked using UV radiation to yield polymer shells containing a dense metal-containing core.
6.7
VESICLES
The preparation and properties of block copolymer vesicles are discussed in Section 2.17. Here we focus on proposed applications. The use of block copolymer assemblies as synthetic cell elements has been discussed.187 Vesicle bilayers are analogous to cell membranes, and giant wormlike micelles may be synthetic models for filaments in the cell. Due to their enhanced toughness and rigidity, block copolymer 'polymersomes' are much more robust than conventional liposomes. The ability to cross-link the chains of course enables significant additional enhancement in mechanical properties, as discussed below. Liu and coworkers have investigated a number of routes to the preparation of hollow polymer nanospheres for controlled release applications using cross-linked
262
Block Copolymers in Solution: Fundamentals and Applications
block copolymer vesicles. 188 They prepared PI-b-PCEMA vesicles in THF/hexane solutions. The PCEMA shell was photo cross-linked. 188 The PI chains were then hydroxylated to give water soluble vesicles with a hydrophilic coating on both sides of the shell (Figure 6.12). These were used to solubilize a dye compound, the
Figure 6.12 Vesicles formed from PI-b-PCEMA diblocks, following PCEMA shell crosslinking and hydroxylation of the PI core.188 Reproduced by permission of American Chemical Society.
release of which was studied via measurements of fluorescence intensity. The potential of nanocapsules with a cross-linked diblock shell for drug delivery applications was highlighted.188 The enhanced stability of cross-linked nanoparticles (for example in the bloodstream) and the ability to tune the capsule size and to incorporate responsive and/or functional moieties are additional advantages offered by the use of polymers. In an alternative approach, micelles of PSMA-b-PDMA were formed in aqueous solution with PDMA coronas and PSMA cores.189 The PDMA coronas were then chemically cross-linked. Finally the cores of the nanospheres were made hydrophilic by hydrolysis of the nitroxide groups in PSMA. It should be noted that this approach does not produce hollow spheres, and it is not clear how the uptake and release of hydrophilic compounds across the cross-linked shell can be decoupled. Methods to prepare hollow nanospheres have also been
Applications
263
developed for ABC triblocks. PI-b-PCEMA-b-PtBS micelles were formed in THF/ methanol mixtures.190 The PCEMA shell was cross-linked, and the PI core removed by ozonolysis to leave PCEMA shells with a PtBS coating. In contrast to the two-stage procedure used by Liu et al. (cross-linking of the shell plus hydrolysis or ozone etching of the core), Meier and coworkers have pioneered a one-step route to prepare nanoshells from block copolymer vesicles.191,192 Vesicles were formed upon dispersion of an ethanolic solution of PMOXA-b-PDMS-b-PMOXA triblocks into aqueous solution. The PMOXA shell bears polymerizable end groups which can then be photo cross-linked,191 as illustrated in Figure 6.13. The resulting nanocapsules are mechanically stable, and also
Figure 6.13 Schematic showing polymerization of PMOXA-b-PDMS-b-PMOXA block copolymer vesicles.191 Reproduced by permission of American Chemical Society.
retain their shape upon dispersion in organic solvents. Free-standing planar membranes with areas of approximately 1 mm2 were also prepared using this chemistry, and the enhanced mechanical properties (stability against electric field-induced rupture) of these films were compared with those of lipid bilayers.193 In a beautiful concept demonstration, pore-forming transmembrane proteins have been incorporated within triblock copolymer planar membranes194 and vesicles (Figure 6.14).195 The functionality of the bacterial porin OmpF is retained after its incorporation into the block copolymer membrane. When incorporated into the membrane, this protein forms channels that exclude molecules with molecular weights above 400 Da. Responsive nanocapsules can be fabricated containing these channels, and these have great potential for targeted drug delivery and diagnostics.192 It has further been shown that such porous nanocapsules can be used to prepare calcium phosphate nanoparticles, where the Ca2+ is transported across the shell through the ion channels.196 Another application uses the cross-linked vesicles as nanoreactors. The
264
Block Copolymers in Solution: Fundamentals and Applications
Figure 6.14 Nanoreactor comprising a cross-linked block copolymer vesicle encapsulating the enzyme ß-lactamase which catalyses the hydrolysis of ampicillin.197 Reproduced by permission of Royal Society of Chemistry.
enzyme ß-lactamase was encapsulated and used to hydrolyse ampicillin, which is a ß-lactam antibiotic.195,197 The concept is illustrated in Figure 6.14. Application of a transmembrane voltage above a certain threshold causes a reversible gating action of the OmpF.195 Polyelectrolyte added to the solution was shown to trigger the gating transition. The Donnan equilibrium established leads to a concentration difference of polyelectrolyte chains either side of the membrane. If this exceeds the gating potential, the channels are closed and the nanoreactor is deactivated. Incorporation of the bacterial channel forming protein LamB into block copolymer vesicle membranes enables DNA to be transported into the nanocontainer, since LamB serves as a receptor for A phage to trigger the ejection of A phage DNA.198 If the compositional asymmetry is large enough, PMOXA-b-PDMS-b-PMOXA triblocks self-assemble into hollow nanotubes in aqueous solution.199 The tube diameter was relatively uniform (40 nm) with a length extending to several tens of microns. As discussed in Section 2.17, Discher et al. prepared block copolymer vesicles using PEO-b-PEE diblocks.200 The nanocapsules possess enhanced toughness and reduced water permeability compared with liposomes. The use of these particles in protein encapsulation has been comprehensively studied.201 The vesicles exhibit good long-term stability in blood plasma, i.e. phagocytosis is inhibited. Cell viability was also shown not to be adversely affected by addition of the polymersomes. In contrast to liposomes, the block copolymer vesicles can withstand harsh thermal treatments, to some extent although the vesicles shrink and leak their contents. It was proposed that these properties result from steric stabilization
Applications
265
conferred by PEG. Compared with PEG-lipids, the polymersomes offer much improved mechanical stability and stability against dilution. Polymer vesicles formed by block copolymers containing polypeptide blocks, i.e. so-called 'peptosomes' have been developed based on PB-b-poly(7-L-glutamic acid) (Figure 6.15).202-204 The helix-coil transition of the peptide does not change
Figure 6.15 Proposed structure of a PB-b-poly(7-L-glutamic acid) vesicle. by permission of Wiley-VCH.
Reproduced
the dimensions of the peptosome. In contrast, Checot et al. showed that the size of their PB-b-poly(7-L-glutamic acid) peptosomes could be reversibly manipulated as a function of both pH and ion strength. The vesicle hydrodynamic radius increased by 50% when the peptide underwent a transition from an a-helix to a random coil conformation, driven by increasing pH.203,204 These authors also demonstrated UVinduced cross-linking of the PB to produce 'shape persistent stimuli-responsive nanocapsules'. Purely peptide block copolymers have also been shown to selfassemble into peptosomes.205 Poly{Nb-2-[2-(2-methoxyethoxy)-ethoxy]acetyl-Llysine}-b-PLys diblocks self-assembled into giant unilamellar vesicles. Vesicle size could be controlled by adjustment of hydrophobic content or overall copolymer chain length. No release of solubilized dye was observed, over a period of weeks. A copolymer was prepared in which 70% of the L-leucine residues of the hydrophobic domain of the block copolymer were replaced in a statistical manner with L-lysine. At high pH, PLys is not water soluble and it adopts an a-helical conformation. Vesicles similar to those formed by the unmodified copolymers were observed. However, a helix-to-coil transition driven by a reduction in pH caused destabilization of the vesicles and release of their contents. Thus pH-responsiveness was
266
Block Copolymers in Solution: Fundamentals and Applications
demonstrated. The cooperative self-assembly of poly(L-lysine)- b-poly(L-cysteine) block copolymers in the presence of silica and gold nanoparticles can be used to prepare hollow spheres coated with a shell of gold and a shell of silica (Figure 6.16). The cysteine block facilitates disulfide links and these act as attachment points for gold nanoparticles. Introduction of silica nanoparticles then drove the formation of a silica coating. These hollow nanoshells were formed without the use of emulsions or a sacrificial core.
Figure 6.16 Schematic suggested for the hierarchical self-assembly of gold and silica nanopartices into hollow spheres with a two-layer shell structure, using poly(L-lysine)-bpoly(L-cysteine) block copolymers.260 Reproduced by permission of American Chemical Society.
Triblock copolymer vesicles which are unstable to oxidation have been prepared with the aim of developing a responsive system for drug delivery.206 PEG-b-PPS-bPEG triblocks were used, PEG being hydrophilic and poly(propylene sulfide) being hydrophobic but readily oxidized to poly(propylene sulfoxide) and ultimately poly(propylene sulfone), both of which are hydrophilic. This conversion results in the breakup of the vesicles. The resulting individual copolymer chains were short enough to be expected to be readily eliminated from the body. Biodegradable polymersomes have been reported based on PEG-b-PLLA diblocks. 207 The poration of PEG-b-PLLA and PEG-b-PCL vesicles (and mixed vesicles) due to hydrolysis has been investigated.208 Encapsulation of the anticancer agent doxorubicin was also studied. Release kinetics from giant vesicles were studied for model hydrophilic fluorescent encapsulants via phase contrast and fluorescence microscopy. The release (over a timescale from hours to days) was found to occur in a quantized manner, as each vesicle porates and disintegrates individually. A mechanism for the poration process is illustrated in Figure 6.17.
Applications
267
Figure 6.17 Mechanism proposed for membrane poration due to hydrolysis of polyester in PEG-b-polyester vesicles.208 Polyester chains are shown in light grey. Reproduced by permission of Elsevier.
Release kinetics can be controlled via the PEG content of the copolymer and the hydrophobicity of the other block (PCL is more hydrophobic than PLLA). Porous polymer microspheres have been prepared by using Pl-b-PAA diblocks to solubilize chloromethane containing PI-b-PtBA and homopolymer PtEA as oil droplets in water.209 The PtBA was then hydrolysed and the resulting PAA extracted to leave porous spheres (diameter 0.3-1 u.m). The shape and connectivity of pores could be controlled through the relative content of PtBA in the homopolymer versus diblock. Conjugation of block copolymers to liposomes leads to two scenarios, as shown in Figure 6.18.210 Addition of PEO-b-PPO-b-PEO block copolymer after lipid vesicle formation leads to adsorption, if the block copolymer does not contain too much hydrophobic block (if it does, micellization of the block copolymer predominates). If the polymer is present at low concentration during vesicle formation, however, it can be incorporated into the vesicle membrane, the PPO blocks being incorporated into the hydrophobic compartment.210 In contrast to this work, Johnsson and coworkers found significant morphological changes in liposome structure upon addition of Pluronic copolymers.211 Cryo-TEM was essential in elucidating morphology of the phosphatidylcholine vesicles upon addition of block copolymer. Pluronics with long PEO blocks were found to induce the formation of bilayer disks, whereas those with shorter PEO blocks tend to promote a reduction in liposome size. The Pluronics were found to increase the porosity of the liposomes, as revealed by fluorescence experiments on solubilized dye. Inclusion of cholesterol into the lipid membrane was found to reduce the incorporation of copolymer.
268
Block Copolymers in Solution: Fundamentals and Applications
Figure 6.18 Schematic for interaction between a unilamellar lipid vesicle and a PEO-bPPO-6-PEO triblock copolymer.210 Reproduced by permission of American Chemical Society.
6.8
SEPARATION MEDIA
The use of amphiphilic copolymers in gel electrophoresis has been examined. The selective partitioning between aqueous and hydrophobic domains in two-phase systems with Pluronic and dextran polymer components leads to a distribution between the two phases depending on the hydrophobicity or hydrophilicity of the protein.212-215 The partitioning has been investigated, with the Pluronic in both the unimer and micellar states. The partitioning was found to depend strongly on temperature, but not on micellization per se. Selective partitioning has also been exploited in capillary gel electrophoresis of nucleotides and DNA fragments, as investigated by the groups of Rill216-218 and Chu.219-221 In concentrated solution, Pluronic F127 was shown to form an fcc cubic micellar phase in the buffer solution used for the electrophoretic separations.219,221 Separation of nucleotides up to a few thousand base pairs in length has been demonstrated using this technique,216 and separation of individual nucleotides from standard mixtures (Figure 6.19).217
6.9
TEMPLATING
Block copolymers have been extensively used to template the formation of 'mesoporous' materials.7,140,222 Since the pore size is typically 5-25 nm, these
Applications
269
Figure 6.19 Separation via gel electrophoresis of double stranded DNA (123 base pairs) using gels of Pluronic F127 (gel concentration indicated).218 Reproduced by permission of AAAS.
materials should probably be termed 'nanoporous'. Compared with conventional nonionic surfactants, amphiphilic block copolymers offer the advantage of larger pore sizes. The templating of silica by amphiphilic block copolymers was first performed using commercially available PEO-b-PPO-b-PEO triblocks.223 A representative TEM image of a hexagonal structure is shown in Figure 6.20. Lamellar or bicontinuous structures have been prepared in a similar manner. Since the pore size and wall thickness can be varied according to the processing conditions, it is evident that the nanoporous silica is not generally a 'cast' of the lyotropic mesophase formed by the triblocks in solution, rather that the copolymers act as structuredirecting agents of the tetraalkoxysilane precursors. Other research teams have investigated the production of mesoporous silica from different PEO-b-PPO-b-PEO triblocks,224-226 and also related PEO-b-PBO-b-PEO copolymers.227,228 The patterning of silica and other metal oxides including TiO2, ZrO2, WO3, etc., was performed in a related procedure, although using chloride salts in ethanol rather than metal alkoxides in acidic aqueous solutions.229,230 It is proposed that complexation occurs between the oxygen atoms in PEO and the multivalent metal
270
Block Copolymers in Solution: Fundamentals and Applications
Figure 6.20 Hexagonal structure of calcined mesoporous silica, templated using a PEO-bPPO-b-PEO triblock copolymer.223 Reproduced by permission of Science.
species in the alkoxides. The production of mesoporous silica in thin films using PS-b-PEO diblocks has been achieved via selective solvent evaporation, which leads to cooperative self-assembly of the diblock and silicate to produce mesoporous silica with hexagonally arranged pores, vesicular structures or rods.231,232 The use of CH3-Si(OCH2CH3)3 with the PS-b-PEO diblock in the synthesis reduces the siloxane condensation rate, allowing cooperative self-assembly of the silica and amphiphile.232 True nanocasting can be achieved starting from a high concentration liquid crystal phase, where the continuous liquid phase undergoes a sol-gel transition. The use of block copolymers in the preparation of mesoporous materials has been discussed in a review of nanocasting methods.233 This route has been explored using block copolymers in which one block is a polyelectrolyte234 or both blocks are non-ionic.235,236 In the former case,234 both cationic and anionic diblocks have been used as structure-directing agents to produce materials with irregular arrays of spherical pores. In the latter case,235,236 diblock copolymers based on PEO have been used. Sol-gel chemistry has also been exploited to prepare metal oxide-containing mesoporous silica with well-defined lamellar and cylindrical structures using Pl-bPEO diblock solutions to template the structure formation.235 It is proposed that the metal alkoxide precursors preferentially swell the PEO block. Condensation then leads to mesoporous silica containing metal oxide. By dissolving the organic component, spherical, cylindrical and plate-like nanoparticles could be prepared.237 These are coated with PEO chains (so-called 'hairy' nano-objects), which can be removed by thermal treatment. A schematic showing the synthetic approach used to prepare these nano-objects is shown in Figure 6.21. In subsequent work, the same team has reported the preparation of a mesoporous silicon/aluminium oxide ceramic
Applications
271
Figure 6.21 Schematic of the sol-gel approach to synthesize PEO-coated nanoparticles of different shape using a sol-gel synthesis of metal alkoxides templated by PI-PEO diblocks.237 The symbol A indicates thermal treatment. Reproduced by permission of Wiley-VCH.
material with a bicontinuous 'plumbers nightmare' pore structure.238 It is interesting to note the analogous approach using the 'ball-at-the-wall' morphology formed by ABC triblocks in the melt to prepare Janus micelles upon crosslinking the PB 'balls', and dissolution of the lamellar matrix,239 as noted in Section 2.12.5. Antonietti and coworkers used PEO-b-PEB diblock copolymers to prepare porous silica.236 Either a disordered or a regular packing of spherical pores was obtained, depending on polymer concentration. In another variation, the lyotropic phase structure can be fixed by cross-linking the gel structure prior to the templating of silica.240 Diblocks containing PB (with PEO) were used for this purpose, since PB can be cross-linked by 7-radiation. In a related approach, photopolymerization of water and oil-soluble monomers in lyotropic mesophases formed by Pluronic copolymers has been studied.241 The influence of the order in the mesophase on the polymerization kinetics was examined. The resulting materials were polymeric hydrogels. A biomimetic route to the preparation of mesoporous silica has been demonstrated by Cha et al.242 They have developed poly(L-lysine)-b-poly(L-cysteine) hydrophilic-hydrophobic diblock copolypeptides that mimic the properties of the protein silicatein, involved in the condensation of tetraethyoxysilane into silica in, for example, sponges and diatoms. The hydrolysis of silica occurs at neutral pH, in
272
Block Copolymers in Solution: Fundamentals and Applications
contrast to conventional methods that rely on acidic conditions, or the use of organic solvents. Porous silica beads and rods were produced, depending on the copolymer composition. Ordering on multiple length scales can be achieved by confining the mesoporous silica synthesis using Pluronic copolymer templates within the interstices between colloid particles,243 using soft lithography with elastomeric stamps (Figure 6.22),243
Figure 6.22 Soft lithographic moulding of a sol-gel block copolymer precursor solution.243 Reproduced by permission of Science.
or within a macroporous polymer foam.244 Using the soft lithography approach, together with the templated synthesis of mesoporous silica, mirrorless lasing has been demonstrated.243 Arrays of mesas (1-3 um wide, 1-2 um high, 2-8 um spacing, several centimetres long) were patterned, containing mesoporous silica. When doped with laser dye, amplification of stimulated emission within the waveguides was noted. Hollow spheres formed by rod-coil diblocks in a selective solvent can aggregate directly into microporous materials provided the solvent casting conditions are properly optimized.245 Figure 6.23 schematically shows the process, and Figure 6.24 shows a fluorescence photomicrograph of an ordered microporous film prepared in this way. Block copolymers have been used to control the growth of calcium carbonate crystals, in either vaterite or calcite forms.246 Various types of diblock with a PEG hydrophilic block and a second hydrophilic polyacid were employed, the latter interacting strongly with the calcium ions. The crystal size and shape could be
Applications
273
Figure 6.23 Schematic showing diblock structure and process of formation of microporous material from hollow microspheres formed by its self-assembly in a coil-selective solvent.245 Reproduced by permission of Science.
controlled by appropriate choice of polymer. A PEG-b-PAsp diblock has similarly been used to control the growth of calcium phosphate crystals,85,86 as discussed in Section 6.3. Ikkala and coworkers have demonstrated hierarchical self-assembly in PS-bP4VP diblocks to which the amphiphile pentadecylphenol is conjugated via hydrogen bonds to the P4VP block.247,248 The surfactant can form a lamellar structure, independent of the microphase separation of the block copolymer. However, as yet, the concept has been investigated for melt systems although it seems natural to extend it to solutions. In an extension of this concept, Valkama et al. have shown
274
Block Copolymers in Solution: Fundamentals and Applications
Figure 6.24 Ordered microporous structure formed in a dried film245 prepared as shown in Figure 6.23. Reproduced by permission of Science.
that mesoporous materials can be prepared by solvent extraction of amphiphiles complexed to PS-b-P4VP diblocks. Zinc dodecylbenzene forms coordination complexes with P4VP. This approach offers the advantage that high molecular weight amphiphiles can be bonded, in comparison to complexes formed via hydrogen bonding. Following microphase separation, the amphiphile was extracted using methanol, leaving a porous lamellar structure, which was mechnically stable due to defects connecting the PS lamellae.
6.10
MEMBRANES
Sulfonated block copolymers have applications as polymer electrolyte membranes for use in fuel cells. Proton conduction occurs through the ionic channels of the membranes, formed by microphase separation between the hydrophilic proton exchange site and the hydrophobic domain. Such membranes have been prepared from sulfonated SEBS, polystyrene-b-poly(ethylene-ran-butylene)-b-polystyrene triblocks.250 These materials complement other well-studied polymers such as Nafion™ (DuPont) which is not a block copolymer, but which is a perfluorosulfonic acid. It has been proposed that this forms a porous structure from a percolating array of ionic clusters swollen by water dispersed in a continuous polymeric matrix. However, Diat and coworkers have suggested a model of elongated polymeric bundles.251,252 Won et al. have shown that cross-linking PS-b-PB-b-PS copolymers prior to sulfonation enables the swelling of the membrane, and hence ionic channel size, to be controlled.250 The proton conductivity was shown to be as good as Nafion, but the undesirable permeability of methanol was reduced. The latter is relevant to direct methanol fuel cells where the transport of methanol across the membrane needs to be minimized.
Applications
275
6.11 OTHER APPLICATIONS Other applications of block copolymers including uses in oil and coal recovery, agricultural applications, emulsification, lubrication and surface treatment among others are discussed elsewhere.2,3 The use of block copolymers as gelators of lipid lamellar phases has been explored by Safinya and coworkers.253'254 Gelation is observed with as little as 0.5% added amphiphilic block copolymer. The block copolymers comprise a hydrophilic PEG chain and hydrophobic double chain 3,4-(alkoxy)benzoic moieties. Gelation occurs for AB diblocks or ABA triblocks,253'254 showing that copolymer bridging is not essential (Figure 6.25). However, swelling in water is hindered in the ABA system due to bridging. The same group have also explored hydrogel formation by addition of PEG lipids to fluid lamellar phases.255,256
Figure 6.25 Lipid membranes decorated with block copolymers.253 (a) Amphiphilic AB diblocks in mushroom conformations, (b) ABA triblock copolymer in bridging conformation, (c) ABA triblock copolymer in looping conformation. Reproduced by permission of American Chemical Society.
As mentioned in Section 6.9, soft lithography has been used to prepare patterned films of porous silica from Pluronic solution precursors.243 Micropatterning of block copolymer solutions using elastomeric stamps has also been investigated by Thomas and coworkers, who point out the necessity of avoiding swelling of the
276
Block Copolymers in Solution: Fundamentals and Applications
mould by the solvent.257 In their case, this was achieved by use of a thin layer of amorphous fluorinated polymer coating the mould surface.
REFERENCES 1. Nace, V. N. Properties of Polyoxyalkylene Block Copolymers. In Nonionic surfactants. Polyoxyalkylene Block Copolymers', Nace, V. N., Ed. Marcel Dekker: New York, 1996; p 145. 2. Edens, M. W. Applications of Polyoxyethylene Block Copolymer Surfactants. In Nonionic Surfactants. Polyoxyalkylene Block Copolymers; Nace, V. N., Ed. Marcel Dekker: New York, 1996; p 185. 3. Edens, M. W.; Whitmarsh, R. H. Applications of Block Copolymer Surfactants. In Developments in Block Copolymer Science and Technology; Hamley, I. W., Ed. John Wiley b Sons, Ltd: Chichester, 2004; p 325. 4. Nace, V. N. (Ed.) Nonionic Surfactants. Polyoxyalkylene Block Copolymers. Marcel Dekker: New York, 1996. 5. Alexandridis, P.; Lindman, B. (Eds) Amphiphilic Block Copolymers: Self-assembly and Applications. Elsevier: Amsterdam, 2000. 6. Liu, S.; Armes, S. P. Curr. Opin. Colloid Interface Sci. 2001, 6, 249. 7. Liu, T.; Burger, C; Chu, B. Prog. Polym.Sci. 2003, 28, 5. 8. Forster, S.; Antonietti, M. Adv. Mater. 1998, 10, 195. 9. Forster, S.; Plantenberg, T. Angew. Chem., Int. Ed. Engl. 2002, 41, 688. 10. Kiserow, D.; Prochzka, K.; Ramireddy, C.; Tuzar, Z.; Munk, P.; Webber, S. E. Macromolecules 1992, 25, 461. 11. Kriz, J.; Masar, B.; Pospisil, H.; Plestil, J.; Tuzar, Z.; Kiselev, M. Macromolecules 1996, 29, 7853. 12. Bronstein, L. M.; Chernyshov, D. M.; Timofeeva, G. I.; Dubrovina, L. V.; Valetsky, P. M. Langmuir 1999, 75, 6195. 13. Matsumoto, K.; Mazaki, H.; Nishimura, R.; Matsuoka, H.; Yamaoka, H. Macromolecules 2000, 33, 8295. 14. Hurter, P. N.; Hatton, T. A. Langmuir 1992, 8, 1291. 15. Paterson, I. R; Chowdhry, B. Z.; Leharne, S. A. Langmuir 1999, 75, 6187. 16. Gadelle, F; Koros, W. J.; Schechter, R. S. Macromolecules 1995, 28, 4883. 17. Kozlov, M. Y; Melik-Nubarov, N. S.; Batrakova, E. V.; Kabanov, A. V. Macromolecules 2000, 33, 3305. 18. Alexandridis, P.; Andersson, K. J. Phys. Chem. B 1997, 707, 8103. 19. Yang, Z.; Pousia, E.; Heatley, F.; Price, C.; Booth, C.; Castelletto, V.; Hamley, I. W. Langmuir 2001, 77, 2106. 20. Fundin, J.; Yang, Z.; Kelarakis, A.; Hamley, I. W; Price, C.; Booth, C. J. Phys. Chem. B 2002,16, 11728. 21. Hamley, I. W.; Castelletto, V.; Fundin, J.; Crothers, M.; Attwood, D.; Talmon, Y. Colloid Polym. Sci. 2004, 282, 514. 22. Fundin, J.; Castelletto, V.; Yang, Z.; Hamley, I. W.; Waigh, T. A.; Price, C. J. Macromol. Sci., Phys. 2004, B43, 893. 23. Nagarajan, R.; Ganesh, K. J. Colloid Interface Sci. 1996, 184, 489.
Applications 24. 25. 26. 27. 28. 29. 30. 31. 32. 33.
34. 35. 36. 37. 38. 39. 40. 41. 42. 43. 44. 45. 46. 47. 48. 49. 50. 51. 52. 53. 54. 55. 56. 57. 58. 59. 60.
277
Nagarajan, R.; Barry, M.; Ruckinstein, E. Langmuir 1986, 2, 210. Kriz, J.; Kurkova, D.; Kadlec, P.; Tuzar, Z.; Plestil, J. Macromolecules 2000, 33, 1978. Jenekhe, S. A.; Chen, X. L. Science 1998, 279, 1903. Chen, X. L.; Jenekhe, S. A. Langmuir 1999, 75, 8007. Creutz, S.; Jerome, R. Langmuir 1999, 15, 7145. Gibanel, S.; Forcada, J.; Heroguez, V.; Schappacher, M.; Gnanou, Y. Macromolecules 2001, 34, 4451. Nagarajan, R.; Ganesh, K. J. Chem. Phys. 1993, 98, 7440. Nagarajan, R.; Ganesh, K. Macromolecules 1989, 22, 4312. Nagarajan, R. Coll. Surf. B 1999, 16, 55. Nagarajan, R. Solubilization of Hydrophobic Substances by Block Copolymer Micelles in Aqueous Solutions. In Solvents and Self-organization of Polymers; Webber, S. E., Munk, P., Tuzar, Z., Eds. Kluwer: Dordrecht, 1997; p 121. Cogan, K. A.; Leermakers, F. A. M.; Gast, A. P. Langmuir 1992, 8, 429. Cogan, K. A.; Gast, A. P. Macromolecules 1990, 23, 745. Vagberg, L. J. M.; Cogan, K. A.; Gast, A. P. Macromolecules 1991, 24, 1670. Tian, M.; Arca, E.; Tuzar, Z.; Webber, S. E.; Munk, P. J. Polym. Sci. B: Polym. Phys. 1995, 33, 1713. Dan, N.; Tirrell, M. Macromolecules 1993, 26, 637. Leibler, L. Makromol.Chem., Macromol.Symp 1988, 16, 1. Wang, Z.-G.; Safran, S. A. J. Phys. France 1990, 51, 185. Cantor, R. Macromolecules 1981, 14, 1186. Hamley, I. W. The Physics of Block Copolymers. Oxford University Press: Oxford, 1998. Pepin, M. P.; Whitmore, M. D. Macromolecules 2000, 33, 8644. Hamley, I. W. The Physics of Block Copolymers. Oxford University Press: Oxford, 1998; Ch. 6. Xing, L.; Mattice, W. L. Macromolecules 1997, 30, 1711. Xing, L.; Mattice, W. L. Langmuir 1998, 14, 4074. Wijmans, C. M.; Zhulina, E. B. Macromolecules 1993, 26, 7214. Jang, J.; Ha, H. Langmuir 2002, 18, 5613. Baines, F. L.; Dionisio, S.; Billingham, N. C.; Armes, S. P. Macromolecules 1996, 29, 3096. Jakobs, B.; Sottmann, T.; Strey, R.; Allgaier, J.; Willner, L.; Richter, D. Langmuir 1999, 15, 6707. Endo, H.; Allgaier, J.; Jakobs, B.; Monkenbusch, M.; Richter, D.; Sottmann, T.; Strey, R. Phys. Rev. Lett. 2000, 85, 102. Mawson, S.; Yates, M. Z.; O'Neill, M. L.; Johnston, K. P. Langmuir 1997, 13, 1519. Rippner, B.; Boschkova, K.; Claesson, P. M.; Arnebrandt, T. Langmuir 2002, 18, 5213. Buhler, E.; Dobrynin, A. V.; DeSimone, J. M.; Rubinstein, M. Macromolecules 1998, 31, 7347. Zhou, S.; Chu, B. Macromolecules 1998, 31, 5300. Zhou, S.; Chu, B. Macromolecules 1998, 31, 7746. Liu, L.-Z.; Cheng, Z.; Inomata, K.; Zhou, S.; Chu, B. Macromolecules 1999, 32, 5836. Triolo, R.; Triolo, A.; Triolo, F.; Steytler, D. C.; Lewis, C. A.; Heenan, R. K.; Wignall, G. D.; DeSimone, J. M. Phys. Rev. E 2000, 61, 4640. Kapellen, K. K.; Mistele, C. D.; DeSimone, J. M. Macromolecules 1996, 29, 495. Canelas, D. A.; Betts, D. E.; DeSimone, J. M. Macromolecules 1996, 29, 2818.
278
Block Copolymers in Solution: Fundamentals and Applications
61. 62. 63. 64. 65. 66. 67. 68. 69. 70. 71. 72.
Johnston, K. P. Curr. Opin. Colloid Interface Sci. 2000, 5, 351. Matsumoto, K.; Kubota, M.; Matsuoka, H.; Yamaoka, H. Macromolecules 1999,32, 7122. Kabanov, A. V.; Kabanov, V. A. Adv. Drug Deliv. Rev. 1998, 30, 49. Pouton, C. W. Adv. Drug Deliv. Rev. 2001, 53, 1. Kataoka, K.; Harada, A.; Nagasaki, Y. Adv. Drug Deliv. Rev. 2001, 47, 113. Kakizawa, Y; Kataoka, K. Adv. Drug Deliv. Rev. 2002, 54, 203. Lavasanifar, A.; Samuel, J.; Kwon, G. S. Adv. Drug Deliv. Rev. 2002, 54, 169. Torchilin, V. P. Adv. Drug Deliv. Rev. 2002, 54, 235. Liggins, R. T.; Burt, H. M. Adv. Drug Deliv. Rev. 2002, 54, 191. Otsuka, H.; Nagasaki, Y; Kataoka, K. Adv. Drug Deliv. Rev. 2003, 55, 403. van Nostrum, C. F. Adv. Drug Deliv. Rev. 2004, 56, 9. Schmolka, I. R. Polyoxamers in the Pharmaceutical Industry. In Polymers for Controlled Drug Delivery; Tarcha, P. J., Ed. CRC Press: Boston, 1991; p 189. Allen, C.; Maysinger, D.; Eisenberg, A. Coll Surf. B 1999, 16, 3. Rosier, A.; Vandermeulen, G. W. M.; Klok, H.-A. Adv. Drug Deliv. Rev. 2001, 53, 95. Torchilin, V. P. J. Controlled Release 2001, 73, 137. Adams, M. L.; Lavasanifar, A.; Kwon, G. S. J Pharm. Sci. 2003, 92, 1343. Duncan, R. Nature Rev. Drug Discov. 2003, 2, 347. Maeda, H.; Wu, J.; Sawa, T.; Matsumura, Y; Hori, K. J. Controlled Release 2000, 65, 271. Kabanov, A. V.; Alakhov, V. Y. Micelles of Amphiphilic Block Copolymers as Vehicles for Drug Delivery. In Amphiphilic Block Copolymers. Self-assembly and Applications; Alexandridis, P., Lindman, B., Eds. Elsevier: Amsterdam, 2000; p 347. Kabanov, A. V.; Chekhonin, V. P.; Alakhov, V. Y; Batrakova, E. V.; Lebedev, A. S.; Melik-Nubarov, N. S.; Arzhakov, S. A.; Levashov, A. V.; Morozov, G., V,; Severin, E. S.; Kabanov, V. A. FEBS Lett. 1989, 258, 343. Kabanov, A. V.; Batrakova, E. V.; Melik-Nubarov, N. S.; Fedoseev, N. A.; Dorodnich, T. Y; Alakhov, V. Y; Chekhonin, V. P.; Nazarova, I. R.; Kabanov, V. A. J. Controlled Release 1992, 22, 141. Kabanov, A. V.; Lemieux, P.; Vinogradov, S.; Alakhov, V. Adv. Drug Deliv. Rev. 2002,54, 223. lijima, M.; Nagasaki, Y; Okano, T.; Kato, M.; Kataoka, K. Macromolecules 1999, 32, 1140. Emoto, K.; Nagasaki, Y; Kataoka, K. Langmuir 1999, 75, 5212. Kakizawa, Y; Kataoka, K. Langmuir 2002, 18, 4539. Kakizawa, Y; Miyata, K.; Furukawa, S.; Kataoka, K. Adv. Mater. 2004, 16, 699. Yokoyama, M.; Miyauchi, M.; Yamada, N.; Okano, T.; Sakurai, Y; Kataoka, K.; Inoue, S. Cancer Res. 1990, 50, 1693. Yokoyama, M.; Okano, T.; Sakurai, Y; Ekimoto, H.; Shibazaki, C.; Kataoka, K. Cancer Res. 1991, 57, 3229. Yokoyama, M.; Okano, T.; Sakurai, Y; Kataoka, K. J. Controlled Release 1994, 32, 269. Kataoka, K.; Ishihara, A.; Harada, A.; Miyazaki, H. Macromolecules 1998, 37, 6071. Jeong, B.; Bae, Y. H.; Lee, D. S.; Kim, S. W. Nature 1997, 388, 860. Jeong, B.; Lee, D. S.; Shon, J.-L; Bae, Y H.; Kim, S. W. J. Polym. Sci., Pan A: Polym. Chem. 1999, 37, 751. Jeong, B.; Bae, Y. H.; Kim, S. W. Macromolecules 1999, 32, 7064. Jeong, B.; Bae, Y H.; Kim, S. W. J. Biomed. Mater. Res 2000, 50, 171.
73. 74. 75. 76. 77. 78. 79.
80.
81.
82. 83. 84. 85. 86. 87. 88. 89. 90. 91. 92. 93. 94.
Applications
279
95. Jeong, B.; Kibbey, M. R.; Birnbaum, J. C.; Won, Y.-Y.; Gutowska, A. Macromolecules 2000, 33, 8317. 96. Li, Y; Kissel, T. Polymer 1998, 39, 4421. 97. Dai, S.; Ravi, P.; Leong, C. Y; Tarn, K. C.; Gan, L. H. Langmuir 2004, 20, 1597. 98. Riley, T.; Stolnik, S.; Heald, C. R.; Xiong, C. D.; Garnett, M. C.; Ilium, L.; Davis, S. S. Langmuir 2001, 17, 3168. 99. Dai, Z.; et al. Colloid Polym. Sci. 2004, 282, 343. 100. Booth, C.; Yu, G.-E.; Nace, V. M. Block Copolymers of Ethylene Oxide and 1,2Butylene Oxide. In Amphiphilic Block Copolymers: Self-assembly and Applications; Alexandridis, R, Lindman, B., Eds. Elsevier: Amsterdam, 2000; p 57. 101. Booth, C.; Attwood, D. MacromolRapid Commun. 2000, 21, 501. 102. Luo, Y.-Z.; Nicholas, C. V.; Attwood, D.; Collett, J. H.; Price, C.; Booth, C.; Chu, B.; Zhou, Z. K. J. Chem. Soc., Faraday Trans. 1993, 89, 539. 103. Rekatas, C. J.; Mai, S.-M.; Crothers, M.; Quinn, M.; Collett, J. H.; Attwood, D.; Heatley, E; Martini, L.; Booth, C. Phys. Chem., Chem. Phys. 2001, 3, 4769. 104. Tanodekaew, S.; Godward, J.; Heatley, E; Booth, C. Macromol. Chem. Phys. 1997, 198, 3385. 105. Ma, Y; Tang, Y; Billingham, N. C.; Armes, S. P.; Lewis, A. L.; Lloyd, A. W.; Salvage, J. Macromolecules 2003, 36, 3475. 106. Ma, Y; Tang, Y; Billingham, N. C.; Armes, S. P.; Lewis, A. L. Biomacromolecules 2003, 4, 864. 107. Castelletto, V.; Hamley, I. W.; Ma, Y; Bories-Azeau, X.; Armes, S. P.; Lewis, A. L. Langmuir 2004, 20, 4306. 108. Shuai, X.; Merdan, T.; Unger, E; Wittmar, M.; Kissel, T. Macromolecules 2003, 36, 5751. 109. Yamasaki, Y; Katayose, S.; Kataoka, K.; Yoshikawa, K. Macromolecules 2003, 36, 6276. 110. Yasugi, K.; Nakamura, T.; Nagasaki, Y; Kato, M.; Kataoka, K. Macromolecules 1999, 32, 8024. 111. Nagasaki, Y; Yasugi, K.; Yamamoto, Y; Harada, A.; Kataoka, K. Biomacromolecules 2001, 2, 1067. 112. Jule, E.; Nagasaki, Y; Kataoka, K. Langmuir 2002, 18, 10334. 113. Jule, E.; Nagasaki, Y; Kataoka, K. Bioconjugate Chem. 2003, 14, 177. 114. Bes, L.; Angot, S.; Limer, A.; Haddleton, D. M. Macromolecules 2003, 36, 2493. 115. Kumar, N.; Ravikumar, M. N. V.; Domb, A. J. Adv. Drug Deliv. Rev. 2001, 53, 23. 116. Otsuka, H.; Nagasaki, Y; Kataoka, K. Biomacromolecules 2000, 1, 39. 117. Jeong, B.; Bae, Y H.; Kim, S. W. J. Controlled Release 2000, 63, 155. 118. Jeong, B.; Kim, S. W.; Bae, Y H. Adv. Drug Deliv. Rev. 2003, 54, 37. 119. Li, S.; Vert, M. Macromolecules 2003, 36, 8008. 120. Grijpma, D. W.; Feijen, J. J. Controlled Release 2003, 72, 247. 121. Kwon, K.-W.; Park, M. J.; Bae, Y. H.; Kim, H. D.; Char, K. Polymer 2002, 43, 3353. 122. Park, M. J.; Char, K. Langmuir 2004, 20, 2456. 123. Aamer, K. A.; Sardinha, H.; Bhatia, S. R.; Tew, G. N. Biomaterials 2004, 25, 1087. 124. Nie, T.; Zhao, Y; Xie, Z.; Wu, C. Macromolecules 2003, 36, 8825. 125. Gan, Z.; Jim, T. E; Li, M.; Yuer, Z.; Wang, S.; Wu, C. Macromolecules 1999, 32, 590. 126. Schild, H. G. Prog. Polym. Sci. 1992, 17, 163. 127. Pelton, R. Adv. Colloid Interface Sci. 2000, 85, 1.
280
Block Copolymers in Solution: Fundamentals and Applications
128. 129. 130. 131. 132. 133. 134.
Xue, W.; Hamley, I. W.; Huglin, M. B. Polymer 2002, 43, 5181. Xue, W.; Hamley, I. W. Polymer 2002, 43, 3069. Neradovic, D.; van Nostrum, C. E; Hennink, W. E. Macromolecules 2001, 34, 7589. Konak, C.; Oupicky, D.; Chytry, V.; Ulbrich, K. Macromolecules 2000, 33, 5318. Topp, M. D. C.; Dijkstra, P. J.; Talsma, H.; Feijen, J. Macromolecules 1997, 30, 8518. Virtanen, J.; Holappa, S.; Lemmetyinen; Tenhu, H. Macromolecules 2002, 35, 4763. Cammas, S.; Suzuki, K.; Sone, C.; Sakurai, Y.; Kataoka, K.; Okano, T. J. Controlled Release 1997, 48, 157. Kohori, E; Sakai, K.; Aoyagi, T.; Yokoyama, M.; Sakurai, Y; Okano, T. J. Controlled Release 1998, 55, 87. Schilli, C. M.; Zhang, M.; Rizzardo, E.; Thang, S. H.; Chong, B. Y. K.; Edwards, K.; Karlsson, G.; Miiller, A. H. E. Macromolecules 2004, 37, 7861. Sugihara, S.; Hashimoto, K.; Okabe, S.; Shibayama, M.; Kanaoka, S.; Aoshima, S. Macromolecules 2004, 37, 336. Park, S. Y; Han, D. K.; Kim, S. C. Macromolecules 2001, 34, 8821. Cohen, R. E. Curr. Opin. Sol. Stat. Mater. Sci. 1999, 4, 587. Hamley, I. W. Nanotechnology 2003, 14, R39. Chan, Y. N. C.; Schrock, R. R.; Cohen, R. E. Chem. Mater. 1992, 4, 24. Chan, Y. N. C.; Schrock, R. R.; Cohen, R. E. J. Am. Chem. Soc. 1992, 114, 7295. Sohn, B. H.; Cohen, R. E. Acta Polym. 1996, 47, 340. Yue, J.; Sankaran, V.; Cohen, R. E.; Schrock, R. R. J. Am. Chem. Soc. 1993, 115, 4409. Cohen, R. E.; Clay, R. T.; Ciebien, J. E; Sohn, B.-H. Metal Nanoclusters (Within Block Copolymer Domains). In Polymeric Materials Encyclopedia; Salamone, J. C., Ed.; CRC Press: Boca Raton, 1996. Sankaran,V;Yue,J.;Cohen,R.E.;Schrock,R.R.;Silbey,R.J.Chem.Mater. 1993,5, 1133. Yue, J.; Cohen, R. E. Supramol. Sci. 1995, 1, 117. Clay, R. T.; Cohen, R. E. Supramol. Sci. 1995, 2, 183. Ahmed, S. R.; Kofinas, P. Macromolecules 2002, 35, 3338. Cummins, C. C.; Schrock, R. R.; Cohen, R. E. Chem. Mater. 1992, 4, 27. Kane, R. S.; Cohen, R. E.; Silbey, R. Chem. Mater. 1996, 8, 1919. Moffitt, M.; Eisenberg, A. Chem. Mater. 1995, 7, 1178. Moffitt, M.; McMahon, L.; Pessel, V.; Eisenberg, A. Chem. Mater. 1995, 7, 1185. Zhao, H.; Douglas, E. P.; Harrison, B. S.; Schanze, K. S. Langmuir 2001, 17, 8428. Spatz, J.; Mossmer, S.; Moller, M.; Kocher, M.; Neher, D.; Wegner, G. Adv. Mater. 1998, 10, 473. Sakai, T.; Alexandridis, P. Langmuir 2004, 20, 8426. Seregina, M. V.; Bronstein, L. M.; Platonova, O. A.; Chernyshov, D. M.; Valetsky, P. M.; Hartmann, J.; Wenz, E.; Antonietti, M. Chem. Mater. 1997, 9, 923. Bronstein, L.; Chernyshov, D.; Valetsky, P. Langmuir 1999, 15, 83. Bronstein, L. M.; Sidorov, S. N.; Valetsky, P. M.; Hartmann, J.; Colfen, H.; Antonietti, M. Langmuir 1999, 75, 6256. Antonietti, M.; Wenz, E.; Bronstein, L.; Seregina, M. Adv. Mater. 1995, 7, 1000. Platonova, O. A.; Bronstein, L. M.; Solodovnikov, S. P.; Yanovskaya, I. M.; Obolonkova, E. S.; Valetsky, P. M.; Wenz, E.; Antonietti, M. Colloid Polym. Sci. 1997, 275, 426. Antonietti, M.; Forster, S.; Hartmann, J.; Oestreich, S. Macromolecules 1996, 29, 3800. Bronstein, L. M.; Valetsky, P. M. J. Inorg. Organomet. Polym. 1994, 4, 415.
135. 136. 137. 138. 139. 140. 141. 142. 143. 144. 145.
146. 147. 148. 149. 150. 151. 152. 153. 154. 155. 156. 157. 158. 159. 160. 161. 162. 163.
Applications
281
164. Bronstein, L. M.; Seregina, M. V.; Platonova, O. A.; Kabachii, Y. A.; Chernyshov, D. M.; Ezernitskaya, M. G.; Dubrovina, L. V.; Bragina, T. P.; Valetsky, P. M. Macromol. Chem. Phys. 1998, 199, 1357. 165. Tsutsumi, K.; Funaki, Y.; Hirokawa, Y; Hashimoto, T. Langmuir 1999, 75, 5200. 166. Sidorov, S. N.; Bronstein, L. M.; Kabachii, Y. A.; Valetsky, R M.; Soo, P. L.; Maysinger, D.; Eisenberg, A. Langmuir 2004, 20, 3543. 167. Bronstein, L. M.; Chernyshov, D. M.; Timofeeva, G. I.; Dubrovina, L. V.; Valetsky, P. M.; Obolonkova, E. S.; Khokhlov, A. R. Langmuir 2000, 16, 3626. 168. Bronstein, L. M.; Chernyshov, D. M.; Timofeeva, G. I.; Dubrovina, L. V; Valetsky, P. M.; Khokhlov, A. R. J. Colloid Interface Sci. 2000, 230, 140. 169. Persigehl, P.; Jordan, R.; Nuyken, O. Macromolecules 2000, 33, 6977. 170. Nuyken, O.; Weberskirch, R.; Bortenschlager, M.; Schonfelder, D. Macromol. Symp. 2004, 215, 215. 171. Bouyer, E; Gerardin, C.; Fajula, E; Putaux, J.-L.; Chopin, T. Coll Surf. A 2003, 217, 179. 172. Gerardin, C.; Sanson, N.; Bouyer, E; Fajula, E; Putaux, J.-L.; Joanicot, M.; Chopin, T. Angew. Chem., Int. Ed. 2003, 42, 3681. 173. Cornelissen, J. J. L. M.; van Heerbeek, R.; Kamer, P. C. J.; Reek, J. N. H.; Sommerdijk, N. A. J. M.; Nolte, R. J. M. Adv. Mater. 2002, 14, 489. 174. Li, Z.; Liu, G. Langmuir 2003, 19, 10480. 175. Yan, X.; Liu, E; Li, Z.; Liu, G. Macromolecules 2001, 34, 9112. 176. Fahmi, A. W.; Braun, H.-G.; Stamm, M. Adv. Mater. 2003, 75, 1201. 177. Gohy, J.-E; Lohmeijer, B. C. G.; Schubert, U. S. Macromolecules 2002, 35, 4560. 178. Gohy, J.-E; Lohmeijer, B. C. G.; Varshney, S. K.; Schubert, U. S. Macromolecules 2002, 35, 7427. 179. Gohy, J.-E; Lohmeijer, B. C. G.; Varshney, S. K.; Decamps, B.; Leroy, E.; Boileau, S.; Schubert, U. S. Macromolecules 2002, 35, 9748. 180. Resendes, R.; Massey, J. A.; Dorn, H.; Power, K. N.; Winnik, M. A.; Manners, I. Angew. Chem., Int. Ed. 1999, 38, 2570. 181. Massey, J. A.; Temple, K.; Cao, L.; Rharbi, Y; Raez, J.; Winnik, M. A.; Manners, I. J. Am. Chem. Soc. 2000, 722, 11577. 182. Cao, L.; Manners, I.; Winnik, M. A. Macromolecules 2001, 34, 3353. 183. Resendes, R.; Massey, J. A.; Temple, K.; Cao, L.; Power-Billard, K. N.; Winnik, M. A.; Manners, I. Chem. Eur. J. 2001, 7, 2414. 184. Raez, J.; Manners, I.; Winnik, M. A. Langmuir 2002, 18, 7229. 185. Massey, J. A.; Winnik, M. A.; Manners, I; Chan, V. Z.-H.; Ostermann, J. M.; Enchelmaier, R.; Spatz, J. P.; Moller, M. J Am. Chem. Soc. 2001, 723, 3147. 186. Resendes, R.; Massey, J. A.; Dorn, H.; Winnik, M. A.; Manners, I. Macromolecules 2000, 33, 8. 187. Dalheimer, P.; Bates, F. S.; Aranda-Espinoza, H.; Discher, D. E. C. R. Phys. 2003, 4, 251. 188. Ding, J.; Liu, G. J. Phys. Chem. B 1998, 702, 6107. 189. Zhang, Z.; Liu, G.; Bell, S. Macromolecules 2000, 33, 7877. 190. Stewart, S.; Liu, G. Chem. Mater. 1999, 77, 1048. 191. Nardin, C.; Hirt, T.; Leukel, J.; Meier, W. Langmuir 2000, 16, 1035. 192. Sauer, M.; Meier, W. Colloidal Nanoreactors and Nanocontainers. In Colloids and Colloid Assemblies; Caruso, E, Ed. Wiley-VCH: Weinheim, 2004. 193. Nardin, C.; Winterhalter, M.; Meier, W. Langmuir 2000, 16, 7708. 194. Meier, W.; Nardin, C.; Winterhalter, M. Angew. Chem., Int. Ed. 2000, 39, 4599.
282 195. 196. 197. 198. 199. 200. 201. 202. 203. 204. 205. 206. 207. 208. 209. 210. 211. 212. 213. 214. 215.
216. 217. 218. 219. 220. 221. 222. 223. 224. 225. 226. 227. 228. 229.
Block Copolymers in Solution: Fundamentals and Applications Nardin, C.; Widmer, J.; Winterhalter, M.; Meier, W. Eur. Phys. J. E 2001, 4, 403. Sauer, M.; Haefele, T.; Graff, A.; Nardin, C.; Meier, W. Chem. Commun. 2001, 2452. Nardin,C.;Thoeni,S.;Widmer,G.;Winterhalter,M.;Meier,W.C/zem.CommM.2000,1433. Graff, A.; Sauer, M.; Van Gelder, P.; Meier, W. Proc. Natl Acad. Sci. USA 2002,99, 5064. Grumelard, J.; Taubert, A.; Meier, W. Chem. Commun. 2004, 1462. Discher, B. M.; Won, Y.-Y.; Ege, D. S.; Lee, J. C.-M.; Bates, F. S.; Discher, D. E.; Hammer, D. A. Science 1999, 284, 1143. Lee, J. C.-M.; Bermudez, H.; Discher, B. M.; Won, Y.-Y; Bates, F. S.; Discher, D. E. Biotechnol. Bioeng. 2001, 73, 135. Kukula, H.; Schlaad, H.; Antonietti, M.; Forster, S. J. Am. Chem. Soc. 2002, 124, 1658. Checot, F; Lecommandoux, S.; Gnanou, Y; Klok, H.-A. Angew. Chem., Int. Ed. Engl. 2002, 41, 1340. Checot, F; Lecommandoux, S.; Klok, H.-A.; Gnanou, Y. Eur. Phys. J. E 2003, 10, 25. Bellomo, E. G.; Wyrsta, M. D.; Pakstis, L.; Pochan, D. J.; Deming, T. J. Nature Mater. 2004, 3, 244. Napoli, A.; Valentini, M.; Tirelli, N.; Miiller, M.; Hubbell, J. A. Nature Mater. 2004, 3, 183. Meng, F; Hiemstra, C.; Engbers, G. H. M.; Feijen, J. Macromolecules 2003, 36, 3004. Ahmed, F; Discher, D. E. J. Controlled Release 2004, 96, 37. Lu, Z.; Liu, G.; Liu, F. J. Appl. Polym. Sci. 2003, 90, 2785. Kostarelos, K.; Tadros, T. F; Luckham, P. F. Langmuir 1999, 15, 369. Johnsson, M.; Silvander, M.; Karlsson, G.; Edwards, K. Langmuir 1999, 75, 6314. Svensson, M.; Joabsson, F; Linse, P.; Tjerneld, F. J. Chromatogr. A 1997, 761, 91. Tani, H.; Matsuda, A.; Kamidate, T.; Watanabe, H. Anal. Sci. 1997, 13, 925. Svensson, M.; Berggren, K.; Veide, A.; Tjerneld, F. J. Chromatogr. A 1999, 839, 71. Svensson, M.; Johansson, H.-O.; Tjerneld, F. Applications of Amphiphilic Copolymers in Separations. In Amphiphilic Block Copolymers. Self-assembly and Applications', Alexandridis, P., Lindman, B., Eds. Elsevier: New York, 2000; p 377. Rill, R. L.; Liu, Y; van Winkle, D. H.; Locke, B. R. J. Chromatogr. A 1998, 817, 287. Liu, Y; Locke, B. R.; van Winkle, D. H.; Rill, R. L. J. Chromatogr. A 1998, 817, 367. Rill, R. L.; Locke, B. R.; Liu, Y; van Winkle, D. H. Proc. Natl. Acad. Sci. USA 1998, 95, 1534. Wu, C.; Liu, T.; Chu, B.; Schneider, D. K.; Graziano, V. Macromolecules 1997, 30, 4574. Wu, C. H.; Liu, T. B.; Chu, B. Electrophoresis 1998, 19, 231. Wu, C.; Liu, T.; Chu, B. J. Non-Cryst. Sol. 1998, 235-237, 605. Raimondi, M. E.; Seddon, J. M. Liq. Cryst. 1999, 26, 305. Zhao, D.; Feng, J.; Huo, Q.; Melosh, N.; Fredrickson, G. H.; Chmelka, B. F; Stucky, G. D. Science 1998, 279, 548. Sato, Y; Nakanishi, K.; Hirao, K.; Jinnai, H.; Shibayama, M.; Melnichenko, Y. B.; Wignall, G. D. Coll. Surf. A 2001, 187-188, 117. Kipkemboi, P.; Fogden, A.; Alfredsson, V.; Flodstrom, K. Langmuir 2001, 17, 5394. Ryan, K. M.; Coleman, N. R. B.; Lyons, D. M.; Hanrahan, J. P.; Spalding, T. R.; Morris, M. A.; Steytler, D. C.; Heenan, R. K.; Holmes, J. D. Langmuir 2002, 18, 4996. Abu-Lebdeh, Y A. L; Budd, P. M.; Nace, V. M. J. Mater. Chem. 1998, 8, 1839. Tattershall, C. E.; Aslam, S. J.; Budd, P. M. J. Mater. Chem. 2003, 12, 2286. Yang, P.; Zhao, D.; Margolese, D. I; Chmelka, B. F; Stucky, G. D. Nature 1998, 396, 152.
Applications
283
230. Yang, P.; Zhao, D.; Margolese, D. I.; Chmelka, B. R; Stucky, G. D. Chem. Mater. 1999, 11, 2813. 231. Yu, K.; Hurd, A. J.; Eisenberg, A.; Blinker, C. J. Langmuir 2001, 17, 7961. 232. Yu, K.; Smarsly, B.; Brinker, C. J. Adv. Functional Mater. 2003, 13, 47. 233. Polarz, S.; Antonietti, M. Chem. Commun. 2002, 2593. 234. Kramer, E.; Forster, S.; Goltner, C.; Antonietti, M. Langmuir 1998, 14, 2027. 235. Templin, M.; Franck, A.; Du Chesne, A.; Leist, H.; Zhang, Y; Ulrich, R.; Schadler, V.; Wiesner, U. Science 1997, 278, 1795. 236. Thomas, A.; Schlaad, H.; Smarsly, B.; Antonietti, M. Langmuir 2003, 19, 4455. 237. Ulrich, R.; Du Chesne, A.; Templin, M.; Wiesner, U. Adv. Mater. 1999, 11, 141. 238. Pinnefrock, A. C.; Ulrich, R.; Du Chesne, A.; Honeker, C. C.; Schumacher, K.; Unger, K. K.; Gruner, S. M.; Wiesner, U. Angew. Chem., Int. Ed. 2001, 40, 1208. 239. Erhardt, R.; Boker, A.; Zettl, H.; Kaya, H.; Pyckhout-Hintzen, W.; Krausch, G.; Abetz, V.; Miiller, A. H. E. Macromolecules 2001, 34, 1069. 240. Hentze, H.-P.; Kramer, E.; Berton, B.; Forster, S.; Antonietti, M.; Dreja, M. Macromolecules 1999, 32, 5803. 241. McCormick, D. T.; Stovall, K. D.; Guymon, C. A. Macromolecules 2003, 36, 6549. 242. Cha, J. N.; Stucky, G. D.; Morse, D. E.; Deming, T. J. Nature 2000, 403, 289. 243. Yang, P.; Deng, T.; Zhao, D.; Feng, P.; Pine, D.; Chmelka, B. F.; Whitesides, G. M.; Stucky, G. D. Science 1998, 282, 2244. 244. Maekawa, H.; Esquena, J.; Bishop, S.; Solans,C.; Chmelka, B. F.Adv. Mater. 2003, 15, 591. 245. Jenekhe, S. A.; Chen, X. L. Science 1999, 283, 372. 246. Colfen, H.; Antonietti, M. Langmuir 1998, 582. 247. Ruokolainen, J.; Makinen, R.; Torkkeli, M.; Makela, T.; Serimaa, R.; ten Brinke, G.; Ikkala, O. Science 1998, 280, 557. 248. Ruokolainen, J.; Saariaho, M.; Ikkala, O.; ten Brinke, G.; Thomas, E. L.; Torkkeli, M.; Serimaa, R. Macromolecules 1999, 32, 1151. 249. Valkama, S.; Ruotsalainen, T.; Kosonen, H.; Ruokolainen, J.; Torkkeli, M.; Serimaa, R.; ten Brinke, G.; Ikkala, O. Macromolecules 2003, 36, 3986. 250. Won, J.; Park, H. H.; Kim, Y. J.; Choi, S. W.; Ha, H. Y; Oh, I.-H.; Kim, H. S.; Kang, K. S.; Ihn, K. J. Macromolecules 2003, 36, 3228. 251. Blachot, J. F.; Diat, O. J. Membr. Sci. 2003, 214, 31. 252. Rubatat, L.; Rollet, A. L.; Gebel, G.; Diat, O. Macromolecules 2003, 35, 4050. 253. Slack, N. L.; Schellhorn, M.; Eiselt, P.; Chibbaro, M. A.; Schulze, U.; Warriner, H. E.; Davidson, P.; Schmidt, H.-W.; Safinya, C. R. Macromolecules 1998, 31, 8503.' 254. Slack, N. L.; Davidson, P.; Chibbaro, M. A.; Jeppesen, C.; Eiselt, P.; Warriner, H. E.; Schmidt, H.-W.; Pincus, P.; Safinya, C. R. J. Chem. Phys. 2001, 115, 6252. 255. Warriner, H. E.; Idziak, S. H. J.; Slack, N. L.; Davidson, P.; Safinya, C. R. Science 1996, 277, 969. 256. Warriner, H. E.; Keller, S. L.; Idziak, S. H. J.; Slack, N. L.; Davidson, P.; Zasadzinski, J. A.; Safinya, C. R. Biophys. J. 1998, 75, 272. 257. Deng, T.; Ha, Y.-H.; Cheng, J. Y; Ross, C. A.; Thomas, E. L. Langmuir 2002,18, 6719. 258. Forster, S.; Konrad, M. J. Mater. Chem. 2003, 13, 2671. 259. Mossmer, S.; Spatz, J. P.; Moller, M.; Aberle, T; Schmidt, J.; Burchard, W. Macromolecules 2000, 33, 4791. 260. Wong, M. S.; Cha, J. N.; Choi, K.-S.; Deming, T. J.; Stucky, G. D. Nano Letters 2002, 2, 583.
Index ABC triblock 45-47, 62-66, 147, 176-177, 182-183 Adsorption kinetics 226, 229-231, 234 Alpha helix 202-204, 206-208, 265 Amyloid 204 Aniansson-Wall theory 53-55 Annealed polyelectrolyte 173 Anomalous micellization 21 Anomalous SAXS 188 Association number 23, 26, 29-30, 31, 43-44, 47-50, 53, 77, 175, 184-185, 189, 194 Associative thickener 60, 62 Atomic force microscopy (AFM) 11-12 Baxter model 14, 51 Beta sheet 202, 204-206, 207, 249 Bicontinuous sponge 116 Bidisperse polymer brush 225 Biocompatibility 224, 251 Biodegradable surface 224, 248 Birefringence 133, 140, 150 Blob model 34, 144, 152 Block copolymer/lipid complex 267-268, 275 Block copolymer/surfactant complex 76-78, 128, 198-199 Block copolymer/surfactant conjugate 273-274 Bowl-shaped micelle 178 Brewster angle microscopy 217-218 Bridging 29, 60-61, 129-131, 162-163, 233 Brush model 33-40, 48, 90-91, 188-190, 216-218, 220, 232, 235-236 Caging 162 Capillary gel electrophoresis 268 Carnahan-Starling equation 16
Chain dynamics 56-60 Coiled coil 203, 206-207 Cold gelation 126 Composition fluctuations 133-135, 144-146, 151 Compound micelles 79, 83, 176 Compound vesicles 82 Computer simulation 44-45, 148-149, 234-236 CONTIN 9 Cooperative diffusivity 159-160 Corona layer thickness 48-49, 184-185, 222, 225 Counterion distribution 187, 188, 191 Creep experiments 154, 158-159 Critical aggregation concentration (cac) 62, 76, 198 Critical gel concentration (cgc) 124-126, 201 Critical micelle concentration (cmc) 20-25, 29, 32-33, 42, 43, 175, 190 Critical micelle temperature (cmt) 24, 32-33, 40, 42-44 Cross-linking 62-66, 68-71, 79, 84-85, 116, 131-132, 206, 249, 259-260, 261-263 Cryo-TEM 7 Crystallization in micelles 90-91 Cubic-cubic phase transition 137-139 Cyclic copolymers 28-30 Daoud-Cotton model 34-35, 48, 187, 235, 243 Debye plot 15 Demicellization 132 Dendrimer-containing block copolymer 32 Density functional theory 136-137, 148 Density profile 35, 48, 217, 221, 246 Detergency 241
Block Copolymers in Solution: Fundamentals and Applications © 2005 John Wiley b Sons, Ltd.
I. W. Hamley
286 Dielectric spectroscopy 119, 130 Differential scanning calorimetry 8 Diffusion coefficient 9 Dilution approximation 106-107, 133-134, 143, 147 Disk-shaped micelle 63, 68, 91 DNA complexation 196-197, 249, 251-252 Domain spacing scaling 115, 140-143, 187 Double hydrophilic copolymers 132, 179, 182, 258 Dynamic density functional theory 149, 195 Dynamic light scattering 8-10 Dynamic modes 56-60, 135, 159-160 Dynamic shear moduli 117-119 Dynamic structure factor 57, 164 Elastic modulus Electric-field alignment 140 Ellipsometry 10 Emulsification 215, 241, 245-246 Enthalpy of gelation 126 Enthalpy of micellization 8, 23-25, 32-33, 77 Enzymatic degradation 253 EPR (enhanced permeability and retention) mechanism 248 Exchange kinetics 53-56 Flower-like micelles 29, 45, 60-61, 129, 163, 261 Fluorescence probe experiments 10, 52, 56, 77, 90 Fluorinated chains 60, 62, 64, 130-131 Form factor 13, 51, 63, 72, 74, 114 Gelation 117-132, 200 Gelation dynamics 160-164 Gel point 119, 163 Gel structure 75, 124-126 Gene therapy 252 Gibbs energy of micellization 23-25, 29 Grafted polyelectrolyte 183 Grain growth 139-140 Guinier equation 13 H-shaped copolymers 31 Halperin model 37, 175, 185 Hard gel 117, 124-126
Index Hard spheres 14, 16, 51, 105, 124, 162, 188 Heterogeneity mode 159-160 Hockey puck micelle 90 Hot gelation 126 Hydrogels 206-207 Hydrogen bonding 74-75, 79, 87, 126, 127, 206 Image charge effect 179 Intermicellar interactions 51-52, 158 Internal mode 57, 151-160 lonically end-capped copolymers 183-184 Jamming 129 Janus micelles 66, 71, 260 Langmuir isotherm 16, 223, 226 Large amplitude oscillatory shear (LAOS) 156-157 Layer sliding 156, 158 Leibler model 37-40, 143 Looping entropy 40 Lower critical solution temperature (LCST) 254 Lyotropic mesophase 105-117, 269 Manning condensation 186 Maxwell-Voigt model 118, 123 Mean field lattice model 115, 147, 191, 234 Membrane protein 263-264 Mesoporous material 269-274 Metal complexation 257-258 Micellar fusion 55-56 Micelle dimensions 27, 28, 33-37, 39, 41-44, 47-50 Micellization kinetics 52-56 Micellization thermodynamics 22-25, 126 Microemulsion 246 Mixed micelles 75-76, 139 Mode-coupling theory 162-163 Monte Carlo simulations 44-45, 149, 234-236 Multiblock ionomer 194 Nagarajan-Ganesh model 43, 49-50, 244 Nanocage 69 Nanocapsule 263 Nanocasting 270
Index Nanodroplet 87-88 Nanoreactor 255-258 Nanotube 260, 264 Nanowire 259-261 Neutron reflectivity 18 Neutron spin echo 59-60 Newtonian flow 120 Nuclear magnetic resonance (NMR) 10-11, 57-60, 164 Octopus structure 82 Onion micelle 63, 70, 79, 116, 198, 243 Order-disorder transition 106, 132-135 Order-order transition 106, 134, 135-143 Ordering kinetics 139-140 Oscillatory shear 122-123, 150, 152, 154-156 Osmotic brush 173, 185, 191, 221 Ozonolysis 69 Parallel orientation 150 Partition coefficient 242-243 Pearl necklace 174 Peptosome 265 Percolation transition 52, 119, 162, 253 Perpendicular orientation 150 pH-responsiveness 132, 173, 179-183, 204, 251, 265 Phase cube 110 Photon correlation spectroscopy, see Dynamic light scattering Pincus regime 173, 191 Pluronic 18-22, 24-27, 32-33, 42-43, 52, 53, 72, 75, 76-78, 86-87, 110-117, 126, 127-128, 131, 139, 146-147, 150-151, 154, 156, 161, 183, 242, 244, 247, 251, 253, 267, 268, 269, 272, 275 Polyampholyte 63, 182 Polydispersity 42 Polyion complex 195-198, 251 Polymer electrolyte 274 Polymersome 83, 87, 261 PRISM theory 146 Quenched polyelectrolyte 173, 193
287 Rheology 11, 117-124 Rod-coil copolymer 66-68, 174, 272 Rod-like micelle 66-68, 71-74, 151-152, 259-261 Salt effect on micellization 32-33, 127, 182, 184-186, 191-194, 197, 234 Salted brush 163, 185, 221, 225 Scaling theory 33-37 Scanning probe microscopy 11-12 Schizophrenic micellization 179-181 Self-consistent mean field theory 40-43, 51, 136, 141, 143, 146-147, 217-218, 22, 234, 246 Self-diffusion coefficient 11 Shape transition 72-74, 177 Shear alignment 86-87, 149-159 Shear banding 156 Shear melting 158 Shell cross-linked knedel (SCK) 68-71, 231 Slip-stick mechanism 153, 158 Slow release system 250 Small-angle neutron scattering (SANS) 12-14 Small-angle x-ray scattering (SAXS) 12-14 Soft gel 117 Soft lithography 272, 275 Sol-gel transition 119, 126, 127, 130, 135, 255 Solubilization 42, 75, 128, 241-245 Solvent distribution 142-143 Solvent effect on micellization 32 Stabilization 246-247 Static light scattering (SLS) 14-16 Sticky hard sphere model, see Baxter model Stokes-Einstein equation 9 Stopped flow experiments 52, 54 Stress plateaux 156-157 Stress relaxation 120-122 Structure factor 13-14, 51-52, 72, 114, 119, 120, 134 Styrenic block copolymer 20 Sugar-coated micelle 70, 252-253 Supercritical CO2 247 Superstrong segregation 47, 64, 177 Surface activity 21-22 Surface forces experiments 231-234
288 Surface micelle 218-219, 221, 224, 226-231, 236 Surface plasmon resonance 223, 229 Surface pressure-area isotherm 16, 215-220 Surface quasi-elastic light scattering 220 Surface tensiometry 16 Surface tension 21-22, 178 Swelling 142 Switch peptide 204 Tapered block copolymer 31-32 Telechelic 4, 60-62, 129-132, 194, 220 definition 4 Ternary phase diagram 111, 115 Tethered chains 45, 90-91, 215 Tetronic 123 T-jump experiments 52-55 Thixotropic stress decay 158 Time-stain separability 122 Toroidal structure 83, 193 Trajectory map 106-110, 143 Transmission electron microscopy (TEM) 7 Tube inversion test 117, 135 Tube theory 120
Index Tubules 79, 83 Ultrasonic absorption 55 Vesicles 68, 79, 83-90, 150-151, 261-268 Virial coefficient 9, 52, 57, 194 Viscometry 17 Viscosity modifier 91 Vitamin A 207 Wilhelmy plate 16-17 Wormlike micelles 64-65, 71-74, 79-80 X-ray photon correlation spectroscopy (XPCS) 59 X-ray reflectivity 17-18 Y junction 80-81 Yield stress 11, 117, 120 Zhulina-Birshtein model 35-37, 50, 175, 185 Zwitterionic copolymer 181, 183