SILVER IN ORGANIC CHEMISTRY
SILVER IN ORGANIC CHEMISTRY
Edited by
MICHAEL HARMATA
Copyright Ó 2010 by John Wiley ...
275 downloads
2607 Views
9MB Size
Report
This content was uploaded by our users and we assume good faith they have the permission to share this book. If you own the copyright to this book and it is wrongfully on our website, we offer a simple DMCA procedure to remove your content from our site. Start by pressing the button below!
Report copyright / DMCA form
SILVER IN ORGANIC CHEMISTRY
SILVER IN ORGANIC CHEMISTRY
Edited by
MICHAEL HARMATA
Copyright Ó 2010 by John Wiley & Sons, Inc. All rights reserved Published by John Wiley & Sons, Inc., Hoboken, New Jersey Published simultaneously in Canada No part of this publication may be reproduced, stored in a retrieval system, or transmitted in any form or by any means, electronic, mechanical, photocopying, recording, scanning, or otherwise, except as permitted under Section 107 or 108 of the 1976 United States Copyright Act, without either the prior written permission of the Publisher, or authorization through payment of the appropriate per-copy fee to the Copyright Clearance Center, Inc., 222 Rosewood Drive, Danvers, MA 01923, (978) 750-8400, fax (978) 750-4470, or on the web at www.copyright.com. Requests to the Publisher for permission should be addressed to the Permissions Department, John Wiley & Sons, Inc., 111 River Street, Hoboken, NJ 07030, (201) 748-6011, fax (201) 748-6008, or online at http://www.wiley.com/go/permission. Limit of Liability/Disclaimer of Warranty: While the publisher and author have used their best efforts in preparing this book, they make no representations or warranties with respect to the accuracy or completeness of the contents of this book and specifically disclaim any implied warranties of merchantability or fitness for a particular purpose. No warranty may be created or extended by sales representatives or written sales materials. The advice and strategies contained herein may not be suitable for your situation. You should consult with a professional where appropriate. Neither the publisher nor author shall be liable for any loss of profit or any other commercial damages, including but not limited to special, incidental, consequential, or other damages. For general information on our other products and services or for technical support, please contact our Customer Care Department within the United States at (800) 762-2974, outside the United States at (317) 572-3993 or fax (317) 572-4002. Wiley also publishes its books in a variety of electronic formats. Some content that appears in print may not be available in electronic formats. For more information about Wiley products, visit our website at www.wiley.com. Library of Congress Cataloging-in-Publication Data: Silver in organic chemistry / [edited by] Michael Harmata p. cm. Includes bibliographical references and index. ISBN 978-0-470-46611-7 (cloth) Printed in Singapore 10987654321
This volume is dedicated to the memories of two outstanding chemists, Dr. Christopher R. Schmid and Dr. Anthony J. Shuker, both of whom succumbed to cancer at an all too early age. Their legacies live on not only in their science but also in those whom they loved, befriended, and inspired.
CONTENTS
Foreword
xv
Preface
xvii
Contributors
xix
1
Silver Alkyls, Alkenyls, Aryls, and Alkynyls in Organic Synthesis
1
Rebecca H. Pouwer and Craig M. Williams
1.1 1.2
Introduction, 2 Csp3-Ag, 2 1.2.1 Synthesis, Stability, and Reactivity of Alkylsilver Compounds, 2 1.2.2 Synthesis and Stability of Perfluoroalkylsilver Compounds, 7 1.2.3 Reactivity of Perfluoroalkylsilver Compounds, 9 1.3 Csp2-Ag, 10 1.3.1 Synthesis and Stability of Arylsilver Compounds, 10 1.3.2 Reactivity of Arylsilver Compounds, 11 1.3.3 Synthesis and Stability of Perfluoroarylsilver Compounds, 12 1.3.4 Reactivity of Perfluoroarylsilver Compounds, 13 1.3.5 Synthesis, Stability, and Reactivity of Alkenylsilver Compounds, 13 1.3.6 Synthesis and Reactivity of Allenylsilver Compounds, 15 1.3.7 Synthesis of Perfluoroalkenylsilver Compounds, 16 1.3.8 Reactivity of Perfluoroalkenylsilver Compounds, 17 vii
viii
CONTENTS
1.3.9
Synthesis and Reactivity of Silver-Substituted Diazomethyl Compounds, 17 1.4 Csp-Ag, 18 1.4.1 Synthesis of Silver Acetylides, 19 1.4.2 Reactivity of Silver Acetylides, 20 1.4.2.1 Addition to Activated Carbonyls and Iminium Ions, 20 1.4.2.2 Nucleophilic Substitution of Activated Heteroaromatics, 24 1.4.2.3 Reaction with Alkyl Halides, 25 1.4.2.4 Coupling Reactions, 27 1.4.2.5 Reactions with Non-carbon Electrophiles, 30 1.4.2.6 Fragmentation, 32 1.4.2.7 Desilylation, 32 1.5 Conclusion, 36 References, 37 2
Cycloaddition Reactions
43
Alex M. Szpilman and Erick M. Carreira
2.1 2.2 2.3
Introduction, 44 [2þ2] Cycloadditions, 44 [3þ2] Cycloadditions, 46 2.3.1 [3þ2] Cycloadditions of Azomethine Ylides, 47 2.3.1.1 Discovery and Development of the Silver-Catalyzed [3þ2] Cycloaddition of Azomethine Ylides, 47 2.3.1.2 Auxiliary-Based Asymmetric [3þ2] Cycloadditions, 50 2.3.1.3 Catalytic Asymmetric [3þ2] Cycloadditions, 58 2.3.1.4 Selected Applications and Extensions of Azomethine [3þ2] Cycloadditions, 66 2.3.2 Other [3þ2] Cycloadditions, 71 2.4 [3þ3] Cycloadditions, 73 2.5 [4þ2] Cycloadditions, 74 2.6 Concluding Remarks, 78 References, 79 3
Sigmatropic Rearrangements and Related Processes Promoted by Silver Jean-Marc Weibel, Aurelien Blanc, and Patrick Pale
3.1 3.2 3.3
Introduction, 84 Wolff and Arndt–Eistert Rearrangements and Related Reactions, 84 Ring Rearrangements, 86
83
ix
CONTENTS
3.3.1 Halogenoamines, 86 3.3.2 Cyclopropane Derivatives, 88 3.3.3 Cubane Derivatives, 92 3.3.4 Halogenocyclopropane Derivatives, 93 3.4 [3,3]-Sigmatropic Rearrangements, 95 3.4.1 With Acyl as Migrating Groups, 95 3.4.2 With Vinyl as Migrating Groups, 98 3.4.3 With Migrating Groups Analogous to Acyl, 101 3.4.4 [3,3]-Sigmatropic Rearrangement and Cyclization Cascades, 102 3.5 [2,3]-Sigmatropic Rearrangements, 107 3.6 [1,2]-Sigmatropic Rearrangements, 108 3.6.1 1,2-Aryl or Alkenyl Migration, 108 3.6.2 1,2-Alkyl Migration, 110 3.6.3 1,2- or 1,5-Alkyl Migration, 110 3.6.4 1,2 versus 3,3 Migrations, 111 3.7 Miscellaneous, 113 3.8 Conclusion, 113 References, 114 4
Silver(I)-Mediated Electrocyclic Processes
117
Tina N. Grant and Frederick G. West
4.1
Introduction, 117 4.1.1 Ring-Opening Reactions of Halocyclopropanes, 118 4.1.2 Silver(I)-Assisted Ring-Opening Reactions, 120 4.2 Nucleophilic Trapping of Cationic Intermediates, 121 4.2.1 Solvolysis Reactions, 121 4.2.2 Intramolecular Trapping with Heteronucleophiles, 124 4.2.3 Diastereoselective Reactions, 127 4.2.4 Carbon–Carbon Bond Formation, 129 4.3 The Silver(I)-Promoted Nazarov Reaction, 132 4.3.1 Development and Initial Findings, 133 4.3.2 Interrupted Nazarov Reactions, 135 4.4 Concluding Remarks, 139 References, 139 5
Silver-Catalyzed Cycloisomerization Reactions Philippe Belmont
5.1 5.2 5.3 5.4 5.5
Introduction, 143 Cycloisomerization Cycloisomerization Cycloisomerization Cycloisomerization
of C¼O onto of C¼O onto of C¼N onto of C¼N onto
C¼C¼C, 144 C:C, 148 C¼C¼C, 152 C:C, 153
143
x
CONTENTS
5.6 Ene–Yne Cycloisomerization: C¼C onto C:C, 157 5.7 Other Transformations, 160 5.8 Conclusion, 162 References, 162 6
Silver-Catalyzed Nitrene Transfer Reactions
167
Zigang Li, David A. Capretto, and Chuan He
6.1 6.2
Introduction, 167 Aziridination, 169 6.2.1 Chloramine-T as Nitrene Precursor, 169 6.2.2 Iminoiodanes as Nitrene Precursors, 169 6.2.3 Heterogenous Silver Catalysis, 172 6.3 Sulfide and Sulfoxide Imination, 172 6.4 Amidation, 173 6.4.1 Intramolecular Amidation, 173 6.4.2 Intermolecular Amination with Phenanthroline Ligands, 174 6.4.3 Intermolecular Amination Based on Pyrazolylborate Ligands, 177 6.5 Conclusion, 180 References, 180 7
Silver-Catalyzed Silylene Transfer Tom G. Driver
7.1 7.2
Introduction, 183 Reactivity and Attributes of Metal Silylenoids and Silylmetal Complexes, 184 7.2.1 Synthesis of Transition Metal Complexes of Silylenes, 184 7.2.2 Reactivity of Transition Metal Silylenoids, 187 7.2.3 Transition Metal Silylenoid Complex–Catalyzed Hydrosilation Reactions, 187 7.2.4 Transition Metal Silylenoid–Catalyzed Atom Transfer Reactions, 189 7.3 Silacyclopropanes as Important Synthetic Intermediates, 190 7.4 Silver-Mediated Transfer of Di-tert-Butylsilylene to Olefins, 192 7.5 Silver-Mediated Transfer of Di-tert-Butylsilylene to Acetylenes, 200 7.6 Silver-Mediated Transfer of Di-tert-Butylsilylene to Carbonyl Compounds, 207 7.7 Silver-Mediated Transfer of Di-tert-Butylsilylene to Imines, 214 7.8 Silver-Mediated Di-tert-Butylsilylene Insertion into C--O Bonds, 219 7.9 Conclusion, 221 References, 222
183
xi
CONTENTS
8
Silver Carbenoids
229
Carl J. Lovely
8.1 8.2 8.3
Introduction, 229 Wolff Rearrangement, 230 Carbene Transfer Reactions to p Bonds, 232 8.3.1 Aziridination, 232 8.3.2 Cyclopropanation, 233 8.4 Formation and Reactions of Ylides, 234 8.4.1 C--Hal Addition–Rearrangement, 234 8.4.2 C--S Addition–Rearrangement, 241 8.5 C--H Insertion, 242 8.6 N--H Insertion, 243 8.7 Ring Expansion Reactions, 250 8.8 Intermediacy of Silver Carbenes, 250 8.9 Miscellaneous Reactions Involving Silver Carbenoids, 253 8.10 Summary, 254 Acknowledgments, 255 References, 255
9
Aldol and Related Processes
259
Masanori Kawasaki and Hisashi Yamamoto
9.1 Introduction, 259 9.2 Allylation Reaction Using Allyltributyltin, 260 9.3 Allylation Reaction Using Allylsilanes, 264 9.4 Aldol Reaction Using Tin Enolates, 268 9.5 Aldol Reaction Using Silyl Enol Ethers 271 9.6 Mannich Reaction, 276 9.7 Nitrosoaldol Reaction, 277 9.8 Aldol Reaction with Azodicarboxylate, 281 9.9 Conclusion, 281 References, 282 10
Coupling Reactions Promoted by Silver Jean-Marc Weibel, Aurelien Blanc, and Patrick Pale
10.1 10.2 10.3 10.4 10.5
Introduction, 286 sp3–sp3 Coupling Reactions Promoted by Silver Salts, 286 sp3–sp2 Coupling Reactions Promoted by Silver Salts, 289 sp3–sp Coupling Reactions Promoted by Silver Salts, 290 sp2–sp2 Coupling Reactions Promoted by Silver Salts, 291 10.5.1 Homocoupling of Vinyl- or Arylsilver Species, 292 10.5.2 Organosilver Species as Nucleophilic Reagents, 293 10.5.3 Silver as a Lewis Acid Reagent, 294 10.5.4 Silver as a Halogen Scavenger, 297 10.5.4.1 Silver in Pd-Catalyzed Couplings, 298
285
xii
CONTENTS
Silver in PdII-Promoted Electrophilic Substitution of Arenes (C--H Activation), 306 10.5.4.3 Silver as Reagent for Decarboxylative Coupling, 309 10.6 sp2–sp Coupling Reactions Promoted by Silver Salts, 310 10.6.1 Organosilver Species as Nucleophilic Reagents, 311 10.6.2 Organosilver Species in Transmetallations, 314 10.6.3 Silver as a Lewis Acid Reagent, 315 10.6.4 Organosilver Species as Intermediates in Catalyzed Enyne or Arylyne Synthesis, 316 10.7 sp–sp Coupling Reactions Promoted by Silver Salts, 322 10.8 Conclusion, 323 References, 324 10.5.4.2
11
Supramolecular Chemistry of Silver
329
Wei-Yin Sun, Zheng-Shuai Bai, and Jin-Quan Yu
11.1 Introduction, 329 11.2 Cage-Like Complexes, 330 11.3 Tube-Like Compounds, 337 11.4 Polycatenanes with Silver(I), 339 11.5 Polyrotaxanes with Silver(I), 342 11.6 Silver(I) Coordination Polymers with Specific Topology, 345 11.7 Conclusion, 350 Acknowledgments, 352 References, 352 12
A Critical Comparison: Copper, Silver, and Gold A. Stephen K. Hashmi
12.1 12.2
Introduction, 358 Reactions Catalyzed by Copper, Silver, or Gold, 358 12.2.1 Aldehyde–Alkyne–Amine Coupling, 358 12.2.2 Carbene Insertion Reactions, 360 12.2.3 In Silico Comparison of Organocopper(I), Organosilver(I), and Organogold(I) -Ate Compounds, 361 12.2.4 Cyclization of ortho-alkynylbenzaldehydes, 362 12.2.5 Allenyl Ketones: The Cycloisomerization to Furans, 362 12.2.6 A Thiol in the Substrate: The Cyclization of a-Thioallenes, 364 12.2.7 Furans by Propargyl Claisen Reaction, 365 12.2.8 Tandem Cyclization/Pinacol Rearrangement, 366 12.2.9 Furanones by Domino Heterocyclization/1,2 Shift, 366 12.2.10 Conia-ene Reaction, 368 12.3 Reactions Catalyzed by Silver or Gold, 368 12.3.1 Cyclization of N-Propargylcarboxamides, 368
357
xiii
CONTENTS
12.3.2 12.3.3 12.3.4 12.3.5 12.3.6 12.3.7
Dake’s Pyrrole Synthesis, 369 Combination with Organocatalysis, 369 Vinylallenes Deliver Cyclopentadienes, 370 a-Pyrones by a Cascade Reaction, 371 Dihydrofurans from Propargyl Esters, 371 Methylene Butyrolactones by Addition of Carboxylates to Alkynes, 372 12.3.8 Hydroarylation of Allenes, 373 12.3.9 Different Products by Silver and Gold Catalysts, 373 12.3.9.1 The Epoxide–Alkyne Reaction, 373 12.3.9.2 The Carbonyl–Alkyne Reaction, 374 12.4 Reactions Catalyzed by Copper or Silver, 374 12.4.1 General Trends, 374 12.4.2 Pyrroles by Hydroamination, 374 12.4.3 Copper/Silver Cocatalysis, 375 12.4.4 Carbonylations, 375 12.5 Conclusion, 376 References, 376 Index
381
FOREWORD
In the last two centuries, the discipline of synthesis has profoundly transformed our world, enabling access to molecules that in former times would be only scarcely or unreliably available from natural sources. Increasingly, synthesis is also being used to access new molecules, designed for function (e.g., catalysts, smart materials, self replicating materials, molecular devices, energy generation and storage systems, diagnostics, drug delivery systems, therapeutics)—many with activities superior to or often different from what nature has produced. We are no longer exclusively reliant on nature for our molecular needs. This too brings new opportunities. Whereas once the challenge in synthesis was simply to make molecules, increasingly that challenge has given away to a more demanding goal: developing strategies that provide molecules in a step economical, and green, if not ideal way. Our ability to meet this goal rests heavily not only on the refinement of existing methodology, but also on the introduction of new reactions and reagents that enable or enhance new synthetic strategies—a focus of this book. This book explores the use of silver in organic synthesis. Silver and its salts and complexes have figured significantly in the history of chemistry, recognized for their special conductive properties, use in photography, and even biological activities. Notwithstanding the importance of these areas the broader use of silver in chemistry, and more specifically in synthesis, has lagged behind that of other coinage metals. That is changing. One now finds silver as a key component of much that is “nano,” including nano-rods, spheres, sheets, clusters, prisms, membranes, plates, pillars, cubes, bowls, fibers, wires, gels, and sensors. Increasing interest is also being directed at its use and that of other coinage metals in improving synthetic procedures and in enabling new ones. This book provides an insightful overview of how silver figures in these new developments. xv
xvi
FOREWORD
Professor Harmata is one of the gifted educators of our time. Through his research and books he has contributed significantly to the advancement of synthesis. For this book, he has assembled a remarkable team of thought leaders who have in their own research contributed significantly to the emerging interest in silver-based reaction science. The resultant product is a must read for those interested in synthesis. It spans impressively from the preparation and use of silver compounds to silver-catalyzed or mediated cycloadditions, rearrangements, isomerizations, group transfers, aldols, and coupling reactions to supramolecular chemistry and comparisons with other metals. It is both an educational and inspirational experience. It has impressive depth and breadth. This contribution to our community sprung in part from frustration with a rejected but clever manuscript title (“All that glitters is not gold”) and the resultant motivation “to do something on behalf of silver.” There is a silver lining to that cloud, as this book on silver in organic chemistry represents a brilliant contribution to the field and an educational experience that is expected to inspire new ideas and glitter for an emerging area of interest. PAUL A. WENDER Stanford University April 2010
PREFACE
It was a dark and stormy night. . .. Editors get to have some fun, don’t they? This book was born out of the recognition that there existed no compilation on the power of silver in organic chemistry, particularly synthesis. I recognized this, and within less than a year, while these reviews were being written, a very nice Chemical Reviews issue appeared dedicated to the coinage metals and their importance to organic chemistry. That’s life! Such is the pace of developments in the area of coinage metals that those reviews, and those contained herein, will need to be updated within the next few years, however. Have I just suggested that I might take on a second edition of this monograph? I must be nuts. This book also came about because I am at times pigheaded and not the teddy bear that I am often perceived to be. Not too long ago, I tried to publish a paper that was partially entitled “All that glitters is not gold,” in an effort to do some cheerleading for the silver cation. A referee thought this was an abomination, and my response was less than that of a gentleman and scholar. Fortunately, cooler heads eventually prevailed, the situation was resolved, and the paper was published: I changed the title. However, I was left with the feeling that I needed to do something on behalf of silver, and this book is the result. My thanks go out to the authors. Through their fine efforts, a very nice monograph has been produced. If this monograph teaches and inspires, even just a little, we will have accomplished our mission. I must thank Wiley and all the fine folks there for their help and support. My thanks go in particular to Ms. Anita Lekhwan, whose confidence in me and the project never waivered. We all need people to believe in us. My family has been very patient with me as I put in the extra effort to bring this book to life. My deepest thanks to Judy, Gail, Diana, and Alexander. xvii
xviii
PREFACE
Finally, whenever I do a project like this, I like to remind the community that I can make time for this because I have a supported research program. When I began this project, I had both NIH and NSF funding. I will retain the latter for the next few years and hopefully regain the former. A synergistic activity like this allows me to produce a teaching and learning tool and affords me the chance to interact with leading colleagues of the day. Hopefully it adds something to the community; it certainly enriches me. MICHAEL HARMATA University of Missouri – Columbia April 2010
CONTRIBUTORS
Zheng-Shuai Bai, Coordination Chemistry Institute, State Key Laboratory of Coordination Chemistry, School of Chemistry and Chemical Engineering, Nanjing National Laboratory of Microstructures, Nanjing University, Nanjing 210093, China Philippe Belmont, Equipe “Chimie Organome´tallique, He´te´rocycles et Cibles Biologiques” (COHCB), Institut Curie et CNRS, 26 rue d’Ulm, 75005 Paris, France Aurelien Blanc, Laboratoire de Synthese et Reactivite Organiques, Institut de Chimie, Universite de Strasbourg, 4 Rue Blaise Pascal, 67000 Strasbourg, France David A. Capretto, Department of Chemistry, University of Chicago, 5735 South Ellis Avenue, Chicago, IL 60637, USA Erick M. Carreira, Laboratory of Organic Chemistry, ETH Zu¨rich, HCI H335, Wolfgang-Pauli-Strasse 10, 8093 Zu¨rich, Switzerland Tom G. Driver, Department of Chemistry, University of Illinois at Chicago, 845 W. Taylor St., Chicago, IL 60607, USA Tina N. Grant, Department of Chemistry, University of Alberta, 11227 Saskatchewan Drive, T6G 2G2 Edmonton, Canada A. Stephen K. Hashmi, Organisch-Chemisches Institut, Ruprecht-Karls-Universita¨t, Heidelberg, Im Neuenheimer Feld 270, 69120 Heidelberg, Germany Chuan He, Department of Chemistry, University of Chicago, 5735 South Ellis Avenue, Chicago, IL 60637, USA xix
xx
CONTRIBUTORS
Masanori Kawasaki, Medicinal Chemistry Research Laboratories, Pharmaceutical Research Division, Takeda Pharmaceutical Company Limited, 17-85, Jusohonmachi 2-chome, Yodogawa-ku, Osaka 532-8686, Japan Zigang Li, Department of Chemistry, University of Chicago, 5735 South Ellis Avenue, Chicago, IL 60637, USA Carl J. Lovely, Department of Chemistry and Biochemistry, The University of Texas at Arlington, 700 Planetarium Place, Arlington, TX 76019, USA Patrick Pale, Laboratoire de Synthese et Reactivite Organiques, Institut de Chimie, Universite de Strasbourg, 4 Rue Blaise Pascal, 67000 Strasbourg, France Rebecca H. Pouwer, School of Chemistry and Molecular Biosciences, University of Queensland, Brisbane 4072, Queensland, Australia Wei-Yin Sun, Coordination Chemistry Institute, State Key Laboratory of Coordination Chemistry, School of Chemistry and Chemical Engineering, Nanjing National Laboratory of Microstructures, Nanjing University, Nanjing 210093, China Alex M. Szpilman, Schulich Faculty of Chemistry, Technion—Israel Institute of Technology, 32000 Technion City, Haifa, Israel Jean-Marc Weibel, Laboratoire de Synthese et Reactivite Organiques, Institut de Chimie, Universite de Strasbourg, 4 Rue Blaise Pascal, 67000 Strasbourg, France Frederick G. West, Department of Chemistry, University of Alberta, 11227 Saskatchewan Drive, T6G 2G2 Edmonton, Canada Craig M. Williams, School of Chemistry and Molecular Biosciences, University of Queensland, Brisbane, 4072 Queensland, Australia Hisashi Yamamoto, Department of Chemistry, University of Chicago, 5735 South Ellis Avenue, Chicago, IL 60637, USA Jin-Quan Yu, Department of Chemistry, The Scripps Research Institute, La Jolla, CA 92037, USA
1 SILVER ALKYLS, ALKENYLS, ARYLS, AND ALKYNYLS IN ORGANIC SYNTHESIS REBECCA H. POUWER
AND
CRAIG M. WILLIAMS
School of Chemistry and Molecular Biosciences, University of Queensland, Brisbane, Australia
1.1 Introduction 1.2 Csp3-Ag 1.2.1 Synthesis, Stability, and Reactivity of Alkylsilver Compounds 1.2.2 Synthesis and Stability of Perfluoroalkylsilver Compounds 1.2.3 Reactivity of Perfluoroalkylsilver Compounds 1.3 Csp2-Ag 1.3.1 Synthesis and Stability of Arylsilver Compounds 1.3.2 Reactivity of Arylsilver Compounds 1.3.3 Synthesis and Stability of Perfluoroarylsilver Compounds 1.3.4 Reactivity of Perfluoroarylsilver Compounds 1.3.5 Synthesis, Stability, and Reactivity of Alkenylsilver Compounds 1.3.6 Synthesis and Reactivity of Allenylsilver Compounds 1.3.7 Synthesis of Perfluoroalkenylsilver Compounds 1.3.8 Reactivity of Perfluoroalkenylsilver Compounds 1.3.9 Synthesis and Reactivity of Silver-Substituted Diazomethyl Compounds 1.4 Csp-Ag 1.4.1 Synthesis of Silver Acetylides 1.4.2 Reactivity of Silver Acetylides 1.4.2.1 Addition to Activated Carbonyls and Iminium Ions 1.4.2.2 Nucleophilic Substitution of Activated Heteroaromatics 1.4.2.3 Reaction with Alkyl Halides 1.4.2.4 Coupling Reactions 1.4.2.5 Reactions with Non-carbon Electrophiles
Silver in Organic Chemistry Edited by Michael Harmata Copyright 2010 John Wiley & Sons, Inc.
1
2
SILVER ALKYLS, ALKENYLS, ARYLS, AND ALKYNYLS IN ORGANIC SYNTHESIS
1.4.2.6 Fragmentation 1.4.2.7 Desilylation 1.5 Conclusion References
1.1
INTRODUCTION
While the coordination and inorganic chemistry of silver compounds have been prolifically documented, the use of organosilver compounds to effect useful synthetic transformations is severely underrepresented in the synthetic organic chemistry literature. This has prompted us to present a review of literature reporting synthetically useful applications of organosilver compounds in the hope of inspiring further development in this field. The majority of the literature covered in this review concentrates on silver(I) organo-species as reagents, although on some occasions silver(II) and silver “ate” complexes will be discussed, in addition to organosilver intermediates. General reviews encompassing all classes of organosilver compounds have appeared previously.1–3 1.2 1.2.1
Csp3 -Ag Synthesis, Stability, and Reactivity of Alkylsilver Compounds
As a result of extremely low thermal stability, alkylsilver compounds have found only a narrow range of use in organic synthesis. Procedures for the synthesis of alkylsilver compounds as anything but fleeting proposed intermediates are limited to a handful. Semerano and Riccoboni first reported the synthesis of methyl-, ethyl-, and propylsilver in 1941 (Scheme 1.1). Reaction of silver nitrate and the corresponding tetralkyllead in alcohol at 80 C gave the compounds as brown precipitates that decomposed rapidly on warming to room temperature to give metallic silver and a mixture of hydrocarbons.4 This methodology has been utilized in a limited number of investigations into the mechanism of decomposition of alkylsilver compounds.5,6 In these cases, the presence of the alkylsilver compound, and its subsequent decomposition, is inferred from the isolation of alkyl dimers. Two plausible mechanistic pathways have been proposed for the thermal decomposition of alkylsilver compounds: either a radically-mediated cleavage of the carbon–silver bond or a process by which the breaking of the silver–carbon bond and formation of the carbon–carbon bond are concerted. Mechanistic studies by Whitesides and coworkers in which the product ratios obtained for the thermal process R4Pb
AgNO3
RAg
Scheme 1.1
R3PbNO3
3
Csp3-Ag
TABLE 1.1. Silver-Catalyzed Dimerization of Alkylmagesium Halides BrCH2CH2Br, AgOTs (1 mol%)
R–MgX
Entry
Substrate
R–R
Product
Yield (%)
MgBr
1
97 10
MgBr
O
2
O O
O
O
80
O MgBr
3
99
14
6
MgBr
4
95
14
Ph
Ph
Ph
were compared to those for known radically-mediated reactions have suggested that a concerted process is more likely, although this has not proved to be general.7–9 The formation of methylsilver and dimethylargentate has been observed in the collision-induced dissociation MS3 spectrum of silver diacetate. Dimethylargentate is stable in the gas phase, and has been isolated for short periods (10 s) without significant decomposition.10 Alkylsilver compounds have been prepared by treatment of Grignard reagents with silver salts,11–19 and similarly undergo oxidative homocoupling to give alkyl dimers.11–13,19,20 Exploitation of this finding has resulted in the development of general methodology for silver-catalyzed alkyl–alkyl homocoupling of Grignard reagents (Table 1.1).21 The catalytic cycle of this reaction is proposed to proceed via the oxidation of metallic silver with 1,2-dibromoethane to generate silver bromide (Scheme 1.2). RMgX
AgOTs
1 – R–R 2
R–Ag RMgX
AgBr
1 – C2H4 2
Scheme 1.2
Ag
1 – BrCH CH Br 2 2 2
4
SILVER ALKYLS, ALKENYLS, ARYLS, AND ALKYNYLS IN ORGANIC SYNTHESIS
TABLE 1.2. Silver-Mediated Ring-Closing Reaction of Bis(Alkylmagnesium Halides) 1. Mg, THF
X
Entry 1 2 3 4 5 6 7 8
2. [IAgPBu3]4
X
n
n
Concentration N 102
Substrate 1,4-Dibromobutane 1,5-Dibromopentane 1,6-Dichlorohexane 1,7-Dibromoheptane 1,8-Dibromooctane 1,9-Dichlorononane 1,10-Dibromodecane 1,12-Dibromododcane
5.0 2.5 2.5 2.5 2.5 2.5 0.77 0.77
Product
Yield (%)
Cyclobutane Cyclopentane Cyclohexane Cycloheptane Cyclooctane Cyclononane Cyclodecane Cyclodecane
83 83 43 23 2 <1 10–15 10–15
Of particular note is the use of this reactivity to form small carbocycles. Whitesides and coworkers have shown that the treatment of primary bis(alkylmagnesium halides) with tributylphosphinesilver iodide produces carbocycles in a range of yields, with a strong dependence on ring size (Table 1.2). The best results were obtained for four-, five-, and six-membered rings. Although it was hoped that the aggregated nature of alkylsilver compounds would facilitate the formation of medium to large rings, compounds of this type were produced with only low yields.8 It has also been shown that treatment of primary bis(alkylmagnesium halides) with silver trifluoromethanesulfonate effects ring closure under mild conditions for a range of substrates, thus highlighting the generality of this reaction for producing small carbocycles (Table 1.3). An equivalent reaction has been achieved via the treatment of hydroborated bisalkenes with alkaline silver nitrate solution (Table 1.4).22,23 This method has been used to synthesize a number of small and medium-size carbocyclic rings in moderate to good yield. The selectivity for terminal cyclization observed for 1,6heptadiene and 1,7-octadiene indicates that, in these cases, hydroboration of each of the alkenes occurs independently to yield acyclic boranes. It has, however, been found that both cyclic and acyclic boranes react under these conditions to yield the ringclosed products (Scheme 1.3). R B
Tx AgNO3/KOH
AgNO3/KOH
54%
82%
OAc B Tx Tx = –C(CH3)2CH(CH3)2
Scheme 1.3
B
OAc
5
Csp3-Ag
TABLE 1.3. Silver-Mediated Ring-Closing Reaction of Bis(Alkylmagnesium Halides) X
n
n
Entry
Substrate
1
2
1. Mg, THF 2. AgOTf
X
Product Br
Br
94 Cl
Cl
Yield (%)
55
3
Cl Cl
82
4
Cl Cl
61
Cl
0
5 Cl
Intermolecular dimerization has also been effected by a comparable protocol.24–26 Treatment of triethylborane with silver nitrate and sodium hydroxide in water at 25 C led to the rapid evolution of n-butane (72%), ethylene (9%), and ethane (9%). Reaction of two different alkylboranes led to statistical mixtures of dimerized and cross-coupled products. Furthermore, this strategy has been used successfully in the synthesis of olefins from dihydroborated internal acetylenes,27 and in polymerizations of bifunctional organoboron compounds.28 The addition of lithium bromide significantly increases the thermal stability of alkylsilver compounds.14 Westmijze and coworkers found that the reaction of nbutylmagnesium bromide, for example, with AgBr2LiBr gave a solution of butylsilver that was stable up to 10 C, which is in stark contrast to the species obtained from the reaction with silver bromide alone, which decomposes at 60 C. This marked stabilization of the alkylsilver compounds allowed for the first meaningful use of these reagents in intermolecular reactions. The Grignard-derived alkylsilver–lithium bromide complexes have been added to conjugated enynes and enynenitriles to give allenes via the intermediacy of allenylsilver reagents (Scheme 1.4).14–18 The alkylsilver reagents generally reacted with the enynenitriles to give 2,3-alkadienenitriles on protolysis, while homoargentates (R2AgMgCl) tended to give 2,4-alkadienenitriles. The homoargentates underwent selective trans addition to the enynenitriles, which is in contrast to the selective cis addition observed for alkylcopper reagents.17
6
SILVER ALKYLS, ALKENYLS, ARYLS, AND ALKYNYLS IN ORGANIC SYNTHESIS
TABLE 1.4. Silver-Mediated Ring Closing of Hydroborated Bisalkenes 1. B2H6 2. KOH/MeOH, AgNO3
n
n
Entry
Substrate
Product
Yield (%)
1
79
2
85
3
66
17
4
67
5
42
6
19
A number of simple alkylsilver compounds have been found to react with enynylsulfinates via a 1,5 addition to give substituted butatrienes in excellent yield (Scheme 1.5).15 Interestingly, a different outcome was obtained for alkylcopper compounds, which underwent 1,3 addition to the enynylsulfinates. Kauffmann and coworkers have described the reaction of alkylsilver compounds with cyclohexenone (Scheme 1.6).29 It was found that the alkylsilver derivatives reacted preferentially, and in some cases exclusively, in the b position. A significant amount of 3-methylcyclohex-2-enol was observed for the reaction of Me2AgMgBr, which has been attributed to the elimination and subsequent reaction of silver(I) hydride. Silver(I) hydride has been observed in the gas-phase fragmentation of other organosilver compounds.30
7
Csp3-Ag
H 2C
RAg⋅2LiBr
H
1. R2AgMgCl 2. H+
R R1
H
R1
R CH2
H+
C
THF–HMPT
R1
R2
R CH2
R1
Ag⋅2LiBr
R
CN
CN
C
R1
CN
H 58–96%
R2
1. RAg 2. H+
R2
H C
R1
H 85–90%
70–90% trans addition
1. RCu 2. H+
R = alkyl R1 = alkyl, phenyl R2 = alkyl, phenyl
NC R2 R1
R
76–90% cis addition
Scheme 1.4 R2
H2C
OS(O)Me R1
R3
RAg⋅3LiBr THF–HMPT 75–95%
R2
RH2C C C R1
R3
R1 = alkyl, R2 = alkyl, R3 = alkyl R = alkyl
t-BuCu⋅LiBr THF–HMPT 90%
R2
t-Bu C
R3 R1 R2 = R3 = CH3
Scheme 1.5
Silver trinitromethane undergoes addition to adamantyl (Ad) halides,31 methyl iodide, and trinitromethyl iodide (Scheme 1.7).32 1.2.2
Synthesis and Stability of Perfluoroalkylsilver Compounds
In comparison to the corresponding alkylsilver compounds, the perfluoroalkylsilver derivatives are significantly more stable and are able to participate in chemistry not
8
SILVER ALKYLS, ALKENYLS, ARYLS, AND ALKYNYLS IN ORGANIC SYNTHESIS
O
O Ag
AgH −AgH
Me2AgMgBr
O
O
OH
R n-BuAg n-Bu2AgMgBr MeAg Me2AgMgBr
61% 60% 21% 20%
HO
R
R
43%
17%
Scheme 1.6 AgC(NO2)3
AdBr
AdC(NO2)3
AgC(NO2)3
MeI
MeC(NO2)3
AgC(NO2)3
IC(NO2)3
(NO2)3CC(NO2)3
Scheme 1.7
known for the alkylsilver derivatives.33 The perfluoroalkyl silver compounds were first synthesized by Miller and coworkers through the nucleophilic addition of silver fluoride to perfluoroalkenes (Scheme 1.8).34,35 In an attempt to avoid the use of silver fluoride, Dyatkin and coworkers developed the reaction of perfluoroalkenes with silver trifluoroacetate in the presence of cesium or potassium fluoride to produce a variety of perfluoroalkyl silver derivatives.36 Perfluoroalkylsilver compounds have also been synthesized by transmetallation of the corresponding cadmium reagents with silver nitrate.37 An interesting synthetic route to the perfluoroalkylsilver derivatives has been reported by Knunyants and coworkers with the production of perfluoroisopropylsilver in excellent yield via the decarboxylation of the silver salt of a-hydroperfluoroisobutyric acid in pyridine (Scheme 1.9).38 Tyrra has described the preparation of trifluoromethylsilver from Me3SiCF3 and silver fluoride in propionitrile (Scheme 1.10).39 On treatment with PNPCl ([PNP] þ ¼ Ph3P¼N--PPh3 þ ) trifluoromethylsilver was trapped as a mixture of argentates, F3C
F
F
F
AgF
CH3CN
Scheme 1.8
Ag–CF(CF3)2
9
Csp3-Ag
CF3 O
CF3 Ag CF3
pyridine
F
F
−CO2
CF3 OAg
100% conversion
Scheme 1.9
AgF
Me3SiCF3
AgCF3
EtCN, r.t.
Scheme 1.10
[PNP][Ag(CF3)2] and [PNP][Ag(CF3)Cl]. The latter was characterized by X-ray crystallography.33,39 1.2.3
Reactivity of Perfluoroalkylsilver Compounds
The known reactions of the perfluoroalkylsilver derivatives generally equate to nucleophilic substitution reactions and the oxidative perfluoroorganylation of group 12–16 elements.39,40 Normant and coworkers have reported the reaction of both carbon dioxide and allyl bromide with perfluoroisopropylsilver (Scheme 1.11).41 Heptafluoro-2-nitrosopropane has been prepared in good yield from perfluoroisopropylsilver by treatment with nitrosyl chloride.42 Dyatkin and coworkers have reported the reaction of perfluoroalkylsilver derivatives with iodine to give perfluoroalkyl iodides, and with elemental sulfur to give the perfluoroalkylthiosilver, which was subsequently alkylated with ethyl iodide to give the thioether (Scheme 1.12).36 Perfluoro-tert-butylmethyl sulfide was synthesized by treatment of the silver derivative with trifluoromethylsulfenyl chloride. The disulfide was also obtained from initial reaction with elemental sulfur, followed by treatment with the sulfenyl chloride (Scheme 1.13).43
Br (F3C)2FC 50%
(CF3)2CFAg
O
1. CO2 2. PhCH2Br
(F3C)2FC
OCH2Ph
60%
Scheme 1.11
10
SILVER ALKYLS, ALKENYLS, ARYLS, AND ALKYNYLS IN ORGANIC SYNTHESIS
Scheme 1.12
CF3SCI
F3C
CF3 Ag CF3
1. S0 2. CF3SCI
F3C
F3C
CF3 SCF3 CF3 57%
CF3 S SCF3 CF3 80%
Scheme 1.13
1.3 1.3.1
Csp2 -Ag Synthesis and Stability of Arylsilver Compounds
The first synthesis of an arylsilver compound was reported by Krause in 1919, who isolated a compound with the formula (PhAg)2AgNO3 as a canary yellow powder from the reaction of silver nitrate with triphenylethyllead, which in the presence of “fuming HNO3 . . . detonates with great violence and a flash of blinding white light.”44 Phenylsilver produced by this method decomposes thermally to a mixture of silver metal, benzene, and biphenyl.6 In addition, a number of procedures were reported for the synthesis of arylsilver compounds from organomagesium compounds as mixtures, contaminated with silver and magnesium salts, and with low thermal stability.45–47 Organolithium reagents react similarly, and in a very few number of cases in which the aryl was substituted with heteratoms, the arylsilver compound could be purified by means of extraction or recrystallization.48–50 It was only in 1972, more than 50 years after the first report by Krause, that a method for obtaining pure phenylsilver was achieved. Experimenting with the methodology of Krause, Beverwijk and Van der Kerk found that pure phenylsilver could be obtained as a white precipitate from the very slow addition of trialkylphenyllead or -tin compounds to ethanolic solutions of silver nitrate (Scheme 1.14).51
11
Csp2-Ag
R3MPh
AgNO3
EtOH −10−15ºC 80– 95%
PhAg
[R3MNO3]
M = Pb or Sn R = alkyl
Scheme 1.14
Ph2Zn
Et2O
AgNO3
0ºC >99%
PhAg
[PhZnNO3]
Scheme 1.15
A remarkable increase in the stability of pure phenylsilver is observed in comparison to that of the silver nitrate complex, with the compound surviving several days at room temperature under a nitrogen atmosphere. Diarylzinc compounds react with silver salts to give arylsilver compounds of high purity and stability (Scheme 1.15).52 Van der Kerk and coworkers synthesized phenylsilver and a number of methyl-substituted arylsilver compounds via this route, and found that ortho-methyl substitution significantly increased the thermal stability of the compound, as is the case for the corresponding arylcopper compounds (Table 1.5).53 As is the case for other organosilver compounds, arylsilver compounds are not monomeric. Cryoscopic molecular weight determination in benzene found phenylsilver to be polymeric and the methyl-substituted arylsilver compounds to be trimeric.53 While cryoscopy of mesitylsilver, prepared from reaction of the corresponding Grignard reagent with silver chloride, gave a molecular weight corresponding to that of a dinuclear entity, determination of the structure by X-ray crystallography determined mesitylsilver to be tetrameric in the solid state.54 1.3.2
Reactivity of Arylsilver Compounds
In a very limited number of cases phenylsilver displays reactivity typical of moderately reactive organometallic compounds.46 Gilman and coworkers reported the reaction of phenylsilver, prepared from phenylmagnesium iodide, with acetyl TABLE 1.5. Temperatures of Decomposition of Arylsilver Compounds Entry 1 2 3 4 5
R--Ag Phenyl 2-Methylphenyl 3-Methylphenyl 4-Methylphenyl 2,6-Dimethylphenyl
Decomposition Temperature ( C) 74 91 82 78 170
12
SILVER ALKYLS, ALKENYLS, ARYLS, AND ALKYNYLS IN ORGANIC SYNTHESIS
O AcCl
24% Br
Ag
30% O
PhCNO
N H Trace
Scheme 1.16
chloride to give acetophenone in 24% yield, with allyl bromide to give allylbenzene in 30% yield, and with PhNCO to give a trace of benzanilide (Scheme 1.16).46 The predominant product in all three cases was biphenyl. 1.3.3
Synthesis and Stability of Perfluoroarylsilver Compounds
As is general for organosilver compounds, perfluorophenylsilver is considerably more stable than phenylsilver.55 The first example of a perfluoroarylsilver compound was reported by Miller and Sun in 1970.49 Reaction of perfluorophenyllithium with silver trifluoroacetate, or of perfluorophenyl bromide with perfluoroisopropylsilver, gave perfluorophenylsilver as colorless crystals (Scheme 1.17). The compound was found to be stable up to 150 C, at which temperature it slowly dimerized to give perfluorobiphenyl and metallic silver. Since this initial report a number of syntheses of perfluorophenylsilver from perfluorophenyllithium followed.56,57 A new convenient synthesis of perfluorophenylsilver has been reported by Tyrra and coworkers. Treatment of trimethyl(perfluorophenyl)silane with silver fluoride in propionitrile gave perfluorophenylsilver in nearly quantitative yield (Scheme 1.18).39,58 Perfluorophenylsilver synthesized by this method, isolated as a 1 : 1 adduct with propionitrile, has been characterized by X-ray crystallography.58
C6F5Li
CF3COOAg
C6F5Br
(CF3)2CFAg
Et2O/hexane −75ºC 78%
CH3CN 25ºC 65%
Scheme 1.17
C6F5Ag
CF3COOLi
C6F5Ag
(CF3)2CFBr
13
Csp2-Ag
F
F
F
F TMS
F
AgF
F
F
EtCN, r.t.
F
Ag⋅EtCN F
F 97%
Scheme 1.18
O O R
CI C6F5
Me3SiCI
F
C6H5CH2Br
F
F
R
C6F5SiMe3
C6F5
C6F5
Ag CH3I
F
F Br2
C6H5I
C6F5CH3
C6F5Br
C6F5 C6H5
Scheme 1.19
1.3.4
Reactivity of Perfluoroarylsilver Compounds
The reactions of perfluorophenylsilver amount to the nucleophilic substitution of a variety of electrophiles and the oxidative perfluoroorganylation of group 12–16 elements (Scheme 1.19).39,40,49,55,59,60 1.3.5
Synthesis, Stability, and Reactivity of Alkenylsilver Compounds
Alkenysilver reagents have found very limited use in organic synthesis. This section focuses on the few methods where alkenylsilver reagents are synthesized with the intention of utilizing, or gaining further understanding of the nature of, the chemistry of these highly reactive species. Alkenylsilver derivatives are relatively less thermally stable than alkynylsilver derivatives, but tend to be significantly more stable than the alkylsilver derivatives. The first investigation into the thermal stability of alkenylsilver reagents was reported by Glockling, who described the synthesis of isobut-1-enylsilver from the reaction of
14
SILVER ALKYLS, ALKENYLS, ARYLS, AND ALKYNYLS IN ORGANIC SYNTHESIS
H
H
AgNO3 EtOH -78ºC
Et3Pb
H
Ag
Ag EtOH H H
H
Et3Pb
Et3PbOEt
EtO H
Scheme 1.20
isobut-1-enyltriethyllead with ethanolic silver nitrate at 78 C.61,62 An orange precipitate resulted, which was found to slowly deposit silver metal on warming to 20 C, and significant isobutene evolution was observed above 10 C. The formation of isobutene was in fact catalytic in silver, suggesting the initiation of a chain reaction whereby an isobutenyl radical may abstract a proton from ethanol, followed by reaction of the incipient ethoxyl radical with isobut-1-enyltriethyllead (Scheme 1.20). In the absence of ethanol, the alkenylsilver reagent decomposed to give dimer and metallic silver upon gradual warming of the solution. An increase in the stability of this alkenylsilver reagent can be effected through coordination. Both triethylamine and 2,20 -dipyridyl had a remarkable effect on stability, whereby deposition of metallic silver from solution occurred only slowly at room temperature.62 Styrylsilver was synthesized from triethyl-a-styryl lead as a deep red precipitate, and proved to be comparatively stable, with decomposition to silver requiring several days at room temperature.5 Unlike isobut-1-enylsilver, the formation of dimer (1,4diphenylbutadiene) was not observed; instead, an insoluble polymeric substance was formed. Alkenylsilver compounds have been synthesized by treating vinyllithium reagents with silver salts. Both cis-1-propenylsilver and its corresponding tributylphosphine adduct were synthesized by this method for the purposes of investigating the outcome of the thermal decomposition of vinylsilver compounds. In both cases the product of dimerization predominated with almost complete retention of stereochemistry, thus suggesting that, as for the alkylsilver compounds, long-lived free propenyl radicals are not intermediates in this reaction (Scheme 1.21).63,64 Kochi and Tamura report a similar finding whereby treatment of cis-propenylmagnesium bromide and methyl
Li
[IAgPBu3]4
AgPBu3
r.t.
Ag
–78ºC
97% yield 96.5% cis,cis
Scheme 1.21
PBu3
15
Csp2-Ag
Li
AgPBu3
IAgPBu3
I
I2
~80% 96% cis
Scheme 1.22
Ph
Li
Ph
Cl
AgCl
Ph Ph
CH3OH Ph Ph
Cl
∆
Ph
Ph
Cl
Ph
Ph Cl ~20%
Ph Ph
Ph C C
Ph ~40%
Ph ~40%
Br2 l
Ph
Cl
Ph
64%
H Ph
I2
H
73%
Ag
Br Cl 77%
Scheme 1.23
bromide with catalytic amounts of silver(I) salt produced cis-but-2-ene in almost quantitative yield.19 The tributylphosphine adduct was found to react with iodine to give cis-1iodopropene (Scheme 1.22).64 1-Chloro-2,2-diphenylvinylsilver has been produced at low temperatures by treating the lithiated derivative with silver chloride.65 The compound decomposed at room temperature to give 1-chloro-2,2-diphenylethene, diphenylacetylene, and tetraphenylbutatriene. The silver derivative does not react with carbon dioxide, but gave the expected products of electrophilic substitution when treated with methanol, iodine, and bromine (Scheme 1.23). 1.3.6
Synthesis and Reactivity of Allenylsilver Compounds
Allenylsilver reagents have been prepared by the reaction of allenyllithium reagents with silver(I) salts,66,67 and by the addition of alkylsilver reagents to conjugated enynes (see alkylsilver section 1.2.1 above), and have been added to a number of electrophiles in high yield (Scheme 1.24). In most cases, this resulted in allene contaminated with a small percentage of alkyne (<10%). Carbon disulfide reacted solely at the propargylic position to give b,g-unsaturated g-dithiolactones in excellent yield (Scheme 1.25),16,68 whereas carbon dioxide reacted at the allenic position to produce allenyl carboxylic acids in good yield.18 The allenylsilver compound did not react with methyl iodide under these conditions.66,68 Allenylsilver compounds appear to enter the palladium coupling catalytic cycle via transmetallation. It was found that tert-butyl allenylsilver, and the corresponding argentate derivative, underwent palladium-catalyzed coupling with iodobenzene to
16
SILVER ALKYLS, ALKENYLS, ARYLS, AND ALKYNYLS IN ORGANIC SYNTHESIS
R
R
H Ag.2LiBr
R1
H
EX
C
R
C
THF–HMPT 80–90%
R1
E R1
E
H
R = Bu, iPr, tBu R1 = H or Me EX = BrCN, I2, MeS-SO2Me, allyl bromide, CO2, CS2, Me3SiCl, Me3GeCl, Me3SnCl, NBS, NCS
Scheme 1.24
R CH2
H
CS2
C R1
R CH2
R1 H
1
R
Ag.2LiBr
SAg.2LiBr
S
Ag⋅2LiBr
RH2C S
S
H+
R1
H
RH2C S
S 80–95%
Scheme 1.25
give the aryl allene in good yield without any detectable amount of alkyne (Scheme 1.26).66 1.3.7
Synthesis of Perfluoroalkenylsilver Compounds
Perfluoroalkenylsilver compounds55 have been synthesized by the treatment of perfluoroalkynes,69 perfluoroallenes,70 and 1-trimethylsilyl perfluoroalkenes40 with silver fluoride; via the metathesis of lithium-substituted perfluoroalkenes with silver trifluoroacetate;71 and by exchange between perfluorophenyl- or perfluoroisopropylsilver and the corresponding iodoperfluoroalkene (Scheme 1.27).49
H
AgBr(LiBr)3
Li
THF –60ºC
C H
H C H
Ag
Scheme 1.26
H
PhI, Pd(PPh3)4 HMPT, THF 75%
C H
17
Csp2-Ag
AgF
F3C
F3C
CF3
F AgF
CF3
F
Ag
F
F
CF3
F
F
Ag
C F F
F
F
F
F
TMS
F
Ag
F
CF3
F
CF3
F
Li
F
Ag
AgF
TMS–F
CF3CO2Ag
F3C
CF3CO2Li
F3C
CF3
CF3
RfAg
RfI F
F
I
Ag
Rf = C6F5 or (CF3)2CF
Scheme 1.27
1.3.8
Reactivity of Perfluoroalkenylsilver Compounds
Although finding relatively wider synthetic use than the alkenylsilver compounds, examples of reactions of the perfluoroalkenylsilver compounds are limited to a handful of reactions, including protolysis, bromination, dimerisation, methylation, and silylation (Schemes 1.28 and 1.29).69,70 1.3.9 Synthesis and Reactivity of Silver-Substituted Diazomethyl Compounds Diazomethyl compounds are an extremely useful class of molecules for synthetic organic chemistry. Silver-substituted diazomethyl compounds have further increased the scope of these reagents. The thermally unstable argentodiazoesters are synthesized in situ by treating the corresponding diazoester with silver oxide. The resulting F3C
CF3
F
CH3
F3C
CF3
CH3I
F3C F
CF3 Ag Br2
F
Scheme 1.28
Br
18
SILVER ALKYLS, ALKENYLS, ARYLS, AND ALKYNYLS IN ORGANIC SYNTHESIS
F3C
H2O
CF2 H F3C
Br2
CF2 Br
F3C
∆
F2C
CF3
F3C
CF2
CF2 Ag
F3C
MeI
CF2 H3C F3C
Me3SiCl
CF2 Me3Si
Scheme 1.29
N2 H
N2
Ag2O
CO2Et
Ag
I CO2Et
N2 CO2Et
Scheme 1.30
argentodiazoesters may then easily participate in electrophilic substitution with SN1 reactive alkyl bromides and iodides (Scheme 1.30).72 Silver(diazomethyl)phosphoryl compounds may also be synthesized by treatment of the corresponding diazophosphoryl compound with silver oxide. The thermally stable phosphoryl derivatives undergo electrophilic substitution with alkyl iodides73 but, unlike the carbonyl-substituted derivatives, also undergo electrophilic diazoalkane substitution with a variety of H€ uckel aromatic salts (Scheme 1.31).74–79 Disilver diazomethane has also been synthesized. When treated with an alkyne, it produces the corresponding silver acetylide.80
1.4
Csp -Ag
The use of silver acetylides to effect useful synthetic transformations has been perhaps more widely reported than the use of any other organosilver compound, as previously reviewed by Pale et al.81 and Sladkov et al.82 Their use, however, in comparison to that of other metal acetylides (Na, Cu, Li), has perhaps been limited by their relatively low reactivity and/or solubility. The silver acetylides are, however, generally comparatively more stable than the more commonly used metal acetylides, as they can be
19
Csp-Ag
N2 Ag H
CI
P(O)R2
R=OMe or Ph DCM 0°C
N2 P(O)R2
H
C P(O)R2 N2
Scheme 1.31
stored for months without significant decomposition, and are prepared under relatively mild conditions.83,84 These characteristics, in conjunction with the often unique chemistry displayed by this class of compounds, allow a number of novel transformations under conditions generally more suitable for sensitive substrates. 1.4.1
Synthesis of Silver Acetylides
Classically, silver acetylides have been prepared by the treatment of an alkyne in ammoniacal silver nitrate solution to give the silver acetylide as a precipitate, which may be easily collected by filtration and stored in the absence of light (Scheme 1.32).83–86 A number of silver acetylides have been synthesized by this method, with a wide range of functionality. Other early methodology includes the synthesis of (1-butynyl)silver from the reaction of silver bromide with the corresponding Grignard reagent,87 and of propynyl- and (phenylethynyl)silver from the corresponding potassium acetylide and silver nitrate in liquid ammonia.88 AgNO3, NH4OH
R
H
H2O, MeOH
R
Ag
AgBr
MgBr
Et2O 60%
Scheme 1.32
Ag
20
SILVER ALKYLS, ALKENYLS, ARYLS, AND ALKYNYLS IN ORGANIC SYNTHESIS Ag
AgOTf
H
−
DIPEA
OTf
Ag
H
Scheme 1.33
Pale and coworkers have demonstrated that silver acetylides may be synthesized with a variety of silver salts in the presence of base in a number of solvents. By treating 1-hexyne in either deuterated benzene or DMF with silver triflate, the group was able to observe the formation of the p-alkyne–silver complex and subsequently the silver acetylide through the use of 1 H, 13 C, and 109 Ag NMR. The incipient p-alkyne–silver complex is rapidly deprotonated on addition of diisopropylethylamine, to give the silver acetylide as a white precipitate (Scheme 1.33).89 More recently, the well-known deprotection of 1-trimethylsilyl-1-alkynes with silver salts has been exploited for the isolation of silver acetylides (Scheme 1.34).90 Treatment of 1-trimethylsilyl-1-alkynes with silver salts in alcohol gives a variety of silver acetylides under relatively very mild conditions. The reaction also proceeds through formation of the p-alkyne–silver complex, thus enabling nucleophilic attack by the solvent to give the silver acetylides. This methodology has the advantage of avoiding strongly basic conditions in the synthesis of sensitive substrates, and allows silver acetylide formation from acetylenic alcohols for which silver-catalyzed cyclization is a possibility. Most silver acetylides are highly insoluble in standard organic solvents because of the formation of coordination polymers, and this characteristic can cause difficulties with characterization and limit reactivity. This can be overcome by the addition of coordinating donor molecules such as triphenylphosphine that perturb the formation of the polymeric structure.91 In the experience of the authors, however, the addition of such ligands can decrease reactivity. 1.4.2 1.4.2.1
Reactivity of Silver Acetylides Addition to Activated Carbonyls and Iminium Ions
The poorly nucleophilic silver acetylides are generally not considered reactive enough to add directly to unactivated carbonyl groups, although a number of Me3Si
R
AgNO3 or AgOTf
Ag
H2O/MeOH 53 – 99%
R
Me3SiOCH3 CH3OH
Ag+ Me3Si
R
Scheme 1.34
21
Csp-Ag
O R Ag
O
X
CCl4, reflux or AlCl3, CCl4, 0°C 26–72%
R X = Br or Cl
Scheme 1.35
examples of addition to activated carbonyls are known. The addition of silver acetylides to acid halides and carbon dioxide has been utilized in the synthesis of acetylenic ketones92,93 and esters,94,95 respectively. Davis and Schreiber reported the first addition of silver acetylides to acid halides to give ynones.83 Silver acetylides bearing an alkyl chain were treated with an acid halide in either refluxing carbon tetrachloride, or at 0 C in the presence of aluminum trichloride to give ynones in moderate to good yield (Scheme 1.35). This methodology has been used by other groups to synthesize a number of acetylenic ketones under relatively mild conditions.93 This methodology was utilized effectively by Inanaga and coworkers in the total synthesis of methynolide.96 Using the silver acetylide in this case allowed for the mild introduction of the acetylenic moiety, avoiding the use of strong base in the presence of the base sensitive ynone product (Scheme 1.36). Even the somewhat menacing-looking peroxyl silver acetylides undergo addition to acid chlorides to give ketones (Scheme 1.37).97 Silver acetylides have been added to iminium salts to give propargylic amines. In the derivitization of cortanine hydrochloride for biological screening, a wide variety O
O
HO O
OH
Methynolide Ag AcO
O Br
O OAc
MeO2C
CCl4, reflux 85%
O O
AcO AcO MeO2C
O O Cl
Cl
Scheme 1.36
22
SILVER ALKYLS, ALKENYLS, ARYLS, AND ALKYNYLS IN ORGANIC SYNTHESIS
O O
Ag
Cl O
O
R
R
R
O
O
R
Scheme 1.37 O N
O
∆
Cl
Ag
R
CH3CN
OMe
O AgCl
N
O OMe
Cortanine HCl R
Scheme 1.38
of silver acetylides were added to the iminium salt to give alkynylated products in good yield (Scheme 1.38).98,99 This same methodology was first used by the same group in the synthesis of noncyclic propargylic amines in a limited number of examples.100 b,g-Alkynyl a-amino acid derivatives have been obtained from the silvercatalyzed addition of alkynes to a-iminoesters in good yield. It is likely that the p-alkyne–silver complex is deprotonated by the iminoester to give the silver acetylide in situ, which may then react with the activated iminium ion (Scheme 1.39).101 An enantioselective version of this reaction has been reported by Rueping et al.102 Treatment of an a-iminoester and an alkyne with silver acetate and a binol phosphate derivative gave propargylic amines with the highest enantiomeric ratio (er) reported as 96 : 4. Although the proposed catalytic cycle invoked the in situ formation of the H
Ag H
OTf Ph
PMP H
Ph
Ag
Ph
PMP
N CO2Et
H
AgOTf
PMP NH
OTf CO2Et
Scheme 1.39
NH CO2Et
Ph
23
Csp-Ag
Ar O O P O OH Ar N MeO2C
PMP
HN
AgOAc
R
MeO2C
60–90%
H
PMP
R er <96 : 4
Scheme 1.40
silver acetylide and subsequent addition to an activated a-iminoester, the authors did not discount the possibility of participation of a chiral silver–phosphate complex (Scheme 1.40). Li and coworkers have reported a novel silver-catalyzed, three-component coupling of an aldehyde, an amine, and an alkyne, and the first example of a direct addition of a silver acetylide to an aldehyde to produce propargylic alcohols (Scheme 1.41).103–105 Treatment of an aldehyde with a silver halide and secondary amine gave propargylic amines in good yields in organic solvent, water, or ionic liquids. The proposed mechanism suggests addition of a silver acetylide to the iminium ion formed in situ. Further screening of conditions revealed that the use of Cy3PAgCl gave only propargylic alcohol rather than propargylic amine. The authors have postulated that, as silver acetylides are generally considered to be too stable to add directly to unactivated carbonyls, the electron-donating effect of the phosphine ligand activates the silver–carbon bond. The phosphine ligand in this reaction is thereby acting as a “chemoswitch,” allowing for selection of either the propargylic amine or alcohol. The reaction is neither water nor air sensitive and provides a “greener” alternative to the corresponding Grignard reaction (Scheme 1.42). OH cat. AgCI/PR33 H2O, 100°C 63–93%
2
R
R1
O H
R1
R2
H
N H N
cat. AgCl H2O, 100°C 47–99%
2
R
R1
Scheme 1.41
24
SILVER ALKYLS, ALKENYLS, ARYLS, AND ALKYNYLS IN ORGANIC SYNTHESIS
O R
R1
Cy3PAg O
H
R
H Cy3PAg
R1
Cy3PAg
O
R R1
i-Pr2NEtH Cl
H2O, CI Cy3PAgCl R1
H
OH
i-Pr2NEt R
R1
Scheme 1.42
Ag
CO2Me
OH
[Cp2ZrCl2] RCHO AgOTf
R CO2Me
Scheme 1.43
Silver acetylides participate in the zirconium-catalyzed nucleophilic addition to carbonyls in the absence of base to produce functionalized g-hydroxy-a,b-acetylenic esters.106 Although this reaction involves a transmetallation rather than a direct addition of a silver acetylide to an unactivated aldehyde, its particular attraction of is the avoidance of strong bases to produce the reactive zirconium species (Scheme 1.43). This methodology has been applied to the total synthesis of the potent anticancer natural product FR901464 (Scheme 1.44, where dr ¼ diastereomeric ratio).107 An extension of the original methodology has seen the synthesis of propargylic alcohols in moderate to good yields from silver acetylides via zirconium acetylides (Scheme 1.45).108 1.4.2.2
Nucleophilic Substitution of Activated Heteroaromatics
With the finding that silver phenylacetylide was soluble in pyridine, Agawa and coworkers attempted to react the acetylide with benzoyl chloride in pyridine as solvent. While the expected ynone was isolated in poor yield, the predominant product of reaction was from initial acylation of the pyridine nitrogen to produce the electrophilic pyridinium salt, followed by addition of the silver acetylide to the 2
25
Csp-Ag
O Ag
H
CO2Me
OH
Cp2ZrCl2, AgOTf
O
O
MeO2C
84% (dr = 6:1)
O
O
O OH
N H
O FR901464
Scheme 1.44 OH O R1
Ag
Cp2ZrCl2, AgOTf
R
33–89%
R
R1
Cp2Zr Cp2ClZr
R1
Cp2(TfO)Zr
O
R1
R1 TfO-
R
R = alkyl, aryl, ester, acetal
Scheme 1.45
position. A variety of pyridinium salts, and of other nitrogen-containing aromatic salts, are known to react with silver acetylides to give selective ethynylation ortho to the nitrogen, enabling a useful route to functionalized pyridines.109,110 Interestingly, in the case of N-methoxypyridinium iodide a ring-opened product was isolated rather than the ethynylated pyridine. The formation of this product is rationalized because the methoxide anion is a poorer leaving group than the benzoate, thus allowing ring opening to compete with aromatization (Scheme 1.46). 1-Aminobenzimidazolium salts react with silver acetylides in refluxing DMF to give pyrazoles in moderate yield (Scheme 1.47).111 1.4.2.3
Reaction with Alkyl Halides
In the synthesis of nucleosides from their corresponding C-glycosyl derivatives, silver acetylides have been proved effective in introducing an acetylenic ester. The ribose
26
SILVER ALKYLS, ALKENYLS, ARYLS, AND ALKYNYLS IN ORGANIC SYNTHESIS
R
PhCOCI
N
Ag
N Cl COPh
N COPh
∆
Ph
70% R
R
R R
PhCOCI
Ag
Cl N OCOPh
N O
∆
N R
25–60%
Ph Ph
Ag
N I OMe
MeO
H
N
N OMe
∆
Ph 70%
Scheme 1.46
N
Ag
H N
N
Ph
Ph DMF ∆
N NH2
N NH2
N
N
Ph
66%
Scheme 1.47
moiety of these molecules may be further modified through 1,3-dipolar cycloaddition of the alkynes with azides or diazoalkanes (Scheme 1.48).112,113 Acetylenic esters are more reactive than ordinary alkynes in 1,3-dipolar cycloadditions, and silver acetylide chemistry provides an efficient alternative route to these compounds. When, however, the alcohols were protected as their benzoate esters, the expected product was not
CO2Me TrO
TrO
O
O
Cl O
O TrO
AgC CCO2Me
42%
O
O CO2Me
Scheme 1.48
O
O
O
27
Csp-Ag TrO
BzO
O
TrO
O
Br OBz
BzO
O
TrO CO2Me
Ag
O
CO2Me
O
Ph BzO
O O
Ph
Scheme 1.49
obtained. The authors have invoked anchimeric assistance in displacement of the bromide by the benzoate ester, followed by attack on the incipient oxonium ion by the silver acetylide to give alkynylation alpha to the phenyl group (Scheme 1.49). Direct bridgehead alkynylation of 1-iodoadamantane has been achieved by treatment with a variety of silver acetylides at reflux in N-methylmorpholine.114 The methodology has been extended to the direct bridgehead substitution of methylbicyclo[2.2.2]octane and a carborate anion (Scheme 1.50).115 In an attempt to elucidate the mechanism, the reaction was performed in the presence of a radical trap (TEMPO¼2,2,6,6-tetramethyl-1-piperidinoxyl), a reversible electron acceptor (dinitrobenzene), or a radical sensitizer (dimethoxybenzene). In all cases the proportion of coupled product was increased, suggesting that both radical and polar mechanisms contribute to the outcome of this reaction. Primary alkyl halides can also be alkynylated by silver acetylides. Isabelle and coworkers reported the reaction of methyl iodide, ethyl iodide and d3-methyl iodide with several silver acetylides to give disubstituted alkynes.116 The authors preferred a non-radical-mediated mechanism for this reaction, as neither methane nor ethane, expected byproducts of a radical reaction, were observed. 1.4.2.4
Coupling Reactions
As is the case for allenylsilver reagents, silver acetylides undergo palladium-catalyzed coupling reactions. Ruitenberg and coworkers showed that (1-trimethylsilylethynyl)
I
R
Ag
∆
R = phenyl = 4-methylphenyl = 4-pentylphenyl = 4-bromophenyl = 2-chlorophenyl = 4-methoxyphenyl
Ag
C I
R
NMM
68% 44% 57% 20% 62% 51%
= 4-ethynylphenyl = 3-pyridyl = n-butyl = t-butyl = cyclohexenyl = TIPS
∆
C
NMM
H11
H11
Scheme 1.50
41%
40% 0% 29% 42% 25% 25%
28
SILVER ALKYLS, ALKENYLS, ARYLS, AND ALKYNYLS IN ORGANIC SYNTHESIS Ag
1
R1 R2
R
or
C R2
X
SiMe3 or
LiAg
H
SiMe3
R1 2
C
Pd(Ph3P)4
X
R2 SiMe3
X = Br, OAc, OS(O)Me, OSO2Me, OP(O)OEt)2
≤98%
Scheme 1.51
silver, and the corresponding lithium argentate, react with 1-allenic or 2-propynic substrates in the presence of catalytic amounts of Pd(Ph3P)4 to give conjugated allenynes (Scheme 1.51).117 During attempts at the total synthesis of neocarzinostatin chromophore, Pale and coworkers sought to develop methodology for the coupling of epoxyalkynes with enol triflates. Standard Sonogashira conditions resulted in disappointing yields due to decomposition of the sensitive epoxyenyne moiety. It was found, however, that replacing copper(I) iodide with silver(I) salts led to the clean formation of product in at least comparable yield (Scheme 1.52 and Table 1.6).118–120 Because of the chemical similarities between silver and copper, silver acetylides were implicated in a catalytic cycle comparable to that proposed for the coppercatalyzed reaction, and the group has produced undeniable evidence for their participation. Authentic samples of silver acetylides were found to couple with enol triflates under the same conditions as for the catalytic reaction to produce enynes in
O O
O
O
O O O NHMe
O
OH OH Neocarzinostatin chromophore O OTf OPG O
Pd(PPh3)4, R3N cocatalyst DMF
Scheme 1.52
OPG
29
Csp-Ag
TABLE 1.6. Silver-Catalyzed Sonogashira-Type Coupling Entry
Cocatalyst
1 2 3 4 5
CuI AgI AgI AgNO3 Ag2CO3
Catalyst Loading (mol%) 10 10 20 10 5
Time (h)
Yield (%)
1 24 20 4 4
44–54 51 78 60 56
high yield.121 Furthermore, 109 Ag NMR spectra of silver triflate in DMF-d7 gave a signal at 80 ppm that shifted to 750 ppm on addition of 1-hexyne, and subsequently to 1050 ppm on addition of diisopropylethylamine. These values are consistent with initial formation of a silver p complex, followed by deprotonation and formation of the silver acetylide (Scheme 1.53).89 In 2006, Li and Wang reported the palladium-free, silver-catalyzed Sonogashiratype coupling of aryl halides and terminal alkynes. The reaction proceeds in high yield in the presence of catalytic silver iodide, triphenylphosphine, and potassium carbonate. Although the mechanism remains unclear, it is evident that the silver acetylide has a role, as the acetylide is formed on mixing of the reagents (Scheme 1.54).122
R OTf Pd0
PdII
PdII OTf
R
AgOTf
H
Ag
R
R Ag
R R3N
R3NH+
Scheme 1.53
Ag
30
SILVER ALKYLS, ALKENYLS, ARYLS, AND ALKYNYLS IN ORGANIC SYNTHESIS
R
R R1
H
AgI, PPh3, K2CO3
X
R1
DMF, 100ºC 62–99%
X = I, Br
Scheme 1.54
Although catalytic amounts of Ag2O have been found to promote the palladiumcatalyzed coupling of aryl boronic acids and terminal alkynes, the authors in this case do not attribute the desired reactivity to the formation, and subsequent transmetallation, of a silver acetylide. Rather, it is proposed that the Ag2O activates the alkynylpalladium complex to allow transmetallation from the boronic acid, and that any competing formation of the silver acetylide results in a homocoupling of the alkynes (Scheme 1.55).123 Silver acetylides undergo oxidative condensation in the presence of cupric chloride to give diynes.124 Dimerization is a reaction common to a number of organosilver compounds, including both alkyl- and arylsilver derivatives. 1.4.2.5
Reactions with Non-carbon Electrophiles
Silver acetylides have also been used to form bonds with non-carbon centers to produce synthetically useful compounds. Sladkov and coworkers report the reaction of silver acetylides to arenediazonium chlorides to give areneazoethynes.125,126 A necessary condition for the reaction was the presence of para or ortho electron acceptor substituents on the aromatic ring of the diazonium salt (Scheme 1.56). As discussed, earlier, silver acetylides appear to undergo transmetallation with Cp2ZrCl2, and with palladium species in Sonogashira-type coupling reactions. Silver acetylides also react efficiently with phosphorous, antimony, and arsenic chlorides to give the ethynyl derivatives in good yield (Scheme 1.57). Ethynyl derivatives of mercury127 and tin128,129 were similarly synthesized. R R
Pd
R R
Pd0
Ag
R
R
H
Pd H
R
R Ph B(OH)2
Ph Ph
R
Pd
Scheme 1.55
31
Csp-Ag
Cl
N N Cl
acetone
Ag
Cl
35%
N N
Scheme 1.56
Ph
EICl3
Ag
El
benzene–THF
Ph
3
El = P, As, Sb HgCl2
Ph
Ag
Hg
THF
Ph
2
Scheme 1.57
CI
CI
CI
CI CI2
Ag CI
CCI4
CI
CI
CI
CI CI
CI
Scheme 1.58
Iodine, bromine, and chlorine react with silver acetylides to give haloalkynes. An early report of such a reaction was applied to the first synthesis of perchlorophenylacetylene (Scheme 1.58).130,131 The silver acetylides of 2-hydroxy-2-ethynyladamantane and 4-trans-ethynyl-cis2,6,6-trimethylbicyclo[3.1.1]heptan-4-ol react with iodine in chloroform to give the corresponding iodoalkynes under mild conditions in good yield.132 Similarly, the bromoalkynes of various n-alkyl acetylenes have been synthesized through treatment of the corresponding silver acetylides with bromine.133 Iodoform has been utilized in one example as an iodinating agent, whereby iodoform and silver phenylacetylide in refluxing dimethoxymethane gave 1-iodo-2-phenylacetylene in 70% yield.127 Alkynes bearing peroxyl functionality have also been facilely halogenated under these conditions (Scheme 1.59).97 The halogenation of alkynes via silver acetylides can be performed catalytically, utilizing a silver(I) salt and an electrophilic halogen source, such as N-bromosuccinimide. Hofmeister and coworkers first developed this highly useful methodology in the synthesis of 17a-haloethynyl steroids. On treatment of the alkynes with catalytic silver nitrate and either N-bromosuccinimide or N-iodosuccinimide in acetone at room temperature, the corresponding halo derivatives were obtained in high yield, Ag
I O
O
O
I2
O Ag
Scheme 1.59
O
O
O
O
I
32
SILVER ALKYLS, ALKENYLS, ARYLS, AND ALKYNYLS IN ORGANIC SYNTHESIS
R
OH
H
O cat. AgNO3,
X N
O
R
OH
X
acetone 51–88%
H
H
Scheme 1.60 Me3SiCl, AgX (cat.), DBU
R
H
R
DCM 75–97%
SiMe3
Scheme 1.61
although use of N-chlorosuccinimide failed to give the chloroalkyne (Scheme 1.60). Thus, this mild procedure obviated the need for strongly alkaline conditions and protection of functionality in the molecule. The catalytic reaction has found general use in the synthesis of natural products and complex molecules, and has been reviewed exhaustively prior to this publication.134 In an analogous reaction, 1-trimethylsilyl-1-alkynes have been synthesized in good to excellent yield on treatment with trimethylsilyl chloride, catalytic amounts of silver salts, and DBU in refluxing dichloromethane (Scheme 1.61).135 1.4.2.6
Fragmentation
a-Acetylenic carbinols may undergo deethynylation to give ketones when treated with silver carbonate or silver oxide in DMSO. The reaction is promoted by intermolecular coordination between the silver of one silver acetylide molecule and the triple bond of another (Scheme 1.62).136 1.4.2.7
Desilylation
Perhaps the most widely reported use of silver acetylides is in the deprotection of 1-trimethylsilyl-1-alkynes. Schmidt and Arens first reported the treatment of 1trimethylsilyl-1-alkynes with silver nitrate to give the silver acetylide in situ, which, when treated with potassium cyanide, is converted to the free alkyne (Scheme 1.63).137 This very mild and selective, but toxic, means for the selective
R
R
OH
R
OH
R
R Ag OH Ag R
R
O
H Ag O
R Ag
R
R
Scheme 1.62
Ag
Ag
33
Csp-Ag
R
Ag+
AgNO3
SiMe3
R
SiMe3 O H
H
R
Ag KCN H2O
–Me3SiOH
R
H
Scheme 1.63
desilylation of alkynes thus provides an alternative to strongly basic conditions and has been utilized in a number of natural product total syntheses. In an important extension to this methodology, published in 2005, protodesilylation was effected with catalytic quantities of silver(I) salt in either acetone with a large excess of water,138 or in a mixture of dichloromethane, methanol, and water.90,139 Selectivity for deprotection of 1-trimethylsilyl-1-alkynes was very high in both cases, with the exception of substrates bearing silyl protected alcohols, which, under prolonged exposure to reaction conditions, gave varying percentages of free alcohol. This fall in selectivity is attributed to the formation of nitric acid during the catalytic cycle, and was remedied in one example by the addition of pyridine to the reaction mixture. The usefulness of this specific and mild means of desilylation has been advantageously extended to include one-pot reaction of the formed silver acetylide. Catalytic desilylation/halogenation of 1-trimethylsilyl-1-alkynes has been utilized in a variety of syntheses, and, as for the halogenation of unprotected alkynes, is generally performed using a silver(I) salt in the presence of an electrophilic halogen source in acetone. A particularly elegant example of this methodology can be found in the hapten synthesis of Forsyth and coworkers, whereby the 1-trimethylsilyl-1-alkynes was cleanly converted in 95% yield to the iodoalkyne in the presence of other silyl protecting groups on the highly functionalized spiro-N,O-acetal core on treatment with AgOCOCF3 and N-iodosuccinimide (Scheme 1.64).140 The versatility of silver-catalyzed desilylation has been further extended by its use in a one-pot catalytic desilylation/Sonogashira-type coupling reaction. In the total
N
Cbz
O
N
OTES H
AgOCOCF3, NIS acetone 95%
TBSO
Cbz
O
OTES H
TBSO
I
TMS
Scheme 1.64
34
SILVER ALKYLS, ALKENYLS, ARYLS, AND ALKYNYLS IN ORGANIC SYNTHESIS O
OBz
OBz
O
TMS
HN
I
O
Ag2CO3, Pd(OAc)2, PPh3 Et3N, Bu4NCl
O OMe
HN
O
O OMe
THF 85%
OMe
OMe
O N
O
O
OMe OMe Lennoxamine
Scheme 1.65
R OTf Pd0
PdII
AgOTf
XMe3Si
PdII OTf
R
Ag
R
R
Me3SiX
K2CO3, MeOH, X = OMe or TBAF.3H 2O, X = F Me3Si
R
Scheme 1.66
35
Csp-Ag
I
Pd(PPh3)4, AgI K2CO3, MeOH
TMS
DMF 70%
OTf
R
Pd(PPh3)4, AgI nBu4NF⋅3H2O
O
O
DMF
OH
OH R = TMS 78% R = H 55%
Scheme 1.67
synthesis of lennoxamine, Koseki and Nagasaka obtained a significantly improved yield of coupled product on treatment of the trimethylsilyl-protected alkyne with Ag2CO3, Pd(OAc)2, PPh3, Bu4NCl, and triethylamine, in comparison to that obtained for standard Sonogashira conditions for the free alkyne (Scheme 1.65).141 The reaction is, in this case, stoichiometric in silver, and although a mechanism is not given, the authors suggest the intermediacy of a silver acetylide. Development of the scope of this reaction showed that a variety of 1-trimethylsilyl-1-alkynes and aryl iodides were coupled under these conditions in good to excellent yield.142 Pale and coworkers provided the first example of combined desilylation/coupling catalytic for silver. They found that 1-trimethylsilyl-1-alkynes in the presence of tetrakis(triphenylphosphine)palladium, a silver(I) salt, and an activator (potassium carbonate in methanol, or TBAF3H2O) in DMF coupled with vinyl triflates and aryl iodides to give enynes good yields (Scheme 1.66).143,144 Although silver(I) salt was not necessary for reaction when TBAF3H2O was used for activation of the carbon–silicon bond, a small to significant improvement was observed for all reported
O I
O
Pd(PPh3)4, AgCl, K 2CO3, MeOH
TMS DMF 69%
N B
O
BH3.THF THF, 54% ee 84% OH
(S)-1-Dehydroxyvirol A
Scheme 1.68
36
SILVER ALKYLS, ALKENYLS, ARYLS, AND ALKYNYLS IN ORGANIC SYNTHESIS
OTf OH
OH TMS
TIPS H
AgCI, Pd(PPh 3)4, DIPEA DMF 82%
TMS TIPS
OTf K2CO3, MeOH, AgCl, Pd(PPh3)4 DMF 80%
OH
OH PhI, Bu4NF.3H2O AgCl, Pd(PPh3)4
Ph
DMF 72%
TIPS
Scheme 1.69
cases on addition of silver(I) salt. Coupling was not observed in the absence of silver (I) salt when potassium carbonate in methanol was used. The advantages of this methodology are particularly evident in the significantly higher yields obtained for sensitive substrates in comparison to that obtained for coupling of the free alkyne (Scheme 1.67). The reaction conditions have been applied successfully to the total synthesis of (S)1-dehydroxyvirol A by Fiandanese et al. (Scheme 1.68, where ee ¼ enantiomeric excess).145 A powerful example of the utility of the Pale group methodology can be found in their selective coupling of molecules bearing a number of reactive alkynes. By careful selection of conditions, they selectively coupled each alkyne of a trialkyne, bearing one TMS group and one TIPS group, to vinyl triflates or aryl iodide (Scheme 1.69).146,147
1.5
CONCLUSION
In conclusion, it is clear that much scope remains for the development of methodology that exploits the unique chemistry of organosilver compounds. From the methods
REFERENCES
37
reviewed, it is evident that it is important to temper the reactivity of the organosilver compounds so as to find a balance between inertness and instability. There is no doubt that future success in this field may be found through the same balancing act.
REFERENCES 1. Silver-organische verbindungen, in Gmelin Handbuch, Der Anoroganischer Chemie Auflage 8, Teil B5, system nummer 61, RJ Meyer, EHE Pietsch, A Kotowski, M Beck-Goechring, K.-C Buschbeck, R. Keim. Ed: Springer-verlag, Berlin, 1975 2. Van Koten, G.; Noltes, J. G., Copper and silver, in Comprehensive Organometallic Chemistry, Vol. 2, Wilkinson, G.; Stone, R. G. A.; Abel, E. W., eds., Pergamon Press, Oxford, 1982, pp. 709–763. 3. Schmidbaur, H.; Bayler, A.; Synthesis and uses of organosilver compounds, in The Chemistry of Organic Derivatives of Gold and Silver, Patai, S.; Rappoport, Z., eds., Wiley, Chichester, UK, 1999, pp. 211–226. 4. Semerano, G.; Riccoboni, L., Chem. Ber. 1941, 74, 1089–1099. 5. Glockling, F.; Kingston, D., J. Chem. Soc. 1959, 3001–3004. 6. Gilman, H.; Woods, L., J. Am. Chem. Soc. 1943, 65, 435–437. 7. Whitesides, G. M.; Bergbreiter, D. E.; Kendall, P. E., J. Am. Chem. Soc. 1974, 96, 2806–2813. 8. Whitesides, G. M.; Gutowski, F. D., J. Org. Chem. 1976, 41, 2882–2885. 9. Whitesides, G. M.; Panek, E. J.; Stedronsky, E. R., J. Am. Chem. Soc. 1972, 94, 232–239. 10. O’Hair, R. A. J., Chem. Commun. 2001, 2002, 20–21. 11. Joseph, L.; Gardner, J. H., J. Org. Chem. 1940, 5, 61–67. 12. Gardner, J. H.; Joseph, L.; Gollub, F., J. Am. Chem. Soc. 1937, 59, 2583–2584. 13. Gardner, J. H.; Borgstrom, P., J. Am. Chem. Soc. 1929, 51, 3375–3377. 14. Westmijze, H.; Kleijn, H.; Vermeer, P., J. Organomet. Chem. 1979, 172, 377–383. 15. Kleijn, H.; Westmijze, H.; Meijer, J.; Vermeer, P., J. Organomet. Chem. 1980, 192, 275–281. 16. Meijer, J.; Ruitenberg, K.; Westmijze, H.; Vermeer, P., Synthesis 1981, 551–554. 17. Kleijn, H.; Westmijze, H.; Meijer, J.; Vermeer, P., J. Organomet. Chem. 1981, 206, 257–264. 18. Westmijze, H.; Kleijn, H.; Bos, H. J. T.; Vermeer, P., J. Organomet. Chem. 1980, 199, 293–297. 19. Tamura, M.; Kochi, J., Synthesis 1971, 303–305. 20. Yamamoto, K.; Nakanishi, K.; Kumada, M., J. Organomet. Chem. 1967, 7, 197–202. 21. Nagano, T.; Hayashi, T., Chem. Lett. 2005, 34, 1152–1153. 22. Murphy, R.; Prager, R. H., Aust. J. Chem. 1976, 29, 617–626. 23. Murphy, R.; Prager, R. H., Tetrahedron Lett. 1976, 463–464. 24. Brown, H. C.; Verbrugge, C.; Snyder, C. H., J. Am. Chem. Soc. 1961, 83, 1001. 25. Brown, H. C.; Snyder, C. H., J. Am. Chem. Soc. 1961, 83, 1002–1003. 26. Brown, H. C.; Hebert, N. C.; Snyder, C. H., J. Am. Chem. Soc. 1961, 83, 1001–1002.
38
SILVER ALKYLS, ALKENYLS, ARYLS, AND ALKYNYLS IN ORGANIC SYNTHESIS
27. Avasthi, K.; Ghosh, S. S.; Devaprabhakara, D., Tetrahedron Lett. 1976, 17, 4871–4874. 28. Chujo, Y.; Tomita, I.; Kozawa, Y.; Saegusa, T., Macromolecules 1993, 26, 2643–2644. 29. Kauffmann, T.; Huelsduenker, A.; Menges, D.; Nienaber, H.; Rethmeier, L.; Robbe, S.; Scherler, D.; Schricke, J.; Wingbermuehle, D., Tetrahedron Lett. 1990, 31, 1553–1556. 30. Wang, F. Q.; Khairallah, G. N.; Koutsantonis, G. A.; Williams, C. M.; Callahan, D. L.; OHair, R. A. J., Phys. Chem. Chem. Phys. 2009, 11, 4132–4135. 31. Lukyanov, O. A.; Savostyanova, I. A.; Gorelik, V. P.; Shlykova, N. I.; Tartakovskii, V. A., Izvest. Akad. Nauk Ser. Khim. 1992, 1798–1803. 32. Hantzch, A.; Caldwell, K. S., Chem. Ber. 1906, 2472–2478. 33. Tyrra, W., Heteroat. Chem. 2002, 13, 561–566. 34. Miller, W. T., Jr.; Burnard, R. J., J. Am. Chem. Soc. 1968, 90, 7367–7368. 35. Probst, A.; Raab, K.; Ulm, K.; Von Werner, K., J. Fluorine Chem. 1987, 37, 223–245. 36. Dyatkin, B. L.; Martynov, B. I.; Martynova, L. G.; Kizim, N. G.; Sterlin, S. R.; Stumbreviciute, Z.; Fedorov, L. A., J. Organomet. Chem. 1973, 57, 423–433. 37. Naumann, D.; Wessel, W.; Hahn, J.; Tyrra, W., J. Organomet. Chem. 1997, 547, 79–88. 38. Polishchuk, V. R.; Fedorov, L. A.; Okulevich, P. O.; German, L. S.; Knunyants, I. L., Tetrahedron Lett. 1970, 3933–3936. 39. Tyrra, W., J. Fluorine Chem. 2001, 112, 149–152. 40. Tyrra, W.; Naumann, D., J. Fluorine Chem. 2004, 125, 823–830. 41. Dubot, G.; Mansuy, D.; Lecolier, S.; Normant, J. F., J. Organomet. Chem. 1972, 42, C105–C106. 42. Banks, R. E.; Dickinson, N.; Morrissey, A. P.; Richards, A., J. Fluorine Chem. 1984, 26, 87–92. 43. Rossman, D. I.; Muller, A. J.; Lewis, E. O., J. Fluorine Chem. 1991, 55, 221–224. 44. Krause, E.; Schmitz, M., Chem. Ber. 1919, 52, 2150–2164. 45. Krause, E.; Wendt, B., Chem. Ber. 1923, 56, 2064–2066. 46. Gilman, H.; Straley, J. M., Recl. Trav. Chim. Pays-Bas 1936, 55, 821–834. 47. Reich, R., C. R. Chim. 1923, 177, 322–324. 48. Wennerstrom, O., Acta Chem. Scand. 1971, 25, 2341–2349. 49. Miller, W. T., Jr.; Sun, K. K., J. Am. Chem. Soc. 1970, 92, 6985–6987. 50. Leusink, A. J.; Van Koten, G.; Noltes, J. G., J. Organomet. Chem. 1973, 56, 379–390. 51. Beverwijk, C. D. M.; Van der Kerk, G. J. M., J. Organomet. Chem. 1972, 43, C11–C12. 52. Boersma, J.; Des Tombe, F. J. A.; Weijers, F.; Van der Kerk, G. J. M., J. Organomet. Chem. 1977, 124, 229–233. 53. Hofstee, H. K.; Boersma, J.; Van der Kerk, G. J. M., J. Organomet. Chem. 1978, 168, 241–249. 54. Gambarotta, S.; Floriani, C.; Chiesi-Villa, A.; Guastini, C., J. Chem. Soc., Chem. Commun. 1983, 19, 1087–1089. 55. Burton, D. J.; Yang, Z. Y.; Morken, P. A., Tetrahedron 1994, 50, 2993–3063. 56. Smith, V. B.; Massey, A. G., J. Organomet. Chem. 1970, 23, C9–C10. 57. Uson, R.; Laguna, A.; Abad, J. A., J. Organomet. Chem. 1983, 246, 341–345. 58. Tyrra, W.; Wickleder, M., Z. Anorg. Allg. Chem. 2002, 628, 1841–1847.
REFERENCES
39
59. Kremlev, M. M.; Tyrra, W.; Naumann, D.; Yagupolskii, Y. L., J. Fluorine Chem. 2005, 126, 1327–1331. 60. Tyrra, W.; Aboulkacem, S.; Hoge, B.; Wiebe, W.; Pantenburg, I., J. Fluorine Chem. 2006, 127, 213–217. 61. Glockling, F., J. Chem. Soc. 1955, 716–720. 62. Glockling, F., J. Chem. Soc. 1956, 3640–3642. 63. Whitesides, G. M.; Casey, C. P., J. Am. Chem. Soc. 1966, 88, 4541–4543. 64. Whitesides, G. M.; Casey, C. P.; Krieger, J. K., J. Am. Chem. Soc. 1971, 93, 1379–1389. 65. Koebrich, G.; Froehlich, H.; Drischel, W., J. Organomet. Chem. 1966, 6, 194–201. 66. Ruitenberg, K.; Kleijn, H.; Meijer, J.; Oostveen, E. A.; Vermeer, P., J. Organomet. Chem. 1982, 224, 399–405. 67. Meijer, J.; Ruitenberg, K.; Westmijze, H.; Vermeer, P., Synthesis 1981, 551–553. 68. Westmijze, H.; Ruitenberg, K.; Meijer, J.; Vermeer, P., Tetrahedron Lett. 1980, 21, 1771–1772. 69. Miller, W. T., Jr.; Snider, R. H.; Hummel, R. J., J. Am. Chem. Soc. 1969, 91, 6532–6534. 70. Banks, R. E.; Haszeldine, R. N.; Taylor, D. R.; Webb, G., Tetrahedron Lett. 1970, 5215–5216. 71. Morken, P. A.; Lu, H.; Nakamura, A.; Burton, D. J., Tetrahedron Lett. 1991, 32, 4271–4274. 72. Schoellkopf, U.; Rieber, N., Angew. Chem. Int. Ed. 1967, 6, 261. 73. Felcht, U.; Regitz, M., Chem. Ber. 1975, 108, 2040–2054. 74. Boehshar, M.; Heydt, H.; Regitz, M., Tetrahedron 1986, 42, 1815–1822. 75. Arenz, S.; Boehshar, M.; Regitz, M., Chem. Ber. 1986, 119, 1755–1765. 76. Bethaeuser, W.; Weber, B.; Heydt, H.; Regitz, M., Chem. Ber. 1985, 118, 1315–1328. 77. Regitz, M.; Weber, B.; Eckstein, U., Liebigs Ann. Chem. 1979, 1002–1019. 78. Heydt, A.; Heydt, H.; Weber, B.; Regitz, M., Chem. Ber. 1982, 115, 2965–2980. 79. Eisenbarth, P.; Regitz, M., Chem. Ber. 1984, 117, 445–454. 80. Blues, E. T.; Bryce-Smith, D.; Irwin, J. G.; Lawston, I. W., J. Chem. Soc. Chem. Commun. 1974, 466–467. 81. Halbes-Letinois, U.; Weibel, J. M.; Pale, P., Chem. Soc. Rev. 2007, 36, 759–769. 82. Sladkov, A. M.; Golding, I. R., Usp. Khim. 1979, 48, 1625–1683. 83. Davis, R. B.; Scheiber, D. H., J. Am. Chem. Soc. 1956, 78, 1675–1678. 84. Royer, E. C.; Barral, M. C.; Moreno, V.; Santos, A., J. Inorg. Nucl. Chem. 1981, 43, 705–709. 85. Glaser, C., Justus Liebigs Ann. Chem. 1870, 154, 137–171. 86. Liebermann, C., Justus Liebigs Ann Chem. 1865, 135, 266–290. 87. Danehy, J. P.; Nieuwland, J. A., J. Am. Chem. Soc. 1936, 58, 1609–1610. 88. Nast, R.; Schindel, H., Z. Anorg. Allg. Chem. 1963, 326, 201–208. 89. Letinois-Halbes, U.; Pale, P.; Berger, S., J. Org. Chem. 2005, 70, 9185–9190. 90. Viterisi, A.; Orsini, A.; Weibel, J. M.; Pale, P., Tetrahedron Lett. 2006, 47, 2779–2781. 91. Bruce, M. I.; Humphrey, M. G.; Matisons, J. G.; Roy, S. K.; Swincer, A. G., Aust. J. Chem. 1984, 37, 1955–1961.
40 92. 93. 94. 95. 96. 97. 98. 99. 100. 101. 102. 103. 104. 105. 106. 107. 108. 109. 110. 111. 112. 113. 114. 115. 116. 117. 118. 119. 120. 121. 122. 123. 124.
SILVER ALKYLS, ALKENYLS, ARYLS, AND ALKYNYLS IN ORGANIC SYNTHESIS
Yerino, L. V.; Osborn, M. E.; Mariano, P. S., Tetrahedron 1982, 38, 1579–1591. Naka, T.; Koide, K., Tetrahedron Lett. 2003, 44, 443–445. Fukue, Y.; Oi, S.; Inoue, Y., J. Chem. Soc. Chem. Commun. 1994, 2091. Tsuda, T.; Ueda, K.; Saegusa, T., J. Chem. Soc. Chem. Commun. 1974, 380–381. Inanaga, J.; Katsuki, T.; Takimoto, S.; Ouchida, S.; Inoue, K.; Nakano, A.; Okukado, N.; Yamaguchi, M., Chem. Lett. 1979, 8, 1021–1024. Yuvchenko, A.; Moiseichuk, K. L.; Dikusar, E. A.; Zhukovskaya, N. A.; Oldekop, Y., Zh. Obshch. Khim. 1990, 60, 1587–1593. Ukhin, L. Y.; Suponitskii, K. Y.; Kartsev, V. G., Chem. Nat. Compd. 2003, 39, 482–488. Ukhin, L. Y.; Golding, I. R.; Kartsev, V. G., Chem. Nat. Compd. 2004, 40, 156–159. Ukhin, L. Y.; Komissarov, N.; Orlova, Z. I.; Tokarskaya, O. A.; Yanovskii, A. I.; Struchkov, Y. T., Zh. Org. Khim. 1987, 23, 1323–1325. Ji, J. -X.; Au-Yeung, T. T. -L.; Wu, J.; Yip, C. W.; Chan, A. S. C., Adv. Synth. Catal. 2004, 346, 42–44. Rueping, M.; Antonchick, A.; Brinkmann, C., Angew. Chem. Int. Ed. 2007, 46, 6903–6906. Wei, C.; Li, Z.; Li, C. -J., Org. Lett. 2003, 5, 4473–4475. Yao, X.; Li, C. -J., Org. Lett. 2005, 7, 4395–4398. Li, Z.; Wei, C.; Chen, L.; Varma, R. S.; Li, C. -J., Tetrahedron Lett. 2004, 45, 2443–2446. Koide, K.; Shahi, S. P., Angew. Chem. Int. Ed. 2004, 43, 2525–2527. Albert, B. J.; Sivaramakrishnan, A.; Naka, T.; Czaicki, N. L.; Koide, K., J. Am. Chem. Soc. 2007, 129, 2648–2659. Albert, B. J.; Koide, K., J. Org. Chem. 2008, 73, 1093–1098. Nishiwaki, N.; Minakata, S.; Komatsu, M.; Ohshiro, Y., Chem. Lett. 1989, 18, 773–776. Agawa, T.; Miller, S. I., J. Am. Chem. Soc. 1961, 83, 449–453. Ukhin, L. Y.; Komissarov, N.; Orlova, Z. I.; Dolgopolova, N. A., Zh. Obshch. Khim. 1984, 54, 1676–1678. De las Heras, F. G.; Tam, S. Y. K.; Klein, R. S.; Fox, J. J., J. Org. Chem. 1976, 41, 84–90. Albrecht, H. P.; Repke, D. B.; Moffatt, J. G., J. Org. Chem. 1974, 39, 2176–2182. Pouwer, R. H.; Williams, C. M.; Raine, A. L.; Harper, J. B., Org. Lett. 2005, 7, 1323–1325. Pouwer, R. H.; Harper, J. B.; Vyakaranam, K.; Michl, J.; Williams, C. M.; Jessen, C. H.; Bernhardt, P. V., Eur. J. Org. Chem. 2007, 241–248. Isabelle, M. E.; Leitch, L. C., Can. J. Chem. 1958, 36, 440–448. Ruitenberg, K.; Kleijn, H.; Westmijze, H.; Meijer, J.; Vermeer, P., Recl. Trav. Chim. PaysBas 1982, 101, 405–409. Bertus, P.; Pale, P., Tetrahedron Lett. 1997, 38, 8193–8196. Bertus, P.; Pale, P., Tetrahedron Lett. 1996, 37, 2019–2022. Bertus, P.; Pale, P., J. Organomet. Chem. 1998, 567, 173–180. Dillinger, S.; Bertus, P.; Pale, P., Org. Lett. 2001, 3, 1661–1664. Li, P.; Wang, L., Synlett 2006, 14, 2261–2265. Zou, G.; Zhu, J.; Tang, J., Tetrahedron Lett. 2003, 44, 8709–8711. Sladkov, A. M.; Golding, I. R., Dokl. Akad. Nauk SSSR 1971, 200, 132–133.
REFERENCES
125. 126. 127. 128. 129. 130. 131. 132. 133. 134. 135. 136. 137. 138. 139. 140. 141. 142. 143. 144. 145. 146. 147.
41
Sladkov, A. M.; Ukhin, L. Y., Izvest. Akad. Nauk SSSR Ser. Khim. 1964, 1552–1553. Sladkov, A. M.; Ukhin, L. Y.; Gorshkova, G. N., Zh. Org. Khim. 1966, 2, 1456–1459. Golding, I. R.; Sladkov, A. M., Izvest. Akad. Nauk SSSR Ser. Khim. 1972 (3), 529–530. Le Quan, M.; Cadiot, P., Bull. Soc. Chim. Fr. 1965, 35–44. Johnson, O. H.; Holum, J. R., J. Org. Chem. 1958, 23, 738–740. Ballester, M.; Castaner, J.; Riera, R.; Tabernero, I., J. Org. Chem. 1986, 51, 1413–1419. Ballester, M.; Castaner, J.; Riera, R.; Tabernero, I.; Cornet, C., Tetrahedron Lett. 1977, 27, 2353–2354. Dikusar, E. A.; Kozlov, N. G.; Kovalskaya, S. S.; Popova, L. A.; Moiseichuk, K. L., Russ. J. Gen. Chem. 2001, 71, 290–293. Dikusar, E. A.; Zvereva, T. D.; Zhukovskaya, N. A.; Moiseichuk, K. L., Russ. J. Gen. Chem. 2001, 71, 917–920. Yamamoto, Y., Chem. Rev. 2008, 108, 3199–3222. Taniguchi, Y.; Inanaga, J.; Yamaguchi, M., B. Chem. Soc. Jpn. 1981, 54, 3229–3230. Vitali, R.; Gladiali, S.; Gardi, R., Gazz. Chim. Ital. 1972, 102, 673–678. Schmidt, H. M.; Arens, J. F., Rec. Trav. Chim. Pay-B. 1967, 86, 1138–1142. Carpita, A.; Mannocci, L.; Rossi, R., Eur. J. Org. Chem. 2005, 1859–1864. Orsini, A.; Viterisi, A.; Bodlenner, A.; Weibel, J. M.; Pale, P., Tetrahedron Lett. 2005, 46, 2259–2262. Forsyth, C. J.; Xu, J.; Nguyen, S. T.; Samdal, I. A.; Briggs, L. R.; Rundberget, T.; Sandvik, M.; Miles, C. O., J. Am. Chem. Soc. 2006, 128, 15114–15116. Koseki, Y.; Nagasaka, T., Chem. Pharm. Bull. 1995, 43, 1604–1606. Koseki, Y.; Omino, K.; Anzai, S.; Nagasaka, T., Tetrahedron Lett. 2000, 41, 2377–2380. Halbes, U.; Pale, P., Tetrahedron Lett. 2002, 43, 2039–2042. Bertus, P.; Halbes, U.; Pale, P., Eur. J. Org. Chem. 2001, 4391–4393. Fiandanese, V.; Bottalico, d.; Cardellicchio, C.; Marchese, G.; Punzi, A., Tetrahedron 2005, 61, 4551–4556. Halbes-Letinois, U.; Vasiliev, A.; Pale, P., Eur. J. Org. Chem. 2005, 2828–2834. Halbes-Letinois, U.; Pale, P., J. Organomet. Chem. 687, 2003, 420–424.
2 CYCLOADDITION REACTIONS ALEX M. SZPILMAN Schulich Faculty of Chemistry, Technion—Israel Institute of Technology, Haifa, Israel
ERICK M. CARREIRA Laboratory of Organic Chemistry, ETH Z €urich, Switzerland
2.1 Introduction 2.2 [2þ2] Cycloadditions 2.3 [3þ2] Cycloadditions 2.3.1 [3þ2] Cycloadditions of Azomethine Ylides 2.3.1.1 Discovery and Development of the Silver-Catalyzed [3þ2] Cycloaddition of Azomethine Ylides 2.3.1.2 Auxiliary-Based Asymmetric [3þ2] Cycloadditions 2.3.1.3 Catalytic Asymmetric [3þ2] Cycloadditions 2.3.1.4 Selected Applications and Extensions of Azomethine [3þ2] Cycloadditions 2.3.2 Other [3þ2] Cycloadditions 2.4 [3þ3] Cycloadditions 2.5 [4þ2] Cycloadditions 2.6 Concluding Remarks References
Silver in Organic Chemistry Edited by Michael Harmata Copyright 2010 John Wiley & Sons, Inc.
43
44
2.1
CYCLOADDITION REACTIONS
INTRODUCTION
In the 2001 book Cycloaddition Reactions in Organic Synthesis,1 the use of silver salts as mediators or catalysts was conspicuous only by its absence. The catalytic effects of silver were briefly mentioned in the chapter on [3 þ 2] cycloadditions. It is refreshing that in 7 years the field has progressed so far as to warrant a number of reviews, each covering various aspects of silver-mediated synthesis.2–5 This chapter seeks to afford the reader with a comprehensive survey of the field from its inception to the end of 2008.
2.2
[2+2] CYCLOADDITIONS
In one of the first contributions to the field, Friedman reported on the ability of silver to influence the reactivity of benzyne toward benzene (Scheme 2.1). In the absence of silver, benzyne 2, generated in situ from 1, reacted with benzene in a [4þ2] manner to afford adduct 3 in addition to small amounts of dimerization product 4.6 However, in the presence of substochiometric amounts of silver ions, the course of the reaction was altered, leading to formation of biphenyl 5 and benzocycloctatriene 6. A contingent of other metal ions including Tl þ , Cu þ , Cu2 þ , Hg22 þ , and Hg2 þ were unable to influence product formation in a similar manner. On the basis of these observations, Friedman6 and Paquette7 proposed the mechanism shown in Scheme 2.2. Thus, silver was suggested to activate benzyne for electrophilic attack to afford silver intermediate 8. Intramolecular proton transfer would account for the formation of 5, while cyclization would result in formation of 9 with subsequent 6e disrotary electrocyclic ring opening to afford 6. Masamune reported that reaction of 10 with AgClO4 in the presence of 3 equiv of maleic anhydride resulted in CO evolution and formation of cycloadduct 11 in 85%
N2+ 1
–
CO2
45°C
3
5 AgClO4
20%
2
4
20% 6 5 : 6 ~ 61 : 35
3 : 4 ~ 88 : 12
Scheme 2.1
45
[2þ2] CYCLOADDITIONS
Ag
Ag+
H
7
8
9 disrotary electrocyclic ring opening
+
H transfer
5
6
Scheme 2.2 O
Me H AgClO4
O Me
CO + Me
Me O
10
O O
H H
O
11 85%
O
Scheme 2.3
yield (Scheme 2.3).8 While no mechanistic studies were performed, it was postulated that ring opening takes place by an ionic mechanism with possible formation of 1,3-dimethylcyclobutadiene. In support of this hypothesis, in situ generation of free 1,3-dimethylcyclobutadiene, from its stable tricarbonyliron complex, in the presence of maleic anhydride leads to the formation of 11 in high yield. Kozmin reported that AgNTf2 is an efficient catalyst for [2þ2] cycloaddition of TIPS alkynyl ethers 12 with electrophilic alkenes (Scheme 2.4).9 The finding arose serendipitously from an attempt to generate TIPSNTf2 in situ. Notably, AgNTf2 was superior to a broad selection of standard Lewis acid (e.g., BF3OEt2 and TiCl4) and other silver salts (e.g., AgOTf). The reaction is of a fairly wide scope with respect to the alkene 13. A range of alkenes with electron-withdrawing substituents are able to participate in the reaction, including acrylonitrile, methyl acrylate, and cyclohexenone. The substituent on the alkyne can be either alkyl or phenyl. Mechanistic studies involving deuterium labeling and NMR spectroscopy indicated that the mechanism is stepwise and is initiated by the complexation of AgNTf2 to alkyne 12.
OTIPS
TIPSO R2
+ R1 12
O
O
R2 R1=Ph, Bn, n-Pr R2=Me, Ph, OMe
AgNTf 2 70–83%
R1
13
14
Scheme 2.4
46
CYCLOADDITION REACTIONS
OBn
R1
R2 15
O Me Me
30°C
45–94% dr 35–135 : 1
16
fod =
OBn
Ag(fod)
N
O F F
Me
N
R2
R1
R1=Ts, Ms R2=Ph, CO2Et, p-CF3C6H4
17
F F F F F
Scheme 2.5
OBn N Ph
18
Amberlyst 15
1. H+/CrO3
NsHN
H2O, 80°C
Ns
CHO Ph 19
86%
2. HSCH2CO2Et Amberlyst A21
90%
H2N
CO2H Ph 20
Scheme 2.6
The Nakamura group reported the [2þ2] cycloaddition of alkoxymethylenecyclopropanes to imines (Scheme 2.5).10 They identified Ag(fod) from an assortment of metal salts including gold (AuBr3), palladium [Pd(OAc)2], and other silver species [Ag(acac)]. Interestingly, the uncatalyzed reaction takes place at 80 C and is a rare case of a thermal [2þ2] cycloaddition. However, the catalytic process offers advantages in terms of higher cis/trans ratios. It should be noted that the trans isomer can be isomerized into the thermodynamically favored cis product by extended heating to 80 C in acetonitrile. The utility of the products was demonstrated by the synthesis of amino acid 20 (Scheme 2.6). In summary, it is clear that silver possesses unique abilities to activate both olefinic species and small strained rings for [2þ2] cycloaddition reactions. Nevertheless, reports of silver-mediated [2þ2] cycloadditions remain rare.
2.3
[3+2] CYCLOADDITIONS
The [3þ2] cycloadditions are the most prolific of all the silver-catalyzed cycloadditions. The unique affinity of silver for imines has facilitated the development of highly efficient and enantioselective cycloadditions of azomethine ylides to alkenes. Judicious choice of reaction conditions is crucial in achieving high yields for different substitution patterns.
47
[3þ2] CYCLOADDITIONS
MeO2C Ph
OMe
N
salt
+
CO2Me
80°C toluene
O 21
Ph
22
N H
CO2Me
23
OMe
Ph
N M 24
O
Scheme 2.7
2.3.1
[3+2] Cycloadditions of Azomethine Ylides
2.3.1.1 Discovery and Development of the Silver-Catalyzed [3+2] Cycloaddition of Azomethine Ylides The first example of the title process was reported by the Grigg group in 1982.11 This work examined the cycloaddition of Schiff base 21, generated from benzaldehyde and glycine, with methyl propiolate 22 in the presence of a metal salt incorporating a Lewis acidic metal and a basic counteranion. The uncatalyzed reaction occurred in d8toluene at 80 C to afford product 23 in 94% yield and with a starting material half-life (t1/2) of 38 h (Scheme 2.7; Table 2.1, entry 1).12 Silver acetate was able to catalyze the reaction, affording the product in 95% yield (entry 3). In contrast, Brønsted acid was found to be detrimental to product formation (entry 2). Zinc acetate bishydrate and lithium acetate bishydrate were also efficient catalysts for the reaction (entries 4 and 5). Solvent effects were found to be substantial for this [3þ2] cycloaddition. Thus, the half-life of the thermal reaction of 21 with maleic anhydride at 105 C was reduced to 59 min in d3-MeNO2 compared with 120 min in d8-toluene. The authors proposed the key notion that a reactive, conformationally rigid metalated azomethine species 24 is responsible for the high diastereoselectivity observed in the reaction. A second pivotal contribution from the Grigg group asserted the positive influence of the use of amine bases in combination with Lewis acids to catalyze the thermal TABLE 2.1. Effects of Acid on the Cycloaddition of Iminoester (21) and Methyl Propiolate (22) Entry 1 2 3 4 5 6 a b
Lewis Acid a
t1/2 of 21 (h)
Yield (%) b
None CH3COOH AgOAc Zn(OAc)22H2O LiOAc2H2O Mg(OAc)2
38 1.8 3.25 3.0 5.5 8.75
94 0 95 88 93 0
Run with 1.5 equiv. Determined by NMR with hexamethyl benzene as internal standard.
48
CYCLOADDITION REACTIONS
CO2Me MeO2C
26
MeO2C R1
2
R 1
R
OMe see Table 1.2
N 25
O
CO2Me R2 CO2Me N H 28
CO2Me 27
MeO2C R1
MeO2C N H
R2 CO2Me 29
Ph
N
Me CO2Me
30
Scheme 2.8 Optimization of the dipolar cycloaddition of iminoesters. See Table 2.2 for details.
reaction.13 An important finding was that a combination of a metal salt and an amine base would minimize stereomutation14 of the putative dipole, and thus improve diastereoselectivity. For this type of protocol it was found that silver acetate is vastly superior to both lithium bromide and zinc bromide. A range of solvents were investigated. This screen confirmed the earlier finding that cycloaddition rates in polar solvents are substantially higher than in nonpolar solvents. Acetonitrile was found to be the solvent of choice. Thus, the reaction of aromatic iminoesters 25 with dipolarophiles 26 and 27 took place in acetonitrile in the presence of silver acetate and triethylamine (Scheme 2.8). The reaction was completed within 0.5–2.5 h at room temperature to afford the pyrrolidine products 28 and 29, respectively, in good yield. In contrast, lithium bromide catalysis resulted in formation of conjugate addition byproducts of type 30, which were isolated in fair yields (35–45%). A key finding was that substochiometric amounts of silver acetate could be used, without significantly lowering the yield (Table 2.2, entry 8). TABLE 2.2. Influence of Metal Salts and Solvent on the Reaction Shown in Scheme 2.8 Entry 1 2 3 4 5 6 7 8 a
R1
R2
Dipolarophile
Lewis Acid
Solvent
Time (h)
Yield (%)
Ph Ph Ph Ph Ph 2-Naphthyl 2-Naphthyl 2-Naphthyl
Me Me Me Me Me H H H
27 27 27 26 26 27 27 27
LiBr (1.5) AgOTs (1.5) AgOAc (1.5) AgOAc (1.5) AgOAc (1.5) AgOAc (1.5) AgOAc (1.5) AgOAc (0.15)
MeCN THF MeCN THF MeCN THF MeCN MeCN
0.5 48 0.5 36 0.67 16 0.5 2.5
55 b 81 80 90 86 c 87 100 87 d
Run with 1.0 equiv of triethylamine unless otherwise noted. With 45% of 30. c 7:1 mixture of diastereoisomers. d 0.1 equiv triethylamine. b
49
[3þ2] CYCLOADDITIONS Me
O N 36 M
Ph
O
31
Me OMe
N
32
Me
Me Me
1.50 equiv AgOAc 1.0 equiv Et3N
MeO2C
MeCN or toluene
O Me
OMe
OMe
N
53–67%
R
N 34 H
MeO2C Me CO2Me + R
35
N H
Me CO2Me
dr 6 : 1 (MeCN) >99 : 1 (toluene)
Me OMe
N O
Me
33
Scheme 2.9
It proved possible to expand the scope of this cycloaddition protocol to aliphatic iminoesters (Scheme 2.9).15 For aliphatic iminoesters like 31 and 33, the existence of enamine-to-imine tautomerization during the reaction can lead to the formation of byproducts. However, the reaction of aldimines 31–33 under the silver acetate/ triethylamine/acetonitrile protocol led to the formation of the products in 53–67% yield. In the case of chiral iminoester 32, stereocontrol was poor (30% de). Lithium bromide also catalyzed the reaction, but at lower rates and produced compounds of type 30 (see Scheme 2.8) as the main product. Paradoxically, although reaction rates were lower, stereoselectivity was higher in nonpolar solvents. Grigg suggested that this was because product formation through azomethine intermediate could take place through either conformation 24 or 36. Since the latter has greater separation of charge, its contribution should be diminished in nonpolar solvents. As has been eluded to in the preceding text, the choice of the metal is crucial. Lithium, zinc, or other metal salts are often capable of catalyzing a specific azomethine cycloaddition reaction. However, silver salts frequently provide the products in higher yields. Furthermore, the silver salt anion is of crucial importance.16 Silver acetate (429.8 CHF/100 g, 99% purity Acros) is usually the salt of choice. For example, the reaction of naphthyl iminoester 25 (R1 ¼ H) with methyl acrylate catalyzed by 1.5 equiv of silver acetate and 1 equiv of triethylamine in acetonitrile affords the product in quantitative yield within 0.5 h. In contrast, catalysis by silver carbonate and silver tartrate affords the product in 67% and 61%, respectively, at considerably longer reaction times. On the other hand, silver nitrate (253.3 CHF/500 g, 99.9% purity, ABCR), is a more cost-efficient alternative and competitive in terms of yields (87%). It should be noted that no asymmetric induction could be observed in the silver tartrate–catalyzed reactions. Grigg has also reported that zeolite-bound Ag2O, CuCl, CuCN, NiCl2(dppe), and NiCl2(PPh3)2 catalyzes the reaction.17 In the case of silver oxide-impregnated zeolites, imine- and DBU-dependent leaching appears to be responsible for the catalytic activity.
50
CYCLOADDITION REACTIONS 1. (a) n-BuLi TMSCl (b) n-BuLi TMSCl
( )n N Boc
2. TFA 3. RI K2CO3
2 equiv AgF
( )n TMS 37
N R
( )n
TMS SO2Ph
N 38 R
MeCN
36
n =1, R = Bn, 73% n =2, R = Me, 92% n =3, R = Me, 48% n =1, R = Bn, 92% n =2, R = Me, 90% n =3, R = Me, 68%
R N ( )n
SO2Ph 39
Scheme 2.10
The Pandey group has developed a silver fluoride–promoted desilylation of tertiary bis(silyl)amines as an interesting alternative method to access azomethine ylides (Scheme 2.10).18 Notably, this method allows the generation of nonstabilized azomethine ylides under essentially neutral conditions. The starting materials are prepared by a three-step process, sometimes coupled into a single operation. For example, Boc-protected pyrrolidine 36 can be sequentially deprotonated and silylated twice in a one-pot reaction (Scheme 2.10). Removal of the Boc group and alkylation of the free amine leads to bis(silyl)amine 37. When this compound is treated with 2 equiv of silver fluoride in the presence of phenyl vinyl sulfone, rapid formation of products 39 as single stereoisomers results. During the reaction silver(0) deposits on the sides of the reaction vessel. The sequence of events leading to the azomethine ylide is unclear. However, evidently single-electron transfer (SET) from the amine to silver takes place either prior or subsequent to fluoride-enabled silyl cleavage. This process is repeated with a second equivalent of silver fluoride resulting in the formation of 38 either in free form or more likely as its silver complex. The scope of the method was expanded to the synthesis of bicylic systems exemplified here by tropinone 42 (Scheme 2.11).19 Pandey has also extended the protocol to the synthesis of tricycloalkanes20 and applied it to a total synthesis of the poisonous frog alkaloid epibatidine.21 2.3.1.2
Auxiliary-Based Asymmetric [3+2] Cycloadditions
As mentioned above, Grigg et al. had performed a brief reconnaissance into chiral silver complexes as inducers of chirality in the [3þ2] cycloaddition.16 This work O 2 equiv AgF
MeN
Raney Ni
MeN
O TMS
N Me 40
TMS
O
SO2Ph PhO2S MeCN
41
Scheme 2.11
42
51
[3þ2] CYCLOADDITIONS
focused on silver salts similar in nature to silver acetate, namely, silver tartrate and silver O-benzoyltartrate. While these reagents were able to promote the cycloaddition reaction, no asymmetric induction was observed. Asymmetric induction in [3þ2] cycloaddition reactions by the action of a chiral metal complex had, however, already been achieved using superstochiometric amounts of a cobalt–ephedrine complex (see discussion below).22 However, in the early 1990s the advent of practical asymmetric catalysis using substochiometric amounts of chiral silver complexes was still an elusive goal. As has historically been the case,23 the focus of earlier studies shifted to the use of auxiliaries to induce stereoselective cycloadditions of azomethine ylides and alkenes. As aliphatic chiral iminoesters gave poor diastereoselectivity in the [3þ2] cycloaddition (Scheme 2.9),15 the appendage of the chiral steering group to the dipolarophile component was attempted. Menthyl acrylate (44) proved promising in this regard (Scheme 2.12).24 However, menthyl acrylate (44) proved much less reactive than methyl acrylate (27). Thus, cycloaddition of 43 with menthyl acrylate required 8 h to reach full conversion compared with 1 h for methyl acrylate. Accompanying the long reaction times were low yields of cycloadduct (36% in the specific case) and the formation of byproducts. Accordingly, efforts were directed at increasing the rate of cycloaddition. The pKa of arylaldimines of a-amino esters is 17–19.5. It was argued that stronger bases should increase the rate of cycloaddition.24 Indeed, the use of DBU in place of triethylamine had already proved itself effective in increasing the yield in lithium salt–catalyzed reactions.15 The rate and yield of the cycloaddition of 43a to menthyl acrylate 44 in toluene were found to be directly related to the basicity of the base used. Thus, yield and rate
Me 2
R 1
R
1.5 equiv AgOAc or LiBr
O OMe
N
+
O Me
a b c d e f g h i j k l
R1
MeCN
O 43
MenthylO2C
Me
45
44 R1
R2
Ph Ph 2-naphthyl 2-naphthyl cyclohexyl 3-pyridyl 2-naphthyl Ph p-MeOC6H4 PhCHCH 3-pyridyl 3-pyridyl
H Me H Me Me Bn Ph Ph Ph Ph Ph Me
N 2 CO2Me H R
46 or 47
Yields:
43a–f: 63–76% (AgOAc/46) 71–78% (AgOAc/47) 43g–l: 64–86% (LiBr/46) 68–84% (LiBr/47)
N t-Bu
N N DBU (46)
Scheme 2.12
Me2N 47
NMe2
52
CYCLOADDITION REACTIONS
Me 43c–g,m–p +
O
H O
AgOAc
O
H
Me
OMenthyl H
R1
47
Me
O
O
1.5equiv
N 2CO2Me H R
71–91%
49
48
Bn
Me
OMe
N
Me
O
OMe
N
O
O
O 43o
43m O
Me
O
Me OMe
N
O
O
43n
OMe
N O
O
43p
Scheme 2.13
increased in the following order: 47 (pKa of the conjugate acid 14.6 in DMSO) > DBU 46 (pKa 12) > Et3N (pKa 9.0). The cycloaddition of 43a with menthyl acrylate was completed in 1 h using 47 as the base and afforded the product as a single diastereomer in 75% yield. For aryl iminoesters 43g–l bearing bulky R2 substituents, lithium bromide was the catalyst of choice. Unsaturated chiral amides were also investigated as dipolarophiles.25 A further development, by the Grigg group, was the use of menthyl acetal 48.26 This chiral acetal reacted with aromatic iminoesters 43c–g in the presence of silver acetate (1.5 equiv) and guanidine base 47 (1.2 equiv) in acetonitrile, to give cycloadducts 49 in good yields and as a single diastereoisomer in each case (Scheme 2.13). In contrast, toluene was the preferred solvents for aliphatic iminoesters 43e and 43m–p. Hiemstra’s chiral hemiaminal lactam 5027 was also investigated by the Grigg group (Scheme 2.14).26 This dipolarophile reacted to afford the cycloadduct 51 in
O 1.5 equiv AgOAc
Me 43c–e +
O
N Ac
O
Me
DBU or Et3N 52–93%
50
H
H
1
R2 N H CO2Me
R
51
Scheme 2.14
Me Me Ac N O
53
[3þ2] CYCLOADDITIONS
tol CO2Me S
tol 43a, c–d + O S
O
1.5 equiv AgOAc DBU or Et3N
OMe
19–77% dr 1 : 2–4 : 1
O S O MeO2C R1 CO2Me + R1 N 2 H R 53 ∆ toluene 68–89%
Me 52
MeO2C R1 55
CO2Me N 2 H R 54 ∆ toluene 73–92%
MeO2C CO2Me N 2 H R
CO2Me N 2 R H
R1 56
Scheme 2.15
moderate to excellent yields. Dichloromethane proved the best solvent for this reaction. The Ruano group pioneered the use of sulfoxide-based auxiliaries.28 Methyl S-para-tolylsulfinyl acrylate 52 underwent reaction with iminoesters 43a, c–d using Grigg’s catalyst system (AgOAc, DBU) in acetonitrile and THF (Scheme 2.15). Silver acetate catalyst loadings ranged from 7.5 mol% to 1.5 equiv. In general, good yields were obtained but diastereoselectivity was highly variable. Interestingly, the ratios of diastereoisomers were highly dependent on the identity of the solvent and base, but varied little as a function of catalyst loading or base/catalyst ratios. The auxiliary was removed by thermally induced syn elimination of para-toluenesulfenic acid, which took place to produce the corresponding 2-carboxy-2,3-pyrrolidines 55 and 56 in high yield. The chemistry of sulfoxide 57 has been studied in the Diels–Alder reaction with dienes.29 Ruano et al. studied its reaction with iminoesters 43a, c–d under a variety of conditions (Scheme 2.16).30 Silver-catalyzed cycloaddition at 0 C or room temperature afforded 1 : 1 mixtures of diastereomers 58 and 59 in variable yields. In contrast, deprotonation in the absence of silver using LiHMDS at –78 C afforded product 58 in 98 : 2 dr and 89–93% yield.30 The authors suggested a stepwise mechanism in which precomplexation of the sulfoxide to the lithium azaenolate would take place, thus allowing conjugate addition to follow.30 This notion was based on Posner’s31 and Paquette’s32 earlier work on conjugate additions of nucleophiles to 57. As before, thermal elimination of toluene sulfenic acid led to the conjugated products (not shown) in 92–93% yields for R2 ¼ methyl.30 However, for R2 ¼ H thermal elimination of arylsulfenic acid did not afford any dihydropyrrole product but rather led to the formation of pyrroles. Treating 58 or 59 with Raney nickel at low temperature led to the rapid formation of 60 and 61, respectively, in 82–86% yield.
54
CYCLOADDITION REACTIONS
O
O 43a, c–d
S
+ Me
57
7.5 mol% AgOAc 0.6 equiv DBU (46) 44–80%, dr 1 : 1
LiHMDS –78ºC 89–93% dr >98 : 2
O
O
MeC6H4OS R1 58
H CO2Me + N 2 H R
MeC6H4OS R1 59
H CO2Me N 2 H R Raney nickel 82–86%
Raney nickel 82–86%
O
O
H
H CO2Me N 2 H R
R1
H R1
60
H CO2Me N 2 H R
61
Scheme 2.16
Najera and Sansano et al. studied the reaction of acryl esters 63 of R- and S- lactate (only S shown; see Scheme 2.17).33 The reaction transpired at room temperature catalyzed by 10 mol% silver acetate in the presence of either 10 mol% KOH or triethylamine. Both methods afforded the endo products in moderate to good yields and excellent stereoselectivity. The auxiliary was cleaved by the action of aqueous base. In collaboration with the Cossio group, these authors have applied the method to synthesize a number of hepatitis C RNA polymerase inhibitors.34
R2 R1
O
CO2R3 +
N
CO2Me
O
62 R1
= Ph, 2-naphthyl R2 = H, Me, t-Bu, Bn R3 = Me, i-Pr, t-Bu
63
10 mol% AgOAc 10 mol% KOH toluene, 1d 10 mol% AgOAc 10 mol% Et3N toluene, 1d 32–70% dr 86 : 1–99 : 1
Scheme 2.17
Me MeO2C
O O R1
64
CO2R3 N 2 R H
55
[3þ2] CYCLOADDITIONS
R1
N
1.5 equiv AgOAc 1 equiv Et3N MeCN, 24–69%
NO2
R2 R3
CO2Me +
R4
65
5 equiv LiClO4 2 equiv Et3N MeCN, 24–66%
66 1
R4
O2N R3 R1 67
CO2Me N 2 R H
R = Ph, o-HO-C6H6, o-MeO-C6H6, 2-pyridyl R2 = H, Me R3 = H, Me R4 = p-Cl-C6H4, p-MeO-C6H4
Scheme 2.18
Cossio and Linden et al. studied the reaction of aryl iminoester-derived azomethine ylides 65 with 2- and 1,2-disubstituted nitroalkenes 66 using silver acetate and triethylamine in acetonitrile (Scheme 2.18).35 The product 3-nitropyrrolidines were formed in yields of 24–69% with excellent regioselectivity. These findings mirrored earlier work by Toke et al.36 Unfortunately, diastereoselectivity was generally in the 1 : 1–4 region. Only in two cases where yields were low were the products formed in a 2 : 98 ratio. Lithium perchlorate was able to catalyze the reaction as well with similar yields. Remarkably, the diastereoselectivity with lithium perchlorate in some cases resembled that obtained with silver acetate and in others was reversed. Cossio also reported the stereoselective reaction of phenyl isocyanates with azomethine ylides 65 under the same conditions.37 These reactants combined to give the expected products as single stereoisomer in 30–50% yield. While this initial excursion into the chemistry of nitroalkenes was disappointing, it proved possible to turn the tables by the use of a propionate aldol auxiliary attached at the 2 position of the nitroalkenes, namely, compounds 69 (Scheme 2.19).38 These chiral nonracemic nitroalkenes reacted stereospecifically with metalated azomethine ylides to afford the corresponding product 70 in 60–92% yield. Only 10 mol% of silver acetate was required to achieve complete conversion within 5 h at room R3
R3 O
R2 R1
N
10 mol% AgOAc 1 equiv Et3N MeCN
4
CO2Me +
R
O2N
68 R1 = Ph, t-Bu, c-Hex, c-Pr R2 = H, Me R3 = Ph, 2-F-C6H4, 2,6-F-C6H3 2,3-F-C6H3,3,5-F-C 6H3 R4 = Me, Et
Me
H
O2N
69
N H 70
1
R
HO R4
H
R2 Me H N
N O 71
Scheme 2.19
R4 R2 Me
R1
60–92%
O2N
O
CO2H
CO2Me
56
CYCLOADDITION REACTIONS
RCHO 72 + CO2Me 2 equiv H2N CO2Me
10 mol% AgOAc 20 mol% Ph3P THF
MeO2C MeO2C
N H
48–95% dr 4 : 1–9 : 1
73
CO2Me R
74 R = alkyl, aryl
+ O 20 equiv
OMe 27
Scheme 2.20
temperature. A full equivalent of triethylamine served as the base. Diastereoselectivity was better than 99 : 1 when R3 was an aryl group and ranged from 99 : 1 to 91 : 9 for aliphatic iminoesters. Compounds 70 were converted into a series of integrin inhibitors 71 by hydrolysis of the methyl ester function and amide formation to glycine. These compounds were shown to inhibit adhesion of cancer cells to microvascular endothelium in vitro and metastasis in vivo. The Garner group has developed a silver acetate/triphenylphosphine (10 mol%) catalyzed one-pot three-component coupling (Scheme 2.20) of aldehydes 72, methyl acrylate (27), and dimethyl-2-aminomalonate (73, 2 equiv).39 Notably, both aliphatic and aromatic aldehydes may participate in the reaction and yields up to 95% can be achieved. A breakthrough was achieved by replacing 2-aminomalonate with glycyl camphorsultam 75 (Scheme 2.21).40,41 This auxiliary allowed the one-pot threecomponent coupling of dimethyl maleate (76) and a variety of aliphatic aldehydes. Following the findings of the Zhou group (see text below), no amine base was used. Furthermore, in addition to dimethyl maleate (76), a variety of other dipolarophiles, including methyl acrylate (27), N-phenylmaleimide, and phenyl vinyl sulfone, participated in the reaction. Yields in the range 59–94% were achieved with stereoselectivity ranging from 7 : 1 to 19 : 1. Only dimethyl fumarate (26) underwent cycloadditions to give a mixture of several diastereoisomers. The sultam auxiliary could be removed with concomitant formation of the octyl thioester in 68% yield by treatment of 77 with lithium octyl thiolate. Garner has also reported that imines formed by the reaction of glycine with Garner’s (serine-derived) aldehyde proceeds in moderate yields and good diastereoselectivity.42 The power of this cycloaddition method is clearly illustrated in its application to the synthesis of 80 (Scheme 2.22), a known intermediate for the total synthesis of cyanocycline A (81) and bioxalomycin b2 (82).43 The condensation of aldehyde 78 and 75 in neat methyl acrylate (27) took place to afford the pyrrolidine 79 in 74% yield as a single stereoisomer.44 This compound was then converted into 80 in 12 steps. The method has also been applied to the synthesis of a number of analogs of 81 and 82 with
57
[3þ2] CYCLOADDITIONS RCHO
72 + Me
O
Me
MeO2C SO2 N
5 mol% AgOAc
1.1 equiv H2N
N O2S
THF 58–76% dr 7 : 1–13 : 1
Me
75 +
O
O
Me
CO2Me R
N H
77
R = alkyl OMe OMe
3 equiv O
76
Scheme 2.21
BocHN BnO MeO
Me
CHO NBnCbz + 1.1 equiv 75 + 27 (neat)
O
10 mol% AgOAc
OMe
HN
78 Me
N O 2S
BocHN H
NC
Me N
HO N H HO
MeO
H N H OMe
MeO
H CO Me 2 NBnCbz
BnO
OMe
79
Me
O
Me
80
HO
H
NC
Me N
H
Me N
O
H
N H
O
O H O
MeO
N
N H
O
H MeO
H
H
N
O
O Me
Me Cyanocycline A (81)
Scheme 2.22
Bioxalomycin β2 (82)
Me
58
CYCLOADDITION REACTIONS
notable biological properties.45 The broad range of aldehydes that may participate in the reaction marks this method as one of the most generally useful to date. 2.3.1.3
Catalytic Asymmetric [3+2] Cycloadditions
As noted above, the first use of a chiral silver salt (silver tartrate) to induce asymmetry in the cycloaddition of azomethine ylides with dipolarophiles was unsuccessful.16 However, the Grigg group, undaunted by this setback, tested chiral bisphosphine 83 as a ligand in the silver triflate–catalyzed reaction of 43d with dipolarophiles phenyl vinyl sulfone and methyl acrylate 27 (Scheme 2.23).46 This reaction proceeded in good yield, 64% for phenyl vinyl sulfone and 83% for methyl acrylate, in both cases in 70% ee. The preliminary nature of these findings was overshadowed by the authors’ success with a cobalt chloride–ephedrine (85) complex, which, while requiring superstochiometric amounts of catalyst, provided the product 86 in 84% yield and 96% ee.22 It was to be almost 8 years until the next silver-based catalytic asymmetric reaction was reported. In their pioneering work,47 the Zhang group screened a number of privileged phosphine ligands including BINAP (87), the Trost ligand (88), Duphos (89), and PennPhos (90) in the reaction of iminoester 103 with dimethyl maleate (76) (see Fig. 2.1, Scheme 2.24, and Table 2.3). However, the best results were achieved with the FAP ligands 91 and 92. Of these, the bulkier xylyl congener 92 provided the best enantioselectivity. The cycloaddition reaction was performed at room temperature and employed substochiometric amounts of H€ unig’s base. A significant finding was that efficient reaction took place with only 3 mol% of silver acetate and a similar amount of ligand. The silver acetate/FAP catalyst system facilitated the reaction of a range of aromatic glycine iminoesters with dimethyl maleate (76) to provide products 104 in excellent yield and enantio- and diastereoselectivity. Aliphatic imines underwent reaction under the same conditions to provide the product in high yield, but enantioselectivity was significantly lower (Table 2.3, entries 12 and 13).
OMe
N
MeO2C
O
Me
AgOTf
OMe
+
Ph 2P
O 43d
N H
PPh2
27
Me CO2Me
84, 83%, 70% ee 83
N Bn MeO2C
OMe
N O 43c
O +
2 equiv CoCl2
OMe
2 equiv
27
Ph H
Me H
HO
N
85
Scheme 2.23
N H 86, 83%, 96% ee
CO2Me
59
[3þ2] CYCLOADDITIONS
O
PPh 2 PPh 2
Me
O
(R, R ) Trost ligand (88)
O Me
P Me Me
PPh2Ph2P (Ra)-BINAP (87)
Me P
NH HN
(R, R )-Me-DuPhos (89)
O NH HN
Me
Fe P
P
Fe
Me Me (R, S, R, S )-PennPhos (90)
OMe PPh2
PAr2 Ar2P
(S, S, Sp )- FAP (91), Ar=Ph (S, S, Sp)- xylyl-FAP (92) Ar=3,5-dimethylphenyl
(Ra)-MOP (93) Me O N N
N Me O
PPh2 PPh2
Me O
(S, S )-DIOP (94)
PPh2
PAr2
Bn N Fe P(p-CF C H ) 3 6 42
PPh2
(Sa)-Quinap (95)
O Fe
Ph
Me
(Sa, R )-[O]-Pinap (96) a, R=H, Ar=Ph 98 b, R=Me, Ar= Ph
NR2 99
(S, Sp )-(97)
a, R=H, Ar=3,5-Me2C6H4 b, R=Me, Ar=3,5-Me2C6H4
Me Me Ph
N O
Ph
Ph St Bu Fe PPh 2
P(o-MeC6H4)2 (S)-PHOX (100)
O P N O
Me Me
Ph (S, Sp )-FeSulphos (101)
(S a, R, R)-Feringa's ligand (102)
Figure 2.1 Chiral ligands used in catalytic asymmetric [3 þ 2] cycloadditions.
60
CYCLOADDITION REACTIONS
R
OMe
N
+ 1.2 equiv 76
CO2Me
MeO2C
10 mol% i-Pr2EtN toluene
O
103
3 mol% AgOTf 3.3 mol% 92
R
N H
CO2Me
104
Scheme 2.24
TABLE 2.3. Enantioselective Cycloaddition of Iminoesters (103) and Dimethyl Maleate (76) (Scheme 2.24) Entry 1 2 3 4 5 6 7 8 9 10 11 12 13
R
Yield (%)
ee (%)
Ph p-Me-C6H4 p-MeOC6H4 p-Cl-C6H4 p-F-C6H4 p-CN-C6H4 o-Xl-C6H4 o-Me-C6H4 1-Naphthyl 2-Naphthyl 3-Pyridyl i-Pr c-Hex
87 93 98 96 96 90 96 97 73 98 98 82 82
87 88 92 92 90 96 86 90 85 97 84 70 81
The Schreiber group reported that a complex between silver acetate and Quinap (95) functioned as a highly active chiral promoter and catalyst for the reaction of arylglycyl imines with tert-butyl acrylate (106) (see Scheme 2.25 and Table 2.4).48 Other phosphine ligands that were tested in the reaction included BINAP (87), the Trost ligand (88), MOP (93), and DIOP (94). Further, screening of starting materials revealed some shortcomings of the system. Iminoesters bearing an a substituent as well as substituted dipolarophiles underwent reaction to give the products in high yield and stereoselectivity but moderate enantioselectivity. Nevertheless, the catalytic system was notable for its ability to promote cycloaddition even at –45 C. While Quinap (95) is an excellent and versatile ligand from which numerous developments in asymmetric catalysis have benefited,49 its synthesis is cumbersome, and hence its price remains high (R enantiomer 100 mg, 337 CHF). Carreira et al. showed that the synthetically readily accessible PINAP ligand class is as versatile as R2 R1
O 105
O
OMe
N
+
Ot-Bu 106
3 mol% AgOTf 3 mol% 95/96 3mol% i-Pr2EtN THF, –45 °C
Scheme 2.25
t-BuO2C R1
N H 107
R2 CO2Me
61
[3þ2] CYCLOADDITIONS
TABLE 2.4. Yield and Enantioselectivity for Formation of 107 Using Ligand 95 (Scheme 2.25) R1
R2
Yield (%)
ee (%)
p-MeOC6H4 p-Br-C6H4 p-NC-C6H4 2-Naphtyl 2-Toluyl Ph Ph
H H H H H Me Bn
93 89 92 89 95 98a 93b
95 95 96 94 89 80 77
Entry 1 2 3 4 5 6 7 a b
Reaction performed with 10 mol% catalyst loading at 20 C for 24 h. Reaction performed with 10 mol% catalyst loading at 20 C for 48 h.
TABLE 2.5. Yield and Enantioselectivity for Formation of 107 Using Ligand 96 (Scheme 2.25) R1
R2
Yield (%)
ee (%)
p-MeOC6H4 p-Br-C6H4
H H
88 94
92 95
Entry 1 2
Quinap (95) in a number of mutually unrelated transition-metal-catalyzed asymmetric reactions.50 Thus, 105 added to tert-butyl acrylate 106 using silver acetate and PINAP ligand 96, under the conditions described in Scheme 2.25, to provide the desired product 109 in excellent enantioselectivity and yield (see Scheme 2.6 and Table 2.5). Notably, the reaction required only 3 mol% of catalyst for complete conversion at –40 C within 36 h. The Zhou group reported the use of silver acetate and N,P-ferrocenyl ligand 97 in the azomethine cycloaddition with dimethyl maleate (76) as shown in Scheme 2.27
R
OMe
N
i-Pr2EtN
O
CO2Me
MeO2C
3 mol% AgOAc 3.3 mol% 97
R
108
CO2Me
N H
109 i-Pr2EtN
AgOAc/L*
AgOAc/L*
HOAc +
−
i-Pr2EtNH , AcO
R
N L*/ Ag
OMe
MeO2C
CO2Me
MeO2C
CO2Me
76 O 110
R 111
Scheme 2.26
N L*/ Ag
OMe O
62
CYCLOADDITION REACTIONS
TABLE 2.6. Yields and Enantioselectivity for Formation of 109 (Scheme 2.27) Entry 1 2 3 4 5 6 7 8 9 10 11
R
Yield (%)
ee (%)
Ph p-MeOC6H4 p-Cl-C6H4 p-F-C6H4 p-CN-C6H4 o-Cl-C6H4 o-Me-C6H4 1-Naphthyl 2-Naphthyl 3-Pyridyl i-Pr
85 94 99 96 91 98 99 85 95 76 56
97 98 97 97 97 97 98 98 98 93 88
(see Table 2.6).51 Zhou proposed the mechanism shown in Scheme 2.26, based on the earlier proposals by Grigg and Zhang.46,47 Zhou proposed that silver acetate should be able to catalyze the reaction without the commonly reported use of strong amine bases (indeed, this had already been shown using excess silver acetate in the pioneering studies of Grigg12). Thus, acetate/ acetic acid was proposed for its ability to function as a proton shuttle for the reaction. This proposal was borne out by experiment. Silver acetate/97 was able to efficiently catalyze the reaction of arylglycyl iminoester 108 with dimethyl maleate (76) at low temperature in ether without the addition of an external base (Scheme 2.27, Table 2.6). High yields and enantioselectivity were achieved. Notably, the reaction time was significantly reduced compared to earlier work. As a corollary, the authors investigated the significance of the anion in the silver salt. As predicted, a range of silver salts with basic counteranions catalyzed the reaction, but silver triflate, for instance, was unable to catalyze the reaction unless H€ unig’s base was added as an external base. Zhou also reported a series of related P,S-ferrocenyl ligands and their use in the [3þ2] cycloaddition of aryl-substituted azomethine ylides with N-phenylmaleimide.52 While these silver complexes were able to efficiently catalyze the reaction, the enantioselectivity was lower than in the protocol described above. Although mechanistic aspects of these reactions have not been investigated in depth, important findings and clues emanate from the many reports summarized in this chapter. An additional intriguing finding was reported by the Zhou group in 2007.53 While screening a series of N,P-ferocenyl ligands of types 98 and 99, they observed that reactions involving bismethylated ligands 98b led to formation of the
R
OMe
N 108
O
+ 1.5 equiv 76
3 mol% AgOAc 3.3 mol% 97 Et2O, −25°C
Scheme 2.27
CO2Me
MeO2C R
N H 109
CO2Me
63
[3þ2] CYCLOADDITIONS CO2Me
MeO2C p-ClC6H4
N H 110
CO2Me
3 mol% AgOTf p-ClC6H4 3.3 mol% 98a Et2O, −25°C 95%, 76% ee
N
111 + 1.5 equiv 76
O
OMe 3 mol% AgOTf 3.3 mol% 98b
MeO2C
Et2O, −25°C 91%, 83% ee
p-ClC6H4
CO2Me
CO2Me N H ent-110
Scheme 2.28
Figure 2.2 Models showing effects of presence or absence of H bonding on silver/ligand complex and dipolarphile. (Taken from ref. 53 with permission.)
products ent- 110 in high yield and enantioselectivity (Scheme 2.28). Curiously, when ligand 98a (bisdesmethyl 98b) was used, the same product was formed in high yield and enantioselectivity, but with the opposite absolute configuration. Clearly the ability of the ligand 98a to form additional hydrogen bonds led to fundamental changes in the structure of the silver azomethine complex, resulting in formation of the enantiomer of the product formed with ligand 98b. This trend proved to be consistent even with other ligands of the same type (e.g., 99a vs. 99b). Computational studies yielded the models shown in Figure 2.2. These models illustrate how the existence or absence of hydrogen bonding between the silver/ligand complex and the dipolarphile may lead to selection of opposite faces in the reaction with the silver-bound azomethine ylide derived from 111. While virtually all of the research described above has focused on the intermolecular cycloaddition of azomethine ylides, the intramolecular process holds considerable promise for the synthesis of polycyclic natural products. The Pfaltz group reported an intramolecular catalytic asymmetric cyclization of aryl iminoesters 112 using a complex of silver acetate with PHOX type ligand 100 (Scheme 2.29, O
R2
N
O
3 mol% AgOAc
CO2Me
3.3 mol% 100 toluene, 0°C
R1 112
H H
CO2Me R2
113
Scheme 2.29
R1
HN
64
CYCLOADDITION REACTIONS
TABLE 2.7. Yields and Enantioselectivity for Formation of 113 (Scheme 2.29) R1
R2
Yield (%)
ee (%)
CO2Me CO2tBu CO2Me 2-Pyridyl
H H Me H
74 66 61 70
96 99 96 83
Entry 1 2 3 4
R2 R1
O
OMe
N O
+
Ot-Bu 106
105
3 mol% AgOTf 3 mol% 95/96
t-BuO2C 1
R
3 mol% i-Pr2EtN THF, −45 °C
N H 107
R2 CO2Me
Scheme 2.30
Table 2.7).54 The tricyclic products 113 were formed in good yield and excellent enantioselectivity. Interestingly, the PHOX–silver complex performed less well in intermolecular cycloadditions. Najera reported that silver perchlorate forms a catalytically active complex with (S)-BINAP (87).55 This complex efficiently catalyzes the reaction of aryl iminoesters 108 with N-methylmaleimide 114 in the presence of the external base triethylamine (Scheme 2.30, Table 2.8). High enantioselectivity is obtained in this reaction. As a consequence of the low solubility of the silver BINAP complex in the reaction medium toluene, it can be quantitatively recovered from the reaction mixture by simple filtration. The recovered catalyst could be recycled at least 4 times with no apparent loss in activity. Carretero reported that N-phenylmaleimide also reacted with
TABLE 2.8. Yields and Enantioselectivity for Formation of 115 (Scheme 2.30) Entry 1 2 3 4 5 6a 7 8 9a 10 a
R
Yield (%)
ee (%)
Ph Ph 2-Naphthyl o-Cl-C6H4 p-Me-C6H4 p-Me-C6H4 p-MeO-C6H4 p-Cl-C6H4 p-Cl-C6H4 2-Thienyl
90 90 89 82 88 88 85 87 87 87
>99 >99 99 82 86 99 80 64 98 90
Reaction performed with 5 mol% DBU instead of triethylamine.
65
[3þ2] CYCLOADDITIONS Me O
106 t-BuO2C 2
R1
N H
R CO2R3
116
i PrO
N H
3
R1
O O
114
R2 OR
N O
5 mol% Et3N or DABCO toluene, −20°C
5 mol% AgOTf 5 mol% 102
O
H
H Me N Ph CO2Me H 117, (R2 = Me)
5 mol% DABCO toluene, 25°C
62
Me N
80%, 98%ee
CO2i Pr
CO2i Pr
2C
Ph
5 mol% AgOTf 5 mol% 102
N
Me CO2Me
119
5 mol% AgOTf 5 mol% 102 5 mol% DABCO
i PrO
2C
118
toluene, 0°C 79%, 82% ee
Scheme 2.31 Cycloaddition of 62 with various dipolarophiles.
a complex formed from silver acetate and Fesulphos ligand (101) and 108 to give products in up to 89% ee. However, in this case copper salts performed better.56 Najera et al. have shown that Feringa’s phosphoramidite ligand 102 is superior to other ligands in the cycloaddition of aryl iminoesters 62 with tert-butyl acrylate (106), N-methylphthalimide 114, and diisopropyl fumarate (118).57 As shown in Scheme 2.31 and Table 2.9, the products 116 and 117 were formed in excellent yields and exquisite endo selectivity. Most of the examples given in the preceding text relied on silver phosphine complexes as catalysts. The Jørgensen group examined the use of cinchona-derived ligands in the silver fluoride–catalyzed reaction of aromatic iminoesters 103 with methylacrylate (27). Yields were excellent, but enantioselectivities were only in the 41–73% range.58 While significant progress has been achieved in catalytic asymmetric cycloaddition of aryl iminoesters with dipolarophiles, the field is still in its infancy. Notably, TABLE 2.9. Yields and Enantioselectivities for Synthesis of 116 (Scheme 2.31) R1
R2
R3
Yield (%)
ee (%)
o-Me-C6H4 o-Cl-C6H4 Ph p-MeOC6H4 Ph 2-Thienyl 2-Thienyl Ph
H H H H Me Me t-Bu Bn
Me Me i-Pr i-Pr Me Me Me Me
83 90 81 79 78 77 70 77
98 >98 98 98 91 92 82 98
Entry a
1 2a 3 a, b 4c 5c 6c 7c 8c a
DABCO as base. Performed at 0 C. c Et3N as base. b
66
CYCLOADDITION REACTIONS
C
OMe
N
1 mol% AgOAc
N H
R2 121 o-BrPhOC
120
85%
+
MeCN
O
123
R2
R1
R1
CO2Me
122 O Br MeN
O N Me
O
1 mol% AgOAc MeCN
4
N
CO2Me
124
dr ∼83 : 17 (C4)
Scheme 2.32
the lack of protocols capable of leading to asymmetric cycloaddition of aliphatic iminoesters marks this as an important direction for future research. Additionally, the application of intramolecular cycloaddition may provide unique solutions in the synthesis of complex molecules. Also understanding of the mechanisms involved is still lacking, although it is likely that advances in this area would catalyze further developments. 2.3.1.4 Selected Applications and Extensions of Azomethine [3+2] Cycloadditions Work from the Grigg group since the late 1990s has focused on extensions of the chemistry detailed above. In one effort, it was shown how the unique reactivity of isocyanides allows them to function as azomethine ylide precursors. The reaction of isocyanide 120 with dipolarophiles 121 proceeds in the presence of silver acetate to afford dehydropyrroles 122 in good yield (Scheme 2.32, Table 2.10). Reaction with dimethyl maleate (76) afforded the corresponding product in excellent yield and diastereoselectivity. In contrast, its stereoisomer, dimethyl fumarate (26), furnished a 6 : 4 mixture of isomers under identical conditions.59 The authors proposed a stepwise mechanism for this reaction (Scheme 2.33). Tepe has reported the synthesis of the structurally related D1-pyrrolines (Scheme 2.34, Table 2.11).60 These structures are found in numerous natural products.
TABLE 2.10. Yields of Compound 122 (Scheme 2.32) Entry 1 2 3 4 5 6
R1
R2
Yield (%)
dr
CO2Me CN CHO COCH3 CO2Me CO2Me
H H H H CO2Me (cis) CO2Me (trans)
67 72 73 82 65 88
— — — — 96 : 4 63 : 37
67
[3þ2] CYCLOADDITIONS
OMe
N
C
CO2Me
+
MeO2C
1 mol% AgOAc MeCN
O 120
O
C
N
OMe
CO2Me
MeO2C
27
Ag
O 125
N
−AgOAc +HOAc
+AgOAc −HOAc
Ag
CO2Me
122
27
OMe
N H
Ag
C
OMe
N O
126
127
Scheme 2.33
This cycloaddition takes place to give the exo cycloadduct, rather than the more common endo cycloadduct, as the major product. Presumably this is a consequence of the intermediate munchnone 131 being locked in an anti orientation as compared to the commonly favored syn orientation (as in 24). Grigg has also shown that trans-substituted nitroolefins react with azomethine ylides in good yield (Scheme 2.35).61 Although trans-disubstituted dipolarophiles often give a mixture of endo and exo products, the diastereoselectivity was generally excellent except for (E)-2-methoxy-4-(2-nitrovinyl)phenol, which reacted with 2-naphthylmethyliminoglycyl ester 103 to give the product in only 42% yield and 2 : 1 endo : exo diastereoselectivity. Low diastereoselectivity also resulted from the cycloaddition of these nitroolefins with homoserine-lactone-derived imines.61 The scope of the silver-catalyzed [3þ2] cycloaddition is not limited to amino acid–derived iminoesters. Grigg has reported the use of aminophosphane-derived azomethine ylides in the cycloaddition to afford the corresponding phosphonatesubstituted pyrrolidines (Scheme 2.36).62 The diastereoselectivity of these cycloadditions is not inferior to those of standard amino acid–derived imines. The incorporation of fluorine is one of the most commonly employed strategies in lead optimization in medicinal chemistry. Franchini has reported that trans methyl 3-fluoroacrylate (137) undergoes reaction with glycine menthyl ester– derived imines 136 and 139 to afford fluorinated prolines 138 and 140 in good yield and diastereoselectivity (Scheme 2.37).63
O
R3
O Ph
N 128
R1
4
R
2. TMSCHN2
R1 R1
THF
129
130
Scheme 2.34
O
R4
R3 1. 10 mol% AgOAc
N
O Ph
CO2Me 131
N Ag
R1
68
CYCLOADDITION REACTIONS
TABLE 2.11. Yields of D1-Pyrrolines R1
Entry
1
Dipolarophile (129)
Me
Ph N
O
Product (130) Ph N
O
O
O
H Ph
2
Me
N
CO2Et
CO2Et
Ph
3
CO2Et
Me Ph
CO2Et
75
Me CO2Me
N
EtO2C
EtO2C
78
H Me CO2Me
EtO2C EtO2C
Yield (%)
75 a
Me CO2Me
N
MeO2C
4
MeO2C
Me
Ph
N
EtO2C
5
EtO2C
Ph
a b
CO2Et
CO2Et
Ph
95 b
Me CO2Me
N
15
Ph CO2Me
2 : 1 mixture of exo/endo epimers. 1 : 1 mixture of regioisomers.
Raghunathan has disclosed a synthesis of spiropyrrolidines 143 based on the Grigg triethylamine protocol (Scheme 2.38).64 The structurally complex spirocycles are formed in a single step, in high yield from simple starting materials. This protocol was extended to the synthesis of b-lactam spirocyclic compounds.65 Alcaide studied the silver acetate/triethylamine-catalyzed reaction of chiral blactam iminoesters 144 with methyl acrylate (27) as shown in Scheme 2.39.66 These R2 1
R
OMe
N O
NO2
+ R3
AgOAc (1.5 equiv) or Ag2O (10 mol%)
R1
Et3N (1.5 equiv)
105
132
toluene
R3
O2N N H
R2 CO2Me
133
42–95%
R1 = 2-naphthyl, 4-biphenyl R2 = H, Me, Bn R3 = 9-anthracyl, 3-indolyl, 4-HO-3-MeOC6H3, 2-furyl, 2-thienyl, 3-pyridyl
Scheme 2.35
69
[3þ2] CYCLOADDITIONS
R2 R1
N
O
P(OR3)2 + O
MeO2C
1.5 equiv AgOAc
OMe
R2 R1
DBU, MeCN 8–98%
27
134
N H 135
P(OR3)2 O
R1 = 2-naphthyl, 2-thienyl, c-Hex, p-NO2C6H4, o -IC6H4, 2-pyridyl R2 = H, Me, Ph, Bn, CH2OH, CH2OTBS R3 = Me, Et
Scheme 2.36
Me
Me O
Ph
Et3N
F
O
Me
AgOAc
OMe
+
O
N
Ph
75%, dr~ 9 : 1
137
O
N H
toluene, −40°C
Me
136
F
MeO2C
O
Me
Me
138
Me O N O
NC
+
O Me
Me
AgOAc
OMe
Et3N
F
toluene, −40°C 82%, dr ~ 9: 1
139
137
Me F
MeO2C
O NC
N H
O
Me
Me
140
Scheme 2.37
R2
R1
O 141
O
OMe
N
2
R
+
15 mol% AgOAc Et3N MeCN
O
68–91%
142
R1 = Ph, p-ClC6H4, p-MeO-C6H4 R2 = H, Cl, OMe, Me, NO 2
Scheme 2.38
O
O R1 N H 143
CO2Me
70
CYCLOADDITION REACTIONS
O N MeO
H
R1
O
R2
N OTBS
OMe
OMe
+ O
AgOAc Et3N
MeO
toluene
O
H N
OTBS CO Me 2 H HN R1
R2 CO2Me
57–78%
144
27
dr >9 : 1
R1 = Bn, CH2CHCH2 R2 = H, Me OTBS CO Me 2 H
H R1 N
N
MeO O
146
145
NaOMe 55–71% MeOH
R2 CO2Me
Scheme 2.39
reactions proceeded to afford the pyrrolidines 145 in 57–78% yield and with better than 9 : 1 diastereoselectivity. The compounds could be cyclized to afford the corresponding indolizidinones 146 in moderate yield. The indolizidinone moiety is another common motif in natural products. P€atzel achieved a stereoselective synthesis of methylcarboxy-substituted pyrrolidines 149 by reaction of phenyl iminoester 147 with chiral enones 148 (Scheme 2.40).67 Silver acetate (15 mol%) and DBU (1.2 equiv) in THF was used as promoters for the reaction. In almost all cases, selectivity better than 95 : 5 was achieved. Bashiardes studied the reaction of sugar-derived chiral nonracemic pyran 151 with aryl iminoesters 150 under Grigg’s conditions (AgOAc, DBU, MeCN).68 The products 152 were formed in 59–66% yield as single diastereoisomers (Scheme 2.41).
O
O R1
OEt
N
Me
+
O
R2
147
R2
Me 15 mol% AgOAc
R1
DBU, THF 72–98%
148
149
N H
dr 72 : 38–95 : 5
R1 = Ph, 3-pyridyl, p-MeIC6H4, C6H11 R2 = Me O
Me O
O
O
HO
OBn
NBn2
NBn2 Bn
Scheme 2.40
CO2Et
71
[3þ2] CYCLOADDITIONS
O
HO
OH
O
R1 Ph
OMe
N O
O
+
MeCN, 25ºC
OEt 150
O H
AgOAc DBU
59–66%
151
Ph 152
OEt H Y N H X
1
R = i-Bu, Ph: X = R1; Y = COOMe
R1 = H, i-Bu, Ph
R1 = H; X = COOMe; Y = H
Scheme 2.41
Grigg has pioneered the coupling of two or more reaction protocols into tandem or sequential one-pot procedures. These efforts have led to successful protocol for a silver-catalyzed azomethine ylide cycloaddition and a subsequent rhodium-catalyzed [2þ2þ2] cycloisomerization,69 as well as a tandem Heck–azomethine ylide cycloaddition.70 The Grigg group has also developed strategies for the rapid assembly of complex polycyclic structures by running an azomethine ylide cycloaddition reaction in sequence with a further cyclization reaction. For example, azomethine cycloaddition followed by a Pictet–Spengler cyclization gives access to a variety of bridged tri-, tetra-, penta-,71 and hexacyclic72 indole derivatives. This protocol has also been carried out on solid phase by employing Wang resin-linked acrylate ester as the dipolarphile in the [3þ2] cycloaddition.73 After the Pictet–Spengler step the products were cleaved of the solid phase using TFA.74 As a part of a program directed toward the synthesis of the potent topisomerase I inhibitors, the lamellarins (e.g., 153 and 154), Porco has reported the silver triflate–catalyzed tandem cycloisomerization–azomethine ylide cycloaddition of 155 (Scheme 2.42).75 The postulated mechanism of this intriguing and highly efficient process is shown in Scheme 2.43. Silver-catalyzed addition of the imine nitrogen to the alkyne results, on subsequent deprotonation, in the formation of an azomethine ylide 160. This ylide participates in [3þ2] cycloaddition with the alkyne component leading to formation of a dehydropyrrole 161. Finally, oxidation by adventitious oxygen leads to formation of the product 162. 2.3.2
Other [3+2] Cycloadditions
Billups reported the cycloaddition of benzocyclopropene 163 to 1,3-butadiene (Scheme 2.44). The reaction proceeded at 0 C with only 1 mol% of AgBF4 to afford product 165 in 64% yield.76 Under these conditions, compound 163 also reacted in a [3þ2] fashion with alkenes, allenes, and disubstituted alkynes but in much lower yields (11–25%) of [3þ2] cycloaddition products. A phenylsilver species has been invoked as an intermediate. The reaction probably proceeds through a cationic mechanism. Saito reported the extension of Billups’ tandem silver-catalyzed ring-opening cycloaddition reaction methodology to addition to imines (Scheme 2.45).77 Naphtho[b]cyclopropene 167 added to substituted aryl imines 168 at room temperature in the
72
CYCLOADDITION REACTIONS
MeO
AcO
O
O N
MeO
N
MeO
O
O
MeO
MeO AcO
OAc
MeO
HO
R4
R1
OSO3Na
Lamellarin α20-sulfate (154)
Lamellarin D-triacetate (153) R3
MeO
R1
R3
10 mol% AgOTf
+ N
2
R
CO2Me
155
toluene
Me
R1 = H, OMe, OCH2O t-Bu R2 = H, OMe, F R3 = alkyl,phenyl,TMS, cyclopropyl R4 = CO2Me, CO2Et, CHO, COMe R5 = CO2Me, CO2Et, C4H9, Ph, TMS
N
R2
60°C
R5 156
157
R4
CO2Me R5
28–83% t-Bu
N
Scheme 2.42
Ag (CH2)6Me N
CO2Me
159 (CH2)6Me
(CH2)6Me N
N
CO2Me 160
158
CO2Me CO2Me
AgOTf CO2Me (CH2)6Me N 162 MeO2C
CO2Me
(CH2)6Me N
[O]
161 MeO2C
CO2Me
Scheme 2.43
CO2Me CO2Me
73
[3þ3] CYCLOADDITIONS
1 mol% AgBF6
+ 163
165, 64%
164 166, 26%
Scheme 2.44
R X N X
+ 167
3 mol% AgBF4
168
N
R
169 a X=O, R=OMe, 62% b X=O, R=NO2, 30% c X=S, R=OMe,14% d X=S, R=NO2, 2% e X=NMe, R=OMe, 0%
Scheme 2.45
presence of 3 mol% AgBF4 to afford cycloadducts 169 in 0–62% yield. The best yields were achieved for electron-rich aryliminyl furans. Notably, no reaction took place under thermal conditions. Catalysis of the reaction by Yb(fod)3 could also be achieved, but this system required 60 C to proceed. Yields were moderate.
2.4
[3+3] CYCLOADDITIONS
Saito also investigated the reaction of naphtho[b]cyclopropene 167 with tropononederived imines 170 (Scheme 2.46). Reaction took place to afford tetracyclic dihydroisoquinolines 171 in modest yields.78 However, the preferred product was the simple addition product 172. The yields of cycloadducts 171 were generally only 8–15%. Ar N + 2 equiv
3 mol% AgBF4
N
Ar
N
+
benzene
167
170
171, 8–15%
Scheme 2.46
172, 15–24%
Ar
74
CYCLOADDITION REACTIONS
The proposed mechanism involves silver-catalyzed attack by the imine function on the cyclopropyl system. This is followed by conjugate addition of the intermediate silver species on the tropone ring system and subsequent isomerization to afford 171.
2.5
[4+2] CYCLOADDITIONS
In an early contribution to the field, Jendralla showed that the inherently unstable trans-cycloheptene could be stabilized by the addition of silver salts (AgClO4 and AgOTf) as 1 : 1 adduct 175 (see Scheme 2.47).79 This complex undergoes facile additions to dienes 176–178 to give the [4þ2] cycloadducts 179–181 in good yields. Interestingly, the complex could be dissociated by addition of trans-cyclooctene. This is due to the higher affinity of silver for trans-cyclooctene, with which it forms 3 : 1 adducts.79 Mukaiyama reported that 4–10 mol% of Ph2Sn¼S and AgClO4 catalyzed the Diels–Alder reaction between but-3-ene-2-one (183) or (E)-1-phenylbut-2-en-1-one (182) and unactivated dienes 184–186 to afford the products 187–192 in 57–95% yield (Scheme 2.48).80 The reaction with cyclic dienes 185 and 186 was highly endoselective (dr 99 : 1). Naphthalene-1,4-dione also participated in the reaction with similar yields. Silver-impregnated solids may also serve as catalysts in [4þ2] cycloadditions. Thus, Mayoral reported that silver ion–exchanged montmorillite K10 functioned as a catalyst for the reaction of methyl 2-cyanocinnamate and cyclopentadiene 185 (Scheme 2.49).81 The use of such heterogeneous catalysts may enhance catalyst recovery, an important consideration for industrial applications. With the relative air and water stability of silver phosphine complexes in mind, the Frost and Weller groups reported the use of a silver(I) carborane triphenylphosphine complex as a catalyst for the aza–Diels-Alder reaction.82,83 These unique catalyst complexes were able to catalyze the reaction of Danishefsky’s diene (195) with 196 in H
173
H
NO N NH2 O
175 AgClO4
NaHCO3
OMe
MeOH
H
174
OMe
O 176
AgClO4
O 179
H
Me OMe
H
180
H
177
OMe
Scheme 2.47
Me Me 181
Me 178
H
H
OMe
75
[4þ2] CYCLOADDITIONS
Me 184 Me
4–10 mol% AgClO4/
Ph
or
R1 187, R1=Me, R2=H, 90% 188, R1=Ph, R2=Me, 57% R2
Me
O
Me
O Me
O
185
Ph2Sn=S
182
R1
189, R1=Me, R2=H, 92% 190, R1=Ph, R2=Me, 90%
R1
191, R1=Me, R2=H, 95% 192, R1=Ph, R2=Me, 83%
R2 O
O Me 183
186 R2
Scheme 2.48 Ph
Ph + 6 equiv 185 NC
K10-Ag 62% 1.7: 1 endo : exo
CO2Me 193
NC
CO2Me
194
Scheme 2.49
CD2Cl2 at room temperature at low catalyst loadings (0.1 mol%) (Scheme 2.50, Table 2.12). Interestingly, catalytic activity was dependent on the presence of 50 mol% water. No reaction took place in its absence. This method was also studied using water as the solvent. The complex (PPh3)Ag(CB11H6Br6) catalyzed the reaction with a turnover frequency (TOF) of 4000 h1. Thus, it was more active than (PPh3)Ag (CB11H12). Both catalysts were more active than AgOTf(Ph3P), AgClO4(Ph3P) and AgBF4(Ph3P). Complexation of the three silver salts Ag(CB11H12), Ag(CB11H6Br6), and Ag(OTf) to polymer bound triphenylphosphine also yielded active catalyst systems. The polymer-bound catalyst could be recycled 3 times with no loss of activity. Dimeric complexes [e.g., [Ag(PPh3)2(CB11H12)]2] were significantly poorer catalysts. Kobayashi examined the efficacy of various silver salts (e.g., AgOTf, AgClO4, Ag2CO3) as catalysts for the aza-Diels–Alder cycloaddition of 195 to imines formed in situ in a 9 : 1 THF/water mixture (Scheme 2.51).84 The silver salts with basic anions (e.g., Ag2CO3) were unable to catalyze the reaction, but the silver salts with nonbasic OMe Ph +
N Ph
TMSO 195
N
See Table 2.12
196
Scheme 2.50
H 2O
O
Ph Ph
197
76
CYCLOADDITION REACTIONS
TABLE 2.12. Yields of 197 Using Various Silver Salts (Scheme 2.50) a Entry 1 2 3 4 5 6 7 a
Catalyst
Yield (%)
[Ag(PPh3)(BF4)] [Ag(PPh3)(OTf)] [Ag(PPh3)(ClO4)] [Ag(PPh3)(CB11H12)] [Ag(PPh3)2(CB11H12)] [Ag(PPh3)(CB11H6Br6)] [Ag(PPh3)2(CB11H6Br6)]
35 70 90 98 99 99 85
Reactions run in water for 60 min isolated yields.
OMe +
R1CHO +
R2NH2
TMSO 195
198
10 mol% AgOTf H2O
199
N
R2 R1
O 200
Scheme 2.51
anions (e.g., AgCl) all promoted the reaction to some extent; AgOTf was superior to all other potential catalysts examined. Further studies revealed that a 10 mol% loading of AgOTf was sufficient to catalyze the reaction of Danishefsky’s diene (195) with a variety of aromatic phenylimines bearing electron-withdrawing/donating substituents, in 57–92% yield within 2–3 h. These studies were carried out in water. Electron-poor imines generally required the use of 3 equiv of diene 195 to obtain satisfactory yields. The method was extended to a one-pot three-component protocol with in situ formation of the imine from the aniline and 1.5 equiv of the aldehyde (Scheme 2.51, Table 2.13). Because of TABLE 2.13. Yields of 200 for Several Aldehydes (Scheme 2.51) a Entry 1 2b 3 b, c 4b 5 6b 7b 8 a
R1
R2
Yield (%)
Ph Ph Ph Ph c-C6H12 c-C6H12 PhCH2CH2 i-Pr-CH2
Ph Ph p-BrC6H4 o-MeOC6H4 Ph Ph Ph Ph
63 80 90 56 70 51 53 72
With 1.5 equiv of aldehyde and diene relative to amine. With 10 mol% Triton X-100 added. c With 3 equiv diene. b
77
[4þ2] CYCLOADDITIONS
10 mol% AgX
OMe Ts + TMSO
O H
195
N
87 or 203
N CO2Et 201
Ts CO2Et
PR2 PR2
−78°C
75–90% 30–43% ee X= SbF6, ClO4, OTf
202 BINAP (87), R=Ph Tol-BINAP (203), R= p-toluyl
Scheme 2.52
the low solubility of the reagents in water, the reaction mixture was heterogenous. The complexity of the physical state of the mixture and its dependence on reactant structure is evident from the fact that in some cases the addition of 10 mol% Triton X-100, a nonionic surfactant, improved yields for aromatic aldehydes while in the case of cyclohexane carbaldehyde the yield was reduced (compare, e.g., entries 1 and 2 with entries 5 and 6 in Table 2.13). Attempts to develop enantioselective protocols for the aza-Diels–Alder reaction were reported simultaneously with those described above. A first contribution in this area was the report by the Jørgensen group,85 who studied the influence of salts of copper, silver, palladium, and zinc. Copper(I) perchlorate provides optimal yields and enantioselectivity, but complexes of BINAP (87) and Tol-BINAP (203) with AgSbF6, AgOTf, and AgClO4 were able to catalyze the reaction, albeit with low enantioselectivity (Scheme 2.52). In contrast, the Hoveyda and Snapper groups reported the efficient catalysis of silver acetate complexes of imine 204.86 These authors first established the superior performance of N-ortho-methoxyaniline-substituted imines in the reaction with Danishefsky diene 195. Because of the modular nature of ligand 206, optimization of its structure could be achieved by surveying a combinatorial library of amino acid derivatives. The l-isoleucine para-methoxyaniline conjugate 206 proved the best. This ligand facilitated full conversion within 2–3 h with high yield and enantioselectivity (Scheme 2.53). Through conjugation to a Wang resin, the ligand could be immobilized as its benzyl congener 207. A complex between 207 and AgOAc (5 mol%) was able to catalyze the reaction between 204 (Ar ¼ 2-naphthyl) and Danishefsky’s diene 195 in 96% yield and 86% ee. This drop in enantiomeric excess, compared to the reaction featuring silver acetate/206 complex, accurately reflects the drop in enantioselectivity observed by using the non-resin-bound analog of 207 and so appears to be unrelated to the solid phase. The resin-bound catalyst AgOAc/207 complex could be recycled at least 5 times with no drop in activity or stereoselection. The group of Yamamoto reported the catalytic enantioselective hetero-Diels– Alder reactions of azo compound 209 and dienes 208 (Scheme 2.54).87 In a ligand screening the use of BINAP (87) gave higher conversion and enantioselectivity than both Segphos (211) and Difluorphos (212). Interestingly, the optimal silver
78
CYCLOADDITION REACTIONS
OMe
MeO 0.1–1.0 mol%
N
AgOAc/
N
206
1.5 equiv 195 + H
Ar
O
4°C, THF 1 equiv H2O
204
Ar 205
78–98% 88–95% ee
Me
OMe
N
H
Et
H
Me
OMe
H N
N O
OMe
206
H
Et H N
H
O
207
Scheme 2.53
R1 Troc +
TIPSO R2
N N
5 mol% (Ra)-BINAP (87) EtCN 78ºC to –40°C 65–87% 84–99% ee
N
208
209
O O O
R1
10 mol% AgOTf
PPh2 PPh2
O (Ra)-Segphos (211)
F
O
F
O
F
O
F
O
N N
TIPSO R2
Troc
N
210
PPh2 PPh2
(Ra)-Difluorphos (212)
Scheme 2.54
triflate–to-BINAP ratio was 2 : 1. This would indicate the presence of a bimetallic catalyst exhibiting cooperative effects of two neighboring silver atoms.
2.6
CONCLUDING REMARKS
Silver salts and phosphine complexes have proved themselves as catalysts and mediators of a range of cycloadditions. No doubt only the tip of the iceberg has been uncovered regarding the possible applications of silver-mediated cycloadditions.
REFERENCES
79
REFERENCES 1. Kobayashi, S.; Jørgensen, K.A., eds., Cycloaddition Reactions in Organic Synthesis, Wiley-VCH, 2001. 2. Pellissier, H., Tetrahedron 2007, 63, 3235–3285. 3. Alvarez-Corral, M.; Munoz-Dorado, M.; Rodriguez-Garcia, I., Chem. Rev. 2008, 108, 3174–3198. 4. Calaza, M. I.; Cativiela, C., Eur. J. Org. Chem. 2008, 3427–3448. 5. Husinec, S.; Savic, V., Tetrahedron: Asymmetry 2005, 16, 2047–2061. 6. (a) Friedman, L., J. Am. Chem. Soc. 1967, 89, 3071–3073; (b) Friedman, L.; Lindow, D. F., J. Am. Chem. Soc. 1968, 90, 2324–2328. 7. Paquette, L. A., Chem. Commun. 1971, 1076–1077. 8. Ona, H.; Sakai, M.; Suda, M.; Masamune, S., Chem. Commun. 1973, 45–46. 9. Sweis, R. F.; Schramm, M. P.; Kozmin, S. A., J. Am. Chem. Soc. 2004, 126, 7442–7443. 10. Nakamura, I.; Nemoto, T.; Yamamoto, Y.; de Meijere, A., Angew. Chem. Int. Ed. 2006, 45, 5176–5179. 11. Grigg, R.; Gunaratne, H. Q. N., Chem. Commun. 1982, 384–386. 12. Grigg, R.; Gunaratne, H. Q. N.; Sridharan, V., Tetrahedron 1987, 43, 5887–5898. 13. Barr, D. A.; Grigg, R.; Gunaratne, H. Q. N.; Kemp, J.; Mcneekin, P.; Sridharan, V., Tetrahedron 1988, 44, 557–570. 14. Grigg, R.; Kemp, J., Tetrahedron Lett. 1980, 21, 2461–2464. 15. Grigg, R.; Montgomery, J.; Somasunderam, A., Tetrahedron 1992, 48, 10431–10442. 16. Amornraksa, K.; Barr, D. A.; Donegan, G.; Grigg, R.; Ratananukul, P.; Sridharan, V., Tetrahedron 1989, 45, 4649–4668. 17. Grigg, R.; Cooper, D. M.; Holloway, S.; McDonald, S.; Millingtona, E.; Sarkera, M. A. B., Tetrahedron 2005, 61, 8677–8685. 18. Pandey, G.; Lakshmaiah, G., Tetrahedron Lett. 1993, 34, 4861–4864. 19. Pandey, G.; Lakshmaiah, G.; Ghatak, A., Tetrahedron Lett. 1993, 34, 7301–7304. 20. Pandey, G.; Sahoo, A. K.; Bagul, T. D., Org. Lett. 2000, 2, 2299–2301. 21. Pandey, G.; Laha; Joydev K.; Lakshmaiah, G., Tetrahedron 2002, 58, 3525–3534. 22. Allway, P.; Grigg, R., Tetrahedron Lett. 1991, 32, 5817–5820. 23. Carreira, E. M.; Kværnø, L., Classics in Stereoselective Synthesis, Wiley-VCH, 2009. 24. (a) Barr, D. A.; Dorrity, M. J.; Grigg, R.; Malone, J. F.; Montgomery, J.; Rajviroongit, S.; Stevenson, P., Tetrahedron Lett. 1990, 31, 6569–6572; (b) Barr, D. A.; Dorrity, M. J.; Grigg, R.; Hargreaves, S.; Malone, J. F.; Montgomery, J.; Redpath, J.; Stevenson, P.; Thornton-Pett, M., Tetrahedron. 1995, 51, 273–294. 25. (a) Nyerges, M.; Bendell, D.; Arany, A.; Hibbs, D. E.; Coles, S. J.; Hursthouse, M. B.; Groundwater, P. W.; Meth-Cohn, O., Synlett 2003, 947–950; (b) Nyerges, M.; Bendell, D.; Arany, A.; Hibbs, D. E.; Coles, S. J.; Hursthouse, M. B.; Groundwater, P. W.; Meth-Cohn, O., Tetrahedron 2005, 61, 3745–3753. 26. Cooper, D. M.; Grigg, R.; Hargreaves, S.; Kennewell, P.; Redpath, J., Tetrahedron 1995, 51, 7791–7808. 27. Koot, W.-J.; Hiemstra, H.; Speckamp, W. N., J. Org. Chem. 1992, 57, 1958–1961. 28. Ruano, J. L. G.; Tito, A.; Peromingo, M. T., J. Org. Chem. 2002, 67, 981–987.
80
CYCLOADDITION REACTIONS
29. (a) Alonso, I.; Carretero, J. C.; Ruano, J. L. G., Tetrahedron Lett. 1989, 30, 3853–3856; (b) Alonso, I.; Carretero, J. C.; Ruano, J. L. G.; Cabrejas, M. L. M.; Lopez-Solera, I.; Raithby, P. R., Tetrahedron Lett. 1994, 35, 9461–9464. 30. Ruano, J. L. G.; Tito, A.; Peromingo, M. T., J. Org. Chem. 2003, 68, 10013–10019. 31. (a) Posner, G. H.; Hulce, M.; Mallamo, J. P., J. Org. Chem. 1981, 46, 5244–5246; (b) Posner, G. H.; Mallamo, J. P.; Miura, K., J. Am. Chem. Soc. 1981, 103, 2886–2888; (c) Posner, G. H.; Asirvatham, E., J. Org. Chem. 1985, 50, 2589–2591. (d) Posner, G. H., Acc. Chem. Res. 1987, 20, 72–78. 32. Paquette, L. A.; Tae, J.; Arrington, M. P.; Sadoun, A. H., J. Am. Chem. Soc. 2000, 122, 2742–2748. 33. Najera, C.; De Gracia Retamosa, M.; Sansano, J. M., Tetrahedron Asym. 2006, 17, 1985–1989. 34. Najera, C.; De Gracia Retamosa, M.; Sansano, J. M.; De Cozar, A.; Cossio, F. P., Eur. J. Org. Chem. 2007, 5038–5049. 35. Ayerbe, M.; Arrieta, A.; Cossio, F. P.; Linden, A., J. Org. Chem. 1998, 63, 1795–1805. 36. Nyerges, M.; Rudas, M.; Toth, G.; Herenyi, B.; Kadas, I.; Bitter, I.; Toke, L., Tetrahedron 1995, 51, 13321–13330. 37. Erkizia, E.; Aldaba, E.; Vara, Y.; Arrieta, A.; Gornitzka, H.; Cossio, F. P., Arkivoc 2005, 189–199. 38. Zubia, A.; Mendoza, L.; Vivanco, S.; Aldaba, E.; Carrascal, T.; Lecea, B.; Arrieta, A.; Zimmerman, T.; Vidal-Vanaclocha, F.; Cossio, F. P., Angew. Chem. Int. Ed. 2005, 44, 2903–2907. 39. Garner, P.; Kaniskan, H. U., J. Org. Chem. 2005, 70, 10868–10871. 40. Garner, P.; Kaniskan, H. U.; Hu, J.; Youngs, W. J.; Panzner, M., Org. Lett. 2006, 8, 3647–3650. 41. Vandewalle, M.; Van Der Eycken, J.; Oppolzer, W.; Vullioud, C., Tetrahedron 1986, 42, 4035–4043. 42. Garner, P.; Kaniskan, H. U., Tetrahedron Lett. 2005, 46, 5181–5185. 43. For the conversion of structure 80 into cyanocycline A, see Fukuyama, T.; Li, L.; Laird, A. A.; Frank, R. K., J. Am. Chem. Soc. 1987, 109, 1587–1589. For conversion of cyanocycline A into bioxolamycin b2, see Fukuyama, T., Adv. Heterocycl. Chem. 1992, 2, 189–249. 44. Kaniskan, H. U.; Garner, P., J. Am. Chem. Soc. 2007, 129, 15460–15461. 45. Kahsai, A. W.; Cui, J.; Kaniskan, H. U.; Garner, P. P.; Fenteany, G., J. Biol. Chem. 2008, 283, 24534–24545. 46. Grigg, R., Tetrahedron: Asymmetry 1995, 6, 2475–2486. 47. (a) Longmire, J. M.; Wang, B.; Zhang, X., Tetrahedron Lett. 2000, 41, 5435–5439; (b) You, S. -L.; Hou, X. -L.; Dai, L. -X; Gao, B. -X.; Sun, J., Chem. Commun. 2000, 1933–1934. 48. Xiaodong Li, C. C.; Schreiber, S. L., J. Am. Chem. Soc. 2003, 125, 10174–10175. 49. Kocovsky, P.; Vyskocil, S.; Smrcina, M., Chem. Rev. 2003, 103, 3213–3245. 50. Kn€opfel, T. F.; Aschwanden, P.; Ichikawa, T.; Watanabe, T.; Carreira, E. M., Angew. Chem. Int. Ed. 2004, 43, 5971–5973. 51. Zeng, W.; Zhou, Y.-G., Org. Lett. 2005, 7, 5055–5058.
REFERENCES
52. 53. 54. 55. 56. 57. 58. 59. 60. 61. 62. 63. 64. 65. 66. 67. 68. 69. 70. 71. 72. 73. 74. 75. 76. 77. 78. 79. 80. 81.
81
Zeng, W.; Zhou, Y.-G., Tetrahedron Lett. 2007, 48, 4619–4622. Zeng, W.; Chen, G.-Y.; Zhou, Y.-G.; Li, Y.-X., J. Am. Chem. Soc. 2007, 129, 750–751. Stohler, R.; Wahl, F.; Pfaltz, A., Synthesis 2005, 1431–1436. Najera, C.; De Gracia Retamosa, M.; Sansano, J. M., Org. Lett. 2007, 9, 4025–4028. Cabrera, S.; Arrayas, R. G.; Martin-Matute, B.; Cossio, F. P.; Carretero, J. C., Tetrahedron 2007, 63, 6587–6602. Najera, C.; De Gracia Retamosa, M.; Sansano, J. M., Angew. Chem. Int. Ed. 2008, 47, 6055–6058. Alemparte, C.; Blay, G.; Jørgensen, K. A., Org. Lett. 2005, 7, 4569–4572. Grigg, R.; Lansdell, M. I.; Thornton-Pett, M., Tetrahedron 1999, 55, 2025–2044. Peddibhotla, S.; Tepe, J. J., J. Am. Chem. Soc. 2004, 126, 12776–12777. Grigg, R.; Kilner, C.; Sarker, M. A. B.; De la Cierva, C. O.; Dondas, H. A., Tetrahedron 2008, 64, 8974–8991. Dondas, H. A.; Durust, Y.; Grigg, R.; Slatera, M. J.; Sarkera, M. A. B., Tetrahedron 2005, 61, 10667–10682. Bonini, B. F.; Boschi, F.; Franchini, M. C.; Fochi, M.; Fini, F.; Mazzanti, A.; Ricci, A., Synlett 2006, 543–546. Subramaniyan, G.; Raghunathan, R., Tetrahedron 2001, 57, 2909–2913. Subramaniyan, G.; Raghunathan, R.; Castro, A. M. M., Tetrahedron 2003, 59, 335–340. Alcaide, B.; Almendros, P.; Redondo, M. C.; Ruiz, M. P., J. Org. Chem. 2005, 70, 8890–8894. (a) P€atzel, M.; Galley, G., Tetrahedron Lett. 1993, 34, 5707–5710; (b) Galley, G.; Liebscher, J.; P€atzel, M., J. Org. Chem. 1995, 60, 5005–5010. Bashiardes, G.; Cano, C.; Mauze, B., Synlett 2005, 587–590. Grigg, R.; Sridharan, V.; Wang, J.; Xu, J., Tetrahedron 2000, 56, 8967–8976. Grigg, R.; Millington, E. L.; Thornton-Pett, M., Tetrahedron Lett. 2002, 43, 2605–2608. Dondas, H. A.; Duraisingham, J.; Grigg, R.; Maclachlan, W. S.; Macpherson, D. T.; Thornton-Pett, M.; Sridharan, V.; Suganthan, S., Tetrahedron 2000, 56, 4063–4070. Grigg, R.; Thornton-Pett, M.; Yoganathan, G., Tetrahedron 1999, 55, 8129–8140. Dondas, H. A.; Grigg, R.; Maclachlan, W. S.; Macpherson, D. T.; Markandu, J.; Sridharan, V.; Suganthan, S., Tetrahedron Lett. 2000, 41, 967–970. Blaney, P.; Grigg, R.; Rankovic, Z.; Thornton-Pett, M.; Xu, J., Tetrahedron 2002, 58, 1719–1737. Su, S.; Porco, J. A., J. Am. Chem. Soc. 2007, 129, 7744–7745. Billups, W. E.; Chow, W. Y.; Smith, C. V., J. Am. Chem. Soc. 1974, 96, 1979–1980. Saito, K.; Ono, K.; Ohkita, M.; Fukaya, M.; Ono, K.; Kondo, Y., Heterocycles 2002, 60, 773–778. Saito, K.; Ono, K.; Ito, N.; Tada, N.; Ando, S., Heterocycles 2002, 57, 235–240. Jendrella, H., Angew. Chem. Int. Ed. Engl. 1980, 19, 1032–1033. Mukaiyama, T.; Watanabe, K.; Shiina, I., Chem. Lett. 1995, 24, 1–2. Fraile, J. M.; Garcia, J. I.; Mayoral, J. A.; Pires, E.; Tarnai, T.; Figueras, F., Appl. Catal. A 1996, 136, 113–123.
82
CYCLOADDITION REACTIONS
82. Hague, C.; Patmore, N. J.; Frost, C. G.; Mahon, M. F.; Weller, A. S., Chem. Commun. 2001, 2286–2287. 83. Patmore, N. J.; Hague, C.; Cotgreave, J. H.; Mahon, M. F.; Frost, C. G.; Weller, A. S., Chem. Eur. J. 2002, 8, 2088–2098. 84. Loncaric, C.; Manabe, K.; Kobayashi, S., Adv. Synth. Catal. 2003, 345, 475–477. 85. Yao, S.; Johannsen, M.; Hazell, R. G.; Jørgensen, K. A., Angew. Chem. Int. Ed. 1998, 37, 3121–3124. 86. Josephsohn, N. S.; Snapper, M. L.; Hoveyda, A. H., J. Am. Chem. Soc. 2003, 125, 4018–4019. 87. Kawasaki, M.; Yamamoto, H., J. Am. Chem. Soc. 2006, 128, 16482–16483.
3 SIGMATROPIC REARRANGEMENTS AND RELATED PROCESSES PROMOTED BY SILVER JEAN-MARC WEIBEL, AURELIEN BLANC,
AND
PATRICK PALE
Laboratory for Organic Synthesis and Reactivity, Institute of Chemistry, University of Strasbourg, France
3.1 Introduction 3.2 Wolff and Arndt–Eistert Rearrangements and Related Reactions 3.3 Ring Rearrangements 3.3.1 Halogenoamines 3.3.2 Cyclopropane Derivatives 3.3.3 Cubane Derivatives 3.3.4 Halogenocyclopropane Derivatives 3.4 [3,3]-Sigmatropic Rearrangements 3.4.1 With Acyl as Migrating Groups 3.4.2 With Vinyl as Migrating Groups 3.4.3 With Migrating Groups Analogous to Acyl 3.4.4 [3,3]-Sigmatropic Rearrangement and Cyclization Cascades 3.5 [2,3]-Sigmatropic Rearrangements 3.6 [1,2]-Sigmatropic Rearrangements 3.6.1 1,2-Aryl or Alkenyl Migration 3.6.2 1,2-Alkyl Migration 3.6.3 1,2- or 1,5-Alkyl Migration 3.6.4 1,2 versus 3,3 Migrations 3.7 Miscellaneous 3.8 Conclusion References
Silver in Organic Chemistry Edited by Michael Harmata Copyright 2010 John Wiley & Sons, Inc.
83
84
3.1
SIGMATROPIC REARRANGEMENTS AND RELATED PROCESSES
INTRODUCTION
Sigmatropic shifts represent a large class of reactions involving the migration of at least one sigma bond. Therefore, such migrations lead to skeletal rearrangements of the carbon frame within the molecule undergoing this reaction. Sigmatropic rearrangements usually involve s bonds adjacent to a p system or a s bond included in a strained system. As other transition metals, but with specific properties due to its d10 electronic configuration, f orbitals and a relativistic effect,1 silver easily interacts with such systems. Silver salts have thus been explored as catalysts to facilitate and promote sigmatropic rearrangements. 3.2 WOLFF AND ARNDT–EISTERT REARRANGEMENTS AND RELATED REACTIONS The Wolff and Arndt–Eistert rearrangements are probably among the earliest known reactions promoted by silver ions.2,3 Discovered at the turn of the nineteenth/ twentieth century, the Wolff rearrangement allows the transformation of a-diazoketones to carboxylic acids,4 while the Arndt–Eistert rearrangement is a similar sequence also leading to carboxylic acids, but including the preparation of a-diazoketones from a shorter acid chloride (Scheme 3.1).5 Numerous conditions have been developed for this transformation, but reproducible yields have usually been obtained by mixing a silver salt with a coreagent, such as silver nitrate associated with wet ammonia, silver oxide with triethylamine or sodium thiosulfate, and silver benzoate with triethylamine. Nonbasic conditions have also been described by Koch and Podlech using silver trifluoroacetate deposited on silica.6 These modifications have been developed for the homologation of Fmoc-protected amino acids. Mechanistic studies suggested that the in situ reduction of silver salts to silver nanoclusters (Agn) produces the true reagent in the Wolff rearrangement.7 As a result, such reactions have been improved by directly using silver nanoclusters.8 The Wolff rearrangement and the Arndt–Eistert homologation sequence are very useful in organic synthesis. One of the most popular applications involves amino acids. An interesting example has been described as a key reaction in the synthesis of a 14 C-labeled amino acid used for deciphering the biosynthesis of penicillin N from glutamic acid (Scheme 3.2).9 This rearrangement proceeds without racemization and can thus be applied in peptide synthesis. Indeed, applied to a-amino acids, the Arndt–Eistert rearrangement leads to b-amino acids with retention of the configuration of the migrating group. Such
O R
N2
AgI
+O
O R
Ag
Ag+ R
Scheme 3.1
O
H2O
R
OH
85
WOLFF AND ARNDT–EISTERT REARRANGEMENTS AND RELATED REACTIONS
NO2
NO2
O
O
HN *
AgOTf, Et3N
HN *
COOH
THF, H2O, 91%
N2
O
O
O
O O
*=
14C
O
O
NO2
NO2
Scheme 3.2 NHBoc Ph
N2 O
Ag2O, Na 2S2O3
Ph
dioxane, water, 85°C 72 %
NHBoc CO2H
Scheme 3.3
compounds can be incorporated in nonproteinogenic peptides, inducing turns and other special secondary structures (Scheme 3.3).10 Since ketene is probably the intermediate of the Wolff rearrangement, the choice of solvents dictates the nature of the product. Indeed, water gave carboxylic acids, whereas alcohols or amines led to esters and amides, respectively. These combinations have been applied to the synthesis of more complex molecules. For example, the total synthesis of carbonolide B, a 16-membered macrolide antibiotic, relied on Arndt–Eistert homologation. In this sequence, a protected furanuronic acid was transformed to the corresponding a-diazoketone, which was then converted to its homologous carboxylic ester. The reaction was achieved using catalytic amounts of silver benzoate and excess of triethylamine in methanol (Scheme 3.4).11 HO2C MeO
O
O
O
MeO2C
(a)–(c)
O
MeO
O O
O Me CHO MeO O Me
O
OH OH
Scheme 3.4 Reagents and conditions: (a) oxalyl chloride, DMF; (b) CH2N2, Et2O, 76% over two steps; (c) PhCO2Ag (0.3 equiv), Et3N (6 equiv), MeOH, 0–25 C, 3 h, 78%.
86
SIGMATROPIC REARRANGEMENTS AND RELATED PROCESSES
O
+O
O Br
Ar
BF3–2MeOH
Ar
O
R
AgI
Ar
R
OMe
MeOH
Ar
R
R
22– 85 %
Scheme 3.5
N Cl
Ag2CO3, MeOH
O
N
reflux, 20 h 72%
+ OMe
N H
Scheme 3.6
Looking for a new route toward arylalkanoic acids, Giordano et al. observed that silver-assisted methanolysis of primary and secondary a-bromoalkyl aryl ketones led to a mixture of substitution and rearrangement products. Following further studies, these authors developed conditions that mainly provided the rearrangement product, an arylalkanoic ester. The yields were better with electron-rich aromatic groups and with alkyl substituents, suggesting a mechanism similar to the Arndt–Eistert rearrangement (Scheme 3.5).12 The silver-induced reactions of a-haloimines have been compared to those of a-haloketones (Scheme 3.6).13 These silver-assisted reactions of a-haloimines have been interpreted in terms of the a-imidoylcarbenium intermediates, which rearrange in a fashion analogous to that of the corresponding ketones.
3.3
RING REARRANGEMENTS
Strained or bridged cycles substituted by at least one halide could be rearranged on treatment with silver salts. On the other hand, strained s bonds directly interact with silver ions, leading to bond breakage and thus initiating rearrangement. 3.3.1
Halogenoamines
N-Chloramines embedded within bicyclic systems can be easily rearranged into azabridged and enlarged cyclic derivatives. N-Chloroisoquinuclidines, specifically, 2azabicyclo[2.2.2]octane derivatives, or 2-azabicyclo[2.2.1]heptane derivatives gave 1-azabicyclo[3.2.1]heptanes on treatment with silver nitrate in methanol (Scheme 3.7).14 Although the mechanism probably involved chloride abstraction by silver and iminium formation after sigmatropic shift, a concerted pathway catalyzed by silver ions could also be operative, as demonstrated by the preservation
87
RING REARRANGEMENTS n
n AgNO3
N CI
N+
MeOH
n
N
60%
(n
OMe
=
0,1)
Scheme 3.7 D D Cl N D
D
AgClO4 MeOH
Cl
N
Scheme 3.8
of chloride in the rearrangement of deuterated N-chloro-2-azabicyclo[2.2.1]heptane derivatives in the presence of silver perchlorate (Scheme 3.8).15 Further studies revealed that the participation of both s and p electrons played an important role in controlling the stereochemistry of such rearrangements.16,17 Of particular interest were the 7-aza analogs, for which the stereochemistry at the nitrogen atom is well defined, due to a higher inversion barrier relative to other amino derivatives. For such compounds, the silver-promoted rearrangement of each stereoisomer at the nitrogen atom (invertomers) proved to be dependent on the configuration at nitrogen. Thus, the syn invertomer of 7-azabenzonorbornadiene gave the normal solvolysis product, resulting from p participation of the alkene moiety in chloride abstraction by silver ion followed by anti addition of methanol (Scheme 3.9, top). In contrast, the anti isomer led to a rearranged product, a 1-azabicyclo[3.2.0]hept-3-ene derivative. The latter arose through p participation of the benzene unit in chloride loss, followed by methanol opening of the thus-formed multibridged intermediate at the position that regenerated benzene aromaticity (Scheme 3.9, bottom).18–20
N
Cl
H N
N
R
R
R
AgNO3 MeOH
MeOH
syn
OMe
R
Cl
MeOH
N
R
R
MeO
N
R
OMe MeO
R
R AgNO3
N
MeOH
R anti
R
Scheme 3.9
H
R
88
SIGMATROPIC REARRANGEMENTS AND RELATED PROCESSES
Z I N
CO2Bn
N
CO2Bn
AgZ
X
X
N
Z = F, OAc
CO2Bn
X
X = OH or F
X= OH 58%, X= F 67%
Scheme 3.10
In a similar way, 6-exo-iodo or bromo-2-azabicyclo[2.2.0]hexanes rearranged into 5,6-difunctionalized 2-azabicyclo[2.1.1]hexanes on treatment with silver or mercury salts. An aziridinium intermediate was probably formed in the presence of silver ions. Opening of this aziridinium then occurred in a stereocontrolled manner by addition of the silver counterion (Scheme 3.10).21 3.3.2
Cyclopropane Derivatives
Strained cyclic systems such as cyclopropane derivatives proved sensitive to silver ions. Indeed, silver usually induced ring opening of cyclopropanes, leading to argento cationic intermediates, which evolved further depending on substituents. Most, if not all, of these ring rearrangements resulted from initial formation of a s complex of the cyclopropane with silver ion. Bicyclo[1.1.0]butanes are typical of such behavior. They can be considered as joined cyclopropanes and as such, readily react in the presence of catalytic amount of silver salts. These compounds usually lead to 1,3-butadienes (Scheme 3.11).22–24 Screening and kinetic investigations of these rearrangements suggested that the argento cationic intermediate formed on interaction of silver ion with one cyclopropane unit could further rearrange through ring opening of the second cyclopropane unit, but that the latter depended on the nature of the substituent. With alkyl or aryl substituents able to stabilize cations, opening seems to proceed in order to form the most stabilized organosilver cation. If the cation is adjacent to the silver–carbon bond, cleavage of this bond gives the more substituted 1,3-butadiene and regenerates the silver ion (Scheme 3.12, top); otherwise the argento cation is stabilized by a
AgBF4 cat. CDCl 3
99% AgBF4 cat. CDCl3
90%
Scheme 3.11
89
RING REARRANGEMENTS R2
R2 Ag
R1
R1 R2
R2 R1
Ag+
R1
R2
Ag
R1
R1
R1
R1
Ag
R1
R1 R2
1
R1
R1
Ag
R1
R1 = H
R2
Scheme 3.12
sigmatropic shift, especially H shifts, leading to another 1,3-butadiene (Scheme 3.12, bottom). It is worth noting that both secondary argento cation intermediates could be in equilibrium through either an H or vinyl shift, depending on the nature of the R1 group (Scheme 3.12). Parallel studies by Masamune et al. led nevertheless to the conclusion that the argento carbonium ion 1 proposed by Paquette et al. may not necessarily be involved in such AgI-catalyzed rearrangements.25,26 Such rearrangements have been exploited to homologate cyclic dienes, to prepare bridged annulenes and cyclohepta-1,3,5-trienes (Scheme 3.13).27,28 Interestingly, combining such fragmentation and rearrangement with hyperconjugative stabilization of the cyclopropylcarbinyl cation when a silyl or comparable substituent at the adequate position allowed the control of reaction course. Thus, 1-substituted norcarenes were obtained mainly after treatment of silylated or germylated tricyclo[4.1.0.02,7]heptanes with a catalytic amount of silver perchlorate (Scheme 3.14). The “classical” rearrangement product, a 1,3-cycloheptadiene in this case, was nevertheless observed.29 This silver-catalyzed ring rearrangement could be explained by preferential interaction of the silver cation with the most substituted edge of the cyclopropane to form an argento cation. In the latter, the presence of C--Si or Ge bond with the right orientation stabilized the cation sufficiently so that it classically evolved through b-H Br
Br
CHBr3
n
t-BuOK
n
Br
CHBr3
MeLi
Br
n
n
Br
t-BuOK
Ag+
MeLi
Br
Br
Br
Scheme 3.13
Ag+
90
SIGMATROPIC REARRANGEMENTS AND RELATED PROCESSES
Me3m AgClO4 cat.
+
PhH, 40°C
mMe3
mMe3 90 : 10
m = Si, Ge
Scheme 3.14
Ag Ag+ mMe3
mMe3
H mMe3
mMe3
Scheme 3.15
elimination (Scheme 3.15, top), leading to the norcarene derivative. Nevertheless, the argentocyclopropyl cation could fragment as mentioned above and thus give a diene derivative (Scheme 3.15, bottom), The highly strained cycloproparenes behave in the same way in the presence of silver salts. For example, the simplest member of this family gave the benzyl ether in a protic, nucleophilic solvent within a few minutes, while in an anhydrous and nonnucleophilic solvent, such as chloroform, dimerization occurred (Scheme 3.16).30 The even more strained alkylidene cycloproparenes gave rise to the same kind of s-complex intermediate with silver ion. In the presence of alcohol, trapping of this intermediate occurred, leading to alkoxystyrene derivatives. Water could also act in the same way, yielding arylmethylketones after keto–enol equilibration. However, if a proton was present on the alkylidene moiety, H shift occurred, leading to an arylalkyne. No dimerization was observed in this case, probably due to steric constraints in such a process (Scheme 3.17).31 As one would expect because of the increasing strain, cyclopropenes readily reacted with silver ions, leading to ring-opening products. For example, dialkylcyclopropenecarboxylates gave mainly E,E-dienoates together with some isomers when
MeOH AgBF4 1 mol% 0°C
Ag CHCl3
Scheme 3.16
OMe
91
RING REARRANGEMENTS OR' Ph Ph
Ag+
Ph R' = Me R' = Et R' = H R
R'OH R
R
Ag
78% 69% 78%
Ph R=H H 62%
Scheme 3.17
R1
AgBF 4 cat.
COOMe
R2
R1
COOMe
or AgClO 4
R2
Scheme 3.18
submitted to a catalytic amount of silver perchlorate or tetrafluoroborate (Scheme 3.18).32 3,30 -Ethylenebicyclopropenyls have also been investigated as substrates for silver rearrangement because they would have elegantly led to Dewar benzenes by silvercatalyzed rearrangement of the bicyclopropenyl moiety. Unfortunately, treatment of such compounds with catalytic amounts of silver perchlorate in dioxane or deuterated chloroform afforded only polymeric materials. Replacing the solvent with methanol or MeOD induced the opening of one or both cyclopropene rings (Scheme 3.19).33 The mechanistic aspects of the silver(I)-promoted rearrangement of cyclopropene derivatives have been investigated, confirming preferential attack of Ag þ on the s bond to give an argentocarbenium ion. This intermediate is trapped by methanol, leading to a vinylsilver intermediate that is deuterated (Scheme 3.20).33,34
AgClO4 cat. MeOH, –20°C, 1 h 28%
OMe
Scheme 3.19 Ag Ag+
Ag
OMe D
O
Scheme 3.20
D D OMe
92
SIGMATROPIC REARRANGEMENTS AND RELATED PROCESSES
R R
AgClO4
+
C6D6
R
+ R
R = H, CH2OAc, or CO2Me
Scheme 3.21
3.3.3
Cubane Derivatives
The skeleta of other strained systems such as cubanes and related compounds can also be reorganized in the presence of silver salts. Rearrangements of cubanes were described in the early 1970s. Eaton et al.35 showed that mono- or disubstituted cubanes gave a mixture of polycyclic regioisomers (Scheme 3.21). No yields were given. Homocubanes were studied or used in synthesis more extensively. Paquette et al.36,37 showed that the rates of homocubane isomerization to norsnoutane in benzene with silver perchlorate followed a second-order catalytic rate law (Scheme 3.22). Kinetic data suggested that a single site was involved in the silver interaction with homocubane. This new reorganization of homocubyl-caged systems led to highly complex carbon frameworks. Heteroatoms could also be incorporated in the molecular skeleton (Scheme 3.23).38 Unfortunately, this reaction did not always lead to norsnoutane structures. Indeed, treatment of a 9-hydroxyhomocubane ester with silver perchlorate in refluxing benzene did not afford the expected norsnoutane derivative, but led instead to a
AgClO4 C6D6
R
R
Scheme 3.22
O P C H 6 5
O AgBF4 CHCl3, reflux 4 days, 80%
Scheme 3.23
P
C6H5
93
RING REARRANGEMENTS
OH
H OH
AgClO4 PhH, reflux 51%
MeO2C
MeO2C
Scheme 3.24
Ag
Ag MeO2C
MeO2C
MeO2C
MeO2C
MeO2C
Scheme 3.25
AgNO3
CO2Me CO2Me
MeOH, 3 days 100%
CO2Me CO2Me
Scheme 3.26
different polycylic framework (Scheme 3.24). The course of the rearrangement is probably dictated by the stereoelectronic effect of the substituents.39 As for the cyclopropane rearrangement, the reaction probably involved the formation of a cationic silver cationic species followed by regioselective s-bond migration, leading preferentially to the more stable carbocation intermediate that is finally intramolecularly trapped by the organosilver species (Scheme 3.25).39 Bishomocubanes have also been used as substrates for silver-catalyzed rearrangements (Scheme 3.26).40 3.3.4
Halogenocyclopropane Derivatives
Although usually described as driven by halide abstraction by silver ion, dibromocyclopropanes readily rearranged in the presence of silver salts in a process similar to those mentioned above. Such rearrangements usually led to ring expansion products. E-Cyclooctene can easily be converted to its dibromocyclopropyl derivative, and the latter reacted with silver perchlorate in protic solvents, efficiently leading to the corresponding alkoxybromocyclononenes (Table 3.1).41 Stereochemical studies starting from ()-E-cyclooctene suggested the intermediate formation of a bromoallyl cation with a well-defined conformation, to which the nucleophile added (Table 3.1).42,43
94
SIGMATROPIC REARRANGEMENTS AND RELATED PROCESSES
TABLE 3.1. Silver-Catalyzed Formation of Alkoxybromocyclononene Br
H AgClO4
Br
H
H OR
ROH
Br
Br
H
H
R H Me Et Ac
H
Time (min)
Yield (%)
10 10 240 30
82 >99 92 80
Complementary investigations revealed that the stereoselectivity of this ring expansion is indeed very high in favor of the E-(cis)-cycloalkenes when starting from 7- to 9-membered cycloalkenes, but is reversed if the ring size increases (from 10- to 13-membered cycloalkenes), especially in polar solvents.44 Combining cyclopropane opening with a nucleophilic attack could also be performed with an internal nucleophile, thus providing a way to form new rings. An interesting example was described by Danheiser et al.45 in the mid-1980s. Dibromocyclopropanes carrying tethered alcohol or carboxylic acid functional group were efficiently converted to bromovinyl tetrahydrofurans or pyrans as well as to the corresponding lactones on treatment by silver salts (Scheme 3.27). A very interesting extension of these bromocyclopropane opening reactions mediated by silver salts has been described by Collet et al.46 Quinolines and isoquinolines could be homologated through a two-step process, including a silver-catalyzed ring expansion. These heterocycles were first alkylated by the anion derived from bromoform after N-benzylation. The resulting tribromomethyl derivatives were then treated with silver nitrate in a water/methanol solution (Scheme 3.28). The silver ion probably promoted the intermediate formation of a dibromoaziridinium ion. In the quinoline series, methanol added at the vinylic position and opened the aziridinium ion in an SN20 reaction, while in the isoquinoline series, the aziridinium ion opening resulted from a direct SN2 reaction. In both cases, hydrolysis led to the
Br
Br
Br R
n
R OH
AgOOCCF 3 HOCH2CF3 r.t.
X X = H, H or O R = H, Me
Scheme 3.27
R R
n
O X 59–97%
95
[3,3]-SIGMATROPIC REARRANGEMENTS
MeO
HOMe H 2O
AgNO3
N
N
MeOH
CBr3
60%
N
Br Ph Br
Ph
Ph
AgNO3
N
Ph
N
MeOH
CBr3 R
MeOH
Br
O
Ph
Br
Ph
H 2O
N R
R MeO
O
R = H, 44% R = Me, 35%, 1:1 mixture of diastereomers
Scheme 3.28 HO
Br
AgNO3 t-BuOH–H2O
OHC
Scheme 3.29
corresponding seven-membered benzolactams. However, the same silver-catalyzed reaction applied to an optically pure N-phenethylisoquinoline derivative (R ¼ Me) yielded a mixture of diastereoisomers, suggesting a more complex mechanism. In a similar process, although not on cyclopropane, bromoisofenchol has been rearranged to campholenaldehyde, most probably by halide abstraction by silver ion (Scheme 3.29).47 This reaction and related ones have been patented for the production of perfumes and essences.48
3.4 3.4.1
[3,3]-SIGMATROPIC REARRANGEMENTS With Acyl as Migrating Groups
Rearrangements of propargyl esters with silver salts were first mentioned by Zakharova in the mid-1940s.49 He described the conversion of 3-chloro-3-methylbut-1-yne into a mixture of acetates in which the allenic acetate, 1-acetoxy-3methylbut-1,2-diene, was the major compound (Scheme 3.30). Although this product could arise from a silver assisted SN20 reaction, it could also be produced from the substitution product through rearrangement, probably catalyzed by silver ions. The first nonambiguous report of rearrangement of esters derived from propargyl alcohols promoted by silver salts was due to Saucy and coworkers at Hoffmann-La Roche.50 Their observations on the transformation of acetate derived from
96
SIGMATROPIC REARRANGEMENTS AND RELATED PROCESSES
Cl
AgOAc
AcO +
AcOH
•
OAc
H + OAc
OAc
Scheme 3.30
dehydrolinalol clearly showed the superiority of silver salts over copper and gold salts and over zinc oxide, which had been used earlier for this rearrangement.51 Although already postulated in 1956 as [3,3]-sigmatropic rearrangement,52 the mechanism, especially the stereochemical outcome of this reaction, was studied only 10 years later. In an elegant studies on C17-ethynyl steroids of biological relevance, Benn showed that treatment of 17a-ethynylandrost-5-ene-3b, 17b-diol with 5 mol% of silver perchlorate in dry acetone at reflux for 96 h cleanly afforded two isomers in 42% and 23% yields, respectively (Scheme 3.31). This poor selectivity was easily explained by equilibration experiments. Indeed, treatment of either pure acetoxyallene with 5 mol% of silver perchlorate in refluxing acetone for 24 h gave an equimolar mixture of both acetoxyallenes.52 Taking into account this lack of stereoselectivity and equilibration between allenic stereoisomers, Benn proposed a mechanism in which silver ion facilitates the acyloxy shift on formation of p complex (or a bridged ionic silver intermediate).52 The thusformed acetoxyallenes would remain coordinated to silver, allowing the formation of an organosilver allyl cation. Free rotation at this stage would provide both acetoxyallene isomers after reversible loss of silver (Scheme 3.32). Further mechanistic work with optically active and labeled compounds confirmed that the rate-determining step in such rearrangement was the silver coordination to the alkyne moiety and revealed that the silver-catalyzed allene epimerization was 2–40 times faster than the propargyl ester rearrangement.53 Interestingly, these authors53 and later others54 showed the importance of the silver counterion in such rearrangements. While silver tetrafluoroborate or silver nitrate gave the [3,3]-sigmatropic shift product mainly or exclusively, silver trifluoroacetate yielded the dienyl acetate through isomerization from the allenic ester (Scheme 3.33). It is worth noting that a silyl group at the propargyl position did not interfere with the silver-catalyzed rearrangement of a propargyl acetate, despite the potential for further evolution of the product (Scheme 3.34).55
H OAc
• Me2CO, reflux
AcO
Scheme 3.31
AcO
AcO
• +
AgClO4 6 mol%
AcO
OAc
H
97
[3,3]-SIGMATROPIC REARRANGEMENTS
O R1
H R2
O
+ Ag+
O
R1
O R1
•
R2
H
Ag
R2 Ag
OAc H
R1
– Ag+
R2
•
OAc H
OAc R1
H
R2
Ag
R1
•
R2
Ag
H OAc
R1
– Ag +
R2
•
H OAc
Scheme 3.32 AgBF4 10 mol%
p-NO2Ph O
•
H OOCPhp-NO2
PhCl
O
PhCl 10 mol% AgOOCCF3
OOCPhp-NO2
Scheme 3.33
AgBF4 10 mol%
O O H
dioxane
H t-BuMe2Si
•
H OAc
t-BuMe2Si
Scheme 3.34
[3,3]-Sigmatropic rearrangements catalyzed by silver salts have been used in total synthesis. The synthesis of three monoterpene compounds that comprise the pheromone of male boll weevil Anthonomus grandis has been achieved using the silvercarbonate catalyzed rearrangement of an acetoxycyclohexyl acetylene as the key step (Scheme 3.35).56 Ley et al. also applied this method to the synthesis of sesquiterpenes through a strategy involving a Diels–Alder reaction. Taking into account the effect of conditions and silver counterion on allene isomerization (see Scheme 33), they obtained the diene partner via isomerization of the acetoxyallene produced on treatment of a propargyl acetate with catalytic amount of silver hexafluoroantimonate (Scheme 3.36).57
98
SIGMATROPIC REARRANGEMENTS AND RELATED PROCESSES
AcO
OAc
H
O
• Ag2CO3 (5 mol%)
hydrolysis
CH3COOH, 90°C
88%
H
Scheme 3.35
AcO
OAc
H •
AgSbF6 (10 mol%)
OAc reflux
PhCl
Scheme 3.36
3.4.2
With Vinyl as Migrating Groups
Reasoning that [3,3]-sigmatropic shifts of other than acyl groups could a priori be induced by silver, Gore and Malacria demonstrated that propargyl and homoallylic alcohols could be converted to hexadienones in the presence of silver salts through a oxy-Cope-type rearrangement (Table 3.2).58 Silver nitrate and especially silver triflate proved to be the most effective promoters for this reaction. On the basis of
TABLE 3.2. Oxy-Cope-Type Rearrangements R1
O AgOTf
HO
THF–H2O
R1
R2
R2
R1
R2
Temperature ( C)
Time (h)
Yield (%)
H H
H n-C5H11
60 40
1 7
55 73
Me
40
20
57
Me
60
30
55
60
24
40
--(CH2)10--
99
[3,3]-SIGMATROPIC REARRANGEMENTS
R1 R2
R1
R4
O
R2 HO
R3
R4 + H+ – Ag+
+ Ag+ R1
R1
R2
R1
R2 – H+
HO
R3
R3 Ag
H O R3
R4
R4
R2 O R3
Ag
R4 Ag
Scheme 3.37
mechanistic considerations, these authors devised a catalytic version of the reaction, using a combination of silver nitrate and potassium nitrate as catalysts. In contrast to the preceding mechanisms proposed for [3,3]-sigmatropic shifts, the mechanism of the silver-catalyzed oxy-Cope rearrangement was proposed as a stepwise process (Scheme 3.37). As usual, the reaction would be initiated by silver coordination to the alkyne moiety. Nucleophilic attack of this complex by the double bond would then lead to a cyclic cationic vinylsilver intermediate. Fragmentation would then give the dienone. This reaction exhibits strong analogy with the silver-promoted cationic aza-Cope rearrangement described by Overman et al. As a new route toward alkaloids, these authors showed that cyanomethylamines carrying unsaturated chain led to pyrrolidine or piperidine derivatives on treatment with silver salts depending on the substituent (Scheme 3.38).59,60 Further mechanistic investigations led to ambiguous conclusions.61 Rearrangement of stereodefined aminoalcohols led to single product, although epimerization sometimes occurred. Aminoalcohols carrying an electron-deficient vinyl group could be rearranged, although in low yield, while aminoalcohols with electron-rich alkene readily reacted (Table 3.3). In both series, the silver counterion seemed to play an important role, favoring (or not) the rearrangement (Scheme 3.39). Some of these results suggested a classical cationic aza-Cope rearrangement, for which silver only
HO
R1
O
R2
R3
Ag+
N R4
CN
R1
R2 R3
R1 R2
R1 R2
SiMe3 Ag+
N R4
R3
Scheme 3.38
N R4
CN
R3
N R4
100
SIGMATROPIC REARRANGEMENTS AND RELATED PROCESSES
TABLE 3.3. Silver-Mediated Aza-Cope Rearrangements HO
O
R AgNO3 1.1 equiv
N
R
b
N
EtOH
H
Temperature ( C)
Time (h)
Yield (%)
40 25 25 25 25
19 1 1 30 30
64 a 78 76 20 20 b
H Me SPh SO2Ph OEt a
CN
R
Performed in CHCl3–pyridine to avoid epimerization. Performed in ether with AgOOCCF3 (see Scheme 3.39).
O
HO
SO2Ph
AgNO3 1.1 equiv
N H
EtOH N
EtOH CN AgOOCCF3 1.1 equiv
HO
SO2Ph
HO
AgNO3 1.1 equiv
OEt
EtOH
PhO2S
Et2O CN AgOOCCF3 1.1 equiv
N
O N
OEt OEt
H O
N
OEt N H
Scheme 3.39
initiated the cascade of events by generating a key iminium intermediate, while others support a more complex role of silver in such rearrangements. Nevertheless, such rearrangements offer a concise route to some alkaloids. For example, the total synthesis of crinine was achieved with the silver-promoted formation of hydroindolone as the key step (Scheme 3.40).59 O
O
O
O
OH OSiMe3
O
HO
O
AgNO3 1.1 equiv EtOH
N
OSiMe3 Ph
CN
H 94%
Scheme 3.40
N
O
N Ph
Crinine
H
101
[3,3]-SIGMATROPIC REARRANGEMENTS
R1
O AgBF 4 10 mol% or
O
OEt
O
•
R2
O
AgSbF 6 5 mol% CH2Cl2, r.t. 100%
R2
R1
OEt
H H
Scheme 3.41
O
H
CHO
AgBF4 (cat.)
OTBS R
PhH, 25°C
98%
R
O
OTBS
MeO2C
CHO AgBF4 (cat.)
R
R=
•
•
PhH, 25°C
CHO slow isomerization
R
R
Scheme 3.42
In another related and well-known [3,3]-sigmatropic shift usually performed under thermal conditions, the propargyl–Claisen rearrangement,62 silver salts were also able to catalyze the reaction. Silver tetrafluoroborate and hexafluoroantimonate proved to be the best catalysts for this reaction, leading quantitatively to allenic b-ketoesters when starting from propargyl ethers derived from b-ketoesters (Scheme 3.41).63,64 To improve the thermal tandem [3,3]-sigmatropic rearrangement–enyne allene cyclization, Grissom et al.65 focused their attention on the transformation of 2propynylvinyl ethers to allenes using silver salts as catalysts. The tetrafluoroborate proved to be the most effective, quantitatively yielding the expected allenyl aldehyde (Scheme 3.42). However, isomerization problems appeared in some cases (see Scheme 3.33). 3.4.3
With Migrating Groups Analogous to Acyl
The ease of silver-promoted [3,3]-sigmatropic rearrangements of propargyl acetates and other esters has led to the investigation of the sigmatropic shift of groups electronically analogous to acyloxy groups. As for the thermal versions, phosphates have been used as such analogs, and they compared favorably to acetates and benzoates (Table 3.4). Sulfonates were also used but seemed less reactive.66 In 2004, Gevorgyan et al. reported that a phosphatyloxy alkynone was also able to smoothly give a phosphatyloxy allenone under silver catalysis. The authors reported only a single example (Scheme 3.43).67
102
SIGMATROPIC REARRANGEMENTS AND RELATED PROCESSES
TABLE 3.4. Silver-Catalyzed [3,3]-Sigmatropic Rearrangements X R2
R1
AgClO4 2.5 mol%
R1
R3
X
•
CH2Cl2, 35°C, 2 h
R2
R3
X
R1
R2
R3
Yield (%)
CH3CO2 — — (C2H5O)2P(O)O — — p-NO2PhCO2 — CF3CO2
Me Me t-C4H9 Me Me t-C4H9 Me Me Me
Me Me Me Me Me Me Me Me Me
Me H H Me H H Me H H
46 68 63 62 54 47 46 37 33
t-Bu
OP(O)(OEt)2 Ph Ph O
AgBF4 5 mol%
(EtO)2(O)PO
Ph
•
t-Bu
CH2Cl2, r.t.
O
56%
Ph
Scheme 3.43
3.4.4
[3,3]-Sigmatropic Rearrangement and Cyclization Cascades
In more sophisticated approaches, [3,3]-sigmatropic rearrangements have been incorporated in cascade reactions in which each step could be catalyzed by silver salts. Usually, the silver promoted [3,3]-sigmatropic shift is combined with a silvercatalyzed cyclization,68 as outlined in the general scheme below for alkynyl compounds (Scheme 3.44). Hiyama et al.69 showed that monoacetylated 1,4-butynediols nicely afforded acetoxy allenols on silver-catalyzed rearrangement. These compounds were ideal substrates for silver-catalyzed cyclization, so the overall sequence could directly be performed in one pot, leading to substituted 2,5-dihydrofurans in high overall yields (Table 3.5). In this process, 5–10 mol% of silver perchlorate or tetrafluoroborate was used in refluxing benzene.
R MG
R
R1 R2
AgI
Nu
[3,3]-sigmatropic shift
MG
• H
MG
AgI cyclization
Nu
Scheme 3.44
R2
R1
R
Nu
R1 R2
103
[3,3]-SIGMATROPIC REARRANGEMENTS
TABLE 3.5. Silver-Catalyzed 2,5-Dihydrofurans Enolacetate Preparation R1 R2
R AcO
Entry
OH
OAc AgClO 4 5 mol% PhH, reflux
R
Starting Material
O
Products
Ph
1
R1 R2
Yield (%)
OAc
AcO
OH
Ph
84
O OAc
2
AcO
99
OH O OAc
3
63 AcO
OH
O OAc
4
80 AcO
OH
O OAc
5
74 AcO
OH
O
PMBO
OAc
6
63
PMBO AcO
O
OH
O
7
OAc
61
O O
AcO
OH
O
O
O
O
This rearrangement–cyclization cascade was applied to natural products synthesis. For example, DDQ oxidation of the appropriate dihydro-3(2H)-furanone enol acetates (Table 3.5, entries 1 and 7) afforded bullatenone and the antitumor agent geiparvarin in a very rapid route (Fig. 3.1).69,70 Toward the total synthesis of ()-ascofuranone (Scheme 3.45), an antibiotic with hypolipidemic, antihypertensive, and antitumor properties, several improvements have been made. A pivaloyl ester was used instead of an acetyl ester.71 More
104
SIGMATROPIC REARRANGEMENTS AND RELATED PROCESSES
O O Ph
O
O O
O
O
Bullarenone
Geiparvarin
Figure 3.1. Synthetic applications of the rearrangement–cyclization cascade toward natural products.
importantly, chirality transfer was studied and showed to be quantitative, indicating that cyclization was more rapid than isomerization of the allene intermediate. Indeed, various enantiomerically enriched tert-butylcarbonyloxybutynols gave the expected 2,5-dihydropyrans with almost the same enantiomeric excess. The chirality was introduced via asymmetric reduction of the starting alkynones with (S)-Alpine Borane (Aldrich-Sigma).72–74 Another cascade reaction was also nicely set up using alkynones instead of alkynols. Allenones were produced in situ through a [3,3]-sigmatropic rearrangement of propargyl acetates and led directly, under these conditions, to substituted furans (Table 3.6).77 Interestingly, the key allenone could be isolated when the reaction was run at room temperature and with a phosphate as the migrating group (see above). In refluxing dichloroethane, this allenone readily cyclized to the corresponding furan. Moreover, both phosphate and sulfonate groups could be engaged in the same reaction, leading to tetrasubstituted furans from substituted allenones (Scheme 3.46). This cascade proved to be quite complex, successively involving a [3,3]-sigmatropic shift of the propargyl acetate, 1,2 migration of this acetate, and cycloisomerization (Scheme 3.47). The silver-catalyzed, propargyl Claisen rearrangement has also been combined with cyclization reactions. Propargyl ethers derived from b-ketoesters were rearranged by silver hexafluoroantimonate and the resulting allenic b-ketoesters cyclized on treatment with base, leading to the formation of 2H-pyrans in moderate to excellent yields (Scheme 3.48).75 The cyclization process was described as a base-catalyzed
OH
1. (S)-Alpine Borane
THPO
2. t-BuCOCl, DMAP
O
OH
H THPO ee 84% (on the diol)
O
t-Bu O
AgBF4
65%
OH O
Cl
O
THPO Me
OH CHO
O ee 84%
Ascofuranone
Scheme 3.45
OCO t-Bu
105
[3,3]-SIGMATROPIC REARRANGEMENTS
TABLE 3.6. Silver-Catalyzed Preparation of Tetrasubstituted Furans AcO t-Bu
R2
AgBF4 5 mol% CH2Cl2, r.t.
R1
R2
• t-Bu
1
Substrate
Product AcO
OAc Ph
t-Bu
Ph
t-Bu
O
2
OAc Ph
t-Bu
OAc Me
t-Bu
4
Me
AcO
5
2
>99
15
73
15
84
15
90
10
86
Ph
O
t-Bu
O
Yield (%)
Me
AcO
OAc Me
t-Bu
Time (min)
Me
O
t-Bu
O
R1
Ph
AcO
Ph
O
Ph
O
t-Bu
O
t-Bu
Ph
AcO
Me
3
R1
O
O
Entry
R2
AcO
AcO
Me O
Me
AcO
t-Bu t-Bu
O
O
isomerization followed by a 6p-oxa electrocyclization, although silver-catalyzed cyclization of the allenone could occur. Again on the basis of the silver-catalyzed propargyl Claisen rearrangement, an interesting one-pot process leading to highly substituted pyrroles was devised by Kirsch et al.64 The allenic b-ketoester obtained through this rearrangement was directly treated with amines and then with a catalytic amount of a gold catalyst. The X t-Bu
X
Ph
•
Ph
AgBF4
Ph
DCE, 60ºC
O
t-Bu
X = –OP(O)(OEt)2, 77% yield X = –OTs, 82% yield
Scheme 3.46
O
Ph
106
SIGMATROPIC REARRANGEMENTS AND RELATED PROCESSES
AcO
t-Bu
R2
AcO
R2 R1
t-Bu
O
O
R1
Ag+ O
O
O •
t-Bu
Ag+
R2 R1
R2 R1
O
t-Bu
O
O
Scheme 3.47 Me
Me CO2Et
O
23°C, CH2Cl2
i-Pr
CO2Et
O
AgSbF6 5 mol%
•
Ph
Ph
Me O
DBU 5 mol%
80%
i-Pr
CO2Et
then
i-Pr
Ph
Scheme 3.48
intermediate enaminoester added to the external double bond of the allenic system in a 5-exo-dig cyclization, giving pyrroles in good to high yields (Scheme 3.49). A cascade reaction relying on the Claisen rearrangement and cyclization, both catalyzed by copper or silver salts, has been described. Phenyl allyl ether rearranged in the presence of catalytic amounts of silver triflate or perchlorate in refluxing toluene, and the resulting 2-allylphenol cyclized to give the corresponding dihydrobenzofuran (Scheme 3.50).76
R1
R1
O
O
1. AgSbF 6 5 mol%
OEt
O
O
CH2Cl2, r.t.
R1 OEt
•
3. (Ph 3P)AuCl 5 mol%
R2
H H
R2
2. R 3NH2
R3
31–90%
Ag+
AgClO 4 (cat.) CHCl3, reflux
N R2
Scheme 3.49
O
COOEt
OH
Scheme 3.50
64%
O
107
[2,3]-SIGMATROPIC REARRANGEMENTS
3.5
[2,3]-SIGMATROPIC REARRANGEMENTS
Following their investigations on nitrene, carbene, and oxo transfer reactions catalyzed by fluorinated silver tris(pyrazoyl)borate (see Chapter 6 on nitrene chemistry), Lovely et al. looked for a combined carbene transfer and [2,3]-sigmatropic rearrangement. On the basis of these mechanistic considerations, these authors showed that diazoacetates, indeed, reacted with allyl halides in the presence of this silver catalyst to give a-halo-g-unsaturated esters (Scheme 3.51).77 The substitution pattern in the product obtained through this sequence was in agreement with the proposed mechanism (Table 3.7). For example, crotyl chloride gave the 2-chloro-3-methyl hex-4-enoates (entries 3 and 4), and propargyl bromides gave the allenic products (e.g., entries 7 and 8). Yields were usually better with the more substituted tert -butyl diazoacetate.
R2O X
N2
R1
O
O X
COOR2
X+
AgI
OR2
R1
R1
Scheme 3.51
TABLE 3.7. Silver-Catalyzed Preparation of a-Haloacetates Entry
R2
Halide
1
Product
Yield (%)
Br
75
Et
Br
2
t-Bu
3
Et
CI
COOR2
65 a
86
CI COOR2
96 a
t-Bu
4 5
Br
Et Br
8 a
Br
t-Bu
7
70 COOR2
Br
6
a
Br
Et
t-Bu
Obtained as a 1 : 1 mixture of syn : anti isomers.
H H
80 a 71
Br
•
COOR2
74
108
SIGMATROPIC REARRANGEMENTS AND RELATED PROCESSES
TABLE 3.8. Silver-Catalyzed Preparation of Allenic Sulfones pTol R1
O S
AgSbF6 , 2 mol%
O R2
R3
pTolSO2
MeNO2, r.t., 2 min
•
R1
R2 R3
Entry
R1
R2
R3
Yield (%)
1 2 3 4 5 6 7 8 9 10 11 12
H H H H Me H H Ph H Et H H
H Me H H H Me Et H Ph H
Me Me H Et H Et Et H H Me
97 97 98 96 99 97 97 99 99 98 98 99
--(CH2)4-H
Me
In a related reaction, Harmata and Huang showed that allenic sulfones were easily prepared from propargylic sulfinates (Table 3.8). Catalyst screening showed that silver nitrate was ineffective, whereas silver tetrafluoroborate and hexafluoroantimonate worked well in dichloromethane, chloroform, or nitromethane at room temperature. All reactions were very rapid, within 2 min, and were very effective in terms of yield.78
3.6 3.6.1
[1,2]-SIGMATROPIC REARRANGEMENTS 1,2-Aryl or Alkenyl Migration
Extending their work on a-bromoaryl ketones (see Scheme 3.5), Giordano et al. have reported a [1,2]-sigmatropic rearrangement assisted by silver ion.79 Indeed, alkylacetals of primary and secondary a-halogenated aryl ketones furnished alkyl esters of a-aryl alkanoic acids in high yields using silver tetrafluoroborate in an alcoholic medium (Table 3.9). More recently, Usui et al. studied the stereochemical aspects of this rearrangement.80 (S)-2-Phenylpropionic acid was stereoselectively obtained by the AgBF4catalyzed rearrangement of (S)-2-chloropropiophenone dimethyl acetal with a complete chirality transfer, suggesting that the rearrangement proceeds with the backside phenyl group participation at the C--Cl bond, that is, with an Ag þ -aided, phenylassisted intramolecular SN2 mechanism (Scheme 3.52).
109
[1,2]-SIGMATROPIC REARRANGEMENTS
TABLE 3.9. Silver-Mediated [1,2]-Sigmatropic Rearrangements R2O OR2 R1 Ar X
R1
AgBF4 1.5 equiv R2OH
+ AgX CO2R2
Ar
Aryl Group
R1
R2
X
Time (h)
Yield (%)
4-Methoxyphenyl 4-Methoxyphenyl 4-Methoxyphenyl 4-Methoxyphenyl 4-Methoxyphenyl 2-Naphthyl
H H H H Me H
Me Me Me Et Me Me
Br I Cl Br Br Br
2 0.25 2 3 1.5 14
98 87 98 98 98 98
MeO OMe
O
AgBF 4 25 mol%
OMe
MeOH
Cl
+
O
OMe 20%
47%–97% ee (S )
98% ee (S )
Scheme 3.52
Similarly, silver hexafluoroantimonate was also able to promote a fast [1,2]sigmatropic rearrangement of trans-4-aryl- or trans-4-alkenyl-3-bromo-4,6-dimethyl3,4-dihydro-2-pyrones in dichloromethane through substitution of the halide by migration of the aryl or alkenyl group to the 3 position. This rearrangement afforded the corresponding 3-substituted 2-pyrones in high yields (Table 3.10).81 It is noteworthy that the debromination–migration process is probably concerted as supported by the fact that cis compounds did not take part in this reaction.
TABLE 3.10. Silver-Assisted Preparation of a-Pyrones Me R
Me
R p-Tolyl p-Anisyl Phenyl 1-Propenyl (Z)-1-Hexenyl
O
Br H O
Me
1. AgSbF 6 1 equiv CH2Cl2 2. 2,6-lutidine
R Me
O
cis : trans
Time (min)
97 : 3 85 : 15 80 : 20 97 : 3 97 : 3
15 15 15 15 15
O
Yield (%) 100 80 47 84 93
110
3.6.2
SIGMATROPIC REARRANGEMENTS AND RELATED PROCESSES
1,2-Alkyl Migration
Kirsch et al.82 have discovered that numerous metal salts, including silver salts, were able to catalyze the rearrangement of a-hydroxyalkynones to 3(2H)-furanones, introducing an original strategy for the construction of such compounds. Interestingly, spirocyclic furanones could be obtained starting from alkynyl cycloalkanones (Scheme 3.53). Although not the best catalyst, AgSbF6 led to the rearranged product in 50% yield. This cascade reaction probably started with the well-known cyclization of the ketone to the alkyne on silver coordination, giving a cyclic oxonium intermediate that rearranged to furanone via an alkyl 1,2-migration (Scheme 3.54).82 A similar reaction was described by Nakamura et al. as recently as 2009. Here again, silver salts were not the best catalysts. Nevertheless, silver triflate was able to rearrange 2-alkynyl tetralones to 1,2-disubstituted naphthalenes (Scheme 3.55).83 O
O
Ph
OH
AgSbF6 5 mol% CH2Cl2, r.t.
O
Ph
50%
Scheme 3.53
3.6.3
1,2- or 1,5-Alkyl Migration
Gevorgyan et al. also observed alkyl migration while studying the cyclization of substituted allenones in the presence of various metal salts. While again not the best catalysts, silver hexafluorophosphate or triflate could be used in toluene or dichloromethane, giving substituted furans in high yields (Scheme 3.56).84 R3
O
R3
O AgSbF6
R1
R1 R2
R2 OH
OH
R3
R3
O
Ag O H R2
O
O
R1 R2
Ag
R1
Scheme 3.54 O R2 R1
R2
Ph AgOTf 10 mol%
O
PhH, 35°C
R1
Scheme 3.55
Ph
111
[1,2]-SIGMATROPIC REARRANGEMENTS
R1
R1
•
R2 Ag+
R1
AgOTf
R3
– Ag+ Ag
•
R3
O
π-complex
Ag
R2
R2
Toluene, 100°C 80 %
O
R
R3
R2
O
R1
1
R2
R3
O
Ag O
R3
[1,2] or [1,5]-shift
Scheme 3.56 Ph Ph • Ph
AgOTf 10%
H O
C6H4-pCN
TTBP 40% DCE 80°C
Ph
O
C6H4-p CN
99% (NMR yield)
Scheme 3.57
From a mechanistic perspective, the reaction is similar to those described above. Silver-catalyzed cyclization of the ketone to the allene gave a cyclic oxonium intermediate. A [1,2]- or [1,5]-alkyl shift modified the sigma skeleton leading to an alkylsilver intermediate, which on elimination gave a trisubstituted furan. In the presence of hindered base, such as 2,4,6-tris(tert-butyl)pyrimidine (TTBP), and in refluxing dichloroethane, this reaction became quantitative (Scheme 3.57).85 3.6.4
1,2 versus 3,3 Migrations
Although acyloxy, phosphatyloxy, and tosyl allenones were obtained from the corresponding propargyl alcohol derivatives via silver catalysis through an overall process that resembled [3,3]-sigmatropic rearrangement (see Sections 3.41 and 3.43), the mechanism was not fully proved and is still in question. The same holds true for the cyclization of these allenones to trisubstituted furans (see Section 3.4.4). More recently, new insights were obtained using labeled substrates.86 Two processes could explain the position of the migrating group on the furan ring: a [3,3]-sigmatropic rearrangement followed by a 1,2 shift or a direct 1,2 shift leading to a carbenoid intermediate (Scheme 3.58). The position of the labeled oxygen atom in the final furan product would confirm the pathway used: double inversion leading to the unlabeled oxygen atom linked to the carbon (pathway A) or simple inversion leading to the labeled oxygen atom link to the carbon (pathway B). Treatment of labeled acyloxy alkynones with silver tetrafluoroborate in dichloromethane at room temperature exclusively gave a single furan, in which the labeled oxygen atom was connected to the ring carbon (Scheme 3.59). This observation was consistent with a 1,2-acyl migration but not with a [3,3]-sigmatropic rearrangement.
112
SIGMATROPIC REARRANGEMENTS AND RELATED PROCESSES
X O
17O
17O
• 17O
X O
t-Bu
t-Bu
[3,3]
R2
O
X
[1,2]
R2
Pathway A
R1
t-Bu
O
R1
O
R2 R1 O
O [1,2]
17 X O
O
R2
17O
R2
X
Pathway B
R1
t-Bu
t-Bu
M O
R1
O
Scheme 3.58 17O 17O
O
t-Bu
Ph
O
O
Ph
AgBF4 5 mol%
Ph
+
CH2Cl2, r.t. 69%
Ph
17O
t-Bu
O
t-Bu
Ph
O
Ph
O
Scheme 3.59
Similarly, a single phosphatyloxy allene in which the labeled oxygen atom was not connected to the allene was obtained from a phosphatyloxy alkyone after treatment with silver tetrafluoroborate. This observation definitively ruled out a [3,3]-sigmatropic rearrangement and supported 1,2 migration. It is worth noting that during cyclization of the phosphatyloxy allenone, the labeled oxygen atom was again not connected to the furan ring. The result strongly suggested 1,2 migration with retention (Scheme 3.60).
t-Bu
OEt OEt 17O P O Ph
OEt P OEt O Ph
17O
OEt O P OEt 17O Ph
AgBF4 5 mol%
+
DCE, 60ºC 60%
Ph
t-Bu
O
Ph
t-Bu
O
O major
AgBF4 5 mol% DCE, r.t.
90%
77%
EtO 17O EtO P O
Ph
• t-Bu
O
Ph
Scheme 3.60
minor
AgBF4 5 mol% DCE, 60ºC
Ph
113
CONCLUSION
O O
O
AgOTf 7.5 equiv PhH, 25–80°C Quantitative
O
Scheme 3.61
These results clearly demonstrated that, depending on substrates, apparent [3,3]sigmatropic shifts actually resulting from 1,2 shifts could occur. Further studies are thus clearly needed to better understand the mechanism of apparent sigmatropic rearrangements catalyzed by silver and other metals. 3.7
MISCELLANEOUS
Looking for new complexes, Yamaguchi et al. observed that a cyclic diaryltetrayne underwent rearrangement in the presence of 2.5 equiv of silver triflate to a difuroperylene, which was isolated in 30% yield. The reaction could be made nearly quantitative by addition of two other portions of silver triflate after 3 and 6 h while stirring (Scheme 3.61). The formation of perylene proceeds by a complex silvermediated process involving sp–sp and sp–sp2 cyclization.87 3.8
CONCLUSION
Silver salts can be used as catalysts or promoters for a variety of reactions involving s-bond migration and skeletal rearrangement. Bonds in strained systems directly interact with silver, leading to bond breaking and formation of a cation. These events lead to skeletal rearrangement through bond migration. In a similar way, halogenated cyclic compounds can also be rearranged on treatment with silver ions. The initiation of such rearrangements is often similar to the preceding, since strained cyclic intermediates are usually produced as the first step. Silver-mediated rearrangement of propargyl esters is a process that has been known since the mid-1940s. Since then, this [3,3]-sigmatropic rearrangement has been extended to other sigmatropic rearrangements with a wide variety of participating groups and different applications. This chemistry has been revamped more recently by its combination with cyclization reactions, most of them also promoted by silver salts.69 More recent years have thus witnessed the development of silver-catalyzed cascade reactions leading to heterocycles under mild conditions.68 This broad range of reactions mediated by silver reveals the importance of silver salts as catalysts or promoters in organic chemistry and especially in synthesis.
114
SIGMATROPIC REARRANGEMENTS AND RELATED PROCESSES
REFERENCES 1. (a) Bagus, P. S.; Lee, Y. S.; Pitzer, K. S., Chem. Phys. Lett. 1975, 33, 408–411; (b) Pyykk€ o, P., Angew. Chem., Int. Ed. 2004, 43, 4412–4456; (c) Gorin, D. J.; Toste, F. D., Nature 2007, 446, 395–403. 2. Wolff, L., Justus Liebigs Ann. Chem. 1902, 325, 129–195. 3. Arndt, F.; Eistert, B., Ber. Deutsch. Chem. Ges. 1935, 68, 200–208. 4. Gill, G. B., in Comprehensive Organic Synthesis, Trost, B. M., Fleming, I., Pattenden, G., eds., Pergamon Press, New York, 1991, Vol. 3, pp. 887–912. 5. Meier, H.; Zeller, K. P., Angew. Chem. Int. Ed. 1975, 14, 32–43. 6. Koch, K.; Podlech, J., Synth. Commun. 2005, 35, 2789–2794. 7. Sudrik, S. G.; Maddanimath, T.; Chaki, N. K.; Chavan, S. P.; Chavan, S. P.; Sonawane, H. R.; Vijayamohanan, K., Org. Lett. 2003, 5, 2355–2358. 8. Sudrik, S. G.; Chaki, N. K.; Chavan, V. B.; Chavan, S. P.; Chavan, S. P.; Sonawane, H. R.; Vijayamohanan, K., Chem. Eur. J. 2006, 12, 859–864. 9. Baldwin, J. B.; Herchen, S. R.; Singh, P. D., Biochem. J. 1980, 186, 881–887. 10. Plucinska, K.; Liberek, B., Tetrahedron 1987, 43, 3509–3517. 11. Tilekar, J. N.; Patil, N. T.; Dhavale, D. D., Synthesis 2000, 395–398. 12. Giordano, C.; Castaldi, G.; Casagrande, F.; Abis, L.; Donegani, I. G., Tetrahedron Lett. 1982, 23, 1385–1386. 13. De Kimpe, N.; Stevens, C., Tetrahedron 1990, 46, 6753–6770. 14. Gassman, P. G.; Fox, B., J. Am. Chem. Soc. 1967, 89, 338–342. 15. Gassman, P. G.; Cryberg, R. L., J. Am. Chem. Soc. 1968, 90, 1355–1356. 16. Davies, J. W.; Malpass, J. R.; Walker, M. P., J. Chem. Soc. Chem. Commun. 1985, 685–686. 17. Durrant, M. L.; Malpass, J. R.; Walker, M. P., J. Chem. Soc. Chem. Commun. 1985, 687–689. 18. Rautenstrauch, V., Chem. Commun. 1969, 1122–1123. 19. Durrant, M. L.; Malpass, J. R., Tetrahedron 1995, 51, 7063–7076. 20. Davies, J. W.; Durrant, M. L.; Naylor, A.; Malpass, J. R., Tetrahedron 1995, 51, 8655–8664. 21. Krow, G. R.; Lin, G.; Yu, F.; Sonnet, P. E., Org. Lett. 2003, 5, 2739–2741. 22. Paquette, L. A.; Wilson, S. E.; Henzel, R. P., J. Am. Chem. Soc. 1972, 94, 7771–7779. 23. Paquette, L. A.; Henzel, R. P.; Wilson, S. E., J. Am. Chem. Soc. 1972, 94, 7780–7788. 24. Paquette, L. A., Angew. Chem., Int. Ed. 1972, 11, 328–329. 25. Masamune, S.; Sakai, M., J. Am. Chem. Soc. 1971, 93, 4610–4611. 26. Masamune, S.; Sakai, M.; Westberg, H. H.; Yamaguchi, H., J. Am. Chem. Soc. 1971, 93, 4611–4613. 27. Paquette, L. A.; Blount, J. F., J. Am. Chem. Soc. 1980, 102, 644–650. 28. Christl, M.; Brunn, E.; Roth, W. R.; Lennartz, H. W., Tetrahedron 1989, 45, 2905–2915. 29. Taylor, R. T.; Paquette, L. A., J. Org. Chem. 1978, 43, 242–250. 30. Billups, W. E.; McCord, D. J.; Maughon, B. R., Tetrahedron Lett. 1994, 35, 4493–4496. 31. Halton, B.; Dixon, G. M.; Forman, G. S., Arkivoc 2006, xii, 38–45. 32. Leftin, J. H.; Gil-Av, E., Tetrahedron Lett. 1972, 13, 3367–3370.
REFERENCES
115
33. Peelen, F. C.; Landheer, I. J.; De Wolf, W. H.; Bickelhaupt, F., Rec. Trav. Chim. Pays-Bas 1986, 105, 326–331. 34. Padwa, A.; Blacklock, T. J.; Loza, R., J. Org. Chem. 1982, 47, 3712–3721. 35. Eaton, P. E.; Cassar, L.; Halpern, J., J. Am. Chem. Soc. 1970, 92, 6366–6368. 36. Paquette, L. A.; Ward, J. S., Tetrahedron Lett. 1972, 13, 4909–4912. 37. Paquette, L. A.; Ward, J. S.; Boggs, R. A.; Farnham, W. B., J. Am. Chem. Soc. 1975, 97, 1101–1112. 38. Turnblom, E. W.; Katz, T. J., J. Am. Chem. Soc. 1973, 95, 4292–4311. 39. Mehta, G.; Ravikrishna, C.; Ravikumar, K., J. Chem. Soc. Chem. Commun. 1994, 2321–2322. 40. Dauben, W. G.; Buzzolini, M. G.; Schallhorn, C. H.; Whalen, D. L., Tetrahedron Lett. 1970, 10, 787–790. 41. Reese, C. B.; Shaw, A., J. Chem. Soc. Perkin Trans. I 1975, 2422–2435. 42. Bach, R. D.; Holubka, J. W.; Willis, C. L., J. Am. Chem. Soc. 1982, 104, 3980–3987. 43. Bach, R. D.; Mazur, U.; Hamama, I.; Lauderback, S. K., Tetrahedron 1972, 28, 1955–1963. 44. Ito, S.; Ziffer, H.; Bax, A., J. Org. Chem. 1986, 51, 1130–1133. 45. Danheiser, R. L.; Morin, J. M.; Yu, M.; Basak, A., Tetrahedron Lett. 1981, 22, 4205–4208. 46. Ludivine, J.-G.; Pauvert, M.; Collet, S.; Guingant, A.; Evain, M., Tetrahedron 2007, 63, 11250–11259. 47. Anhalt, K.; Sprung, I.; Schulze, K., Monatsh. Chem. 2003, 134, 1593–1606. 48. Schulze, K.; Uhlig, H., German (East) Patent (1988), GEXXA8 DD254382 (CAN 111:23747). 49. Zakharova, A. I., Zh. Obsch. Khim. 1945, 15, 429–437. 50. Saucy, G.; Marbet, R.; Lindlar, H.; Isler, O., Helv. Chim. Acta 1959, 42, 1945–1955. 51. Landor, P. D.; Landor, S. R., J. Chem. Soc. 1956, 1015–1019. 52. Benn, W. R., J. Org. Chem. 1968, 33, 3113–3118. 53. Schlossarcyk, H.; Sieber, W.; Hesse, M.; Hansen, H.-J.; Schmid, H., Helv. Chim. Acta 1973, 56, 875–944. 54. Cookson, R. C.; Cramp, M. C.; Parsons, P. J., J. Chem. Soc. Chem. Commun. 1980, 197–198. 55. Sakaguchi, K.; Okada, T.; Shinada, T.; Ohfune, Y., Tetrahedron Lett. 2008, 49, 25–28. 56. Pelletier, S. W.; Mody, N. V., J. Org. Chem. 1976, 41, 1069–1071. 57. Hollinshead, D. M.; Howell, S. C.; Ley, S. V.; Mahon, M.; Ratcliffe, N. M., J. Chem. Soc. Perkin Trans. 1 1983, 1579–1589. 58. Bluthe, N.; Gore, J.; Malacria, M., Tetrahedron 1986, 42, 1333–1344. 59. Overman, L. E.; Mendelson, L. T.; Jacobsen, E. J., J. Am. Chem. Soc. 1983, 105, 6629–6637. 60. Flann, C.; Malone, T. C.; Overman, L. E., J. Am. Chem. Soc. 1987, 109, 6097–6107. 61. Jacobsen, E. J.; Levin, J.; Overman, L. E., J. Am. Chem. Soc. 1988, 110, 4329–4336. 62. Overman, L. E., Angew. Chem., Int. Ed., 1984, 23, 579–586. 63. Suhre, M. R.; Reif, M.; Kirsch, S. F., Org Lett. 2005, 7, 3925–3927. 64. Binder, J. T.; Kirsch, S. F., Org Lett. 2006, 8, 2151–2153.
116
SIGMATROPIC REARRANGEMENTS AND RELATED PROCESSES
65. Grissom, J. W.; Klingberg, D.; Huang, D.; Slattery, B. J., J. Org. Chem. 1997, 62, 603–626. 66. Oelberg, D. G.; Schiavelli, M. D., J. Org. Chem. 1977, 42, 1804–1807. 67. Sromek, A. W.; Kelin, A. V.; Gevorgyan, V., Angew. Chem. Int. Ed. 2004, 43, 2280– 2282. 68. (a) Weibel, J.-M.; A. Blanc, A.; Pale, P., Chem. Rev. 2008, 108, 3149–3173; (b) AlvarezCorral, M.; Munoz-Dorado, M.; Rodriguez-Garcia, I., Chem. Rev. 2008, 108, 3174–3198. 69. Saimoto, H.; Hiyama, T.; Nozaki, H., J. Am. Chem. Soc. 1981, 103, 4975–4977. 70. Saimoto, H.; Hiyama, T.; Nozaki, H., Bull. Chem. Soc. Jpn. 1983, 56, 3078–3087. 71. Saimoto, H.; Kusano, Y.; Hiyama, T., Tetrahedron Lett. 1986, 27, 1607–1610. 72. Shigemasa, Y.; Yasui, M.; Ohrai, S.-I.; Sasaki, M.; Sashiwa, H.; Saimoto, H., J. Org. Chem. 1991, 56, 910–913. 73. Saimoto, H.; Ohrai, S.-I.; Sashiwa, H.; Shigemasa, Y.; Hiyama, T., Bull. Chem. Soc. Jpn. 1995, 68, 2727–2734. 74. Saimoto, H.; Masaru, Y.; Ohrai, S.-I.; Oikawa, H.; Yokoyama, K.; Shigemasa, Y., Bull. Chem. Soc. Jpn. 1999, 72, 279–284. 75. Menz, H.; Kirsch, S. F., Org. Lett. 2006, 8, 4795–4797. 76. Ito, Y.; Kato, R.; Hamashima, K.; Kataoka, Y.; Oe, Y.; Ohta, T.; Furukawa, I., J. Organomet. Chem. 2007, 692, 691–697. 77. Krishnamoorthy, P.; Browning, G. R.; Singh, S.; Siuvappa, R.; Lovely, C. J.; Dias, H. V. R., Chem. Commun. 2007, 731–733. 78. Harmata, M.; Huang, C., Adv. Synth. Catal. 2008, 350, 972–974. 79. Giordano, C.; Castaldi, G.; Casagrande, F.; Belli, A., J. Chem. Soc. Perkin Trans. 1 1982, 2575–2581. 80. Usui, S.; Matsumoto, T.; Ohkubo, K., Tetrahedron Lett. 1998, 39, 9755–9758. 81. Kume, T.; Iwasaki, H.; Yamamoto, Y.; Akiba, K., Tetrahedron Lett. 1988, 29, 3825– 3828. 82. Kirsch, S. F.; Binder, J. T.; Liebert, C.; Menz, H., Angew. Chem. Int. Ed. 2006, 45, 5878–5880. 83. Chan, C. S.; Araki, T.; Nakamura, I.; Terada, M., Tetrahedron Lett. 2009, 50, 216–218. 84. Dudnik, A. S.; Gevorgyan, V., Angew. Chem., Int. Ed. 2007, 46, 5195–5197. 85. Dudnik, A. S.; Sromek, A. W.; Rubina, M.; Kim, J. T.; Kelin, A. V.; Gevorgyan, V., J. Am. Chem. Soc. 2008, 130, 1440–1452. 86. Schwier, T.; Sromek, A. W.; Yap, D. M. L.; Chernyak, D.; Gevorgyan, V., J. Am. Chem. Soc. 2007, 129, 9868–9878. 87. Yamaguchi, Y.; Kobayashi, S.; Wakamiya, T.; Matsubara, Y.; Yoshida, Z.-I., J. Am. Chem. Soc. 2000, 122, 7404–7405.
4 SILVER(I)-MEDIATED ELECTROCYCLIC PROCESSES TINA N. GRANT AND FREDERICK G. WEST Department of Chemistry, University of Alberta, Edmonton, Canada
4.1 Introduction 4.1.1 Ring-Opening Reactions of Halocyclopropanes 4.1.2 Silver(I)-Assisted Ring-Opening Reactions 4.2 Nucleophilic Trapping of Cationic Intermediates 4.2.1 Solvolysis Reactions 4.2.2 Intramolecular Trapping with Heteronucleophiles 4.2.3 Diastereoselective Reactions 4.2.4 Carbon–Carbon Bond Formation 4.3 The Silver(I)-Promoted Nazarov Reaction 4.3.1 Development and Initial Findings 4.3.2 Interrupted Nazarov Reactions 4.4 Concluding Remarks References
4.1
INTRODUCTION
The halophilic character of the silver(I) cation is traditionally exploited to assist in removal of halides from metal complexes in order to open a coordination site and, in effect, initiate catalyst reactivity. Analogously, silver(I) salts can be used to facilitate the removal of halides from organic molecules to reveal reaction pathways that involve cationic intermediates. In the simplest scenarios, silver(I) has been used to Silver in Organic Chemistry Edited by Michael Harmata Copyright 2010 John Wiley & Sons, Inc.
117
118
SILVER(I)-MEDIATED ELECTROCYCLIC PROCESSES
assist in the abstraction of halides during solvolysis or substitution reactions.1 A more intriguing role for silver(I) in organic chemistry is its ability to trigger electrocyclic ring opening of the halocyclopropane moiety, resulting in a reactive cationic species. These cationic processes can be standalone reactions that result in simple, ringopened products, or they can be used to initiate domino reactions that generate more complex and, in some instances, polycyclic products.2 This chapter focuses on the use of silver(I) salts as Lewis acids in the ring-opening reactions of halocyclopropanes. Both the classical and most recent examples of this type of chemistry are examined with emphasis on the use of this simple transformation to build up complex intermediates that are potentially useful in the construction of natural products. 4.1.1
Ring-Opening Reactions of Halocyclopropanes
Halocyclopropanes are readily available substrates that can be utilized in a variety of different transformations, including reduction to furnish monohalocyclopropanes and/or simple cyclopropanes, substitution reactions, and elimination reactions to provide cyclopropene compounds. In all of these cases, the integrity of the threemembered ring is preserved as a result of careful manipulation of the halogen substituents. Although these reactions can be useful for an organic chemist, the most intriguing aspect of halocyclopropane chemistry is the ability of this functionality to undergo facile cationic ring opening under thermal conditions to provide an allyl cation I, which can further undergo nucleophilic trapping or elimination to provide olefinic products (2 and 3, Fig. 4.1). The 2p-electron ring opening process obeys the conservation of orbital symmetry rules described by Woodward and Hoffmann,3 and thereby proceeds in a disrotatory fashion with concomitant loss of a halide anion. The direction of disrotation is strongly influenced by stereoelectronic factors.4 Orbitals that are involved in the breaking of the C--C s bond must rotate in a direction that assists halide departure by overlapping with the s orbital of the dissociating halogen (Fig. 4.2). This stereoelectronic effect implies that chemoselective removal of one of the halogens would result in the preferential formation of one allyl cation over the other. The directionality of ring opening has a distinct impact on ring expansion reactions involving dibromobicyclo[n-1,0]alkanes (Fig. 4.3).5 When medium-size rings are being formed (n ¼ 2–4), the disrotation occurs selectively with loss of the endo-bromide. Rotation in this direction affords the E-isomer, which X X –X
1
X
X
Nu 2
I –H
Nu = X, other
X
3
Figure 4.1. Cationic ring opening of gem-dihalocyclopropanes.
119
INTRODUCTION
Xb
Xb
Xa
R3
R1
R2
R3
R1
–X a
R4
R2
Xb
R4 Xa
Xa
R3
R1
R2
R4
R2
–X b
R4
R1
R3
Figure 4.2. Stereoelectronic factors affecting cationic ring opening. Br
Br OAc
H
H
AgOAc AcOH n = 2-4
(CH2)n
(CH2)n
Br
Br
n = 5-8
H H
OAc
H
H
AgOAc AcOH
Br (CH2)n
(CH2)n
Br (CH2)n
E-isomer
Z-isomer
Figure 4.3. The direction of disrotation in ring expansion reactions.
minimizes ring strain during product formation. However, when larger rings are being generated under the same conditions (n ¼ 5–8), the Z-isomer is preferentially formed, with removal of the more accessible exo-bromide. Formation of the Z-isomer is actually a favorable process if the developing ring strain permits, since untoward A1,3 steric interactions that would arise during ring opening in the other direction are avoided. Porter et al. attempted to manipulate the directionality of ring opening when proposing a general synthesis of polyunsaturated fatty acid hydroperoxides.6 In this work, the vinylcyclopropyl bromide 4 was synthesized and subjected to silver trifluoroacetate and hydrogen peroxide in diethyl ether at room temperature (Scheme 4.1). It was anticipated that disrotatory ring opening would preferentially occur in the direction that assists with loss of the bromide to afford the cationic species II. The authors maintained that trapping of this cation with hydrogen peroxide would occur faster than isomerization, and loss of stereochemistry and, in fact, the reaction
(CH2)7CO2Me C5H11
Et2O
Br 4
1
AgOCOCF3, H2O2
C5H11
3
C5H11
(CH2)7CO2Me
II
5
(CH2)7CO2Me
III
OOH
C5H11 OOH
(CH2)7CO2Me C5H11 5 41%
Scheme 4.1. Porter’s approach to polyunsaturated fatty acid hydroperoxides.
(CH2)7CO2Me 6 44%
120
SILVER(I)-MEDIATED ELECTROCYCLIC PROCESSES
did furnish the predicted hydroperoxides 5 and 6 as the major products in 41% and 44% yields, respectively. These products are the result of trapping the anticipated cationic intermediate III at the C1 and C5 positions. The minor compounds resulting from reaction of this vinylcyclopropyl bromide were also examined and were determined to arise from either product isomerization or ring opening in the less favored direction of disrotation. This “alternate” pathway of ring opening likely occurs with some assistance from the alkenyl substituent on the cyclopropane, since the p bond is situated trans to the departing bromide. Porter’s work illustrates the impact of substituent effects on the ring opening of halocyclopropanes. Although the influence of the alkenyl substituent was weak in this particular example, it has been shown that substitution on the halocyclopropane moiety can either impair or promote electrocyclic ring opening.7 Exploitation of these aspects of ring opening could be valuable for the development of both cascade and asymmetric transformations, and is discussed in greater detail later in this chapter. 4.1.2
Silver(I)-Assisted Ring-Opening Reactions
While cationic ring opening of halocyclopropanes can be induced under strictly thermal conditions, it is most often performed in the presence of a Lewis acid.8 The Lewis acids commonly used in these reactions are silver(I) salts due the inherent halophilicity of the silver(I) cation. The silver(I)-mediated reactions can be carried out at lower temperatures due to activation of the departing halide by coordination to silver(I) in what has been described as a highly concerted “push–pull” mechanism.1a Under these conditions the halide-substituted carbon atom bears slightly increased positive character, which enables the cationic ring opening to proceed under mild conditions (Fig. 4.4). The presence of silver(I) is additionally used to sequester halide anions liberated in the ring-opening process, effectively removing free halide species from the reaction mixture. This sequestration prolongs the lifetime of the haloallyl cation, which can then be intercepted by another nucleophile or undergo rearrangement reactions. In this respect, the thermal and silver(I)-mediated ring opening processes can be used to afford different products from the same starting material. An illustration of this divergent reactivity can be observed in Banwell’s use of gemdibromocyclopropane 7 to construct an intermediate used in an approach to the Amaryllidaceae alkaloids (Scheme 4.2).9 When 7 was heated in the absence of Lewis acid, allyl bromide 8 was produced in quantitative yield via thermal ring opening of the cyclopropane and subsequent trapping of the resultant cation by free bromide in b
c d
a X
b
c d
a
b
X AgX
Ag+
c a
d
Ag
Figure 4.4. Silver(I) assistance in the ring opening of halocyclopropanes.
121
NUCLEOPHILIC TRAPPING OF CATIONIC INTERMEDIATES
Br Br
1. NaN3 2. PPh3, H2O 120ºC 97%
7
3. MeO2CCl (68%, 3 steps)
Br 8
Br
1.1 equiv AgNCO 1,4-dioxane, reflux
MeOH
Br
r.t. 96%
NCO 9
Br NHCO2Me 10
Scheme 4.2. Thermal and silver(I)-assisted ring opening.
the reaction mixture. The allyl bromide was further functionalized (68% over three steps) to the desired carbamate 10. Banwell was able to dramatically improve the synthesis of the carbamate by developing a one-pot reaction, which involved treatment of gem-dibromocyclopropane 7 with silver isocyanate, followed by the addition of methanol. Under these conditions, silver-assisted electrocyclic ring opening provided the haloallyl cation, which was subsequently trapped by isocyanate anion. Interception of the cationic species with isocyanate was successful since bromide was removed from the reaction mixture as a precipitate (AgBr). Finally, treatment of intermediate 9 with methanol furnished the desired carbamate in 96% yield. This example demonstrates the usefulness of the silver(I)-mediated process. Removal of free halide from the reaction mixture affords a long-lived cationic species that can be captured by a different nucleophile, such as solvent, the silver(I) counteranion, or an intramolecular nucleophile. This reactivity has been exploited in many different ways throughout the years and is examined in greater detail later in this chapter.
4.2
NUCLEOPHILIC TRAPPING OF CATIONIC INTERMEDIATES
The construction of complex intermediates from simple and readily available starting materials has been accomplished using the electrocyclic ring-opening reaction of halocyclopropanes. This is typically achieved through interception of the cationic haloallyl intermediate by solvent, the silver(I) counteranion, or some alternate tethered heteroatom or carbon-based nucleophile. Examples of these processes are described below. 4.2.1
Solvolysis Reactions
The earliest report of the solvolysis of a halocyclopropane to an allyl alcohol was presented by Roberts and Chambers in 1951.10 These early examples did not utilize silver(I) to assist in halide departure. Instead, the halocyclopropane was simply dissolved in solvent and allowed to react at elevated temperatures. For example, in
122
SILVER(I)-MEDIATED ELECTROCYCLIC PROCESSES nBuO
Cl
OMe OMe
O
nBuOH,
reflux 82%
Cl
OMe
Cl
11
OMe Cl
12
IV
Scheme 4.3. Unassisted solvolysis of halocyclopropanes.
1959 McElvain and Weyna provided evidence for the formation of a-chloroacrylic esters as a result of subjecting gem-dichlorocyclopropanes, such as 11, to refluxing n-butyl alcohol (Scheme 4.3).11 This reaction was believed to proceed through trapping of the stabilized haloallyl cation (IV) by n-butanol to afford an orthoester that readily collapsed to the observed product. When the solvent participates as a nucleophile in the capture of a cationic species, it may not be necessary for the involvement of silver(I) as a halide sequestering agent since there is a large excess of the nucleophile present; however, silver(I) salts are typically used in order to lower the reaction temperatures and effect a milder solvolysis of halocyclopropanes.12 Unfortunately, in some instances participation by solvent can interfere with a desired reaction pathway and afford unwanted side products. In 2004, Banwell and Sydnes13 disclosed their efforts toward the total synthesis of plant-derived phenanthroquinolizidine alkaloids such as julandine.14 Their initial strategy involved the intermolecular trapping of an allylic cation with 2-piperidinemethanol 14 (Scheme 4.4). The desired gem-dibromocyclopropane precursor 13 was derived from dibromocarbene addition to the respective olefin (12) in 78% yield using standard phase transfer conditions developed by Makosza and Wawrzyniewicz.15 When ring opening of 13 was induced using a variety of silver(I) salts, the desired intermolecular trapping product 15 was not observed. Instead, the cationic species was trapped by either solvent (CF3CH2OH, or (CH3)2CHOH) or the silver(I) counteranion. In an effort to overcome this issue, the authors decided to approach the synthesis of 15 in a different manner, inducing solvolysis of the gem-dibromocyclopropane with silver acetate in the presence of acetic acid to furnish an allylic acetate. The acetate
MeO
OMe
OMe
OMe CHBr3
Br Br 13
OH
Ag(l)
+
50% NaOH aq. 78%
12
MeO
OH
MeO
HN
N
Br 14
15 OMe
OMe H
1. AgOAc, AcOH
+
2. K2CO3, MeOH
N MeO Julandine
TEA DMF 98%
OH
MeO
MeO
82% (2 steps)
Br NCS, DMS 69%
X
16 X = OH 17 X = Cl
HN 14
Scheme 4.4. Steps toward the synthesis of phenanthroquinolizidine alkaloids.
123
NUCLEOPHILIC TRAPPING OF CATIONIC INTERMEDIATES
HO Br Br
AgClO4 acetone–H2O 0ºC, 82%
18
Br 19 E : Z ~ 7 : 93
Scheme 4.5. Silver(I)-promoted solvolysis reaction with ring expansion.
was then smoothly converted to the allylic alcohol 16, with a yield of 82% over the two steps. Subsequently, 16 was treated with N-chlorosuccinimide in dimethyl sulfide to afford allyl chloride 17, which was then coupled with 2-piperidinemethanol to provide the desired product 15 in 98% yield. Although the planned intermolecular trapping process was unsuccessful, solvolysis of gem-dibromocyclopropane 13 provided a useful synthetic intermediate to allow eventual assembly of the desired vinyl bromide. In many instances, however, solvolysis of a halocyclopropane is deliberately accomplished in order to install an essential vinyl halide or hindered olefin. In 2000, Murphy and coworkers performed a silver ion-mediated ring expansion of gemdibromocyclopropane 18 in wet acetone to afford the allylic alcohol 19 in 82% yield (Scheme 4.5).16 Under these conditions the desired cyclononene product was obtained as an inseparable mixture of E- and Z-isomers (7 : 93). Interestingly, two sets of peaks observed in the 1 H NMR spectrum indicated that the Z-isomer existed as two separate conformers at room temperature. This intermediate was subsequently used in Murphy’s approach to the radical-based preparation of tricyclic indoles. One particular advantage to using silver(I) to promote the electrocyclic ring opening of halocyclopropanes is that deliberate selection of the counteranion can result in products derived from cation interception by the silver salt. An example of this mode of trapping was illustrated earlier in the conversion of a halocyclopropane to an allylic carbamate via trapping with silver isocyanate (Scheme 4.2).9,17 One obvious drawback to this mode of trapping is the necessity for a large excess of the silver salt in order to achieve efficient trapping. Nonetheless, an excellent example of the utility of this reaction can be observed in Banwell’s approach to the ABC ring system observed in paclitaxel (Scheme 4.6).18 In this route, the complex intermediate
Br
Br
OMe
OMe Br
10 equiv AgOAc
H O
CH2Cl2, 40ºC 71%
N H
AcO
O
20
H O 21
Scheme 4.6. An approach to paclitaxel.
N H
O
124
SILVER(I)-MEDIATED ELECTROCYCLIC PROCESSES
20 was assembled in five short steps from 5-methoxyindane. The gem-dibromocyclopropane was then subjected to silver acetate in dichloromethane at 40 C to generate the product 21. Under these conditions the silver counteranion was involved in regioselective capture of the haloallylic cation derived from ring opening of the halocyclopropane, resulting in the installation of necessary oxygenation and a very strained bridgehead olefin. 4.2.2
Intramolecular Trapping with Heteronucleophiles
Halocyclopropane compounds that contain an internal nucleophile have proved to be very successful substrates for cascade reactions. On silver(I)-mediated ring opening of the halocyclopropane moiety, the tethered nucleophile can intercept the cationic haloallyl intermediate affording cyclic and/or polycyclic products. Typically these reactions utilize at least one equivalent of the silver(I) salt to assist in ring opening and completely sequester free halide ions from the reaction mixture, which effectively eliminates intermolecular trapping of the cationic species by halide. Early examples of intramolecular trapping in this manner were performed using the silver(I)-mediated ring opening of gem-dibromocyclopropanes tethered to a carboxylic acid or alcohol moiety (Scheme 4.7).19 In this scenario, electrocyclic ring opening afforded a haloallyl cation that was readily trapped by the pendent nucleophile to furnish lactone and cyclic ether products in good to excellent yields. In a similar fashion, gemdihalocyclopropanecarboxylic acids 26 can be used as substrates for intramolecular trapping of the haloallyl cation (Scheme 4.8).20 In this study, the substrates were transformed to the corresponding butenolide products in excellent yields, provided the right combination of silver salt and solvent were utilized. When methanol, isopropyl alcohol, or 2,2,2-trifluoroethanol (TFE), was used as solvent in these reactions, solvolysis products were always observed to some minor extent. When 26b was treated with silver trifluoroacetate in hexafluoroisopropanol (HFP), it underwent ring opening to afford the expected butenolide 28 in 82% yield, along
Br
Br
Br
Br
AgOCOCF3 TFE, 25ºC, 8.5 h 97%
HO2C
O
HO V
22
Br
O
– H+
O
23 Br
Br AgOCOCF3
OH
TFE, 25ºC, 8.5 h 73%
24
O
25
Scheme 4.7. Synthesis of lactones and cyclic ethers from halocyclopropanes.
125
NUCLEOPHILIC TRAPPING OF CATIONIC INTERMEDIATES
CO2H
Br
AgOCOCF3
O
TFE, reflux 93%
Br Br
O
26a
CO2H
27 OH CO2H
Br
AgOCOCF3
O
HFP, reflux
Br
O
Br Br 26b
28 82%
29 11%
CO2H Br 30 7%
CO2H CO2H Br
Br VI
VII
Scheme 4.8. Butenolide formation from gem-dihalocyclopropanecarboxylic acids.
with the hydroxy acid 29 and diene 30. The formation of these minor products can be attributed to intermediacy of two cationic intermediates (VI or VII), only one of which (VI) can proceed immediately to the butenolide product. Cha and coworkers have used the silver(I)-mediated ring opening of a gemdichlorocyclopropane to construct the interesting bicyclo[6.4.0]dodecanone derivative 32 (Scheme 4.9).21 The halocyclopropane compound was synthesized from the corresponding enol ether, which was constructed using an ingenious oxy-Cope rearrangement of cis-1,2-dialkenylcyclopropanols. Silyloxy substitution on a halocyclopropane moiety typically results in the formation of an unsaturated ketone on silver(I)-promoted ring opening; however, the additional oxygen present in Cha’s substrate 31 was poised to participate in the trapping of the intermediate haloallyl cation and afford the ketal 32 in 70% yield. This result implies that the silyloxy substituted allyl cation has a substantial lifetime, allowing trapping processes to occur readily in place of immediate termination to the carbonyl moiety. Nitrogen-based nucleophiles have been thoroughly examined in the context of trapping the cationic intermediate generated from silver(I)-mediated halocyclopro-
Cl
Cl
OTBS
Cl
OTBS
Cl
OTBS
AgNO3 acetone–H2O 70%
O
TBS
O
TBSO 31
VIII
32
Scheme 4.9. Intramolecular ketal formation from ring opening of a halocyclopropane.
126
SILVER(I)-MEDIATED ELECTROCYCLIC PROCESSES Br Br
H 1.1 equiv AgOAc
O
Br
Br
TFE, r.t., 87%
MeO2CHN
NH IX
33
MeO2C
H
N
O
N
H
MeO2C
γ -Lycorane
34
Scheme 4.10. Key step in the synthesis of g-lycorane.
pane ring opening. Banwell has been the most prominent figure in this area of halocyclopropane chemistry and has developed some very elegant ways to utilize intramolecular trapping to construct highly functionalized intermediates toward the synthesis of a number of alkaloid natural products. In 1994, Banwell and Wu presented a racemic total synthesis of g-lycorane that utilized a silver(I)-promoted electrocyclic ring opening as the pivotal step for the formation of an advanced bicyclic intermediate (Scheme 4.10).22 In this reaction, halocyclopropane 33 was subjected to silver acetate in 2,2,2-trifluoroethanol. These conditions induced ring opening to the 2-bromoallyl cation IX and subsequent interception by the nitrogen of a tethered carbamate moiety to afford the bicyclic product 34 in 87% yield. A similar strategy was used in the ensuing syntheses of the crinine-type alkaloids maritinamine and epi-maritinamine.23 The key step in the synthesis of these natural products is depicted in Scheme 4.11, in which the gem-dichlorocyclopropane substrate 35 was subjected to 6 equiv of silver tetrafluoroborate in tetrahydrofuran to afford the cyclized product 36 in moderate yields after reinstallation of the carbamate protecting group. Interestingly, the authors noted that the two epimeric forms of the halocyclopropane compound undergo the cascade cyclization sequence at different rates, with one isomer reacting completely within 8 h and the other requiring prolonged reaction times. The resulting bicyclic product was carried on to complete the synthesis of both maritinamine and its C3 epimer. The intramolecular nitrogen-trapping protocol used by Banwell has also been successfully exploited in the assembly of spirocyclic frameworks relating to the aromatic erythrina alkaloids.24 However, when it was applied to the synthesis of nonaromatic spirocycles, as found in histrionicotoxin, the flexible alkyl tether proved to be problematic.25 In this study, gem-dichlorocyclopropane substrate 37 was initially subjected to silver(I) salts under a variety of conditions only to provide solvolysis and elimination products without any indication of trapping by the pendent Cl Cl
Cl NBoc
6 equiv AgBF4 THF, 40ºC, 21 h;
NHBoc MeO iPrO
35
then (Boc)2O TEA, THF, 25ºC, 15 h 65–75%
MeO iPrO
36
Scheme 4.11. Intramolecular nitrogen trapping en route to maritinamine.
127
NUCLEOPHILIC TRAPPING OF CATIONIC INTERMEDIATES
OR
1.5 equiv AgOAc
Cl
TFE
Cl OR NHCO2Me
NHCO2Me R = CF3CH2 or Ac
LiHMDS;
Cl Cl
Cl
then AgBF4
NHCO2Me
NHCO2Me 37 Cl
1.5 equiv AgOAc
OAc Cl
Cl
DMPU, 100ºC
NHCO2Me
NHCO2Me
Scheme 4.12. Initial attempts at spirocyclization.
nitrogen-based nucleophile (Scheme 4.12). Subsequently, a cis-olefin was incorporated into the tether between the halocyclopropane and carbamate moieties in an effort to introduce an element of rigidity to the substrate (Scheme 4.13). This strategy was successful in inducing spirocyclization to provide a sample of the desired product 39, albeit in poor yield, along with a small amount of polyene 40. The successful utilization of nitrogen-based nucleophiles in the previously described cascade reactions has allowed for the synthesis of complex polycyclic structures from simple and readily available starting materials. The fact that carbamates can participate as nucleophiles has provided the opportunity for development of diastereoselective ring closures onto the halocyclopropane-derived allyl cation. 4.2.3
Diastereoselective Reactions
In 2000, Banwell and coworkers reported their initial attempts toward diastereoselective ring closure onto the cation that results from silver(I)-promoted halocyclopropane ring opening. In this work, a concise approach to the synthesis of bridged seven- and eight-membered nitrogen heterocycles was described in which a chiral auxiliary was used to impart stereoselectivity in the pivotal electrocyclic ringopening/nitrogen-trapping step (Scheme 4.14).26
Cl Cl
NHAlloc LiHMDS, THF; then AgBF4
Alloc
N
Cl
Cl
HN
AllocHN
38
39, 15%
OH
40, 6%
Histrionicotoxin, HTX
Scheme 4.13. Assembly of 1-azaspiro[5.5]undecane framework.
128
SILVER(I)-MEDIATED ELECTROCYCLIC PROCESSES
NHCO2R*
Br Br
10 equiv AgClO4 DME, 3 Å MS, 18ºC 33%
Br
41
42a
+
N CO2R*
Br
N CO2R* 42b
Ph R* =
3 : 11
NHCO2R*
Br Br
10 equiv AgClO4 DME, 3 Å MS, 18ºC 84%
+
N
Br
CO2R*
44a
43
N
Br
CO2R* 44b 2 : 9
Scheme 4.14. Chiral auxiliary approach to diastereoselective cyclizations.
Previously, chiral carbamates were used by Banwell to afford facile separation of diastereomers at appropriate stages in the synthesis of various alkaloid natural products.9,27,28 In those instances, the chiral carbamates were prepared by silver (I)-mediated ring opening of a halocyclopropane, followed by trapping with cyanate and subsequent treatment with ()-menthol, a method analogous to that outlined earlier in this chapter (Scheme 4.2). However, in this example the chiral auxiliary is used for direct control of subsequent ring closure, affording 2-azabicyclo[3.3.1]non7-enes 42 and 6-azabicyclo[3.2.1]oct-3-enes 44 in modest diastereoselectivities. In this work, the best results were obtained when the 8-phenylmenthyl-based auxiliary was used, suggesting that p stacking plays a role and may be further exploited to improve the selectivity of this cascade process. More recently, Banwell has investigated the possible influence of remote stereogenic centers on the intramolecular capture of halocyclopropane-derived allyl cations (Scheme 4.15).29 During the enantioselective total synthesis of ent-erythramine and 3-epi-erythramine, gem-dichlorocyclopropane 45 was prepared in nine steps from known starting materials and then subjected to silver tetrafluoroborate after deprotonation with lithium hexamethyldisilazide. Subsequent deprotection afforded a separable diastereomeric mixture of spirocyclic products 46a and 46b in 26% and 30% yields, respectively. This result implies that the remote stereogenic center is ineffective at controlling the stereoselectivity of the silver(I)-mediated ring opening/ nitrogen-trapping reaction of 45, perhaps due to the lack of steric bulk provided by the methoxy group as well as its distance from the developing spirocenter. NHAlloc O O
Cl Cl
O
1. LiHMDS, THF 2. AgBF4
O NH Cl
O MeO
45
NH Cl
O
3. Pd(PPh3)4 dimedone, THF
MeO
MeO 46a 26%
46b 30%
Scheme 4.15. An attempt to remotely control the stereoselectivity of ring closure.
129
NUCLEOPHILIC TRAPPING OF CATIONIC INTERMEDIATES
1.1 equiv AgClO4
Br Br
Et2O, r.t., 89%
47
+ Br
Br 48 a
48b 1.4 : 1
Br
Br
Br X
XI
XII
Scheme 4.16. Cationic cyclization of halocyclopropane dienes.
4.2.4
Carbon–Carbon Bond Formation
The silver(I)-mediated ring opening of halocyclopropanes has been used to construct complex frameworks through the inter- and intramolecular trapping of cationic intermediates with heteronucleophiles. An obvious extension of this work is the involvement of carbon-based nucleophiles to form new carbon–carbon bonds. In 1996, Kostikov and coworkers reported the participation of aromatic solvents in the capture of halocyclopropane-derived allyl cations even in the absence of silver(I).30 However, this early example of intermolecular attack by a carbon nucleophile is one of very few such reports. In the same year, Gassman et al. reported cationic cyclizations of gem-dibromocyclopropanes tethered to remote diene moieties (Scheme 4.16).31 They found that treatment of diene 47 with 1.1 equiv of silver perchlorate in diethyl ether afforded the cyclized trienes 48a and 48b in a ratio of 1.4 : 1 and 89% yield. The mechanism of this reaction was believed to involve initial ring opening of the gemdibromocyclopropane followed by trapping of the haloallyl cation X to furnish another cationic allyl species. The observed scrambling of alkene geometry in the products can be explained by facile rotation of the carbon–carbon bond that connects the allyl cation to the cyclopentane in XI and XII. These authors also investigated the effect of substitution on the terminal olefin by removing the gem-dimethyl groups. This substrate 49 provided the expected triene in 50% yield, but also resulted in the formation of bicyclic product 51 in 13% yield. This second product is most likely constructed from two successive cation–olefin cyclizations (Scheme 4.17), followed by oxidation to the aromatic compound. In 2004, Banwell and Sydnes reported an unintentional but intriguing result of intermolecular “pseudodimer” formation as a result of carbon–carbon bond formation during their efforts toward the synthesis of phenanthroquinolizidine alkaloids (Scheme 4.18).13
130
SILVER(I)-MEDIATED ELECTROCYCLIC PROCESSES
1.1 equiv AgClO4
Br Br
Et2O, r.t.
49
+ Br 50 50%
51 13% [O]
+ Br
Br
–H+ –HBr
XIII
XIX
XX
Scheme 4.17. Two cation–olefin cyclization events to afford an aromatic indane.
On treatment with silver tetrafluoroborate in acetonitrile, the gem-dibromocyclopropane 13 underwent electrocyclic ring opening and subsequent capture by another molecule of 13 in an SEAr process. This “dimerization” reaction occurred despite the presence of an able, nitrogen-based nucleophile, piperidine 14, in large excess. Despite efforts to effect nitrogen trapping of the cationic intermediates, no solution to this problem could be found, which led to the design of a new approach to the synthesis of the target compounds. Carbon-based nucleophiles have now been thoroughly explored in the design of both inter- and intramolecular trapping reactions of haloallyl cations. Banwell and coworkers have presented examples of both pyrrole and indole participation in these processes, resulting in the construction of important alkaloid scaffolds.32 Banwell’s most recent contribution to this area was described in his work toward the quick assembly of hapalindole and fischerindole structures (Scheme 4.19).33 Indoles 53a and 53b were able to capture the cationic intermediate generated from treatment of gem-dibromocyclopropane 54 with silver tetrafluoroborate in tetrahydrofuran, providing the tricyclic products 55a and 55b in 67–77% yields. The trapping products could be carried on to afford tetracycle 56 (or its cis ring-fused isomer) in four steps. OMe MeO OMe Br Br 13
OMe
HO
MeO +
OMe
AgBF4
HN
MeCN, 27%
Br 14
52
Scheme 4.18. Unexpected carbon–carbon bond formation.
Br Br
131
NUCLEOPHILIC TRAPPING OF CATIONIC INTERMEDIATES
Br N R
+
Br
AgBF4
Br
THF, 18ºC 67–77%
53a R = H 53b R = Me
H H N R
54
N 55a R = H 55b R = Me
56
Scheme 4.19. Intermolecular trapping of the allyl cation by indole nucleophiles.
This strategy allowed for the rapid construction of complex, polycyclic skeletons in good yields from simple starting materials. In the same publication, an enantioselective process was attempted wherein commercially available (2R,3R)-butane-2,3-diol was used to generate the chiral cyclopentene 57, which was cyclopropanated to afford gem-dibromocyclopropane 58 (Scheme 4.20). Unfortunately, when this substrate was subjected to the reaction conditions outlined above, product 59 was obtained as a 1 : 1 mixture of diastereomers. This result implies that selectivity in these trapping processes is unaffected by the presence of a chiral auxiliary on the remote carbon of the cyclopentane framework. The silver(I)-mediated electrocyclic ring opening of halocyclopropanes has been used to induce extensive skeletal rearrangements in gem-dibromospiropentanes, providing rapid construction of naphthalenes and/or indenes (Scheme 4.21).34 Avariety of Lewis acids, Brønsted acids, and solvent effects were carefully examined before optimal conditions were identified. It was found that subjection of spirocycle 60 to silver acetate in trifluoroacetic acid afforded rearrangement products 61 and 62 in moderate to good yields. The proposed mechanism of the reaction is illustrated in Scheme 4.21. Silver(I)-promoted ring opening of the gem-dibromocyclopropane would result in the formation of haloallyl cation XXI, which is the common intermediate in the synthesis of both naphthalene and indene products. It was observed that naphthalene compounds were preferentially formed when electron-withdrawing groups were placed on the aromatic ring. In these substrates, haloallyl cation XXI would be destabilized and therefore readily undergo another ring-opening process to afford cation XXII. This new cation would then be trapped in a Friedel–Crafts reaction to provide the observed naphthalene products 61. In instances where the aromatic ring was substituted with electron-donating groups, the indene products 62 were
O
O
O
CHBr3, DCM, aq-NaOH, 18ºC
O
O +
sonication 91%
57
Br Br
58
AgBF4
N H 53a
Br
O H
THF, 18ºC 59%
N H
Scheme 4.20. Chiral auxiliary approach to diastereoselective trapping processes.
59 (1:1)
132
SILVER(I)-MEDIATED ELECTROCYCLIC PROCESSES
Br
R1
Br
R2 60
Br
Br R1
CF3COOH, -10ºC 61–99%
Br
+
Br
R1
61
62 R2
R2 AgOAc
R1
R2
R2
1.5 equiv AgOAc
Br R1
R2
Friedel–Crafts Br
R1
+
AgBr XXI
60
62
R2 Br R1
R2
Friedel–Crafts then tautomerization
Br R1
+ XXII
61
Scheme 4.21. Skeletal rearrangement of gem-dibromospiropentanes.
preferentially formed. This can be explained in a similar fashion, since these substrates would afford long-lived, stabilized haloallyl cations XXI that could be immediately captured in a Friedel–Crafts reaction instead of undergoing further ring opening processes.
4.3
THE SILVER(I)-PROMOTED NAZAROV REACTION
In 2006, the unique silver(I)-mediated cationic ring opening of gem-dihalocyclopropanes was examined as an interesting entry into the design of a novel Nazarov cyclization.35 The Nazarov reaction is an electrocyclization process that typically transforms a cross-conjugated dienone, 63, into a cyclopentenone product, 64, by means of conrotatory ring closure (Fig. 4.5). Since its initial discovery in 1941,36 the Nazarov reaction has been keenly investigated and important advances have been made that expand both the scope and utility of this process. Halocyclopropane substrates were investigated as novel substrates in the Nazarov reaction since it was well documented that cationic intermediates derived from ring opening were
O
O
63
64
Figure 4.5. Nazarov cyclization.
133
THE SILVER(I)-PROMOTED NAZAROV REACTION
accessible using very mild reaction conditions. It was believed that appropriately substituted gem-dihalocyclopropanes could be used to access a pentadienyl cationic species capable of undergoing Nazarov cyclization. 4.3.1
Development and Initial Findings
In an effort to mimic Nazarov cyclization with nontraditional substrates, the design of a sequential 2p-electron disrotatory ring opening and 4p electrocyclization was undertaken (Fig. 4.6). 1,1-Dihalo-2-(silyloxy)-2-vinylcyclopropanes 65 were determined to be optimal substrates for this investigation, since computational experiments7 have shown that alkenyl substitution on gem-dihalocyclopropanes accelerates the ringopening process relative to hydrogen. Oxygen substitution at the same position on the cyclopropane moiety was also shown to assist in the disrotatory ring opening, which would promote the efficient generation of haloallyl cation XXIII. Because of the presence of the vinyl substituent, this cation might also be viewed as a pentadienyl species XXIV, which is analogous to the cationic intermediate observed during conventional Nazarov cyclizations. It was believed that the pentadienyl cation would undergo conrotatory electrocyclization to furnish a 2-silyloxycyclopentenyl cation (XXV) that would experience traditional elimination and subsequent hydrolysis to provide a-halocyclopentenone products 66 and/or 67. This type of sequential ring opening and electrocyclization involving gem-dihalocyclopropanes has not previously been studied, although conceptually similar protocols have been reported.37,38 Initial experiments to investigate the use of gem-dihalocyclopropanes in the Nazarov reaction were performed on methyl vinyl ketone, a simple and readily available substrate. Methyl vinyl ketone was smoothly converted to 2-triisopropylsilyloxydiene 69 in low yield after immediate purification on an alumina column (Scheme 4.22). Preparations of both gem-dichlorocyclopropane compound 70 and the gem-dibromocyclopropane compound 71 were investigated. While dichlorocyclopropanation proceeded smoothly, the analogous dibromocyclopropanation reaction afforded a mixture of compounds, including the desired product 71, which could not be separated. This observation prompted selection of gem-dichlorocyclopropane compounds for use in development of the new Nazarov methodology. X X R3SiO
R3
R1
–X
OSiR3 X
R1
OSiR3 X
R1
OSiR3 4π electrocyclization
R1
X
(Nazarov)
R2
R2
R3
R2
XXIII
65 O
1. elimination
R1
2. protonation
R3 66
R2 XXV
O X
R2
R3 XXIV
and/or
R1
X R2
R3 67
Figure 4.6. Concept behind gem-dihalocyclopropanes in the Nazarov reaction.
R3
134
SILVER(I)-MEDIATED ELECTROCYCLIC PROCESSES
O
OTIPS
TIPSOTf 2,6-lutidine
CHCl3 50% NaOH aq.
Cl
TIPSO
Cl
TEBA 93%
CH2Cl2, -78ºC 41%
68
69
70
CHBr3 50% NaOH aq.
TIPSO
Br
Br
TEBA
71
Scheme 4.22. The synthesis of 1,1-dihalo-2-(silyloxy)-2-vinylcyclopropanes.
The choice of silyl substituent was based on the durability of a triisopropylsilyl (TIPS) group relative to other, less bulky silyl groups. Other silyl groups, such as tertbutyldimethylsilyl and triethylsilyl, were also investigated during optimization of the reaction conditions, but it became obvious that a more robust silyl substituent was needed to prevent premature termination of the reaction by desilylation to afford 2-chlorodienones such as 73 (Fig. 4.7). Substrate 70 was exposed to a variety of reaction conditions in a qualitative investigation used to assess the practicality of the proposed ring opening and 4pelectrocyclization sequence, in which it was found that treatment with 1.5 equiv of silver tetrafluoroborate in dichloromethane afforded the a-chlorocyclopentenone 74 after prolonged stirring at room temperature (Fig. 4.8). Further investigation showed that this mode of reactivity was, in fact, general, and could be applied to a number of readily accessible gem-dichlorocyclopropane substrates to afford the corresponding a-chlorocyclopentenone products in moderate to good yields. It was later observed that both yields and reaction times for these sequences could be improved by treatment of the starting materials with 1 equiv of silver tetrafluoroborate in refluxing acetonitrile (yields depicted in parentheses in Fig. 4.8).39 The majority of gem-dichlorocyclopropane substrates examined in this study provided the desired a-chlorocyclopentenones as a result of sequential electrocyclic ring opening and Nazarov cyclization. In general, those substrates lacking additional substitution on the cyclopropane moiety provided products 75 selectively as a result of regioselective elimination to deliver the more electron-rich olefin. The mechanism for this transformation is believed to involve disrotatory halocyclopropane ring opening
R3SiO
Cl
O
Cl
SiR3
Ag(I)
O
Cl
Cl –R3Si+
Ph
Ph
Ph 72
XXVI
73
Figure 4.7. Termination of the reaction via desilylation of the haloallyl cation.
135
THE SILVER(I)-PROMOTED NAZAROV REACTION
CI
TIPSO
O
CI
AgBF4 (1.5 equiv)
CI
CH2Cl2 (0.05 M), r.t.
70
74a
O
O CI
O CI
CI
Ph
Ph
Ph
75b, 78% (99%)
74c and 75c, 87%
O
O
O
CI
CI
75d, 45% (86%)
75e,74% (85%)
O CI
CI
74f and 75f, 63%
Figure 4.8. The halocyclopropane Nazarov reaction.
and in situ generation of a pentadienyl cation capable of undergoing Nazarov cyclization, as outlined previously. This proposal is supported by the observation silver(I) salts do not promote cyclization of unactivated a-chlorodienones under the standard reaction conditions (Scheme 4.23). 4.3.2
Interrupted Nazarov Reactions
To further expand the scope of this new silver(I)-mediated reaction sequence, interrupted Nazarov cyclizations were explored using the halocyclopropane chemistry, an investigation that was prompted by the discovery of an intriguing result. It was found that treatment of the phenethyl-substituted compound 76 with 1.5 equiv of silver tetrafluoroborate in dichloromethane provided benzohydrindenone 77 as the sole product, with no apparent formation of the simple a-chlorocyclopentenone (Scheme 4.24). This prompted an examination of appropriately substituted gemdichlorocyclopropane substrates in analogous interrupted Nazarov processes to ascertain the scope of this new cascade reaction sequence. O Cl Ph
AgBF4
no reaction
CH2Cl2 (0.05 M) r.t.
Scheme 4.23. Treatment of chlorodienones with AgBF4.
136
SILVER(I)-MEDIATED ELECTROCYCLIC PROCESSES Cl
TIPSO
Cl
OTIPS
O
1.5 equiv AgBF4
Cl
CH2Cl2, r.t. 63%
76a
XXVII
77a
Scheme 4.24. An interrupted Nazarov reaction using gem-dichlorocyclopropanes.
Although arene trapping has previously been demonstrated in traditional Nazarov reactions,40 the participation of an unactivated aromatic ring was unprecedented. This result implied that 2-silyloxycyclopentenyl cations (XXVII) derived from ring opening of the halocyclopropane moiety were much more reactive than were traditional Nazarov intermediates. This increased electrophilicity may be attributed to the presence of an electron-withdrawing chlorine substituent as well as the absence of stabilizing alkyl substitution on the cyclopentenyl cation XXVII. These cationic intermediates appear to be capable of reacting with even electron-deficient nucleophiles, which presents the possibility for new modes of trapping in the interrupted Nazarov reaction. To assess the generality of this reaction sequence, a number of substrates with one or more pendent arene moieties were synthesized and subjected to the optimized silver(I) reaction conditions (Fig. 4.9).41
Cl
TIPSO
Cl
O
1 equiv AgBF4 MeCN, reflux 80%
76a
77a
OMe MeO O
O
O
O Br
77b, 53%
77d, 79%
77c, 76%
77e, 48%
OH MeO O
O
O Cl
77f, 39%
78g, 89%
Cl
79h, 97%
Figure 4.9. Silver(I)-mediated interrupted Nazarov reactions.
137
THE SILVER(I)-PROMOTED NAZAROV REACTION
In most cases, the desired products were isolated in moderate yields and as a single regioisomer due to electrophilic attack at the least hindered position of the aromatic ring (i.e., para to the methoxy substituent in 77b). Also, all of the tricyclic products (77) had experienced dehydrochlorination and possessed a ring fusion alkene that was conjugated to the aromatic ring and ketone functionalities. Remarkably, a deactivated aromatic ring was capable of trapping the reactive 2-silyloxycyclopentenyl cation to furnish a bromosubstituted benzohydrindenone product (77e). This type of electrondeficient product would not be accessible from the conventional interrupted Nazarov reaction of divinyl ketone substrates, which demonstrates the value of this complementary trapping process. Another interesting observation was made when a silylprotected phenol was subjected to silver tetrafluoroborate in refluxing acetonitrile, prompting deprotection of the labile phenolic protecting group to provide benzohydrindenone 77f in 39% after an extended reaction time. When the effect of a substitution on the vinyl moiety was examined, the importance of an electron-rich aromatic trap was demonstrated. None of the desired trapping product was observed with a deactivated phenyl group (79h) due to stabilization of the 2-silyloxycyclopentenyl cation by alkyl substitution; however, when the 3-methoxyphenyl moiety was used in place of a nonsubstituted phenyl, the chlorocyclopentanone 78g was isolated in very good yield. The cis relationship between the hydrogen and methyl groups at the ring fusion, as well as the relative stereochemistry of the chlorine moiety, were confirmed by single-crystal X-ray crystallography. These results provided important insight into the mechanism of this cascade process (Fig. 4.10). It was proposed that the reaction proceeds through the typical silver(I)-assisted disrotatory ring opening of the gem-dichlorocycloproTIPSO R
Cl
TIPSO R
Cl Ag(I)
Cl
X
X
OTIPS
4π electrocyclization
R
X
AgCl
XXVIII
76 X –H
X R OTIPS
+
XXVII
HBF4
X R O
Cl H XXIX
–HCl / + H
Cl
(R = H)
H
H
78
XXX
X –H
O
+
OH
+
77
Figure 4.10. Proposed mechanism for arene trapping.
Cl
138
SILVER(I)-MEDIATED ELECTROCYCLIC PROCESSES
pane moiety, followed by 4p electrocyclization (Nazarov cyclization) to provide the 2-silyloxycyclopentenyl cation XXVII. This highly reactive cationic species would then be captured by a pendent arene nucleophile to afford the tricyclic intermediate XXIX on rearomatization. Protodesilylation of the silyl enol ether from the leasthindered, convex face of XXIX would then result in the formation of chlorocyclopentanones 78. In the absence of alkyl substitution at the bridgehead position, compounds 78 could experience dehydrochlorination to provide the benzohydrindenone products 77. Since elimination of HCl to generate an olefin is not observed when “R” is a methyl group, it was proposed that double-bond formation proceeds through a second oxyallyl cationic intermediate XXX,42 followed by conventional Nazarovtype elimination and protonation to furnish the final products. In an effort to provide experimental evidence for the formation of a second oxyallyl cationic intermediate, a gem-dichlorocyclopropane substrate was envisaged with two internal nucleophiles: one to participate in the initial interrupted Nazarov reaction, and the other to capture the second cationic species. Surprisingly, when the carefully designed substrate 80 was subjected to the optimized reaction conditions, an alternate mode of trapping occurred to generate the intriguing bridged bicyclic product 81 (Scheme 4.25). This reactivity proved to be a general process, providing the unique products in moderate yields following cyclopropanation and immediate treatment with silver tetrafluoroborate. These structures revealed that a cascade sequence was proceeding stereoselectively in every case to furnish a single product as the result of conrotatory 4p electrocyclization, electrophilic aromatic substitution at the least hindered position on the arene moiety (para to the MeO) in favor of six-membered ring formation, and desilylation with protonation from the exo face of the bicyclic product. Dehydrochlorination to form a second cationic intermediate did not occur in this case, due to structural restrictions imposed by the bridged architecture of 81. The cascade sequences presented herein demonstrate unprecedented modes of reactivity in Nazarov chemistry that are initiated by the silver(I)-promoted ring opening of halocyclopropanes. The ease with which the gem-dichlorocyclopropanes can be prepared, the relatively mild reaction conditions, and the efficiency of these processes make these substrates attractive intermediates for an application in natural product synthesis.
O OTIPS
80
1. CHCl3 50% aq. NaOH BnEt3NCl
OMe
via:
OTIPS
Cl
Cl Ph
2. AgBF4, MeCN reflux (57%; 2 steps)
MeO
81
XXXI
Scheme 4.25. The formation of bridged bicyclic products.
OMe
REFERENCES
4.4
139
CONCLUDING REMARKS
Silver(I) has been used extensively in the development of halocyclopropane chemistry. The electrocylic ring opening of substituted halocyclopropanes can be used to afford many useful synthetic intermediates, both simple and complex, by introduction of a heteroatom- or carbon-based nucleophile to capture the cationic intermediates. Silver(I) can assist in the removal of the halogen from these substrates, inducing ring opening at low temperatures and under mild reaction conditions compared to the unassisted ring-opening processes. In these instances, silver(I) is also helpful in the sequestration of free halide in reaction mixtures, providing long-lived cationic species that can proceed along alternate trapping or rearrangement pathways. This chemistry has also been used in the more recent design of a new approach to the Nazarov reaction, wherein the silver(I)-promoted electrocyclic ring opening of 1,1-dihalo-2(silyloxy)-2-vinylcyclopropanes was used to initiate Nazarov cyclization, leading to the formation of simple cyclopentenones. This sequence was further utilized in the development of novel trapping pathways that can be used to generate benzohydrindenone and bridged architectures in a single transformation. From this discussion, it is clear that the silver(I)-mediated electrocylic ring opening of halocyclopropanes has great potential for future applications in organic synthesis.
REFERENCES 1. (a) Pasto, D. J.; Garves, K., J. Org. Chem. 1967, 32, 778–781; (b) Masuike, T.; Furukawa, N.; Oae, S., Bull. Chem. Soc. Jpn. 1971, 44, 448–450; (c) Cohen, T.; Solash, J., Tetrahedron Lett. 1973, 27, 2513–2516; (d) Cookson, P. G.; Davies, A. G.; Roberts, B. P., J. Chem. Soc. Chem. Commun. 1976, 1022–1023. 2. For reviews on this topic, see (a) Kostikov, R. R.; Molchanov, A. P.; Hopf, H., Small ring compounds in organic synthesis, Top. Curr. Chem. 1990, 155, 41–80; (b) Banwell, M. G.; Reum, M. E., in Advances in Strain in Organic Chemistry, Halton, B., ed., JAI Press, Greenwich, CT, 1991, Vol. 1, pp. 19–64; (c) Fedorynski, M., Chem. Rev. 2003, 103, 1099–1132; (d) Halton, B.; Harvey, J., Synlett 2006, 13, 1975–2000. 3. Woodward, R. B.; Hoffmann, R., Conservation of Orbital Symmetry, Academic Press, New York, 1970. 4. (a) de Puy, C. H.; Schnack, L. G.; Hausser, J. W.; Wiedemann, W., J. Am. Chem. Soc. 1965, 87, 4006–4006; (b) Loozen, H. J. J.; Robben, W. M. M.; Richter, T. L.; Buck, H. M., J. Org. Chem. 1976, 41, 384–385;(c) Loozen, H. J. J.; de Haan, J. W.; Buck, H. M., J. Org. Chem. 1977, 42, 418–422. 5. Ito, S.; Ziffer, H.; Baz, A., J. Org. Chem. 1986, 51, 1130–1133. 6. Porter, N. A.; Ziegler, C. B.; Khouri, F. F.; Roberts, D. H., J. Org. Chem. 1985, 50, 2252–2258. ´ lvarez, R.; de Lera, A ´ . R., J. Org. Chem. 2004, 69, 9002–9010. 7. Faza, O. N.; Lo´pez, C. S.; A 8. Marvell, E. N., Thermal Electrocyclic Reactions, Academic Press, New York, 1980. 9. Banwell, M. G.; Cowden, C. J., Aust. J. Chem. 1994, 47, 2235–2254. 10. Roberts, J. D.; Chambers, V. C., J. Am. Chem. Soc. 1951, 73, 5034–5040.
140
SILVER(I)-MEDIATED ELECTROCYCLIC PROCESSES
11. McElvain, S. M.; Weyna, P. L., J. Am. Chem. Soc. 1981, 10, 2579–2588. 12. (a) Sandler, S. R., J. Org. Chem. 1967, 32, 3876–3881;(b) Parham, W. E.; Yong, K. S., J. Org. Chem. 1968, 33, 3947–3948;(c) Parham, W. E.; Yong, K. S., J. Org. Chem. 1970, 35, 683–685. 13. Banwell, M. G.; Sydnes, M. O., Aust. J. Chem. 2004, 57, 537–548. 14. Hart, N. K.; Johns, S. R.; Lamberton, J. A., Aust. J. Chem. 1968, 21, 2579–2581. 15. Makosza, M.; Wawrzyniewicz, M., Tetrahedron Lett. 1969, 10, 4659–4662. 16. Murphy, J. A.; Scott, K. A.; Sinclair, R. S.; Martin, C. G.; Kennedy, A. R.; Lewis, N., J. Chem. Soc. Perkin Trans. 1 2000, 2395–2408. 17. Banwell, M. G.; Cowden, C. J.; Gable, R. W., J. Chem. Soc. Perkin Trans. 1 1994, 3515–3517. 18. Banwell, M. G.; Gable, R. W.; Peters, S. C.; Phyland, J. R., J. Chem. Soc. Chem. Commun. 1995, 1395–1396. 19. Danheiser, R. L.; Morin, J. M. Jr.; Yu, M.; Basak, A., Tetrahedron Lett. 1981, 22, 4205–4208. 20. Sydnes, L. K.; Mungaroo, R.; Aanesen, B. A., Acta Chem. Scand. 1998, 52, 1386–1391. 21. Lee, J.; Kim, H.; Cha, J. K., J. Am. Chem. Soc. 1995, 117, 9919–9920. 22. Banwell, M. G.; Wu, A. W., J. Chem. Soc. Perkin Trans. 1 1994, 2671–2672. 23. Banwell, M. G.; Harvey, J. E.; Jolliffe, K. A., J. Chem. Soc. Perkin Trans. 1 2001, 2002–2005. 24. Stanislawski, P. C.; Willis, A. C.; Banwell, M. G., Org. Lett. 2006, 8, 2143–2146. 25. Banwell, M. G.; Vogt, F.; Wu, A. W., Aust. J. Chem. 2006, 59, 415–425. 26. Banwell, M.; Edwards, A.; Harvey, J.; Hockless, D.; Willis, A., J. Chem. Soc. Perkin Trans. 1 2000, 2175–2178. 27. Banwell, M. G.; Cowden, C. J.; Mackay, M. F., J. Chem. Soc. Chem. Commun. 1994, 61–62. 28. Banwell, M. G.; Harvey, J. E.; Hockless, D. C. R., J. Org. Chem. 2000, 65, 4241–4250. 29. Stanislawski, P. C.; Willis, A. C.; Banwell, M. G., Chem. Asian J. 2007, 2, 1127–1136. 30. Kostikov, R. R.; Varakin, G. S.; Molchanov, A. P.; Ogloblin, K. A., Zh. Org. Khim. 1996, 32, 367. 31. Gassman, P. G.; Tan, L.; Hoye, T. R., Tetrahedron Lett. 1996, 37, 439–442. 32. Banwell, M. G.; Beck, D. A. S.; Stanislawski, P. C.; Sydnes, M. O.; Taylor, R. M., Curr. Org. Chem. 2005, 9, 1589–1600. 33. Banwell, M. G.; Ma, X.; Taylor, R. M.; Willis, A. C., Org. Lett. 2006, 8, 4959–4961. 34. Wu, L.; Shi, M., Tetrahedron Lett. 2009, 50, 1636–1638. 35. Grant, T. N.; West, F. G., J. Am. Chem. Soc. 2006, 128, 9348–9349. 36. Nazarov, I. N.; Zaretskaya, I. I., Izvest. Akad. Nauk. SSSR Ser. Khim. 1941, 211–224. 37. (a) Paquette, L. A.; Hamme, A. T.; Ku, L. H.; Doyon, J.; Kreuzholz, R., J. Am. Chem. Soc. 1997, 119, 1242–1253; (b) Paquette, L. A.; Liu, Z.; Ramsey, C.; Gallucci, J. C., J. Org. Chem. 2005, 70, 8154–8161. 38. (a) Gee, S. K.; Danheiser, R. L., J. Org. Chem. 1984, 49, 1674–1678;(b) Danheiser, R. L.; Brisbois, R. G.; Kowalczyk, J. J.; Miller, R. F., J. Am. Chem. Soc. 1990, 112, 3093–3100.
REFERENCES
141
39. Grant, T. N., Modified Nazarov Reactions and Ring Expansion Chemistry: Useful Methodologies for the Construction of Carbocyclic and Heterocyclic Compounds, Ph. D. dissertation, Univ. Alberta, Edmonton, AB, 2008. 40. Browder, C. C.; Marms€ater, F. P.; West, F. G., Org. Lett. 2001, 3, 3033–3035. 41. Grant, T. N.; West, F. G., Org. Lett. 2007, 9, 3789–3792. 42. The formation of cyclic oxyallyl cations from dehydrochlorination of chlorocyclopentanones is precedented in the literature; see Harmata, M.; Elomari, S.; Barnes, C. L., J. Am. Chem. Soc. 2004, 82, 375–385.
5 SILVER-CATALYZED CYCLOISOMERIZATION REACTIONS PHILIPPE BELMONT Laboratory for Organometallic Chemistry, Heterocycles and Biological Targets, Unit 176: Conception, Synthesis and Targeting of Biomolecules, Institut Curie and CNRS, Paris, France
5.1 Introduction 5.2 Cycloisomerization of C¼O onto C¼C¼C 5.3 Cycloisomerization of C¼O onto C:C 5.4 Cycloisomerization of C¼N onto C¼C¼C 5.5 Cycloisomerization of C¼N onto C:C 5.6 Ene–Yne Cycloisomerization: C¼C onto C:C 5.7 Other Transformations 5.8 Conclusion References
5.1
INTRODUCTION
The impact of coinage metals (copper, silver, and gold) in organic synthesis is becoming important, especially with the worldwide impact of green chemistry, since their use as catalysts is part of the 12 “principles of green chemistry.”1 Moreover, silver salts are special since silver cations are being considered and evaluated as alternative biocides in water distribution systems, showing the great usefulness and nontoxicity (to humans) of monovalent silver cation.2 Although silver chemistry has a long history in organic chemistry, it has typically been used in stoichiometric amounts and is developed mostly for anion metathesis Silver in Organic Chemistry Edited by Michael Harmata Copyright 2010 John Wiley & Sons, Inc.
143
144
SILVER-CATALYZED CYCLOISOMERIZATION REACTIONS
X
[Ag]
X X Ag
Nu
(hetero)cyclic product
Ag
Scheme 5.1. Cycloisomerization diversity in silver chemistry.
(anion exchange, halogen scavengers) and oxidation reactions. More recent reviews pinpoint the current (r)evolution in silver chemistry.3–10 In fact, with the new gold rush,11–24 silver appeared as a partner to access more reactive cationic species. (Note that silver has also been used with other metals for this purpose: Pd, Cu, Rh, Ru, Pt). More recently catalysis with silver salts has become widespread, and this can be also explained by the s- and p-Lewis acidity properties of silver(I) complexes,25 leading to a variety of chemical transformations. Therefore, exploring new catalytic reactions with silver complexes is of great interest, and this work has been nicely presented in several reviews.3–10,26 The chemical reactions possible with silver catalysis are multiple and cover cycloadditions, cycloisomerizations, allylations, aldol reactions, and even C--H bond activation. Also, asymmetric versions are known, even though they still need to be improved.3–10 Our interest in this chapter is in silver-catalyzed cycloisomerization reactions. Therefore, we shall present different silver-catalyzed cycloisomerization reactions as a function of the nucleophilic and electrophilic moiety. Cycloisomerization reactions including the classical ene–yne cycloisomerization (with X ¼ CHR, Scheme 5.1), and the related heterocyclization reactions with heteroatoms embedded in unsaturated systems (X ¼ NR, O; Scheme 5.1) belong to the same reaction family. In addition, the alkynyl part can be exchanged for an allene unit. Internal or external nucleophiles (Nu) can then stabilize, through cascade reactions, the positive charge created.24 Heterocyclization reactions with saturated moieties (alcohols, amines, thiols, etc.) or acids on unsaturated counterparts (alkenes, allenes, alkynes, etc.) are not covered in this chapter since they are addition, and not isomerization, reactions. Silver is also widely used as an activating agent for producing highly reactive metallic cations (anion metathesis), which, in turn, may catalyze cycloisomerization reactions. This aspect is covered only when the silver control experiments give substantial positive results. A selection of the literature is also necessary in order to give the reader an overview of silver chemistry in the field of cycloisomerization reactions. Therefore, this chapter is not intended to be an exhaustive review of the literature, since more recent specialized reviews can be accessed for that purpose.3–10,26
5.2
CYCLOISOMERIZATION OF C¼O ONTO C¼C¼C
Following preliminary observations on allenyl alcohols from Gore27–31 and Balme32 along with Claesson and Olsnon,33 Marshall et al. demonstrated in a seminal publication34 that allenyl ketones (R3 ¼ alkyl, Scheme 5.2) or allenyl aldehydes
145
CYCLOISOMERIZATION OF C¼O ONTO C¼C¼C
R1
R2
R2
•
R3
AgNO3 or AgBF4
R1
0.5–1 h, 72–99%
O
O
R3
2
1
Scheme 5.2. Marshall’s work.34–38
(R3 ¼ H) 1 could selectively undergo a cycloisomerization reaction to produce various furan rings 2 under mild conditions with rhodium(I) or silver(I) catalysts. This transformation was further studied and the catalyst load could be decreased to 0.2 equiv.35–38 A mechanism was proposed through deuterium incorporation experiments, and the conclusion was that there was no 1,2 shift of the deuterium present in the starting material (A, Scheme 5.3) since exclusive formation of furans D deuterated on position 3 could be explained by the presence of an external source of deuterium (such as D2O). Therefore, it is believed that after silver(I) coordination to the allenyl system (A, Scheme 5.3), the attack by the carbonyl oxygen may lead to an oxo cation intermediate B. Finally, proton lost would generate silver furan C that would lead to furan D after silver release (Scheme 5.3).39 Ag+ H(D)
•
R1
R2 R3
R2
Ag H R1
+ O
R2
Ag
R3
(D)H
3
R2
H+
– H+
R1
O
R3
– Ag+
R1
O
R3
O
A
B
C
D
Scheme 5.3. Marshall’s mechanism.39
This methodology has been nicely used in the total synthesis of several natural products such as racemic kallolide B (3), (Scheme 5.4),37,38 a member of the pseudopterolide diterpene family. Thereafter, Hashmi et al. showed that this transformation (Scheme 5.2) was more efficiently catalyzed with palladium(II)40 and gold(I)41 catalysts. Moreover, following the cycloisomerization reaction, a tandem dimerization reaction is also possible on the same substrates under PdII, AgI, and AuIII catalysis, leading to different substituted furans (4 or 6) depending on the nature of the catalyst used (Scheme 5.5). Indeed, from compound 3 (Scheme 5.5), palladium(II) catalysis led to a 59% yield of 4, whereas silver(I) and gold(III) catalysis led to furans 5 and 6.41 • O
10% AgNO3 on silica gel (0.2 equiv.) hexane, r.t., 84%
TBSO
O
O O
TBSO
Scheme 5.4. Total synthesis of racemic kallolide B.37,38
O (+/–)-kallolide B
146
SILVER-CATALYZED CYCLOISOMERIZATION REACTIONS O
•
5 mol% AgNO3
R O
R
R
R
O
R O
O R
R = 4-(TBSO)C6H4
4
3
5 22%
6
14%
O
Scheme 5.5. Cycloisomerization/dimerization reaction.41
Hashmi’s team41 noticed that the cycloisomerization/dimerization reactions leading to 4 or 6 required, under the same reaction conditions (1 mol% catalyst), over a week for AgNO3, about an hour with PdCl2(MeCN)2 and only about a minute with AuCl3. Gevorgyan et al. in several landmark publications42,43 reported the silvercatalyzed (AgBF4) 1,2 migration of various functional groups such as acyloxy, phosphatyloxy, and sulfonyloxy, positioned on an sp2 carbon from an allenyl unit. The migration was accompanied by a cycloisomerization reaction leading to the efficient synthesis of functionalized furans. Therefore, from acyloxyallene E (Scheme 5.6), the 1,2 migration might generate a dioxolenium F, and cycloisomerization would lead to furan G. R O O
[Ag]
• O E
R + O O
O
O
O
-O F
R
G
Scheme 5.6. Proposed 1,2-migration pathway for furan synthesis.42
So, from acyloxy, phosphatyloxy, and sulfonyloxy alkynyl derivatives 7–9 (Scheme 5.7), on AgBF4 catalysis, allenyl compounds 10–12 are formed, leading to substituted furans 13–15. Note that alkynyl derivative 9 (Scheme 5.7) could not be isolated since during its synthesis, it directly formed allenyl substrate 12. Also, allene intermediate 11 could be isolated (in 56% yield) when the reaction was conducted at room temperature.42 Mechanistic studies provided through 17 O labeling43 of H (Scheme 5.8) showed that the nature of the migrating group was important. The 1,2-migration mechanism proposed earlier was confirmed for the migration of the acyloxy group (Schemes 5.6 and 5.7). Two alternative mechanisms are possible: (1) a direct trapping of the carbenoid (I, Scheme 5.8) by the ketone group leading to the formation of the furan ring J; and (2) a second 1,2-migration mechanism giving the allenyl intermediate K, which via two competitive oxirenium/dioxolenium pathways, affords the furan ring J.43 In addition to this latter work, a silver-catalyzed 1,2-alkyl/aryl shift is also possible with allenyl ketones (16), giving, as before, highly substituted furans (17, Scheme 5.9).44 Not surprisingly, Gevorgyan et al. observed that the migratory aptitude
147
CYCLOISOMERIZATION OF C¼O ONTO C¼C¼C OAc Ph
t-Bu O
AcO
5 mol% AgBF4 CH2Cl2, r.t.
•
t-Bu
Ph
AcO
Ph
t-Bu
Ph O
OP(O)(OEt)2 Ph
t-Bu O
13, >99%
O (EtO)2PO
5 mol% AgBF4 (CH2)2Cl2, 60°C
t-Bu
Ph
Ph
O
10
7
Ph
8
O (EtO)2PO
Ph
•
Ph
11 OTs Ph
t-Bu O
TsO
•
t-Bu
Ph
Ph
1 mol% AgBF4
Ph
t-Bu
(CH2)2Cl2, 60°C
O
9
Ph
14, 65% TsO
Ph
O
t-Bu
O
Ph
O
Ph
15, 82%
12
Scheme 5.7. Access to functionalized furans.42,43
O Y O
R2
O [3,3] migration
O
•
R1
K
O
O
O H
R3
R1
[1,2] migration O
CR 4, P(OR4)2, S(O)R4
R2
O
Y
via oxirenium
Y
Y+ O
1
O
[1,2] migration
R2
R
O
R R1 via oxirenium via dioxolenium
[1,2] migration
Y
1
Y
via dioxolenium R3
Y O+
O
R3
O J
R2
1
R3
R
[M] I
O
Scheme 5.8. Mechanistic pathways based on 17 O labeling.43
of a phenyl versus a methyl group was greater than 100 to 1, which is in agreement with their known migratory aptitudes. p-Philic (AgI, CuI, AuI) and oxophilic (Lewis acid catalysis; e.g., InCl3) activation are also both possible, and this reaction has been thoroughly studied.45 Note that the Me
Ph Ph
20 mol% AgOTf
• Ph O 16
Ph
3
Me
p-xylene, 140°C
Ph
O
Ph
17 79%
Scheme 5.9. [1,2]-phenyl migration.44
148
SILVER-CATALYZED CYCLOISOMERIZATION REACTIONS
R
•
R
R1
R1
R2HN
AgOAc (3 mol%)
O
acetone, r.t.
18
R1
R R
R N
O
R2
R
45–58%
19
O
N R2 20
R = H, R1 = OMe, R2 = alkyl: only 19 R = Me, R1 = H, R2 = alkyl: 19+20
Scheme 5.10. Cycloisomerization versus simple heterocyclization.50 R1 MeO2C Ph
R1
• NH
O 21
AgNO3, 20 mol%, acetone, 70°C, dark or AgBF4, 20 mol%, DCE, 65°C, dark
MeO2C
N
R1 O Ph 22
MeO2C 56–74%
N
O Ph 23
mixture 22 : 23, 3 :1–5:1
Scheme 5.11. Access to oxazines.
51
use of a proton scavenger such as 2,4,6-tris-tert-butylpyrimidine (TTBP) has a significant impact on reactivity, which is a sign of Brønsted acid catalysis during metal-catalyzed transformations, and this has already been noted in the literature.46–49 Brandsma et al.50 reported the cycloisomerization of 2,3-dienamides 18 (Scheme 5.10) to 2-(5H)-furanylidenamines 19 or 1,5-dihydro-2H-pyrrol-2-ones 20. The reaction is catalyzed by AgOAc (3 mol%). AgNO3 is also a good catalyst for this reaction, which is not the case for AgCN and AgBr. Depending on the substituents on 2,3-dienamides 18, one can obtain exclusively furanylidenamines 19 or mixtures containing various amounts of pyrrol-2-ones 20. Brummond et al. published a study51 with substrates homologous to those from Brandsma et al.50 Indeed, using amino acid–derived allenes 21 (Scheme 5.11), under AgBF4 or AgNO3 (20 mol%) catalysis, they obtained oxazines 22 accompanied by minor isomers 23.
5.3
CYCLOISOMERIZATION OF C¼O ONTO C:C
Agrofoglio et al. reported52 the synthesis of furo[2,3-d]pyrimidines 25 via AgNO3catalyzed 5-endo-dig cyclization (Scheme 5.12). Furo[2,3-d]pyrimidines 25 were obtained through activation of the alkynyl moiety of 24 by AgNO3 (0.2 equiv), giving intermediate L, followed by the 5-endo-dig cyclization. Van der Eycken’s group developed a silver(I)-mediated synthesis of substituted furo[2,3-b]pyrazines.53 Starting from p-methoxybenzyl-protected 3,5-dichloropyrazine-2(1H)-ones 26 (Scheme 5.13), after a regioselective microwave-assisted Sonogashira reaction with various terminal alkynes, the cycloisomerization reaction could occur using AgOTf (2 mol%) with trifluoroacetic acid (TFA, 5 equiv) to yield
149
CYCLOISOMERIZATION OF C¼O ONTO C:C
R1
R1
R1
O NH N R
O
Ag
AgNO3 (0.2 equiv)
N
acetone, RT
O
O
N R
24
H
– Ag
N
O
N O R 25 95–98%
L
Scheme 5.12. Access to the furo[2,3-d]pyrimidines.52
PMB R1 Cl
N N
O Cl
R2 Pd(PPh3)2Cl2 (1 mol%) CuI (3 mol %) Et3N/DMF, MW, 80°C 65–99%
PMB R1 Cl
26
O
N
AgOTf (2 mol%)
M
Ag
R2
N
Cl
N
O R2
TFA (5 equiv) CH2Cl2, r.t.
N
R1
27 79%–quant. H+
CF3CO2R1
– Ag
OMe
+ N
O
R1
N
Cl
N O
O
R2 Cl
N N
R2
Ag
Ag
Scheme 5.13. Furo[2,3-b]pyrazine synthesis.53
2-chlorofuro[2,3-b]pyrazines 27. Control experiments showed that both AgOTf and TFAwere needed for high yields and a low catalyst load.53 The PMB deprotection was proposed to occur right after the cyclization as a result of the formation of a cationic intermediate N, which reacts with the TFA anion (M--N--O, Scheme 5.13). Finally, the silver-linked furopyrazine intermediate O regenerates the catalyst on protonolysis and produces furopyrazines 27 in the range of 79% to quantitative (quant.) yield. Porco’s team reported the formation of benzopyrylium salts 29 (Scheme 5.14) from the action of silver and gold catalysts on o-alkynylbenzaldehydes 28.54 Although this reaction proceeds much faster with gold than with silver, the benzopyrylium formation is of interest and will be discussed further. C7H15 HO
HO
C7H15
[M] 0.05 equiv
Me OH
CHO 28
DCE/TFA (10:1), r.t.
O
Me OH
CF3COO
29
[M]: Au(OAc) 3, 1 min, 100% conversion AgNO3, 20 min, 94% conversion
Scheme 5.14. Benzopyrylium synthesis.54
150
SILVER-CATALYZED CYCLOISOMERIZATION REACTIONS
TABLE 5.1. Silver Diversity for a Versatile Reaction55 O 3
cat. (5 mol%)
H
O
MeOH, r.t.
N 2 30
[Ag] Complex AgOTf AgOAc Ag2O
OMe
OMe
N
OMe
Reaction Time 4h 12 h 10 min
31
5-exo-dig product 31 (Conversion %) — 48 100
OMe H
O
+ N 32
OMe H
6-endo-dig product 32 (Conversion %) 100 52 —
pKa <0 4.8 15.7
Belmont et al. worked on a variety of quinoline derivatives bearing a substituted alkynyl group on position 2 of the ring and a carbonyl function on position 3 (e.g., 30, Table 5.1). They accessed diverse furoquinoline and pyranoquinoline cores (e.g., 31 and 32, respectively, Table 5.1).55 One particularly interesting aspect of this work is that, starting from the same quinoline derivatives and depending on the nature of the silver salt used, one could access either a furoquinoline or a pyranoquinoline product. This behavior was rationalized from the value of the pKa of the conjugate acid of the silver counterion (Table 5.1). Three families of silver salts could be found depending on the results of the cyclization reaction. Counterions whose conjugate acids had a negative pKa value led to a 6-endo-dig product (e.g., AgOTf). Those with a pKa value over 10 led to a 5-exo-dig product (e.g., Ag2O). Finally, those with a pKa range between 2 and 5 afforded mixtures of products (e.g., AgOAc, Table 5.1). This observation is not related to traces of base or acid from the silver salts used since control experiments ruled out this possibility. It was known from the literature that the 5-exo-dig versus 6-endo-dig cyclization mode could depend on the nature of the carbonyl group,56,57 of the alkyne substituent,58,59 and of the nature60,61 and oxidation state62 of the metallic source used. Also, work from Yamamoto25 demonstrated the importance of both s- and p-Lewis acidity properties of silver(I) complexes. Therefore, depending on the silver salt used, two mechanistic pathways were proposed (pathways A and B, Scheme 5.15). The reactivity of the first group (e.g., AgOTf, Table 5.1) could be explained by pathway A (Scheme 5.15), where after coordination of the triple bond of 30 (intermediate P) the activated alkyne is attacked by the carbonyl oxygen yielding a pyrylium intermediate Q.63,64 Then, methanol addition on intermediate Q would give intermediate R and, after protonation and regeneration of the catalyst, the 6-endo-dig product 32. The second silver salt group (e.g., Ag2O, Table 5.1), whose oxidative properties should favor direct coordination to the oxygen atom of the carbonyl function, would follow pathway B (Scheme 5.15). Therefore, intermediate S (Scheme 5.15, pathway B), where the silver salt acts as a Lewis acid,25,65 could give acetal T, which on cyclization would give complex U. Expulsion of Ag þ would
151
CYCLOISOMERIZATION OF C¼O ONTO C:C O H [Ag]
O
N
N
MeOH
O
OMe
30
H
H N
OMe
S
P
OMe Ag
Ag O OMe
Ag2O
Pathway B N
+ O
AgOTf
Pathway A
OMe
N
OMe
T
Q
O
O N U
Ag
OMe
OMe
Ag
R O
O N
31
Ag
OMe
OMe
MeOH
OMe
N
OMe
OMe
N
OMe
H
32
H
Scheme 5.15. Proposed reaction pathways.55
yield the 5-exo-dig product 31. The third silver salt group (e.g., AgOAc, Table 5.1) could engage in both pathways, leading to poor regioselectivity. Wu’s team published additional results, related to Porco’s54 and Belmont’s55 previous work on 2-alkynylbenzaldehyde substrates using AgOTf catalysis (see Scheme 5.14 and Table 5.1), but in their case diethyl phosphite [HPO(OEt)2] was used as the nucleophile instead of methanol.66 Padwa et al.67 have used 3-oxo-N-propargylamides and esters to build furopyrrolidinones and furodihydrofuranones via a silver(I) alkyne activation (Schemes 5.16 and 5.17). Me O
O
Me
O AuCl 3
N R
33 R = CH3
OH O N
V
O
O
N
34 59%
NMe
35 78% Ag
2.2 equiv AgNO3 /NaOAc H
Me
Me
R=H
Me
O
Ag+
Me
+ O
Me
Ag+ O
O NMe
H+
Me
O
Me
NMe
HO
NMe
Ag
W
X
Scheme 5.16. Furopyrrolidinone and oxazole synthesis.67
O
Y
152
SILVER-CATALYZED CYCLOISOMERIZATION REACTIONS
O
Me
O
Me
O Cl 36
1.05 equiv AgNO3 1.05 equiv NaOAc
O
O O
37 82%
Scheme 5.17. Furodihydrofuranone synthesis.67 Ag
R
R
AgSbF6, 5 mol%
O
R
O
CH2Cl2, 35°C R3–OH (2 equiv)
O
•
R1
R1 R2–OH
38
Z
2 R1 O–R
39
65–99%
68
Scheme 5.18. Allenylisochromene synthesis.
The authors used silver salts since gold salts catalyzed the reaction with R¼H (giving oxazole 34, Scheme 5.16) but not with R¼Me. Moreover, only traces of the desired furopyrrolidinone were formed with the use of a cationic gold species activated with silver additives. Therefore, silver traces were thought to be the active reagent. Indeed, on activation of compound 33 mediated by AgNO3 in the presence of sodium acetate (Scheme 5.16), the enol moiety V can then accomplish a nucleophilic attack to produce the pyrrolidinone W and after protonolysis give compound X. Pyrrolidinone Y (the enol version of X) can, in turn, be subject to an oxidative cyclization to yield the furopyrrolidinone 35. Two equivalents of silver salts are needed for the activation step and the oxidative cyclization. However, using a chlorinated version of a related ester compound 36 (Scheme 5.17) required only one silver equivalent to prepare pyrrolofuranone 37, since chlorine elimination directly provided the final product without the need of a silver-mediated oxidation. Yamamoto described a cascade cyclization reaction to prepare allene-substituted isochromenes 39 (Scheme 5.18).68 Diynones 38, when submitted to the action of AgSbF6 (5 mol%), formed a benzopyrylium intermediate Z (identified by NMR spectroscopy), which could undergo a 1,4-Michael-type addition with alcohol nucleophiles to produce allenylisochromenes 39 (Scheme 5.18).
5.4
CYCLOISOMERIZATION OF C¼N ONTO C¼C¼C
This transformation has been less widely studied. We chose two publications from Gallagher’s team to show a few examples.69,70 Allenic oximes 40 (Scheme 5.19) were treated with AgBF4 to produce cyclic nitrones 41. Since these nitrones were not sufficiently stable, they were trapped with 1,3-dipolarophiles to give bicyclic or
153
CYCLOISOMERIZATION OF C¼N ONTO C:C
n N OH 40
n
AgBF4 (0.5–1 equiv)
•
N O
DCE or CH2Cl2
n = 1, 2
R1
R
H
H N O
R
with n = 2
R1
41
42
Scheme 5.19. Cyclic nitrones and their 1,3-dipolar adducts.70
tricyclic adducts 42 (Scheme 5.19). Therefore, trans-2,6-disubstituted piperidines and trans-2,5-pyrrolidines could be obtained using this sequence.69,70 This methodology has been used to produce various structures, including hexahydroazepines71 and pyrrolizidine alkaloids.69,72 5.5
CYCLOISOMERIZATION OF C¼N ONTO C:C
Larock’s group reported73 the synthesis of isoquinolines 44 (Scheme 5.20) by the electrophilic cyclization of iminoalkynes 43. This reaction has been extensively studied with various electrophiles such as I2, ICl, ArSCl, and ArSeCl and also under palladium catalysis. Among the electrophiles tested, AgNO3 produced interesting results since several heterocycles 44 (Scheme 5.20) could be prepared on the activation of iminoalkynes 43 with 5 mol% of this reagent (leading to intermediate AA, Scheme 5.20). The tert-butyl group is lost from intermediate BB during the reaction as an isobutene to afford product 44 (Scheme 5.20). Note that this reaction is also possible with pyridyliminoalkynes 43. Porco et al.74 accessed pyrroloisoquinoline heterocycles 46 by employing a domino process. Initially, a metal-catalyzed cycloisomerization of alkynyl N-benzylidene glycinates 45 gave azomethine ylide CC (Scheme 5.21), which underwent
N
(N)
t-Bu
AgNO3 (5 mol%)
N
(N)
t-Bu (N)
N
CHCl3, 50°C
− Ag
N
(N)
R
R
43
t-Bu
AA
Ag
R
Ag
R
44 45–92%
BB
R = alkyls, aryls +H
Scheme 5.20. Isoquinoline and aza derivatives.73
R
AgOTf (10 mol%) DTBMP
R1 N
2
R
45
CO2R5
toluene, 60°C R3
R1
R1
R N
R2
R4
R3
CO2
R5
R
oxidation
R2
R4
CC
Scheme 5.21. Pyrroloisoquinoline synthesis.74
N R3
46
28–78%
CO2R5 R4
154
SILVER-CATALYZED CYCLOISOMERIZATION REACTIONS R
Ag
R1
R1 N
R2
R
10 mol% AgSbF6
Nu
R2
DCE, 40°C
CO2Me
+ N CO2Me
47
R1
R N
R2
DD
48
Nucleophile (Nu) = β-ketoesters, malonates,silylenol ethers
CO2Me
Nu 28–78%
Scheme 5.22. Access to the dihydroisoquinoline core.75
a dipolar cycloaddition with an activated alkyne such as dimethylacetylene dicarboxylate (DMAD), leading to 46 after air oxidation. Both AgOTf and AgSbF6 (10 mol%) were the best catalysts for this transformation, and the reaction was conducted in the presence of an hindered base, 2,6-di-tertbutyl-4-methylpyridine (DTBMP). The azomethine ylide CC intermediate has been identified in a test reaction in the absence of DMAD, via mass spectrometry. The scope and limitations have been thoroughly studied for reaching pyrroloisoquinoline scaffold, which is a substructure of the lamellarin alkaloid family. The same research group developed a tandem cycloisomerization–addition reaction on o-alkynylacylimines 47 (Scheme 5.22) to generate the dihydroisoquinoline core 48.75 This type of reaction had been reported by Asao et al. and will be discussed shortly.76 The best catalyst for this transformation was AgSbF6 (10 mol%), and b-ketoesters, malonates, and silyl enol ethers have been used for the nucleophilic addition on the pyridinium intermediate DD. The dihydroisoquinolines 48 have been further used in several reactions in order to assemble the framework of various alkaloids. One example is given in the formation of dihydroisoquinoline 49, bearing a pendent a,b-unsaturated ketone. Compound 49 can rearrange to the tetracycle 50 (related to the core structure of karachine, Scheme 5.23), using TMSOTf, via a tandem Michael addition–Mannich reaction process (intermediates EE and FF). As mentioned earlier, Asao et al.76 have developed similar systems with particular nucleophiles such as nitromethane or terminal alkynes, along with more classical ones OTMS
OTMS
Me R1
Me
R N
R2
49
Me
TMSOTf CO2Me Me
N Me CO2Me
CH2Cl2, 40°C 76%
EE
N Me CO2Me
FF
Me
Z : E = 1: 3 O
O
Me
Me
O
O O N
MeO MeO
Karachine
50
Scheme 5.23. Toward karachine alkaloid structure.75
N Me CO2Me
155
CYCLOISOMERIZATION OF C¼N ONTO C:C
Nu N
Ph
N
3 mol% AgOTf
Nu
DCE, 80°C
Ph
Ph
51
Ph
52 42–93%
Nucleophile (Nu) = CH3NO2, CH3CH2NO2, terminal alkynes, β-ketoesters, malonates Ph NO2 N
Ph
N
Ph 53
Ph Ph
54
85%
93%
Scheme 5.24. Dihydroisoquinolines synthesis.76
(b-ketoesters, malonates). o-Alkynylaryl aldimines 51 (Scheme 5.24) can thus produce 1,2-dihydroisoquinolines 52 via AgOTf (3 mol%) catalysis. It is worth noting that nucleophiles such as nitromethane and terminal alkynes give interesting products such as 53 and 54, respectively (Scheme 5.24). The mechanism proceeds through activation of the alkynyl moiety (GG, Scheme 5.25), leading to the isoquinolinium intermediate HH. After addition of
N 51
NO2 N
Ph
Ph N
Ph [Ag] Ph
53
Ph
GG
85% OH N O Ph N
Ag
Ph
N
Ph II
H
Ph
Ag
O N
O
Ph
HH
OH N O
Scheme 5.25. Proposed mechanism.76
Ag
156
SILVER-CATALYZED CYCLOISOMERIZATION REACTIONS
N
+ ON
OH AgOTf (5 mol%) CH2Cl2
55
R
r.t. or 70°C
CN TMSCN (1.1 equiv) DBU (1.5 equiv) r.t., 12 h
R 56
N
70–97%
R 57
67–84%
Scheme 5.26. Synthesis of isoquinoline-N-oxides.80
nitromethane, and protonolysis of intermediate II, the silver catalyst is regenerated along with dihydroisoquinoline 53 (Scheme 5.25).76 Wu’s group reported studies related to previous results from the groups of Porco and Asao (see Schemes 5.22 and 5.24, respectively). Indeed, the iminium formed in situ could be the substrate for further addition of diethyl phosphite77 or enamine78 nucleophiles, and could also be reduced with sodium borohydride,79 yielding various 1,2-dihydroisoquinolines. Shin’s group80 also used AgOTf (5 mol%) to catalyze the formation of isoquinolineN-oxide structures 56 (Scheme 5.26) from cycloisomerization of 2-alkynylbenzaldoximes 55. This synthesis is of interest since these structures are usually made by oxidation of the parent nitrogen heterocycles. Other silver salts can efficiently catalyze this reaction at room temperature. These include AgSbF6 and AgNTf2 and also some gold salts such as Au(IMes)OTf. Some of the isoquinoline-N-oxide derivatives 56 (Scheme 5.26) can be further converted in 1-cyanoisoquinolines 57 by reacting with TMSCN and DBU as a base. Wu’s group worked on the same transformation (55–56, Scheme 5.26) with AgOTf (5 mol%) and performed a tandem [3 þ 3] cycloaddition on intermediates 56 with dimethylcyclopropane-1,1-dicarboxylate to generate oxazine-fused dihydroisoquinoline heterocycles.81 An efficient, metal-catalyzed synthesis of N-fused heterocycles 59 (Scheme 5.27) has been designed by Gevorgyan’s team,82,83 without the need of any base or ligand. Indeed, heterocycles bearing a propargyl group 58 (Scheme 5.27) can be smoothly transformed to N-fused heterocycles 59 owing to AgBF4 (3 mol%) catalysis. Some other salts can catalyze efficiently this reaction, including CuCl, AuI, and AgPF6. The substituents on the alkyne (R2) and at the propargylic position (R1) are rather diverse, since R2 can be a (cyclo)alkyl or a (hetero)aryl group and R1 an acetyloxy, an O-TBS, or a diethylphosphatyloxy group (Scheme 5.27). Since the aromatic moiety bearing the propargyl unit can be a pyridine 60, a quinoxaline 61, or a R1 n = 0, 1 A = N or S R1 = OP(O)(OEt)2, OTBS, OAC R2 = H, Alk, Ar, Het, alkenyl, alkynyl
A
n N
R1 R2
AgBF4 (3 mol%)
A
CH2Cl2, r.t.
58
Scheme 5.27. Fused N-containing heterocycles.82,83
n N R2
59
52–95%
157
ENE–YNE CYCLOISOMERIZATION: C¼C ONTO C:C R1
N
R1 AgBF4 (3 mol%)
R2
63
64–95%
R1 AgBF4 (3 mol%)
R2
N
R2
N
N 2
64 R
52–72%
R1
S
CH2Cl2, r.t.
61
R2
R1
R1
N N
N
CH2Cl2, r.t.
60
S
AgBF4 (3 mol%)
62
N
CH2Cl2, r.t.
R2
65
82–89%
Scheme 5.28. Fused-ring diversity.82,83 R4 N
O AgOTf (5 mol%)
R4–NH2
R1 2
R
or Au(PPh3)OTf (5 mol%)
3
R
67
R1
R3
R2
DCE, 50–80°C
68 13–92%
66
R4
N
R1 R2
M
R3
JJ
Scheme 5.29. Access to the pyrrole scaffold.84
thiazole 62, after the cycloisomerization one can obtain indolizines 63, pyrroloquinoxalines 64, and pyrrolothiazoles 65, respectively (Scheme 5.28). Dake’s group84 published an interesting report in which AgOTf and cationic gold (I) complexes were compared for their use in the synthesis of the pyrrole scaffold. From b-alkynyl ketones 66 (Scheme 5.29) and various primary amines 67, the imine intermediates JJ were formed in situ and the intramolecular cyclization produced various pyrroles 68. Both catalysts AgOTf (5 mol%) or Au(PPh3)OTf (5 mol%) were efficient, but the reaction proceeded more rapidly with silver catalysis.
5.6
ENE–YNE CYCLOISOMERIZATION: C¼C ONTO C:C
Few examples of ene–yne cycloisomerization reactions are seen in the literature. The first results for ene-yne cycloisomerizations were with systems bearing an heteroatom (amine or oxygen) next to the alkene counterpart (forming an enamine or an enol ether). Indeed, Dake’s group reported the cyclization of enesulfonamides on alkynes (69–70, Scheme 5.30) under catalysis by platinum and silver salts.85 Catalysis using AgOTf (1–4 mol%) was particularly efficient with systems such as 69 (Scheme 5.30)
158
SILVER-CATALYZED CYCLOISOMERIZATION REACTIONS
EWG
EWG AgOTf (1–4 mol%) CH2Cl2, 40°C or THF/DCE, 60°C
N Ts 69
N major isomer Ts 70 31–99%
Scheme 5.30. Ene-sulfonamides cyclization on activated alkynes.85
CO2Me
CO2Me AgOTf (2 mol%) THF–CH2Cl2, 60°C
N Boc 71
N Boc 72 91%
Scheme 5.31. Access to Boc-protected product.85
bearing an electron-withdrawing group on the alkynyl moiety (e.g., ketones, esters, amides, nitriles). Usually the use of silver catalysts required lower catalyst loading and gave higher yields compared with platinum(II) catalysis. Also, in order to obtain a nitrogen protecting group that would be easier to remove than the p-toluenesulfonyl group (Ts), the tert-butoxycarbamate (BocNR2) derivative 71 (Scheme 5.31) was also engaged in the same reaction to efficiently yield the desired dienyl product 72. The diene products have been used in the Diels–Alder reaction with success. For example, the reaction of eneyne 73 with a-methylacrolein gave the tricyclic compound 74 in a nice 93% yield over two steps (Scheme 5.32). Tanaka et al. published a route to 2-pyridones from the cycloisomerization of N-alkenyl alkynylamides.86 This reaction did not work with N-alkenyl alkynylamide 75 (Scheme 5.33) using AuClPPh3, even in refluxing dichloroethane (DCE). However, the cationic complex AuPPh3BF4 (AuClPPh3 þ AgBF4) gave a high yield of the desired pyridinone 76 at room temperature. The silver salt AgBF4 was the only catalyst tested along with AuPPh3BF4 that was efficient at room temperature.
CO2CH3 N Ts 73
H
CO2CH3
1). AgOTf (1 mol%) CH2Cl2, 40°C 2). Cooled to -78°C 3). α-methylacrolein BF3.OEt2
N CH3 H Ts CHO 74
93%
Scheme 5.32. Tandem cycloisomerization–cycloaddition reaction.85
159
ENE–YNE CYCLOISOMERIZATION: C¼C ONTO C:C
Ph
O
Catalyst system:
Ph
5 mol% cat.
N Bn
DCE
N Bn
AuCl(PPh3), 80°C, 6 h, 0%
O
AuCl(PPh3)/AgBF4, 0.5 h, 82% AgBF4, r.t., 5 h, 52%
76
75
Scheme 5.33. Pyridone synthesis.86
OTBS
OTBS R2
R2
3 AgOTf (5 mol%)
R3
N 2
R3
DCE, 50°C
77
R1
N
R1
78
Scheme 5.34. Polysubstituted acridine synthesis.
72%–quant. 87
Belmont’s group87 reported a cycloisomerization reaction on quinolines 77 (Scheme 5.34) bearing a silyl enol ether group on position 3 and an alkynyl group on position 2, leading to acridine derivatives 78. Although this reaction worked nicely with an activated gold species (AuPPh3SbF6), they studied this reaction under silver catalysis with a broad range of complexes. Of all the silver salts tested, only those having a negative pKa for the conjugate acid of silver counterion were efficient in the reaction: AgSbF6, AgPF6, AgOTf, and AgNO3. All the other silver salts tested (Ag2SO4, AgF, AgOAc, Ag2O, AgO) could not catalyze the reaction. Complexation of the silver salt to the alkynyl OTBS H
R2 3
R
[Ag]
1
N
78
OTBS
R2 R3
R
N
H proton migration O
H
R3
R1
N
LL
complexation
TBS H
R2
77
O
TBS
R2 R3
N
KK
Ag
[Ag]
cycloisomerization
Scheme 5.35. Proposed mechanism pathway.87
R1
R1
160
SILVER-CATALYZED CYCLOISOMERIZATION REACTIONS
O
R2
R1 R2
AgSbF6 (2 mol%)
R1
CH2Cl2, r.t., 30 min
Ar
Ar O 80 71–89%
79
Scheme 5.36. Synthesis of indenyl ketones.89
moiety seems compulsory, and this can explain why some other silver salts are not effective, since they are best known for their oxidative properties and therefore are more oxophilic than p-acidic.25,65 Therefore, after the complexation step on 77, intermediate KK (Scheme 5.35) could undergo an ene–yne cycloisomerization assisted by the alkyne activation and the presence of the oxygen atom to yield the silver complex intermediate LL. Then, proton migration and regeneration of the catalyst would give product 78. A large variety of acridine derivatives were formed with R1 group such as alkyls, (hetero) aryls, and even the ferrocenyl group.87 It is remarkable that Echavarren’s group88 succeeded, using a silver catalysis, in reacting 1,6-ene-ynes bearing an allyl stannane to yield vinyl-stannanes along with tin-free compounds.
5.7
OTHER TRANSFORMATIONS
Liu et al. obtained diverse carbocyclic and heterocyclic frameworks by employing a silver(I)-catalyzed cycloisomerization of epoxyalkynes.89 With AgSbF6 (2 mol%), epoxides 79 were transformed in 3-1H-indenyl ketones 80 in high yields (Scheme 5.36). The reaction has been exemplified with a variety of aryl-substituted alkynes, but it is interesting to note that a similar reaction with an epoxide–ynamide derivative 81 (Scheme 5.37), catalyzed with AgBF4, yielded the tricyclic central core 82 of gibberic acid. These reactions occurred, as in Gevorgyan’s work,44 with a 1,2-alkyl/aromatic shift, and in some cases led to a ring expansion; for example, epoxide 81 gave the tricycle 82 on reaction with AgBF4 (5 mol%) (Scheme 5.37). It was demonstrated through trapping experiments that the reaction catalyzed with silver(I) did not proceed through a carbenoid species but instead through a carbocation intermediate. O
H AgBF4 (5 mol%)
O CH2Cl2, r.t., 20 min
NMeTs 81
NMeTs O 82
57%
CO2H Gibberellic acid
Scheme 5.37. Access to the central core of gibberic acid.89
161
OTHER TRANSFORMATIONS O
AgSbF6 (2 mol%)
Ph
Ph
Me Ph
Me
Ph
Ph
Ag
MM
83
Ph Me
O
O Ag
Me
O
Ph
Ag Ph
OO
NN
Ph Me
84
Ph O 78%
Scheme 5.38. Proposed mechanism pathway.89
[M] cat.
O 85
O
O
O
86
87
88
Scheme 5.39. Chemical diversity under silver and gold catalyses.90
Therefore, from epoxide 83 (Scheme 5.38), a 1,2-phenyl shift in intermediate MM would provide oxonium derivative NN, which would furnish, after rearrangement of the oxetane intermediate OO, the desired ketone 84. It is interesting to note that on the same substrates cationic gold derivatives (activated with silver salts) did not lead to the same final compounds, showing the unique reactivity of silver salts. Liu’s group90 also reported the metal-catalyzed formation of 2-cyclopentenone 86 from cis-2,4-dien-1-al 85 (Scheme 5.39). It is interesting to note that with the same starting material, AuClPPh3 and AuCl catalyzed the formation of pyrano derivative 87 (32% and 63%, respectively), whereas the cationic gold derivatives AuPPh3SbF6 (AuClPPh3 þ AgSbF6) gave only traces of 3-cyclopentenone 88 (15%). Therefore, silver salts were tested alone and, indeed, AgOTf efficiently catalyzed the formation of 2-cyclopentenone 86 (91%, Scheme 5.39). Wu et al.91 published a tandem addition–cyclization reaction using 2-alkynylbenzenamines 89 (Scheme 5.40). The reaction proceeded under AgOTf catalysis R3-NCS (2 equiv) AgOTf, 5 mol%
NH2 R1
H N R1
89
R3
S
THF, 25°C
R2
H N
Ag
H N R1
S
R2
PP
Scheme 5.40. Thiourea additions on alkynes.91
N
R2
90
55–98%
R3
162
SILVER-CATALYZED CYCLOISOMERIZATION REACTIONS
(from 1 to 5 mol%) via thiourea intermediates PP. Because of the high electrophilicity of the isothiocyanate group, the tandem reactions provided a family of 2,4dihydro-1H-benzo[d][1,3]thiazines 90 (Scheme 5.40) in fair to high yields (55–98%) regardless of the electronic nature of the R3 group on the isothiocyanate partner. Very interesting work from Kirsch’s group was not reported here since silver(I) was used to catalyze a propargyl–Claisen rearrangement on ene–yne systems leading to allenyl ketones and not the following cycloisomerization step where gold(I)92 or DBU93 were used.
5.8
CONCLUSION
This overview of silver-catalyzed cycloisomerization reactions shows the great flexibility of the method and its usefulness in reaching the core of numerous heterocycles. On comparison with other metallic species, we are able to detect areas where efforts are still needed, such as the cycloisomerization of simple nonactivated ene–yne systems, as well as C--H bond activation. About two-thirds of the applications for silver in organometallic homogeneous catalysis reported in the literature involve activation of other metallic sources through anion metathesis to generate cationic species. Nevertheless, we can clearly see that silver has gained attention more recently not only for halogen scavenging or oxidation purposes but also for its own reactivity, which is sometimes very distinct from that of other related metals such as copper and gold. Further research is needed to carefully establish the reactivity of the metallic center and therefore clearly eliminate the possibility of simple Brønsted acid catalysis due to catalyst instability. We may wonder whether, following today’s “gold rush” in homogeneous catalysis, silver will be the next in line!
REFERENCES 1. Anastas, P.; Warner, J., Green Chemistry: Theory and Practice, Oxford Univ. Press; New York, 1998. 2. Wu, M. Y.; Suryanarayanan, K.; van Ooij, W. J.; Oerther, D. B., Water Sci. Technol. 2007, 55, 413–419. 3. Lipshutz, B. H.; Yamamoto, Y., Chem. Rev. 2008, 108, 2793–2795. 4. Naodovic, M.; Yamamoto, H., Chem. Rev. 2008, 108, 3132–3148. 5. Weibel, J. -M.; Blanc, A.; Pale, P., Chem. Rev. 2008, 108, 3149–3173. 6. Alvarez-Corral, M.; Munoz-Dorado, M.; Rodriguez-Garcia, I., Chem. Rev. 2008, 108, 3174–3198. 7. Yamamoto, Y., Chem. Rev. 2008, 108, 3199–3222. 8. Dias, H. V. R.; Lovely, C. J., Chem. Rev. 2008, 108, 3223–3238. 9. Diaz-Requejo, M. M.; Perez, P. J., Chem. Rev. 2008, 108, 3379–3394.
REFERENCES
10. 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. 34. 35. 36. 37. 38. 39. 40. 41. 42. 43. 44. 45. 46.
163
Patil, N. T.; Yamamoto, Y., Chem. Rev. 2008, 108, 3395–3442. Hashmi, A. S. K., Chem. Rev. 2007, 107, 3180–3211. Hashmi, A. S. K., Angew. Chem. Int. Ed. 2005, 44, 6990–6993. Hashmi, A. S. K., Gold Bull. 2004, 37, 51–65. Li, Z.; Brouwer, C.; He, C., Chem. Rev. 2008, 108, 3239–3265. Arcadi, A., Chem. Rev. 2008, 108, 3266–3325. Jimenez-Nunez, E.; Echavarren, A. M., Chem. Rev. 2008, 108, 3326–3350. Gorin, D. J.; Sherry, B. D.; Toste, F. D., Chem. Rev. 2008, 108, 3351–3378. Muzart, J., Tetrahedron 2008, 64, 5815–5849. Jimenez-Nunez, E.; Echavarren, A. M., Chem. Commun. 2007, 333–346. Zhang, L.; Sun, J.; Kozmin, S. A., Adv. Synth. Catal. 2006, 348, 2271–2296. Asao, N., Synlett 2006, 1645–1656. Dyker, G., Angew. Chem. Int. Ed. 2000, 39, 4237–4239. Hashmi, A. S. K.; Rudolph, M., Chem. Soc. Rev. 2008, 37, 1766–1775. Kirsch, S. F., Synthesis 2008, 3183–3204. Yamamoto, Y., J. Org. Chem. 2007, 72, 7817–7831. Halbes-Letinois, U.; Weibel, J. -M.; Pale, P., Chem. Soc. Rev. 2007, 36, 759–769. Chilot, J. J.; Doutheau, A.; Gore, J., Tetrahedron Lett. 1982, 23, 4693–4696. Arseniyadis, S.; Gore, J., Tetrahedron Lett. 1983, 24, 3997–4000. Chilot, J. J.; Doutheau, A.; Gore, J.; Saroli, A., Tetrahedron Lett. 1986, 27, 849–852. Bluthe, N.; Gore, J.; Malacria, M., Tetrahedron 1986, 42, 1333–1344. Bernard, D.; Doutheau, A.; Gore, J., Tetrahedron 1987, 43, 2721–2732. Balme, G., Ph.D. thesis, Univ. Claude Bernard, Lyon, 1979. Olsson, L. I.; Claesson, A., Synthesis 1979, 743–745. Marshall, J. A.; Robinson, E. D., J. Org. Chem. 1990, 55, 3450–3451. Marshall, J. A.; Wang, X. J., J. Org. Chem. 1991, 56, 4913–4918. Marshall, J. A.; Wang, X. J., J. Org. Chem. 1992, 57, 3387–3396. Marshall, J. A.; Wallace, E. M.; Coan, P. S., J. Org. Chem. 1995, 60, 796–797. Marshall, J. A.; Sehon, C. A., J. Org. Chem. 1995, 60, 5966–5968. Marshall, J. A.; Bartley, G. S., J. Org. Chem. 1994, 59, 7169–7171. Hashmi, A. S. K.; Rupert, T. L.; Knoefel, T.; Bats, J. W., J. Org. Chem. 1997, 62, 7295–7304. Hashmi, A. S. K.; Schwarz, L.; Choi, J. -H.; Frost, T. M., Angew. Chem. Int. Ed. 2000, 39, 2285–2288. Sromek, A. W.; Kel’in, A. V.; Gevorgyan, V., Angew. Chem. Int. Ed. 2004, 43, 2280–2282. Schwier, T.; Sromek, A. W.; Yap, D. M. L.; Chernyak, D.; Gevorgyan, V., J. Am. Chem. Soc. 2007, 129, 9868–9878. Dudnik, A. S.; Gevorgyan, V., Angew. Chem. Int. Ed. 2007, 46, 5195–5197. Dudnik, A. S.; Sromek, A. W.; Rubina, M.; Kim, J. T.; Kel’in, A. V.; Gevorgyan, V., J. Am. Chem. Soc. 2008, 130, 1440–1452. Hashmi, A. S. K., Catal. Today 2007, 122, 211–214.
164
SILVER-CATALYZED CYCLOISOMERIZATION REACTIONS
47. Li, Z.; Zhang, J.; Brouwer, C.; Yang, C. -G.; Reich, N. W.; He, C., Org. Lett. 2006, 8, 4175–4178. 48. Rosenfeld, D. C.; Shekhar, S.; Takemiya, A.; Utsunomiya, M.; Hartwig, J. F., Org. Lett. 2006, 8, 4179–4182. 49. Rhee, J. U.; Krische, M. J., Org. Lett., 2005, 7, 2493–2495. 50. Nedolya, N. A.; Schlyakhtina, N. I.; Zinov’eva, V. P.; Albanov, A. I.; Brandsma, L., Tetrahedron Lett. 2002, 43, 1569–1571. 51. Mitasev, B.; Brummond, K. M., Synlett 2006, 3100–3104. 52. Aucagne, V.; Amblard, F.; Agrofoglio, L. A., Synlett 2004, 2406–2408. 53. Ermolat’ev, D. S.; Mehta, V. P.; Van der Eycken, E. V., Synlett 2007, 3117–3122. 54. Zhu, J.; Germain, A. R.; Porco, J. A. JJr., Angew. Chem. Int. Ed. 2004, 43, 1239–1243. 55. Godet, T.; Vaxelaire, C.; Michel, C.; Milet, A.; Belmont, P., Chem. Eur. J. 2007, 13, 5632–5641. 56. Barluenga, J.; Vazquez-Villa, H.; Ballesteros, A.; Gonzalez, J. M., J. Am. Chem. Soc. 2003, 125, 9028–9029. 57. Yue, D.; Della Ca, N.; Larock, R. C., Org. Lett. 2004, 6, 1581–1584. 58. Jong, T. T.; Leu, S. J., J. Chem. Soc. Perkin Trans. 1 1990, 423–424. 59. Genin, E.; Toullec, P. Y.; Antoniotti, S.; Brancour, C.; Genet, J. -P.; Michelet, V., J. Am. Chem. Soc. 2006, 128, 3112–3113. 60. Gulias, M.; Rodriguez, J. R.; Castedo, L.; Mascarenas, J. L., Org. Lett. 2003, 5, 1975–1977. 61. Alonso, F.; Beletskaya, I. P.; Yus, M., Chem. Rev. 2004, 104, 3079–3159. 62. Patil, N. T.; Yamamoto, Y., J. Org. Chem. 2004, 69, 5139–5142. 63. Asao, N.; Aikawa, H.; Yamamoto, Y., J. Am. Chem. Soc. 2004, 126, 7458–7459. 64. Asao, N.; Takahashi, K.; Lee, S.; Kasahara, T.; Yamamoto, Y., J. Am. Chem. Soc. 2002, 124, 12650–12651. 65. Asao, N.; Nogami, T.; Takahashi, K.; Yamamoto, Y., J. Am. Chem. Soc. 2002, 124, 764–765. 66. Yu, X.; Ding, Q.; Wang, W.; Wu, J., Tetrahedron Lett. 2008, 49, 4390–4393. 67. Verniest, G.; Padwa, A., Org. Lett. 2008, 10, 4379–4382. 68. Patil, N. T.; Pahadi, N. K.; Yamamoto, Y., J. Org. Chem. 2005, 70, 10096–10098. 69. Lathbury, D. C.; Shaw, R. W.; Bates, P. A.; Hursthouse, M. B.; Gallagher, T., J. Chem. Soc. Perkin Trans. 1 1989, 2415–2424. 70. Lathbury, D.; Gallagher, T., Tetrahedron Lett. 1985, 26, 6249–6252. 71. Shaw, R.; Lathbury, D.; Anderson, M.; Gallagher, T., J. Chem. Soc. Perkin Trans. 1, 1991, 659–660. 72. Lathbury, D.; Gallagher, T., J. Chem. Soc. Chem. Commun. 1986, 1017–1018. 73. Huang, Q.; Hunter, J. A.; Larock, R. C., J. Org. Chem. 2002, 67, 3437–3444. 74. Su, S.; Porco, J. A., Jr., J. Am. Chem. Soc. 2007, 129, 7744–7745. 75. Su, S.; Porco, J. A., Jr., Org. Lett. 2007, 9, 4983–4986. 76. Asao, N.; Yudha, S. S.; Nogami, T.; Yamamoto, Y., Angew. Chem. Int. Ed. 2005, 44, 5526–5528. 77. Sun, W.; Ding, Q.; Sun, X.; Fan, R.; Wu, J., J. Comb. Chem. 2007, 9, 690–694.
REFERENCES
78. 79. 80. 81. 82. 83. 84. 85. 86. 87. 88. 89. 90. 91. 92. 93.
165
Ding, Q.; Wu, J., Org. Lett. 2007, 9, 4959–4962. Ding, Q.; Yu, X.; Wu, J., Tetrahedron Lett. 2008, 49, 2752–2755. Yeom, H. -S.; Kim, S.; Shin, S., Synlett 2008, 924–928. Ding, Q.; Wang, Z.; Wu, J., Tetrahedron Lett. 2009, 50, 198–200. Seregin, I. V.; Schammel, A. W.; Gevorgyan, V., Org. Lett. 2007, 9, 3433–3436. Seregin, I. V.; Schammel, A. W.; Gevorgyan, V., Tetrahedron 2008, 64, 6876–6883. Harrison, T. J.; Kozak, J. A.; Corbella-Pane, M.; Dake, G. R., J. Org. Chem. 2006, 71, 4525–4529. Harrison, T. J.; Dake, G. R., Org. Lett. 2004, 6, 5023–5026. Imase, H.; Noguchi, K.; Hirano, M.; Tanaka, K., Org. Lett. 2008, 10, 3563–3566. Godet, T.; Belmont, P., Synlett 2008, 2513–2517. Porcel, S.; Echavarren, A. M., Angew. Chem., Int. Ed. 2007, 46, 2672–2676. Lin, G. -Y.; Li, C. -W.; Hung, S. -H.; Liu, R. -S., Org. Lett. 2008, 10, 5059–5062. Lo, C. -Y.; Lin, C. -C.; Cheng, H. -M.; Liu, R. -S., Org. Lett. 2006, 8, 3153–3156. Ding, Q.; Wu, J., J. Comb. Chem. 2008, 10, 541–545. Binder, J. T.; Kirsch, S. F., Org. Lett. 2006, 8, 2151–2153. Menz, H.; Kirsch, S. F., Org. Lett. 2006, 8, 4795–4797.
6 SILVER-CATALYZED NITRENE TRANSFER REACTIONS ZIGANG LI, DAVID A. CAPRETTO,
AND
CHUAN HE
Department of Chemistry, University of Chicago, Chicago, Illinois
6.1 Introduction 6.2 Aziridination 6.2.1 Chloramine-T as Nitrene Precursor 6.2.2 Iminoiodanes as Nitrene Precursors 6.2.3 Heterogeneous Silver Catalysis 6.3 Sulfide and Sulfoxide Imination 6.4 Amidation 6.4.1 Intramolecular Amidation 6.4.2 Intermolecular Amination with Phenanthroline Ligands 6.4.3 Intermolecular Amination Based on Pyrazolylborate Ligands 6.5 Conclusion References
6.1
INTRODUCTION
Nitrogen is a prevalent element in naturally occurring molecules, and chemical formation of new carbon–nitrogen bonds has broad applications in both industry and academic research. Despite more recent advances in carbon–nitrogen bond formation, new facile methods are still sought after. Great advances have been made since the late 1990s, particularly in the development of the efficient hydroamination of
Silver in Organic Chemistry Edited by Michael Harmata Copyright Ó 2010 John Wiley & Sons, Inc.
167
168
SILVER-CATALYZED NITRENE TRANSFER REACTIONS
Scheme 6.1. Buchwald–Hartwig coupling chemistry for C--N bond formation.
unsaturated carbon–carbon bonds and coupling reactions between amines (even ammonia) and haloarenes (Scheme 6.1).1 The nitrene transfer reaction is another promising method for installing new C--N bonds.2 The most commonly used nitrene precursors are a class of molecules called iminoiodanes (such as PhI¼NNs, where Ns ¼ 4-nitrobenzenesulfonyl), which are typically stabilized with electron-withdrawing groups. Iminoiodanes are relatively stable and can be easily synthesized and stored for reasonable periods of time at lower temperatures; reactions using iminoiodanes are safe in general.3 Other commonly used nitrene precursors include organic azides and chloramines such as TsN NaCl. However, the preparation of azides and anhydrous chloramines carries a risk of explosion.4 More reactive nitrene precursors, such as TsN NaBr and PhBr¼NNs, are also being increasingly used.5 The aziridination of olefins, which forms a three-membered nitrogen heterocycle, is one important nitrene transfer reaction. Aziridination shows an advantage over the more classic olefin hydroamination reaction in some syntheses because the threemembered ring that is formed can be further modified. More recently, intramolecular amidation and intermolecular amination of C--H bonds into new C--N bonds has been developed with various metal catalysts. When compared with conventional substitution or nucleophilic addition routes, the direct formation of C--N bonds from C--H bonds reduces the number of synthetic steps and improves overall efficiency.2 After early work on iron, manganese, and copper,6 Muller, Dauban, Dodd, Du Bois, and others developed different dirhodium carboxylate catalyst systems that catalyze C--N bond formation starting from nitrene precursors,7 while Che studied a ruthenium porphyrin catalyst system extensively.8 The rhodium and ruthenium systems are
Figure 6.1. Rhodium–carboxylate catalysts for C--N bond formation.
AZIRIDINATION
169
currently the most intensely studied, while copper,9 nickel,10 iron,11 cobalt,12 and gold catalysts13 also show nitrene transfer activity (Fig. 6.1). In addition to rhodium and ruthenium, silver catalysts have also been investigated, and this chapter discusses these silver-based nitrene transfer reactions.2a In discussing the work chronologically, we hope that the readers can get an idea of the evolution of thinking in the research process. 6.2 6.2.1
AZIRIDINATION Chloramine-T as Nitrene Precursor
As stated in the introduction, chloramine-T (where T denotes three crystalline water molecules) is a commonly used nitrene precursor, which is commercially available and costs less than do most other nitrene sources. The benefit of a silver salt in nitrene transfer reactions with chloramine-T is surprisingly simple. Because silver chloride is insoluble in most solvents, substoichiometric amounts of silver salts (like silver nitrate) can be used to remove the chloride from chloramine to facilitate the release of a free nitrene radical, which can aziridinate olefins. Since the amount of silver is near stoichiometric, it should not be called silver-based catalysis, although turnover numbers (TONs) higher than 1 have been observed in some cases. In 2001, Rai first reported such a transformation.14 The work specifically noted that the reaction favors aprotic solvents such as THF. This statement was challenged by Komatsu shortly thereafter, who found no reactivity in THF but high conversions in CH2Cl2 or benzene for the same reaction.15 Solvent grade or quality may play an important role; less likely but still possible, the grade or quality of silver nitrate used may also be a factor. In Komatsu’s report, a stoichiometric amount of AgNO3 and chloramine-T produced 43–92% yields of the aziridine products from different olefin substrates. An advantage for this reaction is that the chloramine-T could be directly used without any dehydration. An important mechanistic study using 1,6-dienes was also carried out, which yielded bicyclic pyrrolidines. The formation of bicyclic pyrrolidines together with the reaction sensitivity toward oxygen, a good radical inhibitor, support a free nitrene radical mechanism (Scheme 6.2). The formation of AgCl precipitate and subsequent generation of a nitrene radical intermediate is well accepted; however, whether the metal salt plays any other role is not clear. 6.2.2
Iminoiodanes as Nitrene Precursors
Silver has three synthetically useful oxidation states: Ag(I), Ag(II), and Ag(III).16 Free Ag(III) is very unstable, and most currently known Ag(III) complexes are stablized with electron-donating and/or sterically demanding ligands.17 It is known that Ag(I) can be oxidized to Ag(II) with strong oxidants such as persulfates. Nitrogen-based ligands such as pyridines are commonly used to stabilize highvalence metal ions.18 In 2003, He and coworkers utilized a pyridine-supported silver catalyst and reported the first silver-catalyzed aziridination of olefins.19
170
SILVER-CATALYZED NITRENE TRANSFER REACTIONS
Scheme 6.2. Olefin aziridination using chloramine-T as a nitrene source.
With the tridentate 4,40 ,400 -tri-tert-butyl-2,20 :60 ,200 -terpyridine (tBu3tpy) ligand, a unique disilver complex crystallized with high yield in acetonitrile. NMR studies showed that the complex forms within 5 min when a silver salt and ligand are mixed together in CD3CN at ambient temperature. The catalyst structure was fully characterized to show a unique dinuclear silver(I) [Ag2(tBu3tpy)2(NO3)](NO3) (Fig. 6.2). The Ag--Ag distance is 2.842(2) A, which indicates a fairly strong silver(I)–silver(I) interaction. Both silvers are five-coordinate if the silver-silver interaction is included. The aziridination catalysis was carried out in situ with mixed ligand and silver salt, which gives results similar to those obtained using the pure [Ag2(tBu3tpy)2(NO3)] (NO3) crystal. The tBu3tpy ligand shows superior reactivity over other pyridine
Figure 6.2. Structure of disilver nitrene transfer catalyst.
171
AZIRIDINATION
ligands, while the choice of silver salt does not play an important role as long as the salt is soluble in the solvent used. Interestingly, this disilver(I) complex has some structural features similar to those of di-Rh(II) compounds, which were previously used to catalyze a similar reaction. Whether the structural similarity implies the same reaction mechanism with rhodium or other systems is presently unclear. One aspect of this research hampering the elucidation of mechanism in all these systems is the insolubility of iminoiodanes in common organic solvents. The aziridination works for both aromatic and aliphatic olefins, including less active linear terminal olefins. Most reactions proceed in good yield at room temperature. The use of cis-stilbene at 0 C gives predominately cis aziridine product in about 90 : 10 cis : trans ratio (Table 6.1). The conservation of cis structure suggests that a discrete silver nitrene intermediate is involved in the reaction path. Because of the unique disilver structure and unlikely formation of a silver(III) species, the authors suspect that a bridged nitrene intermediate between the two silver atoms may be responsible for this transformation in which each silver atom donates one electron to the nitrenoid. However, further research is necessary to prove this hypothesis and a fast radical reaction mechanism cannot be eliminated on the basis of current evidence.
TABLE 6.1. Substrate Scope for Olefin Aziridination by Silver–Terpyridine Catalyst
Olefin +
PhI=NTs
Olefin
5 mol% [Ag2(t Bu3tpy) 3(NO3)](NO3) CH3CN, r.t.
Aziridine
Isolated Yield (%)
NTs
91
NTs
74
Cl Cl NTs
NTs
Ph
Ts N
Ph Ph
a
Reaction performed at 0 C.
71 81
86 a Ph
172
SILVER-CATALYZED NITRENE TRANSFER REACTIONS
Scheme 6.3. Silver–terpyridine catalyst can catalyze olefin aziridination using an in situ prepared nitrene.
The reaction was further investigated with electrospray ionization–mass spectrometry (ESI-MS). Two different iminoiodanes were used, PhI ¼ NTs and PhI ¼ NSO2Ph, and the results suggest an intermediate containing a disilver core plus a nitrene moiety (¼NTs or ¼NSO2Ph). The masses corresponded to these species, and the 14 D weight difference between the two nitrene sources further proved the involvement of the nitrene moeity. This catalyst can also mediate olefin epoxidation, although at very low efficiencies. Although the synthesis of PhI¼NTs is relatively easy, it is more attractive to directly use the sulfonamide precursor and some inexpensive oxidant to generate the nitrene source in situ. Rhodium- or copper-based catalysts are able to perform this one-pot reaction, and silver was able to as well. [Ag2(tBu3tpy)2(NO3)](NO3) catalyzed the direct aziridination of styrenes using TsNH2 and PhI(OAc)2. Unfortunately, aliphatic olefins could only be converted using separately prepared iminoiodanes (Scheme 6.3). 6.2.3
Heterogeneous Silver Catalysis
This chapter focuses mostly on homogenous silver catalysis, so only limited space is spent on discussion of heterogeneous silver catalysis. However, heterogenous systems are important. Heterogeneous silver-based catalysts are well known in industrial ethylene epoxidation and methanol/formaldehyde oxidation.20 Friend et al. showed that the silver(110) single-crystal surface can mediate the oxidative addition of ammonia to styrene to form phenylaziridines. They used oxygen as the oxidant, which shows the promising potential of heterogeneous silver catalysts for olefin functionalization (Scheme 6.4).21
6.3
SULFIDE AND SULFOXIDE IMINATION
Sulfilimines and sulfoximines are useful synthetic building blocks for chiral ligands and pseudopeptides. Compared to other preparation methods, the direct imination of corresponding sulfides and sulfoxides is a more convenient and straightforward synthetic route. Different metal complexes have shown the ability to catalyze these imination reactions, such as rhodium, copper, and iron.22 In 2005, Bolm found that the disilver(I) complex described in Section 6.2.2 catalyzes the imination of sulfides and sulfoxides
173
AMIDATION
Scheme 6.4. Heterogenous silver can facilitate C--N bond formation.
R
O S
R
+
RNH2
+
PhI(OAc)2
4–8 mol% catalyst from Scheme 6.4
O
CH3CN, r.t. or reflux, 16 h
R
NR S R
7 different sulfoxides 4 different amines 79–98% yield
Scheme 6.5. Sulfoxide imination using silver–terpyridine catalyst.
with good to excellent yields.23 Simple sulfonamides can be used directly with commercially available oxidants such as PhI(OAc)2 in acetonitrile at room temperature. Sulfoxides can give sulfoximine products, and sulfides can afford sulfilimines, while double iminations were not detected (Scheme 6.5). In addition, the imination of an enantioenriched sulfoxide yielded a sulfoximine (after deprotection) with the same enantiomeric excess value. This finding also suggests a conserved nitrene transfer process and that the disilver(I) 1 system possibly could be utilized in other catalytic nitrene transfer chemistry.
6.4
AMIDATION
When compared to aziridination or sulfoxide imination, efficient amidation of C--H bonds is a more formidable task since it involves the activation of relatively inert C--H bonds. With an increase in the difficulty of the reaction comes an increase in the value of products, as many are derived from cheap simple hydrocarbons. In addition to making more valuable materials in fewer steps, a better mechanistic understanding of nitrene transfer can also be gleaned, leading to the development of efficient catalysts. 6.4.1
Intramolecular Amidation
Intramolecular amidation provides a useful way to directly install a new C--N bond at a designated position and is a powerful method for introducing new synthons.
174
SILVER-CATALYZED NITRENE TRANSFER REACTIONS
Generally, hydrocarbons tethered with sulfamate esters or carbamates are suitable substrates and afford five- or six-membered ring products through intramolecular C--H amidation. For reasons not yet understood, carbamates and sulfamate esters prefer to form five- and six-membered rings, respectively. In general, this important and useful selectivity can avoid formation of product mixtures, a major drawback for similar organic transformations. After successful application of the silver catalyst shown in olefin aziridination (Section 6.1.1), He and coworkers showed that intramolecular amidation was possible with both hydrocarbon-tethered carbamates and sulfamate esters.24 They found that only the tBu3tpy silver complex could catalyze efficient intramolecular amidation, while other pyridine ligands gave either dramatically lower yields or complicated product mixtures. In an interesting control study, both copper and gold were also tested in this reaction. Both the copper and gold tBu3tpy complexes can mediate olefin aziridination, but only silver can catalyze intramolecular C--H amidation, indicating that the silver catalyst forms a more reactive metal nitrene intermediate. Both tertiary and secondary aliphatic C--H bonds can undergo this transformation to give products in good to excellent yields. In some cases, 4-tBu-pyridine was added to both increase the conversion and afford a cleaner reaction. Notably, no mixed fiveand six-membered ring products were observed for any substrates (Table 6.2). To gain mechanistic perspective, the amidation reaction was performed with a carbamate substrate derived from (S)-2-methyl-1-butanol. The resulting product completely retained its (S) configuration, which suggests a conserved nitrene insertion process. However, a short-lived radical process cannot be ruled out. Unfortunately, the t Bu3 tpy–silver complex could not catalyze intermolecular C--H amination. This lack of activity led to the search for and discovery of a more reactive silver catalyst that can perform intermolecular C--H amination. 6.4.2
Intermolecular Amination with Phenanthroline Ligands
Intermolecular amination is highly sought after and is the most straightforward way to install a new nitrogen-based functional group. However, because of the difficulty of this reaction, published explorations into this field were almost nonexistent in the past. More recently, the field has grown quickly as several different groups independently reported their breakthroughs on the topic. Several examples of intermolecular C--H amination of simple hydrocarbons with copper, rhodium, gold, and silver systems have been reported.7–9,13,25 The He group continued to search for more active silver-based nitrene transfer catalysts. Considering the reactivity, metal nitrene intermediates are generally electrophilic and less donating ligands may enhance the electrophilicity of the intermediate. With this in mind, various pyridine-type ligand systems were tested. In agreement with their hypothesis, the fused-pyridine bidentate ligand bathophenanthroline [4,7-diphenyl-1,10-phenanthroline (BP)] showed the best reactivity. Regular phenanthroline showed good reactivity, but solubility was poor, making the diphenyl version more acceptable. This catalyst allowed intramolecular amidation to
175
AMIDATION
TABLE 6.2. Intramolecular C–H Amidation by Silver-Terpyridine Catalyst 5 mol% [Ag2(tBu3tpy)3(NO3)](NO3)
Substrate + Phl(OAc) 2
Substrate
CH3CN, 82ºC
Product O
O
NH2 O
Isolated Yield (%)
HN O
81
O NH2 O
85
O HN O H N
O O
NH2
O O
48
O
73
O H N
NH2
O
O
proceed at 50 C and improved yields significantly when compared to the terpyridine system using the same substrates (see Section 6.4.1). More importantly, this silver system catalyzes the intermolecular amination of hydrocarbons, as shown in Table 6.3. In addition to aminating weaker benzylic C--H bonds, stronger aliphatic C--H bonds such as those in cyclohexane were also reactive. Although yields with more inert hydrocarbons were modest with the bathophenanthroline system, the discovery of the first silver-catalyzed intermolecular amination opens opportunities for further developments. This reaction favored tertiary cyclic sp3 C--H bonds over secondary cyclic sp3 C--H bonds, and showed limited success with simple linear alkanes. No conversion was observed with any aromatic C--H bonds. The compound NsNH2 was tested as the nitrene precursor with different oxidants. The use of PhI(OAc)2 as oxidant gave the expected amination product with a lower yield, while persulfate and peroxides showed no reactivity. Similar to that of the t Bu 3 tpy–silver catalyst, the crystal structure of the bathophenanthroline–silver complex has a disilver(I) core and two BP ligands stacked over each other, similar to the corresponding palladium(II)–BP complex.
176
SILVER-CATALYZED NITRENE TRANSFER REACTIONS
TABLE 6.3. Silver–Bathophenanthroline Catalysis of a Mild Intermolecular C--H Amination 5 mol% Ag2(BP)2OTf
Substrate + Phl=NNs
CH2Cl2, 50ºC, 4 Å MS
Substrate
Product
Isolated Yield (%)
NHNs
70
NHNs
57
NHNs NHNs
NHNs
25
40
35
NHNs
33
NHNs
39
The silver–silver distance is 3.386 A, indicating a weak silver–silver interaction (Fig. 6.3). Observing the disilver structure for the second time, the authors studied the importance of the disilver structure on reactivity. By testing different phenanthrolines, they found that 2,9-substituted phenanthrolines did not catalyze the amination reaction. This result was supported by structural evidence—the silver-2,9-dimethylphenanthroline complex did not form disilver structure; two ligands coordinated one silver in a tetrahedral geometry, blocking the approach of any other molecule to the silver center. Even one bulky mesityl group in the 2 position inhibits the reactivity completely (Scheme 6.6).
AMIDATION
177
Figure 6.3. X-ray crystallographic structure of silver–bathophenanthroline catalyst system: Ag1A is coordinated with two triflate molecules, while Ag2A is coordinated with a water molecule.
The development of new phenanthroline ligands with electron-withdrawing modifications on the 4 and 7 positions may be a way to enhance the reactivity further. 6.4.3
Intermolecular Amination Based on Pyrazolylborate Ligands
Shortly after the discovery by He and coworkers, Dıaz-Requejo and Perez et al. reported an intriguing silver–pyrazolylborate system that catalyzes the amination of aliphatic hydrocarbons.26 Earlier, they had developed a similar copper–pyrazolylborate system that could also catalyze the C--H amination reaction. However, a different ligand was required with the silver catalyst system.
Scheme 6.6. Ligand effects on intermolecular C--H amination catalysis.
178
SILVER-CATALYZED NITRENE TRANSFER REACTIONS
Figure 6.4. Pyrazolylborate ligands used for silver-based intermolecular C--H amination.
They chose four different pyrazolylborate ligands to test ligand effects on the catalyst, as shown in Figure 6.4, and found TpBr3 to be the ligand with the most electron-withdrawing groups and TpMs the one with the most steric hindrance, while Tp ,Br and Tp were intermediate in electronics and sterics. In the silver case, the Tp ,Br ligand generally gave the highest yield, while TpBr3 was the least efficient. Interestingly, they found that when simple linear alkanes were used, the amination products from primary sp3 C--H bond and secondary sp3 C--H bond activation were detected. The activation of a primary sp3 C--H bonds is unique and rarely detected with other catalysts. The results are shown in Table 6.4. The AgTpMs complex gave the highest ratio of primary sp3 C--H bond activation, possibly due to the sterically demanding mesityl group, which should generate the smallest catalytic pocket and thus prefer smaller, primary CH3 groups. The conversion of simple hydrocarbons into corresponding amine products is significantly higher than in previous reports. The authors also noted that NMR experiments showed that catalyst peaks reappeared after the completion of the reaction, demonstrating the robustness of the catalyst. A high TON may be achieved with this silver–pyrazolylborate system with further reaction optimization. When 2 equiv (relative to catalyst) of 2,6-di-tert-butylhydroxytoluene (BHT, a radical reaction inhibitor) was added to the reaction, the yield of the reaction decreased by half. A radical trap experiment using cyclohexane was carried out in CCl4 as well, and together with 40% amination product, 12% of chlorocyclohexane was also detected. This result also suggests that a radical pathway is operative. When a 1 : 2.3 mixture of cis- and trans-2 pentene was reacted with PhI¼NTs (Scheme 6.7), a 1 : 1.5 cis : trans aziridine product was isolated, which suggests the mechanism may
179
AMIDATION
TABLE 6.4. Substrate Scope for Silver–Pyrazolylborate-Catalyzed Aliphatic C--H Amination R–H + PhI=NTs
Substrate
5 mol% AgTp*,Br no solvent, 80ºC, 16 h
R–NHTs
Isolated Yield (%)a
Product(s) 31 59 10
35
60
65
5
70
NHTs 90
75
10
80
NHTs
NHTs
a
80
90
Yield indicates combined yield. Arrows refer to percent product at the denoted position.
be part stepwise radical process and part discrete nitrene transfer. Generally, transaziridines are more stable and these cis : trans ratios were cited as given from the original report. This observation fits with results previously described by Che et al. with the porphyrin ruthenium nitrenoid. Even when isolated ruthenium nitrenoid was used,
Scheme 6.7. Mechanistic probes of catalytic nitrene transfer.
180
SILVER-CATALYZED NITRENE TRANSFER REACTIONS
Scheme 6.8. Proposed mechanism for silver–pyrazolylborate-catalyzed nitrene transfer.
trans-aziridine products were still detected from cis-olefin substrates, and sometimes as the predominant product. Current results on silver-catalyzed nitrene transfer reactions, indicate that silver probably can interact with iminoiodanes to generate a silver nitrene precursor. This precursor can lead to reactions via either a concerted metal nitrene or a stepwise radical pathway, depending on the substrate and reaction conditions (Scheme 6.8).
6.5
CONCLUSION
Since the original report on aziridination in 2003, silver-catalyzed nitrene transfer reactions have been developed into both intramolecular C--H amidation and intermolecular C--H amination. Currently, three promising ligand groups have been successful, each showing unique reactivity. All are worth further investigation, but new ligand groups will undoubtedly be discovered. Current knowledge indicates nitrogen-based or sterically demanding ligands to be most promising. How to convert these systems into synthetically meaningful enantioselective systems is a challenging task. With the relatively low price of silver salts, the stability of silver complexes, and high yields observed with current systems, the future for silver-catalyzed nitrene transfer is bright.
REFERENCES 1. For reviews, see (a) Hartwig, J. F., Nature 2008, 455, 314–322; (b) Hartwig, J. F., Acc. Chem. Res. 2008, 41, 1534–1544; (c) Hartwig, J. F., Angew. Chem. Int. Ed. 1998, 37, 2046–2067; (d) Surry, D. S.; Buchwald, S. L., Angew. Chem. Int. Ed. 2008, 47, 6338–6361;
REFERENCES
2.
3.
4.
5.
6.
7.
8.
9.
10. 11. 12. 13. 14. 15. 16. 17.
181
(e) Wolfe, J. P.; Wagaw, S.; Marcoux, J.-F.; Buchwald, S. L., Acc. Chem. Res. 1998, 31, 805–818; (f) Correa, A.; Bolm, C., Adv. Synth. Catal. 2008, 350, 391–394. For reviews, see (a) Li, Z.; He, C., Eur. J. Org. Chem. 2006, 4313–4322; (b) Watson, I. D. G.; Yu, L.; Yudin, A. K., Acc. Chem. Res. 2006, 39, 194–206; (c) Du Bois, J., Chemtracts 2005, 18, 1–13; (d) Dauban, P.; Dodd, R. H., Synlett 2003, 1571–1586; (e) Mueller, P.; Fruit, C., Chem. Rev. 2003, 103, 2905–2919; (f) Hilt, G., Angew. Chem. Int. Ed. 2002, 41, 3586–3588. (a) Evans, D. A.; Faul, M. M.; Bilodeau, M. T., J. Org. Chem. 1991, 56, 6744–6746; (b) Heuss, B. D.; Mayer, M. F.; Dennis, S.; Hossain, M. M., Inorg. Chim. Acta 2003, 342, 301–304; (c) Taylor, S.; Gullick, J.; McMorn, P.; Bethell, D.; Bulman, P.; Philip, C.; Hancock, F. E.; King, F.; Hutchings, G. J., Top. Catal. 2003, 24, 43–50. (a) Ando, T.; Kano, D.; Minakata, S.; Ryu, I.; Komatsu, M., Tetrahedron 1998, 54, 13485–13494; (b) Vyas, R.; Chanda, B. M.; Bedekar, A. V., Tetrahedron Lett. 1998, 39, 4715–4716; (c) Ando, T.; Minakata, S.; Ryu, I.; Komatsu, M., Tetrahedron Lett. 1998, 39, 309–312. (a) Vyas, R.; Gao, G.-Y.; Harden, J. D.; Zhang, P. X., Org. Lett. 2004, 6, 1907; (b) Vyas, R.; Chanda, B. M.; Bedekar, A. V., Tetrahedron Lett. 1998, 39, 4715–4716; (c) Vyas, R.; Gao, G.-Y.; Harden, J. D.; Zhang, P. X., Org. Lett. 2004, 6, 1907–1910; (d) Ochiai, M.; Kaneaki, T.; Tada, N.; Miyamoto, K.; Chuman, H.; Shiro, M.; Hayashi, S.; Nakanishi, W., J. Am.Chem. Soc. 2007, 129, 12938–12939. (a) Breslow, R.; Gellman, S. H., J. Chem. Soc. Chem. Commun. 1982, 24, 1400–1401; (b) Breslow, R.; Gellman, S. H., J. Am. Chem. Soc. 1983, 105, 6728–6729; (c) Li, Z.; Quan, R. W.; Jacobsen, E. N., J. Am. Chem. Soc. 1995, 117, 5889–5890. (a) Guthikonda, K.; Du Bois, J., J. Am. Chem. Soc. 2002, 124, 13672–13673; (b) Dauban, P.; Saniere, L.; Tarrade, A.; Dodd, R. H., J. Am. Chem. Soc. 2001, 123, 7707–7708; (c) Espino, C. G.; Du Bois, J., J. Am. Chem. Soc. 2002, 124, 12950–12951; (d) Espino, C. G.; Fiori, K. W.; Kim, M. J.; Du Bois, J., J. Am. Chem. Soc. 2004, 126, 15378–15379; (e) Liang, C.; Collet, F.; Robert-Peillard, F.; M€uller, P.; Dodd, R. H.; Dauban, P., J. Am. Chem. Soc. 2008, 130, 343–344; (f) Fiori, K. W.; Du Bois, J., J. Am. Chem. Soc. 2007, 129, 562–568. (a) Liang, J. -L.; Yuan, S. -X.; Huang, J. -S.; Yu, W. -Y.; Che, C. -M., Angew. Chem. Int. Ed. 2002, 41, 3465–3468; (b) Liang, J. -L.; Huang, J. -S.; Yu, X. Q.; Zhu, N.; Che, C. -M., Chem. Eur. J. 2002, 8, 1563–1572. (a) Dıaz-Requejo, M. M.; Belderraıen, T. R.; Nicasio, M. C.; Trofimenko, S.; Perez, P. J., J. Am. Chem. Soc. 2003, 125, 12078–12079; (b) Fructos, M. R.; Trofimenko, S.; DıazRequejo, M. M.; Perez, P. J., J. Am. Chem. Soc. 2006, 128, 11784–11791. Mindiola, D. J.; Hillhouse, G. L., Chem. Commun. 2002, 17, 1840–1841. Mahy, J. P.; Battioni, P.; Mansuy, D., J. Am. Chem. Soc. 1986, 108, 1079–1080. Gao, G. -Y.; Harden, J. D.; Zhang, X. P., Org. Lett. 2005, 7, 3191–3193. (a) Li, Z.; Ding, X.; He, C., J. Org. Chem. 2006, 71, 5876–5880; (b) Li, Z.; Capretto, D. A.; Rahaman, R. O.; He, C., J. Am. Chem. Soc. 2007, 129, 12058–12059. Kumar, K. A.; Rai, L. K. M.; Umesha, K. B., Tetrahedron 2001, 57, 6993–6996. Minakaa, S.; Kano, D.; Fukuoka, R.; Oderaotoshi, Y.; Komatsu, M., Heterocycles 2003, 60, 289–298. Greenwood, N. N.; Earnshaw, A.; Chemistry of the Elements Elsevier Oxford 1997. Furuta, H.; Maeda, H.; Osuka, A., J. Am. Chem. Soc. 2000, 122, 803–804.
182
SILVER-CATALYZED NITRENE TRANSFER REACTIONS
18. For reviews, see (a) Ishii, T.; Aizawa, N.; Kanehama, R.; Yamashita, M.; Sugiura, K. -I.; Miyasaka, H., Coord. Chem. Rev. 2002, 226, 113–124; (b) Srochinski, D.; Dziegiec, Y.; Grzejdziak, A., Russ. J. Coord. Chem. 1997, 23, 447–460; (c) Levason, W.; Spicer, M. D., Coord. Chem. Rev. 1987, 76, 45–120. 19. Cui, Y.; He, C., J. Am. Chem. Soc. 2003, 125, 16202–16203. 20. Anonymous, Chem. Eng. News 2006, 84, 59–62. 21. Liu, X.; Madix, R. J.; Friend, C. M., J. Am. Chem. Soc. 2006, 128, 14266–14267. 22. Mancheno, O. G.; Bolm, C., Chem. Eur. J. 2007, 13, 6674–6681. 23. Cho, G. Y.; Bolm, C., Org. Lett. 2005, 7, 4983–4985. 24. Cui, Y.; He, C., Angew. Chem. Int. Ed. 2004, 43, 4210–4212. 25. Li, Z.; Capretto, D. A.; Rahaman, R. H.; He, C., Angew. Chem. Int. Ed. 2007, 46, 5184–5186. 26. Gomez-Emeterio, B. P.; Urbano, J.; Diaz-Requejo, M. M.; Perez, P. J., Organometallics 2008, 27, 4126–4130.
7 SILVER-CATALYZED SILYLENE TRANSFER TOM G. DRIVER Department of Chemistry, University of Illinois, Chicago, Illinois
7.1 Introduction 7.2 Reactivity and Attributes of Metal Silylenoids and Silylmetal Complexes 7.2.1 Synthesis of Transition Metal Complexes of Silylenes 7.2.2 Reactivity of Transition Metal Silylenoids 7.2.3 Transition Metal Silylenoid Complex–Catalyzed Hydrosilation Reactions 7.2.4 Transition Metal Silylenoid–Catalyzed Atom Transfer Reactions 7.3 Silacyclopropanes as Important Synthetic Intermediates 7.4 Silver-Mediated Transfer of Di-tert-Butylsilylene to Olefins 7.5 Silver-Mediated Transfer of Di-tert-Butylsilylene to Acetylenes 7.6 Silver-Mediated Transfer of Di-tert-Butylsilylene to Carbonyl Compounds 7.7 Silver-Mediated Transfer of Di-tert-Butylsilylene to Imines 7.8 Silver-Mediated Di-tert-Butylsilylene Insertion Into C--O Bonds 7.9 Conclusion References
7.1
INTRODUCTION
Silver salts are well established to promote atom transfer reactions.1–5 In 1946 Bachmann and coworkers reported that silver oxide facilitated the Wolff rearrangement6 of a-diazocarbonyl compounds.7 After this initial report, several other silver(I) reagents (including AgNO3 and AgO2CPh)8,9 were identified to provide higher yields, Silver in Organic Chemistry Edited by Michael Harmata Copyright 2010 John Wiley & Sons, Inc.
183
184
SILVER-CATALYZED SILYLENE TRANSFER
greater reproducibility, and lower temperatures for formation of the rearrangement product.10 In contrast to the decades of research interest in silver-mediated carbene transfer reactions, the use of silver salts to promote reactions involving silylenes remained undiscovered until 2002.11 Ever since silylenes were proposed as reactive intermediates in 1964, substantial research effort has been directed toward understanding and harnessing their reactivity.12 These divalent silicon reactive intermediates can be generated thermally or photochemically from a number of different sources and participate in a range of atom transfer reactions, including insertion and addition reactions. In addition to interest surrounding the chemistry involving free silylenes, the involvement of transition metal silylenoids in important industrial processes continues to inspire research efforts into their synthesis and reactivity.13 The development of synthetic processes that exploit the reactivity of silylene and metal silylenoid reactive intermediates resulted from these research programs. This chapter examines the synthesis and reactivity of transition metal silylenoids and silylmetal complexes to provide context for the focus of this chapter: silvermediated silylene transfer reactions.
7.2 REACTIVITY AND ATTRIBUTES OF METAL SILYLENOIDS AND SILYLMETAL COMPLEXES 7.2.1
Synthesis of Transition Metal Complexes of Silylenes13,14
The proposal of metal silylene complexes as reactive intermediates in the mechanism of industrial processes such as hydrosilylation15 and dehydrogenative coupling of silanes to polysilanes16 stimulated interest in the synthesis of these subvalent silicon compounds.13,14 For example, certain hydrosilylation catalysts such as (Ph3P)2Pt (C2H4) and (Ph3P)3RhCl were found to be silane-specific; while Et2SiH2 was an acceptable substrate, Et3SiH was not. This specificity was attributed to a metal silylenoid intermediate [Et2Si¼M].17 Similarly, the polymerization of PhSiH3 by Cp2TiMe2 was proposed to occur via a titanium silylenoid intermediate [Cp2Ti¼SiHR] that formed whenever another silyl group was added to the growing polymer.16 Isolable early transition metal silylenoids could be synthesized by photolysis of silylmetal complexes.18–30 Photolysis of iron disilyl complex 1 triggered a 1,2-silyl shift to produce a complex 2 as a 2 : 1 mixture of diastereomers (Scheme 7.1).19 Ogino and coworkers interpreted the nearly planar relationship of the atoms at each silicon and the short Fe--Si bonds (2.207 and 2.222 A) as evidence of bis(silylene) character. They concluded that complex 2 is best described as the combination of the two mesomers 3 and 4. Light-induced silicon–silicon bond cleavage was also used for the synthesis of tungsten silylene complexes;31–33 the exposure of tungsten complex 5 to ultraviolet radiation leads to the loss of CO and formation of tungsten silylene complex 7.33 When the starting tungsten complex bore a cyclopentadiene ligand or when pentamethyldisilane was used, a base-stabilized (THF) tungsten complex was formed.32 Late transition metal silylenoid complexes were first synthesized by exploiting the lability of silicon–triflate bonds (Scheme 7.2).34–37 Tilley and coworkers reported that
185
REACTIVITY AND ATTRIBUTES OF METAL SILYLENOIDS MeO Me Me
Si
Me Fe Si OC Si OMe OC Me OMe
hν
Fe
Me
Me
HSiMe2Si(Me)Mes2 hν, C6D6
OC W Me OC CO
CO + CH4
5
Si
Me
Me 4
1,2-silyl shift
Si Me
40%
Mes
OMe
OC
Me
W Si OC OC
Si
Si Me 3
Me
Fe
OMe
OC
Si Me 2
1
MeO
Si Fe
OMe
OC
CO
Me
MeO
Mes
Si
Mes
Mes
Me
W SiMe3 OC CO 7
6
Scheme 7.1. Light-induced 1,2-silyl shift to synthesize early transition metal silylenoid complexes.
Me3SiOTf
OC Me3P
Ru Si(SEt)3
NaBPh4
Ru
OC Me3P
Si EtS
8
OTf
OC Me3P
SEt
Ru
SEt BPh 4
9
10
PCy3
H H
PCy3
Me3SiOTf
Pt Cy3P
Si(SEt)3
11
57%
PCy3
H
NaBPh4
Pt Cy3P
SEt
Si
Si(SEt)2OTf
Pt Cy3P
56%
Si
SEt
SEt
BPh4
13
12
Scheme 7.2. Synthesis of late transition metal silylenoid complexes.
ruthenium silylenoid could be accessed in two steps from the silyl ruthenium cyclopentadienyl complex 8. Ethyl thiolate abstraction by trimethylsilyl triflate generated 9, which formed the silylenoid complex (10) on exposure to sodium tetraphenylborate.38 A similar strategy was employed for the synthesis of platinum silylenoid 13.38 Downfield 29 Si {1 H} NMR resonances (d 251 and 309 ppm for 10 and 13, respectively) characterized these metal silylenoid complexes. In the presence of a stabilizing base (such as Et2O),39 an upfield shift to d 89 ppm in the 29 Si {1 H} NMR was observed for a platinum silylenoid. Tilley and coworkers reported a more direct procedure for the synthesis of late transition metal silylenoid complexes by reaction with silanes [equation (7.1)].36,37 Exposing iridium complex 14 to dimesitylsilane afforded iridium silylenoid 15.36 In addition to dimesitylsilane, other silanes could be used, including diphenyl-, diethyl-, or dimethylsilane. Primary silanes such as mesitylsilane and 2,4,6-triisopropylphenylsilane were also tolerated as substrates.37 Ph B
PPh2 Ph2 P Ir P Ph2 H 14
Ph H2SiMes2 95°C 100%
B
PPh2 Ph2 Mes P Ir Si P Ph2 H Mes 15
ð7:1Þ
186
SILVER-CATALYZED SILYLENE TRANSFER
t-Bu N Si N t-Bu
t-Bu N + Ni(CO)4
54%
OC OC
Si
N t-Bu t-Bu
Ni Si N
N
Cp*
Cp*
Cp*
+ Ph3PAuCl
Si
100%
Cp* 18
Cl
Si
AuPPh3 19
t-Bu 17
16
Scheme 7.3. Generation of transition metal silylenoid complexes from stable silylenes.
The coordination of stable silylenes to metal complexes was also reported to produce transition metal silylenoids.40,41 Exposure of unsaturated silylene 16 to Ni(CO)4 resulted in the formation of the disilylene-substituted tetrahedral nickel complex 17 (Scheme 7.3).42,43 Similarly, mixing Ph3PAuCl with decamethylsilicocene 18 produced the silylgold complex 19.44 The 29 Si {1 H} NMR spectrum of 19 (d 78 ppm) revealed its silylenoid character. In addition to nickel and gold, other metals, including tungsten,41 platinum,45 iron,40 and ruthenium,46 have been utilized to form silylmetal complexes of stable silylenes. Metal-catalyzed silylene extrusion from hexamethylsilirane 20 was also used to generate transition metal silylenoids [equation (7.2)].47 In the presence of a metalloporphyrin, [Os(TTP)]2 (where TTP ¼ the dianion of meso-tetra-p-tolylporphyrin), hexamethylsilirane 20 extruded dimethylsilylene to produce the THF-stabilized osmium silylenoid complex 21. Despite the coordinated THF molecule, crystallographic analysis of the silylosmium complex 21 revealed the sp2 nature of the silicon atom (Os--C--Si bond angles 121 and 117 ). Osmium silylenoid 21 was also synthesized from the reaction of [Os(TPP)]2 and dimethyldichlorosilane.47 Me Me Si Me Me Me Me 20 10 equiv
Me [Os(TTP)]2 52%
(TTP)2Os
Me
Me
Si
+
Me
THF
Me Me
ð7:2Þ
21
The structure and bonding of metal silylene complexes varied from those of their carbon analogs. While Fischer-type metal carbene complexes without solvent adducts have been extensively characterized,48 most metal silylenoid complexes contain a bound solvent molecule or counterion on the silicon atom. The bond energy for donor silicon complex 22 was determined to be between 15 and 20 kcal/mol:49,50 NMe2
NMe2
(OC)5Cr
Si
∆G‡ ~ 15–20 kcal·mol
(OC)5Cr
Si
Me2N 22
23
NMe2
ð7:3Þ
187
REACTIVITY AND ATTRIBUTES OF METAL SILYLENOIDS
The bonding picture of a metal with a divalent silicon atom emerged through computational studies to explain the need for the incorporation of a Lewis base at silicon.51,52 Similar to the bonding scheme of a metal carbene, the Fischer-type silylene (OC)5Mo¼SiH2 contains an sp s donor and p p acceptor. Silicon differs from carbon in that p bonding is less efficient. This less efficient p bonding leads to a high polarity of the metal silicon double bond and a nearly barrier-free rotation.51 7.2.2
Reactivity of Transition Metal Silylenoids
The reactivity of early transition metal silylenoid complexes is still emerging. An example of the chemistry that these complexes can participate in is the sila-Wittig reaction [equation (7.4)].53 In this transformation, a metathesis occurred between the chromium silylenoid 24 and the dimethyl carbonate to afford a new Schrock carbene, 25, and the trimerized product 26. This methodology allowed access a new carbene complex that eluded previous synthetic efforts. O
Me (OC)5Cr
MeO
Si
OMe OMe
(OC)5Cr
12%
Me
+
C
Me2Si O
OMe
24
25
O
SiMe2 O
Si Me2 26
ð7:4Þ
The reactivity of platinum silylenoid 27 was explored with traditional silylene trapping reagents. While the silylenoid did not react with triethylsilane or 2,3dimethyl-1,3-butadiene, phenylacetylene was a viable substrate, providing the metallocyclohexadiene 29 (Scheme 7.4).54 The formation of platinum complex 29 was hypothesized to occur via platinum cyclobutene intermediate 28, which formed on insertion of the acetylene into the platinum–silicon bond. A second molecule of phenylacetylene was then inserted into the remaining platinum–silicon bond to provide the observed product. 7.2.3 Transition Metal Silylenoid Complex–Catalyzed Hydrosilation Reactions Cationic ruthenium silylenoid complexes function as efficient hydrosilation catalysts. Tilley and Glaser reported that 1 mol% of 30 catalyzed the hydrosilation of
(Et3P)2Pt
27
Ph
Si
H
Et3P Pt Si Me 46%
60ºC
Me
Ph
Et3P Me
Me
Ph
(Et3P)2Pt
Si
Me Me
+ Me2SiH2
Ph
28
29
Scheme 7.4. Reactivity of platinum dimethylsilylenoid toward phenylacetylene.
188
SILVER-CATALYZED SILYLENE TRANSFER
1-methylcyclohexene with phenylsilane [equation (7.5)].55 This process occurred in a diastereo- and regioselective fashion to produce silane 31 as a single isomer. Other petroleum adducts, including 1-hexene, ethene, and styrene, were efficiently hydrosilated. In addition to phenylsilane, n-hexylsilane could be used without reduction of catalytic efficiency. While this reaction tolerated mono-, di-, or trisubstituted olefins, it was limited to primary silanes; no reaction was observed with diethyl- or diphenylsilane, and the hydrosilation product (e.g., 31) was not converted to the tertiary silane.
Ru
i-Pr3P
Si
H
H H Ph B(C6F5)4 30
Me +
H3SiPh
Me
(1 mol%)
SiH2Ph
60°C, 18 h
ð7:5Þ
70%
31
In addition to ruthenium, Tilley and coworkers also reported that cationic iridium silylenoid complexes were efficient olefin hydrosilation catalysts [reaction (7.6)].56 This silylene complex catalyzes the hydrosilation of unhindered mono- or disubstituted olefins with primary silanes to produce secondary silanes with anti Markovnikov selectivity. Iridium catalyst 32 exhibited reactivity patterns similar to those of ruthenium 30; only primary silanes were allowed as substrates. In contrast to 30, cationic iridium 32 catalyzed the redistribution of silanes. Exposing phenylsilane to 5 mol% of 32 in the absence of olefin produced diphenylsilane, phenylsilane, and silane. i-Pr2 H P Si Mes N Ir H P B(C6F5)4 i-Pr2
32 Ph
+
H3SiMes
(5 mol%) BrC6D5, 60 °C 77%
Ph
SiH2Mes
ð7:6Þ
33
Iridium silylenoids were also reported to catalyze the hydrosilation of ketones [equation (7.7)].57 The room temperature reduction of acetophenone was achieved using only 5 mol% of silylenoid 34 to afford alkoxysilane 35 and ethyl benzene. In addition to acetophenone, acetone was also reported as a competent substrate for hydrosilation.
189
REACTIVITY AND ATTRIBUTES OF METAL SILYLENOIDS
Me3P Ir Si Ph H Ph B(C6F5)4
34
O + Ph
Ph2SiH2
Ph
Me
7.2.4
OSiPh2H
(5 mol%)
+
Ph
Me
Me 35 54%
ð7:7Þ
20%
Transition Metal Silylenoid–Catalyzed Atom Transfer Reactions
While trisubstituted silanes are inactive reagents for the above-mentioned hydrosilation processes, under appropriate conditions they can be sources of silylene. Disilanes, in particular, form silylenoids on exposure to platinum complexes and carbonyl compounds (Scheme 7.5).58 Hayashi and Okamoto reported that exposure of pentamethyldisilane (HSiMe2SiMe3) to 3 mol% of (cod)PtCl2 and b,b-dimethylsubstituted phenylenone 36 produced a 10 : 1 mixture of oxasilacyclohexene 38 and oxasilacyclopentene 39. While the process works well for b,b-disubstituted phenylenones, replacement of one of the b-methyl groups with a hydrogen atom changes the product distribution to favor formation of the oxasilacyclopentene (42 or 43). Higher yields of oxasilacyclopentene were obtained when pentaphenyldisilane 41 was used as the source of silylene.59 Tetrasubstituted silanes are also sources of silylene. Suginome and coworkers reported that palladium catalyzed the transfer of dimethylsilylene, formed from silylborane 44, to alkynes [equation (7.8)].60 Exposure of silylborane 48 and alkyne 49 to substoichiometric amounts of palladium and P(t-Bu)2(2-biphenyl) afforded 2,4-disubstituted silole 50. This process tolerates a variety of functionality including silyl ether-, dimethylamino-, and trifluoromethyl groups. In addition to aliphatic terminal acetylenes, arylacetylenes were also competent substrates. For
Me
O
Me
+ Ph
36
H
+ Ph
40
Me
R R Si SiR 3 H 37 (R = Me) or 41 (R = Ph) (1.2 equiv)
Me Si
(cod)PtCl2 (3 mol%) PhMe, 40°C
O
Me
+ Ph
38 72%
37 (1.2 equiv) O
Me
Me Me Si SiMe 3 H
(cod)PtCl2 (3 mol%) PhMe, 40°C
Me Me Si O Me Me 39 7%
Ph
R R Si O Me Ph 42 (R = Me) 67% 43 (R = Ph) 89%
Scheme 7.5. Platinum-catalyzed formation of oxasilacyclohexenes or oxasilacyclopentenes.
190
SILVER-CATALYZED SILYLENE TRANSFER
arylacetylenes, triphenylphosphine could be employed as the phosphine ligand without attenuation of the regioselectivity. Pd(dba)2 (1 mol%)
(CH2)3Cl
P(t-Bu)2(2-biphenyl)
Me Et3N Si Bpin + Cl Me
H
(1.2 mol%) PhMe, 25°C 78%
Cl(H2C)3
Si Me Me 46
45 (2.4 equiv)
44
Cl(H2C)3
(CH2)3Cl
+
91:9
Si Me Me 47
ð7:8Þ Palladium(0) complexes also catalyze the transfer of di-tert-butylsilylene from cis-dimethylsilacyclopropane 48 to alkynes (Scheme 7.6).61 In the presence of 5 mol % of (Ph3P)2PdCl2, the formation of dimethylsilacyclopropene 51 could be achieved at 110 C. In the absence of the palladium catalyst, silylene transfer from silacyclopropane 48 to 3-hexyne occurred over 3 days at 130 C. While a reasonable mechanism was proposed involving palladacycle 49 as an intermediate, an alternative mechanism could involve palladium silylenoid intermediate 50. t-Bu t-Bu Si Me Et
Me 48
Et t-Bu t-Bu
Et
(Ph3P)2PdCl2 (0.5 mol%) 117°C, 17 h
Si Me 49
t-Bu
Et PdLn Me
LnPd OR
50 +
t-Bu t-Bu Si
Si
Me
t-Bu 86%
Et
Et Me
51
Scheme 7.6. Palladium-catalyzed di-tert-butylsilylene transfer to an internal alkyne.
7.3 SILACYCLOPROPANES AS IMPORTANT SYNTHETIC INTERMEDIATES To develop new methods for organic synthesis, Woerpel and coworkers exploited the inherent reactivity of di-tert-butylsilacyclopropanes to create new carbon–carbon bonds in a stereoselective fashion (Scheme 7.7).62 They discovered that transition metal salts catalyze the insertion of carbonyl compounds into the strained carbon– silicon bond to form oxasilacyclopentanes. The regioselectivity of insertion could be controlled by the identity of the catalyst. Copper promoted the insertion of crotonaldehyde into the more substituted C--Si bond of 52 to afford oxasilacyclopentane 53,63 whereas zinc catalyzed the insertion of butyraldehyde into the less substituted bond of 52 to provide the complementary product, 54.64 Oxasilacyclopentanes (e.g., 55) could be transformed into useful synthetic intermediates through oxidation of the C--Si bond,65,66 which provided diol 56 with three contiguous stereocenters. Despite these advances, the conditions required for the construction of silacyclopropanes limited their synthetic utility (Scheme 7.8). These heterocycles can be
191
SILACYCLOPROPANES AS IMPORTANT SYNTHETIC INTERMEDIATES
Me
t-Bu t-Bu Si O
CHO
CuBr2 (10 mol%) 70%
t-Bu Si t-Bu
Me i-Pr 53 (dr 98 : 2) >99 : 1 regioselectivity
i-Pr 52
Me
t-Bu t-Bu Si O
CHO
ZnBr2 (10 mol%) 70%
t-Bu t-Bu Si O
Me 54 (dr 55 : 45) >99 : 1 regioselectivity OH OH
t-BuOOH, CsOH
Ph
Me
i-Pr
Me
n-Bu4NF, DMF
Ph Me
71%
Me 55
56
Scheme 7.7. Synthetic utility of silacyclopropanes.
prepared by reaction of an olefin with a silylene or a silylenoid species. Strongly reducing conditions [Li(0) and t-Bu2SiCl2] are needed to access the requisite lithium silylenoid (57).67 Alternatively, free silylene 60 could be generated thermally or photochemically from cyclic silanes, such as 58 or 59.12j,68–73 Each of these methods, however, limited the functionality that could be introduced on the silacyclopropane. To overcome these limitations, Woerpel and coworkers set out to develop conditions that generated a silylenoid species at low temperatures and exhibited broad functional group tolerance.
t-Bu2SiCl2
t-Bu
Li(0) sonication
t-Bu
Li Si Cl
57
t-Bu Si t-Bu
t-Bu t-Bu Si t-Bu or t-Bu Si Si t-Bu
n-Bu t-Bu
∆
61 Si
or h ν
t-Bu
t-Bu 58 59 cyclic silane examples
t-Bu2SiR2
t-Bu Si t-Bu
n-Bu
60
LnMXm
?
t-Bu
t-Bu Si t-Bu
R
Si
MLn Xm
t-Bu
?
R
62 Goals: low temperature, broad functional group tolerance
Scheme 7.8. Synthesis of silacyclopropanes.
63
192
SILVER-CATALYZED SILYLENE TRANSFER
7.4 SILVER-MEDIATED TRANSFER OF DI-tert-BUTYLSILYLENE TO OLEFINS Exploration of the reactivity of cyclohexene silacyclopropane led Woerpel and coworkers to discover that the inclusion of metal salts enabled silylene transfer to monosubstituted olefins at reduced temperatures (Table 7.1).11,74 A dramatic reduction in the temperature of transfer was observed when cyclohexene silacyclopropane was exposed to copper, silver, or gold salts. Silver salts were particularly effective at decomposing 58 (entries 6–11). The use of substoichiometric quantities of silver triflate enabled n-hexene silacyclopropane 61 to be formed quantitatively at 27 C (entry 6). The identity of the counterion did affect the reactivity of the silver salt. In general, better conversions were observed when noncoordinating anions were employed. While the reactivity differences could be attributed to the solubility of the silver salt in toluene, spectroscopic experiments suggested that the anion played a larger role in stabilizing the silylenoid intermediate. Under these optimized conditions, di-tert-butylsilylene could be transferred to a range of acyclic and cyclic olefins (Schemes 7.9 and 7.10).11,74 The method was not sensitive to the steric nature of the R substituent; nearly quantitative silylene transfer to olefins bearing n-butyl, isopropyl, or tert-butyl groups was observed. Vinylsilanes were also tolerated as substrates. Olefins containing silyl ether, benzyl ether, and pivolate substituents were all effective traps of di-tert-butylsilylene. Silver-catalyzed silylene transfer to disubstituted olefins was also possible (Scheme 7.10).11,74 The transformation was stereospecific; cis-2-butene afforded cis-dimethylsilacyclopropane 66a and trans-2-butene generated trans-dimethylsilaTABLE 7.1. Effect of a Metal Salt on the Temperature of Di-tert-Butylsilylene Transfer t-Bu + n-Bu
Si t-Bu 58
Entry 1 2 3 4 5 6 7 8 9 10 11 a
t-Bu Si t-Bu +
MXn (10 mol%) PhMe
n-Bu
(1.3 equiv)
61
MXn
T ( C)
Time
Yield (%) a
Zn(OTf)2 Cu(OTf)2 Cu(OTf)2 (CuOTf)2PhH CeCl3 AgOTf AgO2CCF3 AgSbF6 AgOBz AgBF4 Ag3PO4
55 25 0 0 25 27 27 27 0–25 25 25
12 h 17 min 2h 2h 15 h 2h 2h 5h 8h 45 min 16 h
60 96 62 62 87 90 86 >95 >95 99 94
As determined using H1 NMR spectroscopy.
193
SILVER-MEDIATED TRANSFER OF DI-tert-BUTYLSILYLENE TO OLEFINS
+
R
n-Bu 63a 90%
t-Bu Si t-Bu
t-Bu
O
PhMe2Si
Bn 63d 90%
t-Bu Si t-Bu TIPSO 63f 82%
63e 88%
t-Bu Si t-Bu
t-Bu Si t-Bu MeO MeO
O
t-Bu
OTIPS
63h 67%
63g 71%
t-Bu Si t-Bu
t-Bu Si t-Bu
t-Bu Si t-Bu
O
+
R 63
t-Bu 63c 96%
i-Pr 63b 99%
BnO
t-Bu
PhMe-d8 –27°C, 2 h
t-Bu Si t-Bu
t-Bu Si t-Bu
t-Bu Si t-Bu
AgOTf (10 mol%)
58
64 (1.3 equiv) t-Bu Si t-Bu
t-Bu Si
O 63j 83%
63i 72%
Scheme 7.9. Silver-catalyzed di-tert-butylsilylene transfer to monosubstituted olefins.
cyclopropane 66b as single isomers. Di-tert-butylsilylene transfer was also diastereoselective. Exposure of 2-substituted cyclopentene 65d or a-pinene to reaction conditions afforded 66d and 66e as single diastereomers. Diastereoselective silylene transfer was not limited to cyclic olefins; silacyclopropane 66f was formed from 65f as a 9 : 1 mixture of diastereomers. One drawback of this methodology was the sensitivity of the product silacyclopropane to the reaction conditions, which prevented isolation of the product. Woerpel and coworkers were able to overcome this debilitating limitation by changing the anion of silver catalyst to trifluoroacetate and reducing the catalyst load to <1 mol% (Scheme 7.11).74 The combination of these two modifications enabled the isolation of the silacyclopropane product after careful Schlenk filtration. Isolation of the silacyclopropane allowed for either zinc-64,75 or copper-mediated functionalization of the silacyclopropane to produce 1,2- or 1,3-disubstituted oxasilacyclopentanes.63 Alternatively, the silacyclopropane could be functionalized in situ through the addition of 30 mol% of zinc bromide (or zinc iodide) and a carbonyl compound to the reaction mixture, which produced oxasilacyclopentane 70 (Scheme 7.12).11,74
R1
R2
+
Me 66a 99%
t-Bu
PhMe-d8 –27°C, 2 h
t-Bu Si t-Bu
t-Bu Si t-Bu TIPSO
66b 92%
Bu 66c >95%
R2 66
Me t-Bu Si t-Bu
PhMe2Si Me
+
R1
58
t-Bu Si t-Bu Me
Me
AgOTf (10 mol%)
Si
65 (1.3 equiv) t-Bu Si t-Bu
t-Bu
Me OTIPS Si
Et 66d 85% (dr >95 : 5)
t-Bu Si t-Bu
t-Bu Me
t-Bu 66e 72% (dr >95 : 5)
Me
Me
66f 90% (dr 89 : 11)
Scheme 7.10. Silver-catalyzed di-tert-butylsilylene transfer to disubstituted olefins.
194
SILVER-CATALYZED SILYLENE TRANSFER t-Bu Si t-Bu
58, AgO2CCF3 (<1 mol%)
Me
Me
PhMe –27°C, 2 h 67%
Me (10 equiv)
Me 66a
(1.3 equiv)
Ph
O
CHO
n-Bu
PhMe –27°C, 2 h 90%
t-Bu Si
MXn
Me
PhMe –27°C, 2 h 85%
(10 equiv)
t-Bu Ph Si t-Bu
58, AgO2CCF3 (<1 mol%)
n-Bu
Me
t-Bu Si t-Bu
58, AgO2CCF3 (<1 mol%)
Me
t-Bu t-Bu Si O
+
t-Bu
n-Bu
63a
Me 66b
Ph
n-Bu
67
68
MXn Yield, (%) 67 : 68 CuI ZnBr2
85% 92%
96 : 4 1 : 99
Scheme 7.11. Formation of either 1,2- or 1,3-disubstituted oxasilacyclopentanes.
This functionalization was limited to the formation of 1,3-substituted oxasilacyclopentanes, as exposure of the in situ–generated silacyclopropane to substoichiometric amounts of copper salts did not produce the complementary 1,2-disubstituted oxasilacyclopentanes.
58, AgOTf (10 mol%)
R1
PhMe –27°C, 2 h
t-Bu Si t-Bu R1
n-Bu
t-Bu t-Bu Si O
OMe
i-Pr
70a 87% (dr 76 : 24)
t-Bu t-Bu Si O
OMe
BnO
OMe
t-Bu t-Bu Si O
70i 78% (dr 64 : 36)
Bn
t-Bu t-Bu Si O
OMe
t-Bu t-Bu Si O OMe MeO OTIPS 70g 82% (dr 70 : 30)
t-Bu t-Bu Si O Me
Bn
R1
70c 75% regioselectivity 74 : 26 (major dr 70 : 30) (minor dr 91 : 9)
70f 61% (dr 70 : 30)
t-Bu t-Bu Si O
ZnX2 (20 mol%)
t-Bu
70b 92% (dr 70 : 30)
TIPSO 70e 61% (dr 64 : 36)
t-Bu t-Bu Si O
R3
69
(1.3 equiv) t-Bu t-Bu Si O
O R2
t-Bu t-Bu Si O O
70j 80% (dr 65 : 35)
Bn 70k 57%
Et Et
R3 R2
70
t-Bu t-Bu Si O Bn
OMe
70d 80% (dr 70 : 30)
t-Bu OMe t-Bu Si O
Ph
n-Bu 70h 73% (dr 65 : 35) t-Bu t-Bu Si O Bn
Me Ph
70l 71% (dr 55 : 45)
Scheme 7.12. In situ functionalization of silacyclopropanes derived from monosubstituted olefins.
195
SILVER-MEDIATED TRANSFER OF DI-tert-BUTYLSILYLENE TO OLEFINS
O Me Me
3
2
OH OH
Me O O
Me
1'
Et
2
3
Me
Me 1'-epi -stegobinone
O
OH
Me
+ Me
t-Bu 58
Et Me Me 72
71
Si
O
Me
Me
t-Bu Si t-Bu
t-Bu
Me
Me
1'
t-Bu t-Bu Si O
Me
t-Bu t-Bu Si O Me
Me 65a
OAc Me 73
Scheme 7.13. Retrosynthesis of 10 -epi-stegobinone.
The synthetic utility of silacyclopropanes was illustrated in the total synthesis of 10 -epi-stegobinone (Scheme 7.13).76 Woerpel and coworkers envisioned that every stereocenter in this pheromone77 could be derived from the stereochemistry of the starting cis-dimethylsilacyclopropane 65a. They believed that 10 -epi-stegobinone could be synthesized from polypropionate 71. This fragment could be produced from an anti,syn-selective aldol reaction of ethyl ketone 72, which is available from the functionalization of oxasilacyclopentane 73. This oxasilacyclopentane is readily available from silacyclopropane 65a, produced from di-tert-butylsilylene transfer from 58 to cis-2-butene. Implementation of the Woerpel methodology at the outset of the total synthesis would illustrate the power of the methodology to move large amounts of material through the silicon atom transfer step. Gram quantities of oxasilacyclopentane 73 were prepared from cis-2-butene through silver-catalyzed di-tert-butylsilylene transfer from cyclohexene silacyclopropane 58 followed by copper-mediated formamide insertion (Scheme 7.14). After acetolysis, the resulting oxasilacyclopentane (73) participated in a nucleophilic substitution reaction with silyl enol ether 74 to produce ethyl ketone 72 diastereoselectively.62,78–80 After protection of the ketone, the hindered C--Si bond was oxidized using conditions developed by Woerpel and coworkers to afford 75.66,81,82 The resulting polypropionate fragment was elaborated in 10 steps to 10 -epi-stegobinone. This first report of using silacyclopropanes as synthetic intermediates in total synthesis clearly demonstrates their potential as polyproprionate synthons. Insight into the mechanism of silver-catalyzed silylene transfer from cyclohexene silacyclopropane to an olefin was obtained using bistriphenylphosphine silver triflate as a catalyst.83 Woerpel and coworkers chose to employ ancillary ligands on silver to address the both the poor solubility of silver triflate as well as its propensity to decompose to afford a silver(0) mirror or precipitate. The addition of triphenylphosphine, however, attenuated the reactivity of the silver catalyst. For example, the reaction temperature needed to be raised from 27 C to 10 C to obtain a moderate rate of reaction (Scheme 7.15).
196
SILVER-CATALYZED SILYLENE TRANSFER OSiMe3
t-Bu Si t-Bu
1. 58, AgO2CCF3 (<1 mol%)
Me Me
t-Bu
2.
Me
PhMe –27°C, 2 h
(excess)
OHC
Bn
N Me
, CuI
3. CuSO4 (aq. soln.) 4. Ac2O, DMAP
Me 65a
Me
t-Bu t-Bu Si O
Et 74
OAc SnBr4
Me Me 73 74% from 65a O
t-Bu Si O
Me
OH OH
1. Ph3PCH3Br n-BuLi, 79%
O
2. PhMe2C(OOH), KH, CsF, 94%
Et Me Me 72 95% (dr 97 : 2 : 1)
Me
Me
3
Me
2
Me O
Et Me 75
O
Scheme 7.14. Total synthesis of 10 -epi-stegobinone.
t-Bu Si 58
Ph
t-Bu
Ln AgX (5 mol%)
Ph
63d
T (°C) t1/2 (approx.)
Ln AgX
t-Bu Si t-Bu
AgOTf
–27
24 min
(Ph3P)2AgOTf
10
60 min
Scheme 7.15. Effect of phosphine ligand on rate of silylene transfer.
The relative reactivities of various silacyclopropanes were compared to gain insight into the reversibility of the reaction (Scheme 7.16).83 Woerpel and coworkers observed that the reactivity of the silacyclopropane toward the silver catalyst depended on the ring size, and that cyclohexene silacyclopropane was the most reactive. Notably, benzyl-substituted silacyclopropane 63d did not react when exposed to the silver complex. This lack of reactivity was interpreted as evidence in support of the irreversibility of silylene transfer to monosubstituted olefins. t-Bu Si t-Bu R
(5 mol%)
Ph
CD2Cl2
(10 equiv) t-Bu Si
58, 10°C t1/2 ~ 1 h
Ph
Si t-Bu
77, 25°C t1/2 ~ 25 h
t-Bu 78, 60°C t1/2 ~ 300 h
R
76
t-Bu
t-Bu Si
t-Bu
t-Bu Si t-Bu +
(Ph3P)2AgOTf
+
1'
Et Me 1'-epi -stegobinone
Me
Ph
t-Bu Si t-Bu
63d, 60°C no reaction
Scheme 7.16. Relative reactivities of silacyclopropanes.
197
SILVER-MEDIATED TRANSFER OF DI-tert-BUTYLSILYLENE TO OLEFINS
t-Bu
Ph
Si
(Ph3P)2AgOTf CD2Cl2
t-Bu
Component Kinetic order 58 1 styrene <1 (Ph3P)2AgOTf Saturation Ph3P <1 Cyclohexene <1
t-Bu Si t-Bu +
Ph 76 58 ‡ ‡ ∆H = 30(1) kcal/mol ∆S = 31(7) eu
Scheme 7.17. Kinetic studies on silver-catalyzed silylene transfer.
To elucidate the mechanism of silver-catalyzed silylene transfer, kinetic studies were performed by Woerpel and coworkers (Scheme 7.17).83 The reaction of cyclohexene silacyclopropane 58 and styrene in the presence of 5 mol% of (Ph3P)2AgOTf was followed using 1 H NMR spectroscopy. The kinetic order in cyclohexene silacyclopropane 58 was determined to be 1. In contrast to the rate acceleration observed with increasing the concentration of 58, inhibition of the rate of the reaction was observed when styrene, cyclohexene, or triphenylphosphine concentrations were increased. Saturation kinetic behavior in catalyst concentration was observed. Activation parameters were determined to be DHz ¼ 30(1) kcal/mol and DSz ¼ 31(7) eu (entropy units). Similar activation parameters were observed in toluene-d8, which suggests that the mechanism is not dependent on the solvent. A catalytically active silylsilver intermediate was observed using low-temperature 29 Si NMR and IR spectroscopy (Scheme 7.18).83 Exposure of cyclohexene silacyclopropane 58 to the silver complex produced cyclohexene as well as a new species, which exhibited two doublets at 97 ppm (JAgSi ¼ 260 and 225 Hz) in the 29 Si {1 H} NMR spectrum. The 29 Si {1 H} NMR spectra of this species are consistent with a Lewis base–stabilized metal silylenoid. Tilley and coworkers have reported that (Et3P)2Pt(H)--Si(St-Bu)2(OTf) appears at 52 ppm,39 and J€utzi et al. observed
t-Bu
(Ph3P)2AgOTf
Si
CD2Cl2, –40°C
t-Bu 58 (2 equiv)
t-Bu t-Bu Si
t-Bu Si t-Bu
Ph
AgLn
40%
Ph
TfO
63d
77 : 97 ppm J = 260, 225 Hz
29Si {1H} δ
Related metal silylenoid complexes H
Et3P Pt
Cl OTf
Et3P
Si
t-BuS
Cp* Cp*
St-Bu
78 (Ref. 39) 29Si {1H}
Ph3P Au Si
δ 52 ppm
19 (Ref. 44) 29Si {1H}
δ 78 ppm
St-Bu
Cy3P Pt Cy3P
Si St-Bu
79 (Ref. 34) 29Si {1H}
δ 367 ppm
Scheme 7.18. Spectroscopic attributes of silylsilver intermediate 77.
198
SILVER-CATALYZED SILYLENE TRANSFER
Ph3PAu--SiCp2Cl at 78 ppm.44 In contrast, neutral silylenoids appear much further downfield: Tilley et al. reported that (Cy2P)2Pt¼SiMes2 appeared at 367 ppm.34 Infrared spectroscopy was also used to gain insight into the structure of the intermediate as ion-paired triflate salts exhibit stretches at 1270 and 1043 cm1,84 whereas covalent t-BuMe2SiOTf produces a spectrum with peaks at 1212 and 970 cm1. Low-temperature IR spectroscopy of the intermediate in CH2Cl2 revealed strong stretches at 1204 and 972 cm1, which is consistent with a monodentate binding of the triflate anion to the silicon atom. On the basis of these observations, the structure of the silver silylenoid intermediate was assigned to be silylsilver 77. The electronic nature of silylsilver intermediate was interrogated through intermolecular competition experiments between substituted styrenes and the silylsilver intermediate (77).83 The product ratios from these experiments correlated well with the Hammett equation to provide a r value of 0.62 using sp constants (Scheme 7.19). Woerpel and coworkers interpreted this r value to suggest that this silylsilver species is electrophilic. Smaller r values were obtained when the temperature of the intermolecular competition reactions was reduced [r ¼ 0.71 (8 C) and 0.79 (8 C)]. From these experiments, the isokinetic temperature was estimated to be 129 C, which meant that the product-determining step of silver-catalyzed silylene transfer was under enthalpic control. In contrast, related intermolecular competition reactions under metal-free thermal conditions indicated the product-determining step of free silylene transfer to be under entropic control. The combination of the observed catalytically active silylsilver intermediate and the Hammett correlation data led Woerpel and colleagues to conclude that the silver functions to both decompose the sacrificial cyclohexene silacyclopropane as well as transfer the di-tert-butylsilylene to the olefin substrate. Woerpel and coworkers analyzed the experimental data and constructed a catalytic cycle to describe silver-catalyzed di-tert-butylsilylene transfer from cyclohexene silacyclopropane to monosubstituted olefins (Scheme 7.20).83 Dissociation of the oligomeric silver phosphine complex followed by ligand exchange generates the postulated catalytically active species, 81, which contains only one triphenylphosphine and two styrene ligands. Coordination of cyclohexene silacyclopropane followed extrusion of t-Bu2Si, generates silylsilver 83. This extrusion was postulated to occur via transition state 82. This transition state forms on coordination of triflate to the electrophilic silicon atom of 58, which triggers the heterolysis of the strained C--Si bond in a transmetallation-like mechanism. This heterolysis is followed by b-silyl elimination to produce 83 and cyclohexene. Turnover-limiting cyclization occurs to form silacyclopropane 76 and generate the coordinatively unsaturated silver catalyst.
+
Ar (10 equiv)
58 (1 equiv) Ln AgX (5 mol%)
Ph (10 equiv)
toluene-d8, 25°C
t-Bu Si t-Bu + Ph
t-Bu Si t-Bu Ar
76
80
Ln AgX
ρ value
–0.59(2) AgOTf (Ph3P)2AgOTf –0.62(2)
Scheme 7.19. Hammett correlation of intermolecular competition experiments.
199
SILVER-MEDIATED TRANSFER OF DI-tert-BUTYLSILYLENE TO OLEFINS t-Bu Si t-Bu 58
(Ph3P)2AgOTf Ph
t-Bu t-Bu TfO Si [Ag] 82
Ph3P Ph
t-Bu t-Bu Si TfO
Ph3PAg(styrene)2OTf 81
Ph3PAg(styrene)3 OTf
Ph3P
76
83
Ph
RDS
t-Bu t-Bu Si
Ph
silver oligomers
Ph Ag(PPh3)
OTf t-Bu Si t-Bu 84
Ph
[Ag] Ph
Ph
Scheme 7.20. Potential catalytic cycle for silver-mediated di-tert-butylsilylene transfer from cyclohexene silacyclopropane 58 to styrene.
An unexpected product containing the incorporation of two molecules of silylene was observed in the reaction of homoallylic ethers at elevated temperatures (Scheme 7.21).85 At reduced temperatures, the reaction of homoallylic ethers with cyclohexene silacyclopropane 58 produced the expected silacyclopropane (Scheme 7.9, cf. 63g and 63h). Increasing the temperature, however, resulted in a rearrangement to afford allylsilane 86. This transformation proved general and a range of silylmethyl allylic silanes could be accessed from mono- or geminal disubstituted homoallylic ethers. The resulting allylic silane was demonstrated to be a competent nucleophile through reaction with N-chlorosulfonyl isocyanate. The mechanism of this transformation was probed using regiospecifically labeled substrates and silacyclopropanes (Scheme 7.22).85 No crossover was observed in the reaction of 87 and 88. Further, incorporation of deuterium into the silane and a position revealed that reorganization of the substrate had occurred. Monosubstituted silacyclopropanes were established as potential reactive intermediates by readily reacting in the presence of cyclohexene silacyclopropane 58 to give the silylmethylsilane 93. Control experiments established that these monosubstituted silacyclopropanes did not extrude di-tert-butylsilylene under reaction conditions. R2 OR3 R1
R4 85
t-Bu +
Si t-Bu 58 (2 equiv)
t-Bu t-Bu 3 R1 Si OR
AgO2CCF3 (0.5 mol%)
25–50°C
(t-Bu)2HSi
R4 86
R2
t-Bu t-Bu t-Bu t-Bu t-Bu t-Bu Si OSiMe2t-Bu t-Bu Me Si OSiMe2t-Bu t-Bu Si OC10H21 Me Si OMe (t-Bu)2HSi (t-Bu)2HSi (t-Bu)2HSi (t-Bu)2HSi i-Pr Me Me 86a 86c 86b 86d 85% 82% 82% 100% dr 89 : 11
Scheme 7.21. Formation of silylmethyl allylic silanes from gem-disubstituted homoallylic ethers.
200
SILVER-CATALYZED SILYLENE TRANSFER
t-Bu t-Bu H Si OBn OBn
(t-Bu)2DSi
t-Bu Me Si t-Bu
D D 87
Me
90 80% +
89
+
AgO2CCF3 (1 mol%)
OC10H21 88
(t-Bu)2HSi
D
t-Bu t-Bu H Si OC10H21 91 85%
t-Bu t-Bu Si
t-Bu OPr
+
Si
AgO2CCF3 (1 mol%) 50°C
t-Bu
Me 92
t-Bu t-Bu R1 Si OPr (t-Bu)2HSi Me
58
93 47%
Scheme 7.22. Mechanistic experiments.
t-Bu t-Bu Si Me OBn 94
t-Bu t-Bu Si BnO 95
t-Bu t-Bu Si Me TfO 96
AgL3
t-Bu Me t-Bu Si BnO Si(t-Bu)2 H
t-Bu t-Bu Si OBn Me
97
SiH(t-Bu)2 98
Scheme 7.23. Potential mechanism for formation of silylmethyl allylic silanes from gemdisubstituted homoallylic ethers.
From these observations, Woerpel and Cleary proposed a mechanism to account for allylic silane formation (Scheme 7.23).85 Silacyclopropane 94 is formed from cyclohexene silacyclopropane 58 through silylene transfer. Coordination of the Lewis basic benzyl ether to the electrophilic silicon atom86–88 generates pentacoordinate siliconate 95 and increases the nucleophilicity of the apical Si--C bond.89 Electrophilic attack by silylsilver triflate 96 forms silyl anion 97. Intramolecular deprotonation and elimination then affords the silylmethyl allylic silane.
7.5 SILVER-MEDIATED TRANSFER OF DI-tert-BUTYLSILYLENE TO ACETYLENES Transition metal complexes were known to facilitate the addition of silylene to acetylenes from a variety of different sources.60,61,90,91 These conditions, however, often required heating, and the initially formed silacyclopropene often incorporated a second molecule of the acetylene to afford a silole.92,93 With their discovery of lowtemperature silver-mediated di-tert-butylsilylene transfer conditions from cyclohexene silacyclopropane 58 to olefins, Woerpel and coworkers set out to investigate the
201
SILVER-MEDIATED TRANSFER OF DI-tert-BUTYLSILYLENE TO ACETYLENES t-Bu Si R1
58 R2
t-Bu R
t-Bu Si t-Bu 1
Ag-cat. ?
O R3
t-Bu t-Bu Si O
R4
MXn ?
R2
R4
H+
R3
R1 R2 100
99
R2 R1
OH R3
R4
101
Scheme 7.24. Di-tert-butylsilylene transfer to alkynes.
possibility of transferring di-tert-butylsilylene to mono- and disubstituted acetylenes at low temperatures (Scheme 7.24). If conditions were found to prevent insertion of a second molecule of alkyne, then the resulting silacyclopropenes might be functionalized in situ with a ketone or an aldehyde to produce oxasilacyclopentenes.94 Protonolysis of the C--Si bond would produce allylic alcohol 101. Formally, the resulting method would represent a three-component reductive coupling reaction to produce an allylic alcohol from an alkyne and a carbonyl compound. Woerpel and Clark identified silver phosphate as the optimal catalyst to promote di-tert-butylsilylene transfer from cyclohexene silacyclopropane to a variety of substituted alkynes (Scheme 7.25).95 While this silver salt exhibited attenuated reactivity as compared to silver triflate or silver trifluoroacetate, it exhibited greater functional group tolerance. Both di- and monosubstituted silacyclopropenes were easily accessed. Terminal alkynes are traditionally difficult substrates for silylene transfer and typically insert a second molecule of the starting acetylene.61,90–93 Consequently, the discovery of silver-mediated silylene transfer represents a significant advance as it enables further manipulation of monosubstituted silacyclopropenes. For enyne substrates, silylene transfer the alkynyl group was solely observed. The chemoselectivity of the formation of 99f was attributed to ring strain as theoretical calculations suggest that silacyclopropenes are less strained than silacyclopropanes.96,97
t-Bu Si R1
58 R2
Me
t-Bu Si t-Bu
Me
t-Bu Si t-Bu OTIPS
Ph 99a 91%
99b 94%
t-Bu Si t-Bu
t-Bu
R1
Ag3PO4 (10 mol%)
t-Bu Si t-Bu Ph 99c 95%
R2 99
t-Bu Si t-Bu n-Bu 99d 92%
t-Bu t-Bu Si t-Bu Si t-Bu Me N Bn OTIPS 99e 99f 97% 81%
Scheme 7.25. Silver-catalyzed di-tert-butylsilylene transfer to di- or mono-substituted alkynes.
202
SILVER-CATALYZED SILYLENE TRANSFER
Si R1
t-Bu t-Bu
58 R2
t-Bu t-Bu Si O
H
Ag3PO4 (10 mol%)
t-Bu t-Bu Si O
t-Bu t-Bu Si O
H
t-Bu t-Bu Si O
n-Pr
nPr
SiMe3 100f 78% CuI
100g 72% CuI
Ph Bn N Me
R4 R3
R1
R2 100
Me
t-Bu t-Bu Si O
t-Bu t-Bu Si O
Me
H OEt
Ph 100d 80%
Ph 100c 83% Me
t-Bu t-Bu Si O
R4
Ph
Ph 100b 79% t-Bu t-Bu Si O
O R3
CuBr2 (15 mol%)
R2 99
Ph Ph 100a 79%
t-Bu Si t-Bu 1 R
t-Bu t-Bu Si O Me Ph 100h 94% Cu(OTf)2
Me Ph
t-Bu t-Bu Si O Me Ph 100i 54% Cu(OTf)2
Ph 100e 68%
Ph
nPr
t-Bu t-Bu Si O
Me Ph
Me
OTIPS 100j 90% Cu(OTf)2
Scheme 7.26. Regioselective in situ functionalization of silacyclopropene.
The resulting silacyclopropenes were particularly reactive and challenging to isolate. To overcome this limitation, Clark and Woerpel developed a procedure for in situ functionalization of 99 (Scheme 7.26).95 Copper(II) bromide was used to promote the regioselective insertion of aldehydes and ketones into the more substituted C--Si bond of 99 to afford oxasilacyclopentenes 100 in one flask from the acetylene starting material. Copper-catalyzed ring expansion proved to be quite general, tolerating a range of carbonyl compounds and functional groups. Regioselective oxasilacyclopentane formation (100g–100i) was even observed for disubstituted silacyclopropenes (99g–99i). The origin of selectivity was attributed to steric interactions between the silacyclopropene substituent and the tert-butyl groups, which arise in the transition state when insertion occurs into the less substituted C--Si bond.63 The resulting oxasilacyclopentenes (100) were amenable to further functionalization (Scheme 7.27).95 They could be transformed into allylic alcohol 101h through protodesilylation or 1,3-diol 103 through hydrogenation followed by oxidation.65 The observed diastereoselectivity was rationalized using the Woerpel model for the reaction of reagents with five-membered rings.98 Fewer destabilizing steric interactions occur from addition of hydrogen to the face opposite the pseudoaxial tert-butyl group (TS-102). When the oxasilacyclopentene bears an olefin substituent (e.g., 100k), a [4 þ 2] cycloaddition dramatically increases the molecular complexity. This latter functionalization showcases the potential of silacyclopropenes as synthetic intermediates by producing 105, which contains five stereocenters, from enyne 104 in two steps. Copper salts were found to be successful catalysts for the insertion of nitriles into silacyclopropenes (Scheme 7.28).99,100 A screen of different copper salts identified (CuOTf)2PhMe as the optimal catalyst. This reaction occurs readily at room temperature for monosubstituted silacyclopropenes to afford azasilacyclopentadienes
203
SILVER-MEDIATED TRANSFER OF DI-tert-BUTYLSILYLENE TO ACETYLENES t-Bu t-Bu Si O
Me Ph
Me
n-Pr
1. H2, Pd(OH)2
Me
n-Bu4NF 84%
Ph 100h t-Bu t-Bu Si O
OH Me Ph Ph
KOt-Bu
101h
Me Ph
Me
t-Bu O n-Bu
Si
Ph 100i
OH
2. t-BuOOH, KH
Me
n-Bu4NF 75°C
t-Bu
103 75% (dr 99 : 1)
1. 58, Ag3PO4 (10 mol%)
O
Me
Ph N
t-Bu t-Bu O H Si O Me Ph Ph N
O
Ph 130°C
2. CuI, PhCOMe
OTIPS
O H
OTIPS 100k 58%
104
n-Pr Ph
TS-102
t-Bu t-Bu Si O
OH
OTIPS 105 69% (dr 4 : 1)
Scheme 7.27. Functionalization of silacyclopentenes to access important synthetic intermediates.
106 as single enamine tautomers. Only when sterically more congested nitriles (e.g., i-PrCN) were employed was the imine tautomer (106f) obtained. Higher temperatures were required for insertions into disubstituted silacyclopropenes (99g–99j). These t-Bu Si t-Bu
R1
+
R3
t-Bu H t-Bu Si N
(CuOTf)2·PhMe (5 mol%)
CN
PhMe
H
R2 106
t-Bu H t-Bu Si N
H
i-Pr
106f 64%
OTBDMS
t-Bu H t-Bu Si N
t-Bu H t-Bu Si N
H Ph 106g 85%
regioselectivity 10 : 1
t-Bu H t-Bu Si N
i-Pr 106h
Ph 106e 84%
106d 65% Me
OTIPS
96% regioselectivity 3 : 2
t-Bu H t-Bu Si N Me i-Pr 106i
Ph
t-Bu H t-Bu Si N
H
n-Bu
OTIPS
83% regioselectivity 3 : 1
Scheme 7.28. Insertion of nitriles into silacyclopropenes.
Ph H
H
H
Me
t-Bu H t-Bu Si N
SiMe
106c 83% Ph
Ph
H Ph
106b 84%
Me Ph
t-Bu H t-Bu Si N
Ph
106a 82% t-Bu t-Bu Si N
Ph H
Ph
H
R1
R2 99 t-Bu H t-Bu Si N
R3
OSiEt3
Ph H
i-Pr 106j
OTIPS
96% regioselectivity 10 : 1
204
SILVER-CATALYZED SILYLENE TRANSFER t-Bu Si 28
H Ph
t-Bu H
t-Bu Si t-Bu
Ag3PO4 (10 mol%)
Ph
Ph
t-Bu H t-Bu Si N
CN
Cu(OTf)2 (10 mol%)
H Ph 106b 51%
99c t-Bu
1.
Si 58
Et
t-Bu
Cu(OTf)2 (10 mol%)
Et
Ph
2. Ph
CN
t-Bu H t-Bu Si N
Ph H
Et Et 106k 67%
Scheme 7.29. Two-step, one-flask transformation of alkynes into azasilacyclopentadienes.
reactions were not quite as regioselective, although good selectivity was observed for 99g and 99j. It is apparent from the selectivity observed for the formation of 106g–106j that both electronic and steric factors influence which C--Si bond the nitrile is inserted into. Isolation of the air-sensitive silacyclopropene was avoided by development of a two-step, one-flask procedure, which transformed alkynes into the desired azasilacyclopentadienes (Scheme 7.29).99 For terminal alkynes, silver phosphate was employed for di-tert-butylsilylene transfer and copper(I) triflate was used to promote nitrile insertion. These conditions successfully transformed phenylacetylene into azasilacyclopentadiene 106b. For internal alkynes, copper(I) triflate catalyzed both silylene transfer to 3-hexyne as well as nitrile insertion to produce enamine 106k. The resulting enamines could be readily transformed into allylic amines (Scheme 7.30).99 Reduction of the imine and methyl carbamate protection of the resulting amine produced vinylsilane 107; protodesilylation of 107 afforded amine 108. Woerpel and Anderson also exploited the inherent nucleophilicity of the enamine to participate in conjugate addition and hydroboration/oxidation reactions; 1,4 addition of azasilacyclopentadiene 106a to methyl acrylate afforded 109 as a single tautomer, and hydroboration of enamine 106b followed by oxidation with H2O2 and NaOH produced allylic 1,2-aminoalcohol 110 as a single stereoisomer. The mildness of the di-tert-butylsilylene transfer was exploited in the synthesis of heteroatom-substituted silacyclopropenes (Scheme 7.31).101 Woerpel and Clark observed that di-tert-butylsilylene could be transferred to alkynes bearing alkoxy, amino, thio, or boryl functional groups. When silylene transfer was attempted to alkynes substituted with halides or sulfones, however, silacyclopropene formation was not observed.101 Instead, acetylenic silanes 113 were observed (Scheme 7.32). Treatment of silacyclopropene 112c (or 112d) with acetophenone and substoichiometric amounts of CuI also induced alkyne formation.
205
SILVER-MEDIATED TRANSFER OF DI-tert-BUTYLSILYLENE TO ACETYLENES O
t-Bu H t-Bu Si N
Ph
CSA, NaBH4
HO(t-Bu)2Si
HN
Ac2O, DMAP
H
O KOt-Bu, n-Bu4NF
Me Ph
Me Ph
55%
82%
Ph
HN
DMSO/THF
Ph 107
106b t-Bu H t-Bu Si N
H
Ph 108
Zn(OTf)2 (10 mol %)
H
H Ph 109
88%
Ph 106a t-Bu H t-Bu Si N
Ph
HO(t-Bu)2Si
1. BH 3·SMe2, THF
H
CO2Me
t-Bu H t-Bu Si N
CO2Me
NH2 Ph
2. NaOH, H 2O2
Ph
Ph 110
85%
106b
OH
Scheme 7.30. Formation of allylic amines from azasilacyclopentadienes. t-Bu Si 58
X
t-Bu X
Ag3PO4 (10 mol%)
R2
R2 112
111
EtO
t-Bu Si t-Bu
TIPSO
H 112a 89%
t-Bu Si t-Bu n-Bu
112b 86%
Ts N Bn
t-Bu Si t-Bu
t-Bu Si t-Bu
MeS
t-Bu Si t-Bu
pinB
n-C6H13
SiMe3 112c 95%
t-Bu Si t-Bu n-Bu 112e 77%
112d 90%
Scheme 7.31. Heteroatom-substituted silacyclopropene synthesis. t-Bu Si X
58 R2 111
Ts N Bn
t-Bu Si t-Bu SiMe3 112c
t-Bu
Ag3PO4 (10 mol%) PhH
t-Bu t-Bu Si X R2 113
111
X
f
Br
R Yield (%) n-Bu 79
g SO2Tol H
91
O Me
Ph
CuI, 50°C 94%
t-Bu t-Bu Si Ts N Bn 114
SiMe3
Scheme 7.32. Intramolecular rearrangement of silacyclopropenes to afford silyl-substituted alkynes.
206
SILVER-CATALYZED SILYLENE TRANSFER
Ag-catalyzed silyl-substituted alkyne formation mechanism Br
+
t-Bu t-Bu Si
n-Bu
t-Bu L t-Bu Si AgLn
L
111f
115
Br
t-Bu t-Bu Si L
AgLn n-Bu 116
AgLn
t-Bu t-Bu Si Br n-Bu
Br
n-Bu
L
117
113a
Cu-catalyzed silyl-substituted alkyne formation mechanism Ts N Bn
t-Bu Si t-Bu SiMe3 112c
CuI
t-Bu t-Bu L O O Si CuLn S N Tol Bn n-Bu 118
t-Bu t-Bu O Si L O S CuLn N Tol Bn n-Bu 119
t-Bu t-Bu Si O O S N CuLn
Tol
Bn 120
114 n-Bu L
Scheme 7.33. Potential mechanism for intramolecular rearrangement.
On the basis of crossover experiments, Woerpel and Clark attributed the formation of alkynes to an intramolecular rearrangement mechanism (Scheme 7.33).101 For silver-promoted alkyne formation, insertion of the acetylene into the Ag--Si bond produced 116. Then b elimination of bromide occurred more rapidly than silacyclopropene formation to generate silyl-substituted alkyne 117. Substitution of L with bromide then afforded the observed product (113a). The mechanism for copper-mediated alkyne formation was proposed to occur in a similar fashion.101 On formation of silacyclopropene 112c, copper-promoted transmetalation produced 118. The close proximity of the tosyl group to the silicon enabled a covalent interaction to produce 119,102 which transformed the sulfonamide into a good leaving group. Then b elimination generated alkyne 120, which still had the tosyl group covalently bound to the silicon. Rearrangement of 120 to the thermodynamically favored sulfone tautomer then afforded 114.103 An alternative mechanism involving the generation of a metal silylenoid from 112c was eliminated, as silylene transfer from silacyclopropene to another alkyne was never observed. In contrast to aforementioned silacyclopropenes, copper-catalyzed ring expansion was achievedwithsilyloxy-orboryl-substitutedsilacyclopropenes(Scheme7.34).101,104 Simply changing the ether substituent from an alkyl group to a silyl group enabled oxasilacyclopentene formation from 112b. These results reveal that the reactivity modes of the silacyclopropene are controlled by the identity of the substituent. Oxasilacyclopentenes were shown to be competent substrates for a scandium triflate–catalyzed Mukaiyama aldol process (Scheme 7.35).104 Exposure of silacyclopentene 121 and benzaldehyde to substoichiometric amounts of scandium triflate produced ketone 122 diastereoselectively.105 This ketone was proposed to form by addition of enolate 123, resulting from desilylation of 121,106 to benzaldehyde. A 1,3-Brook rearrangement then afforded 122 from 124.107 This ketone could be further functionalized through lithium aluminum hydride reduction followed by deprotection to afford triol 125 containing four contiguous stereocenters. Thus, the molecular complexity of silyloxyalkynes can be increased dramatically in just three operations.
207
TRANSFER OF DI-tert-BUTYLSILYLENE TO CARBONYL COMPOUNDS t-Bu Si 58
TIPSO
t-Bu
Ag3PO4 (10 mol%)
n-Bu
TIPSO
t-Bu Si t-Bu
n-BuCHO
t-Bu 58
pinB
O
t-Bu
Ag3PO4 (10 mol%)
n-Bu
n-Bu TIPSO 121 86%
112b
Si pinB
t-Bu Si t-Bu n-Bu 112e 77%
111e
t-Bu O Si t-Bu
CuI, 50°C
n-Bu
111b
n-Pr
Me
Ph
Cu(OTf)2 22°C
Me
t-Bu O Si t-Bu
Ph
n-Bu
pinB 122 63% (regioselectivity 96 : 4)
Scheme 7.34. Copper-mediated ring expansion to afford oxasilacyclopentenes. t-Bu t-Bu Si O
n-Pr + Ph
n-Bu OTIPS
t-Bu t-Bu Si O O
Sc(OTf)3 (20 – 25 mol%) CSA (1 – 3 mol%)
O H
2 : 1 C7H15 : CH2Cl2 –78 – 22°C
Ph n-Bu
121
1. LiAlH4, –98°C 90% (dr . 99 : 1) n-Pr
O 122
2. n-Bu4NF, THF 82%
OH
OH n-Pr
Ph
n-Bu OH 125
89% (dr 92 : 6 : 2) t-Bu t-Bu Si O
t-Bu t-Bu Si O
n-Pr
aldol
n-Bu O 123
(TfO)2Sc O Ph n-Bu O 124
n-Pr
1,3-Brook
Scheme 7.35. Mukaiyama aldol functionalization of oxasilacyclopentenes.
7.6 SILVER-MEDIATED TRANSFER OF DI-tert-BUTYLSILYLENE TO CARBONYL COMPOUNDS In addition to participating in [2 þ 1]-cycloaddition reactions, divalent reactive intermediates can form ylides in the presence of carbonyl or other Lewis basic functionalities.108 These ylides participate in cycloaddition or other pericyclic reactions to furnish products with dramatically increased complexity. While carbenes (or metal carbenoids) are well known to participate in these pericyclic reactions, silylenes, in contrast, were reported to react with aldehydes or ketones to form cyclic siloxanes109,110 or enoxysilanes.111,112 Reaction of silylene with an a,b-unsaturated ester was known to produce an oxasilacyclopentene.109,113,114 By forming a silver silylenoid reactive intermediate, Woerpel and coworkers enabled involvement of divalent silylenes in pericyclic reactions involving silacarbonyl ylides115 to afford synthetically useful products.82,116,117 While exploring the scope of silver-catalyzed silylene transfer to olefins, Woerpel and Calad observed that the electrophilic silver silylenoid species reacted preferentially with the enolizable ester group to produce vinylsilane 129 instead of a
208
SILVER-CATALYZED SILYLENE TRANSFER t-Bu
1. Si O
58
O
t-Bu t-Bu Si O
t-Bu
AgOTf (1 mol%)
i-Pr
O 127
126
t-Bu t-Bu Si OH Me O Me 129 45%
2. H 2O
i-Pr
t-Bu
t-Bu
Si
AgL3
t-Bu Si t-Bu O
L
115 ?
O 128
i-Pr
Scheme 7.36. Silver-catalyzed di-tert-butylsilylene transfer to an enolizable ester.
silacyclopropane (Scheme 7.36).82 They attributed vinylsilane formation to the intermediacy of oxasilacyclopropane 127, which might have formed in a stepwise fashion from 126 via silacarbonyl ylide 128.118 Woerpel and Calad tested for the formation of the silacarbonyl ylide by interrogating the behavior of the electrophilic silver silylenoid intermediate 115 toward a,b-unsaturated carbonyl compounds (Scheme 7.37).82 They hypothesized that formation of silacarbonyl ylide 131 might trigger a 6p-electrocyclization to form oxasilacyclopentene 132. As anticipated, exposure of cyclohexene silacyclopropane 58 to substoichiometric amounts of silver trifluoroacetate in the presence of a,b-unsaturated carbonyl compounds 130 produced oxasilacyclopentenes 132. The reaction tolerated a substitution at the a and/or b position and was general for both esters and ketones. The high diastereoselectivity observed by Woerpel and Calad in the hydrolysis of oxasilacyclopentene 132 (Scheme 7.38) prompted further evaluation of these intermediates in stereoselective synthetic transformations.82 Exposure of esters containing
t-Bu t-Bu Si t-Bu O
Si O
58
R1
R3 R2
t-Bu
AgO2CCF3 (1–5 mol%)
R1
t-Bu t-Bu O Si
t-Bu t-Bu O Si BnO 132a 98%
6π -electrocyclization
R1
R2 131
130
BnO
R3
t-Bu t-Bu O Si R3
t-Bu t-Bu O Si Me BnO
Me 132b 93%
Me 132c 99%
R2 132
t-Bu t-Bu O Si
t-Bu t-Bu O Si Me Ph
EtO 132d 98%
Me 132e 71%
t-Bu t-Bu O Si Ph Me 132f 58%
Scheme 7.37. Silver-catalyzed di-tert-butylsilylene transfer to a,b-unsaturated carbonyl compounds.
209
TRANSFER OF DI-tert-BUTYLSILYLENE TO CARBONYL COMPOUNDS t-Bu
1. Si O EtO
Me Me
t-Bu t-Bu O Si Me
t-Bu
58
AgO2CCF3 (1–5 mol%)
Me 2. H 2O
EtO2C
Si(t-Bu)2OH Me
EtO Me 132g
133 70% (dr 95 : 5)
t-Bu t-Bu O Si Me
t-Bu t-Bu O Si Me Ph O
130g t-Bu O
Si
O
Me Me
58
t-Bu
AgO2CCF3 (1 mol%)
Ph
O
Ph 130h
Ireland– Claisen rerrangement
Me
Me 135 95% single isomer
TS-134
Scheme 7.38. Exploitation of oxasilacyclopentenes in diastereoselective synthetic transformations.
an allyl substituent (e.g., 130h) to reaction conditions facilitated an Ireland–Claisen rearrangement after the initial electrocyclization. This process was highly diastereoselective; the reaction of ester 130h produced silalactone 135 with three contiguous stereocenters as a single isomer. Woerpel and Calad rationalized the stereochemistry of the [3,3]-sigmatropic rearrangement as arising from a chair-like transition state TS134, in which the allyl fragment approaches the oxasilacyclopentene from the face opposite the methyl substituent. The synthetic utility of this cascade reaction was underscored by the facile transformation of crotyl tiglate into 1,3-diol 137 (Scheme 7.39).82 Silver-catalyzed silylene transfer to crotyl tiglate produced silalactone 136 nearly quantitatively as a 97 : 3 mixture of diastereomers. Diol 137 was generated from silalactone after reduction of the lactone moiety with lithium aluminum hydride followed by oxidation of the C--Si bond.65,66,81
t-Bu
1.
O
Si
O
Me Me Me 130i
58
t-Bu
AgO2CCF3 (1 mol%)
t-Bu t-Bu O Si Me Me O Me
OH Me
2. LiAlH4, THF 3. KH, Ph(Me) 2COOH CsF, NMP/THF
Me Me
OH
87%
95%
136 (dr 97 : 3)
137
Scheme 7.39. Demonstration of the synthetic utility of electrocyclization/Ireland–Claisen rearrangement cascade reaction.
210
SILVER-CATALYZED SILYLENE TRANSFER
Ph
O
t-Bu Si t-Bu O
58
O
Me Me
130j
AgO2CCF3 (cat.)
O
t-Bu t-Bu O Si Me
unselective O
6π-electrocyclization
Me Me Ph
O O
Me Me
130k
AgO2CCF3 (cat.)
slow
O
[3,3]
Me
H
O Me TS-141
Ph 140 79% (dr 42 : 58)
t-Bu O Si t-Bu Me
Me
t-Bu O Si t-Bu
+ O
Me
Me 143
142 (dr 80 : 20)
Scheme 7.40. Scope and limitations of cascade electrocyclization/[3,3]-sigmatropic rearrangement.
The limitations of each step of the cascade reaction emerged on interrogation of esters with varied substitution patterns (Scheme 7.40).116 The reaction of ester 130j revealed that the facial selectivity of the 6p-electrocyclization could not be controlled. Exposure of (Z)-allyl ester 130k to reaction conditions produced a mixture of silalactones 142 and 143. The formation of 143 was attributed to hydrolysis of the initially formed oxasilacyclopentene. Observation of the hydrolysis product indicates that the rate of [3,3]-sigmatropic rearrangement is much slower for (Z)-allyl ethers. The authors attribute this rate difference to the axial placement of the methyl substituent in transition state TS-141, which incurs destabilizing steric interactions with the enoxysilane moiety. The unselective 6p-electrocyclization can be avoided with substrates that lack a b-olefin substituent (Scheme 7.41).116 When chiral, nonracemic versions of these substrates are employed, Woerpel and Calad demonstrated that the cascade reaction efficiently transfers the chiral information. While silver salts were competent catalysts for the diastereoselective formation of silalactone 144 from 130l, on scale-up, copper(II) triflate proved to be more efficient. The authors attribute the poorer performance by silver to product inhibition. In eight steps, silalactone 144 was transformed into the antibiotic ( þ )-epi-acetomycin, to further highlight the synthetic utility of the cascade process. t-Bu
t-Bu O Si t-Bu
Si i-Pr O Me
58
O 130l (>99% ee)
Me
t-Bu
AgO2CCF3 (cat.) or Cu(OTf)2 (5 mol%) 97%
O Me
i-Pr
Me
144 (dr 97 : 3)
Me
Ph 139
t-Bu t-Bu O Si Me Me
58
Me
[3,3]
Me
138
Me
t-Bu O Si t-Bu
O
O
8 steps Me
O O
Me Me
OAc
(+)-epi-acetomycin
Scheme 7.41. Efficient transfer of chirality in the cascade reaction.
211
TRANSFER OF DI-tert-BUTYLSILYLENE TO CARBONYL COMPOUNDS
t-Bu t-Bu O Si O i-Pr
Me
Me
O
144 major product
O Me
Me
149 not observed
t-Bu t-Bu Si O i-Pr
Me
t-Bu O t-Bu Si Me
O
O Me
Me
i-Pr 147 not observed
Me TS-146
i-Pr
Me TS-148
i-Pr
t-Bu O Si t-Bu
O Me
Me
Me
Me TS-145
t-Bu O t-Bu Si
t-Bu t-Bu Si O i-Pr
Me
t-Bu t-Bu O Si O i-Pr
t-Bu O Si t-Bu
O
Me
151 i-Pr minor product
TS-150
Figure 7.1. Origin of diastereoselectivity in Ireland–Claisen step of the cascade reaction.
The origin of diastereoselectivity in the cascade process was rationalized through a comparison of four different transition states (Fig. 7.1).116 The major diastereomer was posited to originate from transition state TS-145, which minimizes the developing diaxial-1,3 interactions through placement of the isopropyl group in a pseudoequatorial position. Because only (E)-olefins were obtained, the origin of the minor diastereomer was attributed to the boat-like transition state TS-150. Woerpel and Howard further extended the synthetic utility of this cascade reaction in the stereoselective synthesis of a-hydroxy acids from a-ketoesters (Scheme 7.42).119 In every case, the a-hydroxy acid was formed with excellent diastereoselectivity (dr 98 : 2). Their method was tolerant of a range of substitution patterns, although diminished yields were obtained as the size of the R1 substituent increased (compare 154a and 154c). In contrast, no decrease in yield was observed over a range of R2 substituents. t-Bu 1. Me Si t-Bu O
Me R2
O
R1
Si 89
O
CO2H
Et
HO
2. HF · pyr
HO
O
R2
153 CO2H
HO
CO2H
HO
CO2H
Ph
Ph
Ph
154a
154b
154c
84%
54%
47%
154 HO
CO2H
Ph
Ph
t-Bu
i-Pr
CO2H
R1
R2
152 HO
O
R1
AgOTs (10 mol%)
O
t-Bu
t-Bu
n-Bu
HO
CO2H
Ph OBn
154d
OTBDMS 154e
154f
72%
71%
75%
Scheme 7.42. Synthesis of a-hydroxy acids from a-ketoesters through a silver-promoted cascade reaction.
212
SILVER-CATALYZED SILYLENE TRANSFER
O O
Me
n-Bu
O
1. 89 AgOTs (10 mol%)
Me 152g
t-Bu O Me
O
R
Ph
O
1. 89 AgOTs (10 mol%)
OBn O
t-Bu
t-Bu t-Bu Si O H O n-Bu O Me Me TS-155
O
t-Bu
t-Bu
O
Si
O [3,3]
Ph
O
CO2H S
R
Me
n-Bu
n-Bu
154g 77% (>97% ee)
t-Bu O OBn 2. HF · pyr
R S
HO Me
Me
153g
Me
Si
2. HF · pyr
O R S
OBn Me TS-156
152h
O
[3,3]
t-Bu
Si
O
HO Me R
Me
CO2H R
S
OBn Ph
Ph
153h
154h 63% (80% ee)
Scheme 7.43. Di-tert-butylsilylene transfer to chiral, nonracemic substrates.
When chiral substrates were employed in the cascade reaction, good transfer of the chiral information was observed (Scheme 7.43).119 Excellent diastereoselectivity was obtained with 152g, containing an allylic methyl group; and even 152h, which contained a benzyl ether substituent that would lie outside the chairlike transition state (TS-156), controlled the facial selectivity of the [3,3]-sigmatropic rearrangement. In addition to a-ketoesters, a-iminoesters were successful as substrates, allowing the synthesis of a-amino acid derivatives [equation (7.9)].119 Exposure of 157 to gemdimethylsilacyclopropane 89 and substoichiometric amounts of silver tosylate produced azasilalactone 158. In contrast to the above–mentioned synthesis of 154, hydrolysis of 158 was not observed on aqueous extraction. The authors attributed the enhanced robustness of 158 to the steric bulk of the anisidine group. t-Bu Me
MeO
Me N O
H O 157
Ph
Si t-Bu 89
AgOTs (10 mol%)
MeO t-Bu t-Bu N Si O
ð7:9Þ
O Ph 158
In the absence of an internal olefin, silver-catalyzed di-tert-butylsilylene transfer to aldehydes did not produce an oxasilacyclopropane (Scheme 7.44).117 Instead, the reaction produced a dioxosilacyclopentane 161. Woerpel and Bourque attributed the formation of 161 to the intermediacy of silacarbonyl ylide 159. Subsequent intermolecular1,3-dipolarcycloadditionwithasecondmoleculeofaldehydewouldgeneratethe observed product. On the basis of this mechanistic hypothesis, Woerpel and Bourque sought to trap the silacarbonyl ylide intermediate with an acetylenic dipolarophile to generate an oxasilacyclopentene. To achieve the formation of 162, the dipolarophile would have to react with 159 faster than the aldehyde did. To identify potential dipolarophiles, a series of competition experiments were performed (Scheme 7.45).117 While inclusion of terminal acetylenes resulted in preferential silacyclopropene 99c or dioxosilacyclopentane 161 formation, the desired oxasilacylopentene (163) was obtained when diethylacetylene dicarboxylatewas added
213
TRANSFER OF DI-tert-BUTYLSILYLENE TO CARBONYL COMPOUNDS t- Bu Si O
58
Ph
H
AgOTf (1 mol%) PhMe, 25°C
t -Bu
t -Bu
t- Bu O Ph
Si
O
PhCHO
t -Bu
t -Bu
Si O
k1
Ph
H
Ph 161 Single diastereomer
159
R t -Bu t -Bu
R
t -Bu Si O Ph
O k2
H
R
Ph
Find R such that k 2 >> k 1
160
t -Bu
Si
R 162
Scheme 7.44. Silver-promoted 1,3-dipolar cycloaddition reactions.
to the reaction mixture. Substitution of the carboethoxy groups with methyl and phenyl groups, unfortunately, resulted in dioxosilacyclopentane 161 formation. The cycloaddition reaction with diethyl acetylenedicarboxylate tolerated a range of carbonyl compounds (Scheme 7.46). Facile reaction was obtained with aldehydes, t-Bu t-Bu O Si O
Me
Ph
H
AgOTf (1 mol%) PhMe, 25°C
Ph Ph 161
AgOTf (1 mol%) PhMe, 25°C
82%
t-Bu t-Bu O Si CO 2Et
+ Ph
Ph 99c
89%
t-Bu
O COEt 2
t-Bu Si t-Bu
Ph
Si
H
t-Bu 58 H
EtO 2C
Ph CO 2Et 163
t-Bu t-Bu O Si O
COEt 2
AgOTf (1 mol%) PhMe, 25°C
AgOTf (1 mol%) PhMe, 25°C
85%
79%
Ph Ph 161
Scheme 7.45. Intermolecular competition reactions that examine the relative rates of cycloaddition and cyclization. COEt2
t-Bu Si 58
O R1
H Me
R2
AgOTf (1 mol%) PhMe, 25°C
t-Bu t-Bu O Si CO2Et CO2Et 165a 82%
t-Bu t-Bu O Si CO2Et
H EtO
t-Bu Si O t-Bu
t-Bu
CO2Et 165b 64%
R1
EtO2C R1
R2
R2 164 t-Bu t-Bu O Si CO2Et
H Bn N CO2Et Me 165c 68%
t-Bu t-Bu O Si CO2Et
t-Bu t-Bu O Si CO2Et
Ph Me
CO2Et 165
MeO
t-Bu t-Bu O Si CO2Et CO2Et
CO2Et 165d 64%
165e 74%
Scheme 7.46. Scope of silver-catalyzed silacarbonyl ylide cycloaddition.
214
SILVER-CATALYZED SILYLENE TRANSFER
formates, ketones, and formamides. The slower rate of reaction of formamide required heating and prolonged time (50 C, 12 h). 1,3-Dipolar cycloaddition was even obtained in the reaction of methyl acrylate to produce 165e, although excess ester and silacyclopropane was required to compete with silylene transfer to the olefinic moiety. 7.7 SILVER-MEDIATED TRANSFER OF DI-tert-BUTYLSILYLENE TO IMINES The nitrogen analog of oxasilacyclopropane, azasilacyclopropane (or silaziridine), has been postulated as a potential reactive intermediate in photochemical- or thermal processes involving silylenes. These strained N-heterocycles can be accessed from the reaction of isocyanides and photochemically generated silenes, but this method is limited to the formation of methylene silaaziridines 168 (Scheme 7.47).120–122 Isolation of these N-heterocycles proved challenging, as they rapidly inserted another molecule of the isocyanide starting material on warming to room temperature.121 Woerpel and coworkers believed that the mild nature of silver-catalyzed di-tertbutylsilylene transfer might be exploited to provide general access to these strained Nheterocycles (170) and enable a detailed study of their reactivity. Toward this end, Woerpel and Nevarez examined the possibility of di-tertbutylsilylene transfer from cyclohexene silacyclopropane 58 to imine 169a (Scheme 7.48).123 Thermolysis produced a mixture of silaaziridine 170a and an imine–dimer byproduct (171). The results by Brook and coworkers120–122 suggested that if the temperature of silylene transfer were lowered, isolation of 170a without formation of byproduct 171 would be possible. As anticipated, exposure of cyclohexene silacyclopropane 58 to imine 169a in the presence of substoichiometric amounts of silver triflate produced only 170a. This silaazridine could be purified by bulb-to-bulb distillation to afford the product in 80% yield. Copper salts required
t-Bu
O (Me3Si)3Si
360 nm
t-Bu
C6D6, 10°C
Me3Si
R–NC
Me3Si
OSiMe3 Si SiMe3
166
Me3Si
OSiMe3 168 Isolable only when R = t-Bu or Mes
t-Bu Si N
R1 R2 169
58
t-Bu
AgOTf (cat.)
?
R N
t-Bu
167
R3
Si
t-Bu R3 N Si t-Bu R 1 R2 170
Scheme 7.47. General method for accessing silaaziridines.
215
SILVER-MEDIATED TRANSFER OF DI-tert-BUTYLSILYLENE TO IMINES t-Bu Si Bn H
58
N
t-Bu Bn N Si t-Bu
t-Bu
conditions
Ph
Bn +
t-Bu t-Bu N Si N Bn
Ph
Ph
100°C
72
AgOTf (cat.)
90
63 : 37 100 : 0
Ph 171
170a
169a
Yield (%) 170a : 171
Conditions
Scheme 7.48. Di-tert-butylsilylene transfer to imines.
higher temperatures to promote silylene transfer from 169a to 170a. Yields were significantly diminished using copper catalysts. Silver-catalyzed di-tert-butylsilylene transfer to imines proved general (Scheme 7.49).123 Exposure of alkyl- or arylimines with N-benzyl or N-aryl groups to cyclohexene silacyclopropane 58 in the presence of 1 mol% of silver triflate produced silaaziridines 170. Even ketimine 169f was tolerated as a substrate. In contrast to the results presented above, silver-mediated silylene transfer to a,b-unsaturated imines or imines with N-carbamate groups did not produce silaaziridines(Scheme7.50).123 Thesesubstratesinsteadformedenamine 173oroxazoline175. The authors attributed these products to a 6p-electrocyclization of ylide 172 and 174. While enamine 173 and oxazoline 175 were observed using 1 H NMR spectroscopy, they could not be isolated. Instead, dimeric 177 and silaoxazolidinone 179 were obtained (Scheme 7.51).123 The formation of azasilacyclopentane 177 was rationalt-Bu Si R3
t-Bu Bn N Si t-Bu
58
R1 R2 169
AgOTf (1 mol%)
MeO
t-Bu N Si t-Bu
i-Pr
Me
t-Bu N Si t-Bu
Ph
88% (84%)
F3C
t-Bu N Si t-Bu
Ph
170c 76%
170b
t-Bu R3 N Si t-Bu R1 R2 170
t-Bu
N
Ph
t-Bu Bn N Si t-Bu Et i-Pr 170f 74%
170e 70%
170d 74% (73%)
Scheme 7.49. Scope of Ag-catalyzed di-tert-butylsilylene transfer to aldimines and ketimines. t-Bu Si
Me
58 N
t-Bu
Me
AgOTf (1 mol%)
H 169g
Me
H 169h
t-Bu Si t-Bu O
58
N
AgOTf (10 mol%)
Ph
Me
H 172
O t-BuO
t-Bu Si t-Bu N
t-BuO
N 174
Ph
Me
6π-electrocyclization
6π-electrocyclization
t-Bu t-Bu N Si Me H 173 80% t-Bu t-Bu O Si Ph N t-BuO 175 46%
Scheme 7.50. Silver-catalyzed di-tert-butylsilylene transfer to imines.
216 H 176
H
H
N
Me
t-Bu t-Bu O Si Ph N t-BuO Ph 175
H+
Boc
H
173
t-Bu t-Bu
Si
N Ar
H N Ar
t-Bu t-Bu O Si Ph N t-BuO Ph HN Boc 178 H2O
Si Me t-Bu t-Bu 177
H
t-Bu t-Bu Si N Me H
Me
t-Bu t-Bu O Si Ph N O Ph HN Boc 179 40%
Me 177 40%
HH
t-Bu t-Bu Si N Me
H N Ar Si t-Bu t-Bu
Me
Scheme 7.51. Decomposition of silaaziridines derived from a,b-unsaturated imines or carbamate-protected imines.
H 173 Ar = 4-MeC6H4
t-Bu t-Bu Si N Ar Me
t-Bu t-Bu Si N Ar Me
217
SILVER-MEDIATED TRANSFER OF DI-tert-BUTYLSILYLENE TO IMINES H
t-Bu Bn
N Si
t-Bu H
i-Pr Ph 183 90% t-Bu Bn
N Si
t-Bu
Ph
PhCHO AgOTf (1 mol %)
Pd(PPh3)4 (1 mol %)
t-Bu Bn N Si t-Bu H
Me2N
t-BuNC
Pd(PPh3)4 (1 mol %)
i-Pr 182 87%
Ph N Bn 180 >95% (dr 91: 9) i-Pr
i-Pr 169b
H
t-Bu t-Bu Si O
NMe2
Bn
t-Bu N Si t-Bu
i-Pr
Nt-Bu 181 >95%
Scheme 7.52. Synthetic utility of silaziridines.
ized by the authors as occurring through a series of nucleophilic additions to iminium ions. These additions were diastereoselective as a single isomer of 177 was isolated. A similar addition mechanism was postulated to account for carbamate 179 formation; nucleophilic addition of oxazoline 175 to imine 169h would produce 179 after hydrolysis. Woerpel and Nevarez demonstrated the synthetic potential of silaaziridines by selective insertion reactions (Scheme 7.52).123 Silver-catalyzed aldehyde insertion into the Si--N bond of 169b produced the N,O-cyclic acetal 180 as the cis isomer. In contrast to this process, insertion of tert-butyl isocyanide occurred into the weaker C--Si bond to afford imine 181. The authors rationalized the chemoselectivity for these two processes on the basis of Pearsons hard–soft acid–base theory;124–126 the more ionic Si--N bond reacted with harder benzaldehyde electrophile, whereas the more covalent Si--C bond reacted with the softer isocyanide. Azasilacyclopentenes could also be produced from silaaziridines through the use of a palladium(0) catalyst (Scheme 7.52).123 This process inserted the terminal acetylene into the weaker Si--C bond to produce 182 and 183 as single regioisomers. The authors asserted that the regioselectivity of the insertion was controlled through minimization of steric interactions between the tert-butyl groups and the alkyne substituent. A range of functional groups were tolerated in this process, including silyl ethers, silanes, and tertiary amines. Protodesilylation reveals that azasilacyclopentenes are masked allylic amines: t-Bu Bn
N Si
t-Bu
i-Pr Ph 183
H
KOt-Bu n-Bu4NF 72%
Bn
H N
Ph
ð7:10Þ
i-Pr 184
In contrast to the formation of an N,O-acetal in the reaction of N-benzylsilaaziridine 169b with isobutyraldehyde, the reaction of N-arylsilaaziridines 169g
218
SILVER-CATALYZED SILYLENE TRANSFER
Me
t-Bu N Si t-Bu
t-Bu O t-Bu Si N
PhCHO 23°C >95%
i-Pr
i-Pr 185 single isomer
169g
Me
Ph H
t-Bu N Si t-Bu
t-Bu t-Bu Ar Si N O
PhCHO
Ph
23°C 89%
Ph
+
H
169d
Me
t-Bu t-Bu Ar N Si O Ph
Ph
63 : 37 186 187 single isomer single isomer
Scheme 7.53. Dearomatization reactions of N-arylsilaziridines and benzaldehyde.
and 169d with benzaldehyde produced new C--C bonds through dearomatization processes (Scheme 7.53).127 The reaction of silaziridine 169g with benzaldehyde produced aziridine 185, which contains four contiguous stereocenters, as a single isomer. An alternative dearomatization reaction occurred with silaaziridine 169d; exposure to benzaldehyde resulted in the break of the aromaticity of the phenyl substituent instead of the N-aryl group to produce 186 as a single isomer. Insertion of benzaldehyde into 169d to form 187 was competitive with this dearomatization mechanism. Woerpel and Nevarez postulated a stepwise mechanism for the formation of 185 (Scheme 7.54).127 Coordination of benzaldehyde to the Lewis acidic silicon atom of 169g86 triggers an attack by the pendent N-aryl group to form zwitterion 189. Isomerization of 189 to azomethine ylide 190128 enabled a 4p-electrocyclization to produce aziridine 185. The thermodynamic driving force for this reaction was attributed to the relief of the silaaziridines ring strain. Me
Ph 169g
PhCHO
Me
H
O N Si t-Bu t-Bu i-Pr 189
i-Pr 188
Ph
H Ph
O
N Si t-Bu t-Bu
t-Bu Si N t-Bu
Me H Ph
Me
PhCHO
t-Bu t-Bu Si N Tol O Ph
H
O N Si t-Bu t-Bu i-Pr 190
4π -electrocyclization
t-Bu
H Si
t-Bu
H Ph
i-Pr
O N Si t-Bu t-Bu
185 single isomer
Tol N H H
Me
[3,3]
O
t-Bu t-Bu Ar Si N O H H
Ph
Ph 169d
191
TS-192
Scheme 7.54. Potential mechanisms for dearomatization.
186 single isomer
219
SILVER-MEDIATED DI-tert-BUTYLSILYLENE INSERTION INTO C--O BONDS
The authors suggested a slightly different mechanism for the formation of 186.127 Instead of triggering an electrophilic aromatic substitution reaction, initial coordination of benzaldehyde to 169d enabled a [3,3]-sigmatropic rearrangement through the boat-like transition state TS-192129 to form the observed product with the requisite cis-stereochemical arrangement of hydrogens. 7.8 SILVER-MEDIATED DI-tert-BUTYLSILYLENE INSERTION INTO C--O BONDS Transition-metal-catalyzed [2,3]-sigmatropic rearrangement of an oxonium ylide derived from an allylic ether is a well established method for the synthesis of new C--C bonds in a stereoselective fashion (Scheme 7.55).130–133 Competitive [1,2]-rearrangements, however, do occur if the rate of the [2,3] rearrangement is slow (see 194–196).134,135 Since silver-catalyzed di-tert-butylsilylene transfer to Lewis basic carbonyls was established to afford ylides, Woerpel and coworkers investigated the potential of silylene transfer to allylic ethers 197.136 The resulting oxonium ylide (198) could rearrange through a [2,3]-sila-Wittig rearrangement to afford allylsilane 199 or undergo a [1,2]-Stevens rearrangement to afford allylsilane 200. The latter rearrangement would constitute a formal insertion of silylene into a C--O bond. While investigating trisubstituted allylic ethers, Woerpel and coworkers observed insertion of silylene into the allylic C--O bond (Scheme 7.56).136 The product formation was dependent on the identity of the catalyst. Silver trifluoroacetate promoted the formation of allylic disilane 201, whereas copper(I) triflate O CO2Me
[2,3] sigmatropic rearrangement
O
O
N2
CO2Me
Rh2(O2CR)4 (2 mol%)
Me
Ref. 134
O Me 195 52% (dr 49 : 51)
O
Me CO2Me [Rh] Me
O
O
Me
Me
194
193
CO2Me
[1,2] Stevens rearrangement
O Me
R2 R 3
t-Bu Si R2 R1O
58 R3
t-Bu
AgX (cat.)
Me 196 11%
R1
t-Bu Si t-Bu O R2 R3
197 198
?
Si(t-Bu)2OR1 199 + R2 R1O(t-Bu)2Si 200
R3
Scheme 7.55. [2,3]-Sigmatropic rearrangement versus [1,2]-rearrangements of oxonium ylides.
220
SILVER-CATALYZED SILYLENE TRANSFER 58
Me BnO
AgO2CCF3 (0.5 mol%) 81%
t-Bu t-Bu Me Si Si Me t-Bu t-Bu 201
BnO
Me 197a
Me
Me
197b
AgO2CCF3 (0.5 mol%) 74%
204
AgO2CCF3 (0.5 mol%) 65%
BnO
Me
Me
58
BnO
Si t-Bu t-Bu
(CuOTf)2·PhMe (5 mol%) 68%
58
Me
BnO
202
Me
58
BnO
Si t-Bu t-Bu 203
Me Me
t-Bu t-Bu Si O Ph Si t-Bu t-Bu 205
Scheme 7.56. Transition metal-catalyzed insertion of di-tert-butylsilylene into C--O bonds.
catalyzed the formation of allylic silane 202. The copper-catalyzed reaction was not as efficient, requiring higher catalyst loading and increased reaction time to afford 202. The identity of the product also depended on the steric environment of the allylic ether, silver-catalyzed silylene transfer to secondary allylic ethers afforded allylic silane 203, the product of a single silylene insertion. The insertion of a silylene into the benzylic C–O bond of 204 was also observed to afford disilane 205. Insight into the mechanism for the formation of allylic disilane 201 and silane 202 was obtained from a series of control experiments (Scheme 7.57). Submission of allylic silane 202 to reaction conditions did not produce disilane 201. Woerpel and coworkers interpreted this result to indicate that disilane formation does not occur through subsequent silylene insertions. Crossover experiments established that no dissociation of the alkoxy group occurs during the reaction to suggest that silylene insertion is intramolecular. Sequential insertion control experiment
Me BnO
Si t-Bu t-Bu 202
58
Me
AgO2CCF3 (0.5 mol%)
Me t-Bu t-Bu Si BnO Si Me t-Bu t-Bu 201
Crossover experiment
Me BnO
Me
Me
BnO
Me 197b 58
+ Et Tol
O 197c
Me Me
(CuOTf)2·PhMe (5 mol%)
Me
Si Me t-Bu t-Bu 206 81% + Et Me Tol O Si Me t-Bu t-Bu 207 75%
Scheme 7.57. Mechanistic experiments.
221
CONCLUSION Si(t-Bu)2OBn Me
OBn Me
58 (CuOTf)2·PhMe (5 mol %)
i-Pr
i-Pr
58 (CuOTf)2·PhMe (5 mol %)
210 (>95% ee)
79%
209 (dr 50 : 50)
Si(t-Bu)2OBn Me +
OBn Me
i-Pr
76%
208 (dr 50 : 50)
i-Pr
Me i-Pr
211
68 : 32
Si(t-Bu)2OBn 212
Scheme 7.58. Examination of the stereochemical course of the reaction.
Interrogation of the stereochemical course of the mechanism was obtained through submission of allylic ethers 208 and 210 (>95% ee) to reaction conditions (Scheme 7.58). The reaction of a 1 : 1 mixture of allylic ether 208 produced the allylic silane as a 1 : 1 mixture of diastereomers. Exposure of 210 to substoichiometric amounts of copper(I) triflate and cyclohexene silacyclopropane produced cis-substituted allyl silanes 211 and 212 to reveal that C--Si bond formation occurs on the same face as the C--O bond that is cleaved. The loss of enantiopurity, however, indicates that the rate of allylic transposition is competitive with the insertion process. Woerpel and coworkers interpreted the results of these mechanistic experiments as evidence that the insertion of silylene into the C--O bond occurs through a [1,2]Stevens rearrangement of oxonium ylide 198 and that a competitive [2,3]-sigmatropic rearrangement of 198 could account for allylic transposition.
7.9
CONCLUSION
Woerpel and coworkers discovered that silver salts promote di-tert-butylsilylene extrusion from a sacrificial silacyclopropane (58 or 89, Scheme 7.59). The resulting silver silylenoid reacts with a variety of different functional groups to afford three- or five-membered heterocycles. These silacycles are easily transformed into a variety of important synthetic intermediates, including polypropionates, allylic ethers, allylic amines, and a-hydroxy acids. The synthetic utility of silver-catalyzed silylene transfer was showcased in the total syntheses of 10 -epi-stegobinone and ( þ )-epiacetomycin. The mild nature of silver-catalyzed silylene transfer allows the isolation of particularly reactive monosubstituted silacyclopropenes and silaaziridines to enable their use in stereoselective methods for organic synthesis. The use of silver to mediate silylene transfer reactions has resulted in the development of useful organic transformations that exploit the inherent reactivity of silanes to produce important synthetic intermediates.
222
SILVER-CATALYZED SILYLENE TRANSFER
t-Bu t-Bu Me Si BnO Si Me t-Bu t-Bu
allylic ethers
t-Bu Si
t-Bu
Ph Me
58
O
Si
AgX (cat.) alkynes
Me
compounds
t-Bu t-Bu Si
t-Bu
alkynyl
esters
α,β-unsat. carbonyl compounds
Ph
t-Bu t-Bu O Si Me Ph O
bromide
Me
imines
t-Bu t-Bu O Si Me
BnO Me
t-Bu t-Bu Si Br n-Bu
homoallylic ethers
α,β-unsat.
carbonyl compounds
t-Bu Si t-Bu
AgL3
α-keto-
Me
Me
OTIPS
X
O OBn
O
Me
89
carbonyl
CO2Et
t-Bu
Me
t-Bu Si t-Bu
Me
Si t-Bu
Me
or
t-Bu t-Bu t-Bu O Si CO2Et
OTIPS
olefins
t-Bu t-Bu Me Si OMe (t-Bu)2HSi
t-Bu Bn N Si t-Bu i-Pr
i-Pr
Scheme 7.59. Silver-catalyzed di-tert-butylsilylene transfer.
REFERENCES 1. Ye, T.; McKervey, M. A., Chem. Rev. 1994, 94, 1091–1160. 2. Doyle, M. P.; McKervey, M. A.; Ye, T., Modern Catalytic Methods for Organic Synthesis with Diazo Compounds: From Cyclopropanes to Ylides, Wiley, New York, 1998. 3. Kirmse, W., Eur. J. Org. Chem. 2002, 2193–2256. 4. Li, Z.; He, C., Eur. J. Org. Chem. 2006, 2006, 4313–4322. 5. For silver-promoted C--H bond functionalization or cyclopropanation reactions, see (a) Burgess, K.; Lim, H.-J.; Porte, A. M.; Sulikowski, G. A., Angew. Chem., Int. Ed. Engl. 1996, 35, 220–222; (b) Dias, H. V. R.; Browning, R. G.; Polach, S. A.; Diyabalanage, H. V. K.; Lovely, C. J., J. Am. Chem. Soc. 2003, 125, 9270–9271; (c) Dias, H. V. R.; Browning, R. G.; Richey, S. A.; Lovely, C. J., Organometallics 2004, 23, 1200–1202; (d) Lovely, C. J.; Browning, R. G.; Badarinarayana, V.; Dias, H. V. R., Tetrahedron Lett. 2005, 46, 2453–2455; (e) Urbano, J.; Belderrain, T. R.; Nicasio, M. C.; Trofimenko, S.; Dıaz-Requejo, M. M.; Perez, P. J., Organometallics 2005, 24, 1528–1532; (f) Thompson, J. L.; Davies, H. M. L., J. Am. Chem. Soc. 2007, 129, 6090–6091; (g) Dias, H. V. R.; Lovely, C. J., Chem. Rev. 2008, 108, 3223–3238. 6. Wolff, L.; Justus Liebigs Ann. Chem. 1902, 325, 129–195. 7. Bachmann, W. E.; Struve, W. S., Org. React. 1946, 1, 38. 8. Newman, M. S.; Beal, P. F., J. Am. Chem. Soc. 1950, 72, 5163–5165. 9. Lee, V.; Newman, M. S., Org. Synth. 1970, 50, 77.
REFERENCES
223
10. For leading mechanistic studies on silver-mediated Wolff rearrangement, see (a) Yukawa, Y.; Tsuno, Y.; Ibata, T., Bull. Chem. Soc. Jpn. 1967, 40, 2613–2617; (b) Yukawa, Y.; Tsuno, Y.; Ibata, T., Bull. Chem. Soc. Jpn. 1967, 40, 2618–2623; (c) Takebayashi, M.; Ibata, T., Bull. Chem. Soc. Jpn. 1968, 41, 1700–1707; (d) Duggleby, P. M.; Holt, G.; Hope, M. A.; Lewis, A., J. Chem. Soc. Perkin Trans. 1972, 1, 3020–3024; (e) Agosta, W. C.; Wolff, S., J. Org. Chem. 1975, 40, 1027–1030; (f) Sudrik, S. G.; Maddanimath, T.; Chaki, N. K.; Chavan, S. P.; Chavan, S. P.; Sonawane, H. R.; Vijayamohanan, K., Org. Lett. 2003, 5, 2355–2358; (g) Julian, R. R.; May, J. A.; Stoltz, B. M.; Beauchamp, J. L., J. Am. Chem. Soc. 2003, 125, 4478–4486. ´ irakovic, J.; Driver, T. G.; Woerpel, K. A., J. Am. Chem. Soc. 2002, 124, 9370–9371. 11. C 12. For reviews, see (a) West, R., Pure Appl. Chem. 1984, 56, 163–173; (b) Raabe, G.; Michl, J., Chem. Rev. 1985, 85, 419–509; (c) West, R., Angew. Chem. Int. Ed. Engl. 1987, 26, 1201–1211; (d) Barrau, J.; Escudie, J.; Satge, J., Chem. Rev. 1990, 90, 283–319; (e) Jutzi, P., J. Organomet. Chem. 1990, 400, 1–17; (f) Lappert, M. F.; Rowe, R. S., Coord. Chem. Rev. 1990, 100, 267–292; (g) Satge, J., J. Organomet. Chem. 1990, 400, 121–147; (h) Tsumuraya, T.; Batcheller, S. A.; Masamune, S., Angew. Chem. Int. Ed. Engl. 1991, 30, 902–930; (i) Weidenbruch, M., Coord. Chem. Rev. 1994, 130, 275–300; (j) Weidenbruch, M., Chem. Rev. 1995, 95, 1479–1493; (k) Brook, A. G.; Brook, M., Adv. Organomet. Chem. 1996, 39, 71–158; (l) Hemme, I.; Klingebiel, U., Adv. Organometal. Chem. 1996, 39, 159–192; (m) Driess, M., Adv. Organomet. Chem. 1996, 39, 193–229; (n) Okazaki, R.; West, R., Adv. Organomet. Chem. 1996, 39, 231–273; (o) Baines, K. N.; Stibbs, W. G., Adv. Organomet. Chem. 1996, 39, 275–324; (p) Power, P. P., J. Chem. Soc. Dalton Trans. 1998, 2939–2951; (q) Weidenbruch, M., Eur. J. Inorg. Chem. 1999, 373–381; (r) Tokitoh, N.; Okazaki, R., Coord. Chem. Rev. 2000, 210, 251–277; (s) Haaf, M.; Schmedake, T. A.; West, R., Acc. Chem. Res. 2000, 33, 704–714; (t) Gaspar, P. P., Sci. Synth. 2002, 135–158; (u) Ottosson, H.; Steel, P. G., Chem. Eur. J. 2006, 12, 1576–1585. 13. Waterman, R.; Hayes, P. G.; Tilley, T. D., Acc. Chem. Res. 2007, 40, 712–719. 14. Lickiss, P. D., Chem. Soc. Rev. 1992, 271–279. 15. Brown-Wensley, K. A., Organometallics 1987, 6, 1590–1591. 16. Hengge, E.; Weinberger, M., J. Organomet. Chem. 1993, 443, 167–173. 17. Ojima, I.; Inaba, S.-I.; Kogure, T.; Nagai, Y., J. Organomet. Chem. 1973, 55, C7–C8. 18. Zybill, C.; M€uller, G., Angew. Chem. Int. Ed. Engl. 1987, 26, 669–670. 19. Ueno, K.; Tobita, H.; Shimoi, M.; Ogino, H., J. Am. Chem. Soc. 1988, 110, 4092–4093. 20. Tobita, H.; Ueno, K.; Shimoi, M.; Ogino, H., J. Am. Chem. Soc. 1990, 112, 3415–3420. 21. Sharma, H. K.; Pannell, K. H., Organometallics 2001, 20, 7–9. 22. Pannell, K. H.; Sharma, H. K.; Kapoor, R. N.; Cervantes-Lee, F., J. Am. Chem. Soc. 1997, 119, 9315–9316. 23. Ueno, K.; Masuko, A.; Ogino, H., Organometallics 1999, 18, 2694–2699. 24. Zhang, Y.; Cervantes-Lee, F.; Pannell, K. H., J. Organomet. Chem. 2001, 634, 102–108. 25. Zhang, Y.; Cervantes-Lee, F.; Pannell, K. H., Organometallics 2003, 22, 2517–2524. 26. Zhang, Y.; Pannell, K. H., Organometallics 2003, 22, 1766–1770. 27. Zhang, Y.; Cervantes-Lee, F.; Pannell, K. H., Organometallics 2002, 21, 5859–5867. 28. Zhang, Y.; Pannell, K. H., Organometallics 2002, 21, 503–510. 29. Sato, T.; Okazaki, M.; Tobita, H.; Ogino, H., J. Organomet. Chem. 2003, 669, 189–199. 30. Tobita, H.; Sato, T.; Okazaki, M.; Ogino, H., J. Organomet. Chem. 2000, 611, 314–322.
224
SILVER-CATALYZED SILYLENE TRANSFER
31. 32. 33. 34.
Ueno, K.; Masuko, A.; Ogino, H., Organometallics 1997, 16, 5023–5026. Ueno, K.; Sakai, M.; Ogino, H., Organometallics 1998, 17, 2138–2140. Ueno, K.; Asami, S.; Watanabe, N.; Ogino, H., Organometallics 2002, 21, 1326–1328. Feldman, J. D.; Mitchell, G. P.; Nolte, J.-O.; Tilley, T. D., J. Am. Chem. Soc. 1998, 120, 11184–11185. Grumbine, S. K.; Mitchell, G. P.; Straus, D. A.; Tilley, T. D.; Rheingold, A. L., Organometallics 1998, 17, 5607–5619. Peters, J. C.; Feldman, J. D.; Tilley, T. D., J. Am. Chem. Soc. 1999, 121, 9871–9872. Feldman, J. D.; Peters, J. C.; Tilley, T. D., Organometallics 2002, 21, 4065–4075. Straus, D. A.; Grumbine, S. D.; Tilley, T. D., J. Am. Chem. Soc. 1990, 112, 7801–7802. Mitchell, G. P.; Tilley, T. D., J. Am. Chem. Soc. 1998, 120, 7635–7636. West, R.; Denk, M. K., Pure Appl. Chem. 1996, 68, 785–788. Schmedake, T. A.; Haaf, M.; Paradise, B. J.; Millevolte, A. J.; West, R., J. Organomet. Chem. 2001, 636, 17–25. Denk, M.; Hayashi, R. K.; West, R., J. Chem. Soc. Chem. Commun. 1994, 33–34. Schmedake, T. A.; Haaf, M.; Paradise, B. J.; Powell, D.; West, R., Organometallics 2000, 19, 3263–3265. Theil, M.; Jutzi, P.; Neumann, B.; Stammler, A.; Stammler, H.-G., J. Organomet. Chem. 2002, 662, 34–42. Gehrhus, B.; Hitchcock, P. B.; Lappert, M. F.; Maciejewski, H., Organometallics 1998, 17, 5599–5601. Amoroso, D.; Haaf, M.; Yap, G. P. A.; West, R.; Fogg, D. E., Organometallics 2002, 21, 534–540. Woo, L. K.; Smith, D. A.; Young, V. G., Organometallics 1991, 10, 3977–3982. (a) Crabtree, R. H.; in The Organometallic Chemistry of the Transition Metals, Wiley, New York, 1994, pp. 270–310; (b) Bernasconi, C. F., Chem. Soc. Rev. 1997, 26, 299–307; (c) de Meijere, A.; Schirmer, H.; Duetsch, M., Angew. Chem. Int. Ed. 2000, 39, 3964–4002. Probst, R.; Leis, C.; Gamper, S.; Herdtweck, E.; Zybill, C.; Auner, N., Angew. Chem. Int. Ed. Engl. 1991, 30, 1132–1135. Handwerker, H.; Leis, C.; Probst, R.; Bissinger, P.; Grohmann, A.; Kiprof, P.; Herdtweck, E.; Bluemel, J.; Auner, N.; Zybill, C., Organometallics 1993, 12, 2162–2176. Marquez, A.; Fernandez Sanz, J., J. Am. Chem. Soc. 1992, 114, 2903–2909. Cundari, T. R.; Gordon, M. S., Organometallics 1992, 11, 3122–3129. Leis, C.; Wilkinson, D. L.; Handwerker, H.; Zybill, C.; Mueller, G., Organometallics 1992, 11, 514–529. Yamashita, H.; Tanaka, M.; Goto, M., Organometallics 1992, 11, 3227–3232. Glaser, P. B.; Tilley, T. D., J. Am. Chem. Soc. 2003, 125, 13640–13641. Calimano, E.; Tilley, T. D., J. Am. Chem. Soc. 2008, 130, 9226–9227. Klei, S. R.; Tilley, T. D.; Bergman, R. G., Organometallics 2002, 21, 3376–3387. Okamoto, K.; Hayashi, T., Org. Lett. 2007, 9, 5067–5069. Okamoto, K.; Hayashi, T., Chem. Lett. 2008, 37, 108–109. Ohmura, T.; Masuda, K.; Suginome, M., J. Am. Chem. Soc. 2008, 130, 1526–1527.
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.
REFERENCES
61. 62. 63. 64. 65.
66. 67. 68. 69. 70. 71. 72. 73. 74. 75. 76. 77. 78. 79. 80. 81. 82. 83. 84. 85. 86. 87. 88. 89. 90. 91. 92. 93.
225
Palmer, W. S.; Woerpel, K. A., Organometallics 1997, 16, 4824–4827. Franz, A. K.; Woerpel, K. A., Acc. Chem. Res. 2000, 33, 813–820. Franz, A. K.; Woerpel, K. A., J. Am. Chem. Soc. 1999, 121, 949–957. Franz, A. K.; Woerpel, K. A., Angew. Chem. Int. Ed. Engl. 2000, 39, 4295–4299. For reviews, see (a) Tamao, K., in Advances in Silicon Chemistry, JAI Press, Greenwich, CT, 1996, Vol. 3, pp. 1–62; (b) Fleming, I., Chemtracts Org. Chem. 1996, 9, 1–64; (c) Jones, G. R.; Landais, Y., Tetrahedron 1996, 52, 7599–7662. Smitrovich, J. H.; Woerpel, K. A., J. Org. Chem. 1996, 61, 6044–6046. Boudjouk, P.; Samaraweera, U.; Sooriyakumaran, R.; Chrusciel, J.; Anderson, K. R., Angew. Chem. Int. Ed. Engl. 1988, 27, 1355–1356. Seyferth, D.; Annarelli, D. C., J. Am. Chem. Soc. 1975, 97, 7162–7164. Sch€afer, A.; Weidenbruch, M.; Peters, K.; von Schnering, H.-G., Angew. Chem. Int. Ed. Engl. 1984, 23, 302–303. Boudjouk, P.; Black, E.; Kumarathasan, R., Organometallics 1991, 10, 2095–2096. Belzner, J.; Ihmels, H.; Kneisel, B. O.; Gould, R. O.; Herbst-Irmer, R., Organometallics 1995, 14, 305–311. Belzner, J.; Dehnert, U.; Ihmels, H., Tetrahedron 2001, 57, 511–517. Driver, T. G.; Franz, A. K.; Woerpel, K. A., J. Am. Chem. Soc. 2002, 124, 6524–6525. ´ irakovic, J.; Driver, T. G.; Woerpel, K. A., J. Org. Chem. 2004, 69, 4007–4012. C Franz, A. K.; Woerpel, K. A., Angew. Chem. 2000, 112, 4465–4469. ´ irakovic, J.; Woerpel, K. A., J. Org. Chem. 2007, 72, 1027–1030. Calad, S. A.; C Kuwahara, Y.; Fukami, H.; Howard, R.; Ishii, S.; Matsumura, F.; Burkholder, W. E., Tetrahedron 1978, 34, 1769–1774. Shaw, J. T.; Woerpel, K. A., J. Org. Chem. 1997, 62, 442–443. Shaw, J. T.; Woerpel, K. A., Tetrahedron 1997, 53, 16597–16606. Shaw, J. T.; Woerpel, K. A., Tetrahedron 1999, 55, 8747–8756. Peng, Z.-H.; Woerpel, K. A., Org. Lett. 2002, 4, 2945–2948. Calad, S. A.; Woerpel, K. A., J. Am. Chem. Soc. 2005, 127, 2046–2047. Driver, T. G.; Woerpel, K. A., J. Am. Chem. Soc. 2004, 126, 9993–10002. Lawrance, G. A., Chem. Rev. 1986, 86, 17–33. Cleary, P. A.; Woerpel, K. A., Org. Lett. 2005, 7, 5531–5533. Damrauer, R.; Crowell, A. J.; Craig, C. F., J. Am. Chem. Soc. 2003, 125, 10759–10766. Belzner, J.; Ihmels, H., Adv., Organomet. Chem. 1999, 43, 1–42. Belzner, J.; Dehnert, U.; Ihmels, H.; H€ubner, M.; M€ uller, P.; Uso´n, I., Chem. Eur. J. 1998, 4, 852–863. DePuy, C. H.; Damrauer, R.; Bowie, J. H.; Sheldon, J. C., Acc. Chem. Res. 1987, 20, 127–133. Palmer, W. S.; Woerpel, K. A., Organometallics 1997, 16, 1097–1099. Palmer, W. S.; Woerpel, K. A., Organometallics 2001, 20, 3691–3697. Ishikawa, M.; Sugisawa, H.; Harata, O.; Kumada, M., J. Organomet. Chem. 1981, 217, 43–50. Seyferth, D.; Shannon, M. L.; Vick, S. C.; Lim, T. F. O., Organometallics 1985, 4, 57–62.
226
SILVER-CATALYZED SILYLENE TRANSFER
94. 95. 96. 97. 98. 99. 100.
Seyferth, D.; Vick, S. C.; Shannon, M. L., Organometallics 1984, 3, 1897–1905. Clark, T. B.; Woerpel, K. A., J. Am. Chem. Soc. 2004, 126, 9522–9523. Goller, A.; Heydt, H.; Clark, T., J. Org. Chem. 1996, 61, 5840–5846. Gordon, M. S., J. Am. Chem. Soc. 1980, 102, 7419–7422. Shaw, J. T.; Woerpel, K. A., J. Org. Chem. 1997, 62, 6706–6707. Anderson, L. L.; Woerpel, K. A., Org. Lett. 2009, 11, 425–428. For the analogous photochemical reaction of silacyclopropenes with nitriles, see Sakurai, H.; Kamiyama, Y.; Nakadaira, Y., J. Chem. Soc. Chem. Commun. 1978, 80–81. Clark, T. B.; Woerpel, K. A., Organometallics 2005, 24, 6212–6219. The sulfonyl oxygen has been demonstrated to be nucleophilic. For leading reports, see (a) Braverman, S.; Duar, Y., J. Am. Chem. Soc. 1983, 105, 1061–1063; (b) Carretero, J. C.; Garcia Ruano, J. L.; Martinez, M. C.; Rodriguez, J. H., Tetrahedron 1987, 43, 4417–4423; (c) Lucchini, V.; Modena, G.; Pasquato, L., J. Chem. Soc. Chem. Commun. 1992, 293–294; (d) Saginova, L. G.; Polyakova, O. V.; Shabarov, Y. S., J. Org. Chem. USSR 1992, 28, 938–942; (e) Leca, D.; Song, K.; Amatore, M.; Fensterbank, L.; Lacote, E.; Malacria, M., Chem. Eur. J. 2004, 10, 906–916. For the bond dissociation enthalpies of S¼N, S¼O, Si--N, and Si--O, see (a) Parsons, S.; Passmore, J., Inorg. Chem. 1992, 31, 526–528; (b) Roux, M. V.; Temprado, M.; Jimenez, P.; Davalos, J. Z.; Notario, R.; Guzman-Mejia, R.; Juaristi, E., J. Org. Chem. 2003, 68, 1762–1770; (c) Johnson, E. R.; Clarkin, O. J.; DiLabio, G. A., J. Phys. Chem. A 2003, 107, 9953–9963. Clark, T. B.; Woerpel, K. A., Org. Lett. 2006, 8, 4109–4112. For leading reviews on the Mukaiyama aldol reaction, see (a) Carreira, E. M., in Comprehensive Asymmetric Catalysis I–III; Jacobsen, E. N.; Pfaltz, A.; Yamamoto, H., eds., Springer-Verlag, Berlin, 1999, pp. 997–1065; (b) Mahrwald, R., Chem. Rev. 1999, 99, 1095–1120. (a) Lin, S.; Bondar, G. V.; Levy, C. J.; Collins, S., J. Org. Chem. 1998, 63, 1885–1892; (b) Hollis, T. K.; Bosnich, B., J. Am. Chem. Soc. 1995, 117, 4570–4581; (c) Carreira, E. M.; Singer, R. A., Tetrahedron Lett. 1994, 35, 4323–4326. For examples of 1,3-Brook rearrangements, see (a) Chuprakov, S.; Malyshev, D. A.; Trofimov, A.; Gevorgyan, V., J. Am. Chem. Soc. 2007, 129, 14868–14869; (b) Smith, A. B.; Pitram, S. M.; Boldi, A. M.; Gaunt, M. J.; Sfouggatakis, C.; Moser, W. H., J. Am. Chem. Soc. 2003, 125, 14435–14445; (c) Naganuma, K.; Kawashima, T.; Okazaki, R., Chem. Lett. 1999, 28, 1139–1140; (b) Shinokubo, H.; Miura, K.; Oshima, K.; Utimoto, K., Tetrahedron 1996, 52, 503–514; (c) Yamamoto, K.; Kimura, T.; Tomo, Y., Tetrahedron Lett. 1985, 26, 4505–4508; (d) Wilson, S. R.; Georgiadis, G. M., J. Org. Chem. 1983, 48, 4143–4144; (e) Hudrlik, P. F.; Hudrlik, A. M.; Kulkarni, A. K., J. Am. Chem. Soc. 1982, 104, 6809–6811. For reviews involving carbenes and metallo carbenoids, see (a) Padwa, A.; Hornbuckle, S. F., Chem. Rev. 1991, 91, 263–309; (b) Padwa, A.; Weingarten, M. D., Chem. Rev. 1996, 96, 223–270; (c) Padwa, A.; in Topics in Current Chemistry, Springer, Berlin/Heidelberg, 1997, Vol. 189, pp. 121–158; (d) Padwa, A., Pure Appl. Chem. 2003, 75, 47–62. Gehrhus, B.; Hitchcock, P. B.; Lappert, M. F., Organometallics 1997, 16, 4861–4864. (a) Ando, W.; Ikeno, M.; Sekiguchi, A., J. Am. Chem. Soc. 1978, 100, 3613–3615; (b) Seyferth, D.; Lim, T. F. O., J. Am. Chem. Soc. 1978, 100, 7074–7075; (c) Belzner, J.; Ihmels, H.; Pauletto, L.; Noltemeyer, M., J. Org. Chem. 1996, 61, 3315–3319; (d) Jutzi, P.;
101. 102.
103.
104. 105.
106.
107.
108.
109. 110.
REFERENCES
111. 112. 113. 114. 115. 116. 117. 118. 119. 120. 121. 122. 123. 124. 125. 126. 127. 128. 129.
130. 131. 132. 133. 134. 135. 136.
227
Eikenberg, D.; Bunte, E.-A.; M€ohrke, A.; Neumann, B.; Stammler, H. -G., Organometallics 1996, 15, 1930–1934; (e) Sakai, N.; Fukushima, T.; Minakata, S.; Ryu, I.; Komatsu, M., Chem. Commun. 1999, 1857–1858. Sch€afer, A.; Weidenbruch, M.; Pohl, S., J. Organomet. Chem. 1985, 282, 305–313. (a) Ando, W.; Ikeno, M.; Sekiguchi, A., J. Am. Chem. Soc. 1977, 99, 6447–6449; (b) Ishikawa, M.; Nakagawa, K. -I.; Kumada, M., J. Organomet. Chem. 1977, 135, C45–C49. Heinicke, J.; Gehrhus, B., J. Organomet. Chem. 1992, 423, 13–21. Gehrhus, B.; Lappert, M. F., J. Organomet. Chem. 2001, 617–618, 209–223. Ando, W.; Hagiwara, K.; Sekiguchi, A., Organometallics 1987, 6, 2270–2271. Calad, S. A.; Woerpel, K. A., Org. Lett. 2007, 9, 1037–1040. Bourque, L. E.; Woerpel, K. A., Org. Lett. 2008, 10, 5257–5260. For the isolation of oxasilacyclopropane-derived ketones, see Ando, W.; Hamada, Y.; Sekiguchi, A.; Ueno, K., Tetrahedron Lett. 1982, 23, 5323–5326. Howard, B. E.; Woerpel, K. A., Org. Lett. 2007, 9, 4651–4653. Brook, A. G.; Kong, Y. K.; Saxena, A. K.; Sawyer, J. F., Organometallics 1988, 7, 2245–2247. Brook, A. G.; Saxena, A. K.; Sawyer, J. F., Organometallics 1989, 8, 850–852. Brook, A. G.; Azarian, D.; Baumegger, A.; Hu, S. S.; Lough, A. J., Organometallics 1993, 12, 529–534. Nevarez, Z.; Woerpel, K. A., Org. Lett. 2007, 9, 3773–3776. Pearson, R. G.; Songstad, J., J. Am. Chem. Soc. 1967, 89, 1827–1836. Pearson, R. G., J. Org. Chem. 1989, 54, 1423–1430. Pearson, R. G., Acc. Chem. Res. 1993, 26, 250–255. Nevarez, Z.; Woerpel, K. A., J. Org. Chem. 2008, 73, 8113–8115. Nielsen, I. M. B., J. Phys. Chem. A 1998, 102, 3193–3201. For leading studies on the boat-like nature of [3,3]-sigmatropic rearrangements of divinylsubstituted three- and four-membered rings, see (a) Hammond, G. S.; DeBoer, C. D., J. Am. Chem. Soc. 1964, 86, 899–902; (b) Trecker, D. J.; Henry, J. P., J. Am. Chem. Soc. 1964, 86, 902–905;(c) Piers, E., in Comprehensive Organic Synthesis, Trost, B. M.; Fleming, I.; Paquette, L. A., eds.; Pergamon Press, Oxford, 1991, Vol. 5, pp. 971–998; (d) Bronson, J. J.; Danheiser, R. L., in Comprehensive Organic Synthesis, Trost, B. M.; Fleming, I.; Paquette, L. A., eds.; Pergamon Press, Oxford, 1991, Vol. 5, pp. 999–1035; ¨ zkan, I.; _ Zora, M., J. Org. Chem. 2003, 68, 9635–9642; (f) Zora, M., J. Org. Chem. (e) O 2005, 70, 6018–6026. Roskamp, E. J.; Johnson, C. R., J. Am. Chem. Soc. 1986, 108, 6062–6063. Doyle, M. P.; Bagheri, V.; Harn, N. K., Tetrahedron Lett. 1988, 29, 5119–5122. Doyle, M. P.; Forbes, D. C.; Vasbinder, M. M.; Peterson, C. S., J. Am. Chem. Soc. 1998, 120, 7653–7654. Clark, J. S.; Fretwell, M.; Whitlock, G. A.; Burns, C. J.; Fox, D. N. A., Tetrahedron Lett. 1998, 39, 97–100. Kitagaki, S.; Yanamoto, Y.; Tsutsui, H.; Anada, M.; Nakajima, M.; Hashimoto, S., Tetrahedron Lett. 2001, 42, 6361–6364. Kawachi, A.; Doi, N.; Tamao, K., J. Am. Chem. Soc. 1997, 119, 233–234. Bourque, L. E.; Cleary, P. A.; Woerpel, K. A., J. Am. Chem. Soc. 2007, 129, 12602–12603.
8 SILVER CARBENOIDS CARL J. LOVELY Department of Chemistry and Biochemistry, The University of Texas at Arlington, Arlington, Texas
8.1 Introduction 8.2 Wolff Rearrangement 8.3 Carbene Transfer Reactions to p Bonds 8.3.1 8.3.2
Aziridination Cyclopropanation
8.4 Formation and Reactions of Ylides 8.4.1 8.4.2
C--Hal Addition–Rearrangement Reactions C--S Addition–Rearrangement Reactions
8.5 C--H Insertion 8.6 N--H Insertion 8.7 Ring Expansion Reactions 8.8 Intermediacy of Silver Carbenes 8.9 Miscellaneous Reactions Involving Silver Carbenoids 8.10 Summary References
8.1
INTRODUCTION
Although the other coinage metals, silver and gold, are in the same group as copper, which has a rich history in organic synthesis, the utility of Ag and Au as homogeneous catalysts in organic synthesis has only just begun to be realized since the early 2000s.1 Silver salts and complexes have seen extensive use in organic synthesis as oxidants, as halide sinks, and as heterogeneous catalysts in the Wolff rearrangement.2 Since the late 1990s, silver salts have begun to emerge as viable homogeneous catalysts for a variety of transformations, few of which are feasible Silver in Organic Chemistry Edited by Michael Harmata Copyright 2010 John Wiley & Sons, Inc.
229
230
SILVER CARBENOIDS
Me Me N
Me Ag Cl
N
Me
Me Me 1
Figure 8.1. NHC silver complex.
with other transition metal complexes, and as a result provide new avenues for investigation, in particular mechanistic issues and derived applications. One might anticipate, on the basis of periodic trends, that copper and silver would exhibit some similarities in their chemistries, and therefore it is not surprising that the potential of silver salts to function as group transfer catalysts has been explored.3 This chapter summarizes the current status of net silver-catalyzed carbene transfer processes primarily with diazo precursors.4,5 The term net here refers to the fact there is evidence to suggest that in some cases the reaction does not proceed via a silver carbene, but rather that the silver complex functions in a different capacity, usually as a Lewis acid. In other situations, the reactivity profile suggests that silver carbenes are the reactive intermediates, but this has not been established unequivocally. It should be noted that this discussion does not include stable silver carbenes (NHC ¼ N-heterocylic carbene) of the type exemplified in Figure 8.1,6,7 which have been used to transfer the imidazolylidene fragment to other metal complexes.8
8.2
WOLFF REARRANGEMENT
The Wolff rearrangement constitutes the first step of the well-known Arndt–Eistert homologation of carboxylic acids, which can be performed under a number of conditions, many of which include the presence of silver salts.2 While the mechanistic details of these catalyzed reactions remain to be fully elucidated, there is some evidence to suggest that silver carbenes may serve as intermediates in certain circumstances. In a combined mass spectrometry and computational study, Beauchamp and Stoltz have demonstrated the formation and rearrangement of a silver carbene species derived from dimethyl diazomalonate.9 In this study, the first stable species observed was the b-diketo complex 2 (for comparison see complex 21 in Fig. 8.2), which, on activation using low-energy collision-activated dissociation (CAD), loses the acetonitrile ligand, producing 3 (Scheme 8.1). Further activation leads to loss of nitrogen and a mole of carbon monoxide, leading to the formation of a species with a molecular mass consistent with the silver carbene 4. An additional round of CAD leads to the expulsion of an additional mole of carbon monoxide and the formation of a second silver carbene species 5. Additional support for this
231
WOLFF REARRANGEMENT NCMe Ag O O MeO
Ag
– MeCN
O MeO
OMe
– N 2, CO
O
– CO
O
MeO
OMe
MeO
OMe
OMe +
Ag
+
N2
N2
Ag
2
3
4
5
Scheme 8.1
decomposition pathway was provided through the use of labeled substrates (deuterium or 13 C), although these experiments were performed with the corresponding copper system, which underwent similar chemistry.9 Chavan and coworkers provide evidence that the Wolff rearrangement is facilitated by the formation of silver nanoclusters, which initiate electron transfer to the diazo compound providing 8. While the precise fate of this species remains to be firmly established, they suggest a multicycle process involving the intermediacy of a silver carbene 10 (Scheme 8.2).10–12 Decomposition of the silver carbene to the free carbene 14 precedes rearrangement to ketene 13, which is then trapped with water to provide the carboxylic acid 15 (Scheme 8.2). Agn (0)
O
+ R 8
N2
N2
+
Agn
O
O
Agn (0) +
R
R +
7
9
N2
R
+ 11
O Agn (0)
O R 6 N2
+
Agn
+
O Agn
O Agn(0)/Agn+
R
R
12
+
Agn nOH-
+
Agn
10
n/2Ag2O + n/2H2O + ne O
O R
R 14
13
H2O
R
CO2H 15
Scheme 8.2
232
8.3 8.3.1
SILVER CARBENOIDS
CARBENE TRANSFER REACTIONS TO p BONDS Aziridination
Jørgensen and coworkers reported the preparation of N-tosyl aziridines 19–20 by the net carbene addition (via a diazo compound) to N-tosyl iminoesters with either copper or silver catalysts.13,14 It was noted that the copper catalysts were generally superior, although a catalyst derived from AgSbF6 and (R)-Tol-BINAP provided the corresponding aziridine 19 from 16 and trimethylsilyl diazomethane 17 (R ¼ TMS) in excellent chemical yield with high levels of diastereoselectivity, but unfortunately the enantioselectivity was poor (Scheme 8.3). This success with trimethylsilyldiazo-
N EtO2C
Ts +
(R)-TolBINAP, THF, –78 ºC
R
16
Ts N
AgSbF6 (10 mol%)
N2
EtO2C
R
19: R = TMS, 88% (dr = 20:1; ee = 12%) 20: R = CO2Et, 32% (dr = 1:1; ee = 20%)
17: R = TMS 18: R = CO2Et
Scheme 8.3
methane prompted Jørgensen to extend this reaction with ethyl diazoacetate (EDA, 18, R ¼ CO2Et). Unfortunately, these reactions were less successful, with reduced yield, reduced diastereoselectivity, but interestingly a slight improvement in the enantioselectivity for the chiral trans isomer (the cis-isomer is meso). In the course of this study, Jørgensen observed that no diethyl fumarate or maleate, common byproducts from reactions involving metallocarbenes and EDA (18), were formed in the presence of the silver catalyst, whereas such products were observed with copper catalysts. This observation led them to speculate that the reaction did not involve the formation of a metallocarbene with silver salts, but was in fact Lewis acid-catalyzed.14 H F3C
CF3 CF3
B N
N
N
N
H F3C
N N
Ag CF3 CF 3
F3C
O
O
F3C OMe
MeO N2
21
B N
N
N
N
CF3 CF3 N N
Ag CF3 CF 3 THF
AgTp(CF3)2. (THF) (22)
Figure 8.2. Silver tris(pyrazolyl)borato complexes.
CARBENE TRANSFER REACTIONS TO p BONDS
8.3.2
233
Cyclopropanation
The Dias group has developed an interest in the synthesis and characterization of putative reactive intermediates in a variety of transition-metal-mediated processes, particularly those involving the coinage metals.4,15–17 As a part of this study, the stable AgTp–diazomalonate [Tp ¼ tris(pyrazolyl)borato] complex 21 was prepared in an attempt to obtain a discrete, and characterizable, silver carbene complex.18 However, rather than the desired carbene complex, the b-diester complex 21 (Fig. 8.2) was isolated and characterized by X-ray crystallography. It is of note that coordination occurs via the ester carbonyl oxygens, and thus this may be considered akin to nonproductive resting state in carbene transfer reactions in which diazomalonate participates. A related species was identified in the Stoltz–Beauchamp mass spectrometry investigation discussed above (see compound 3, Scheme 8.1) as formed in high abundance, which then formed the silver carbene on collisional activation. To further probe this issue, an attempt was made to prepare and characterize the corresponding complex with the more reactive EDA (18) and 22; however, no characterizable silver complex was obtained. Interestingly, the evolution of a gas was observed, indicating that a reaction occurred, presumably the formation of a reactive silver–carbene with the release of nitrogen gas. On the basis of these observations, the experiment was repeated in the presence of a carbene trap to intercept the putative silver carbene. An initial attempt at cyclopropanation with styrene 23 was not successful in the clean formation of the corresponding cyclopropane 24. Instead, a rather complex mixture of products was obtained containing the expected cyclopropane 24 (Scheme 8.4), the ring expanded product 25, the solvent insertion product 26 and other products (possibly 27).17 This is in contrast to the corresponding copper complex, which smoothly cyclopropanoates a variety of olefins, including styrene.19 Ultimately it was found that by conducting the reaction with benzene in CH2Cl2 or in neat benzene, the ring expansion product 30a, via the norcaradiene 29a (B€uchner reaction), was obtained in good yield (Scheme 8.5).17 Evaluation of the substrate
CO2Et CO2Et
N2
CO2Et
18
24
25
22, CH2Cl2
CO2Et
Cl 23
CO2Et CO2Et
Cl 26
Scheme 8.4
27
234
SILVER CARBENOIDS
N2
CO2Et
CO2Et
22 , CH2Cl2
R 28a–f
CO2Et
18
R
R 29a–f
30a–f
Scheme 8.5
scope revealed that other aromatic derivatives, 28b–f, participated in the rearrangement, but generally the yields of the ring-expanded product 30b–f were lower than with benzene (Table 8.1). In some cases (30b and 30f), competitive benzylic C--H insertion was observed, providing 31 and 32, respectively (Fig. 8.3). Perez and coworkers have observed similar B€ uchner chemistry, as well as cyclopropanation of styrene with the polybrominated Tp–silver complex 33 (Fig. 8.3), although no chemical yields were reported in either case.20 The corresponding Cu–scorpionate complexes function well as cyclopropanation catalysts.19,21–23 Davies and Thompson have demonstrated that several donor-acceptor diazo compounds (35 and 37) participate in selective silver-catalyzed cyclopropanation of a variety of olefins with AgSbF6 as catalyst at reflux in CH2Cl2 (Schemes 8.6 and 8.7).24 Interestingly, it had been reported previously that this same group of olefins undergoes predominantly C--H insertion if Rh2(OAc)4 is employed as the catalyst. Very high levels of chemoselectivity were obtained, with typically 15 : 1 selectivity for cyclopropanation over C--H insertion (Tables 8.2 and 8.3). Lower selectivity was observed in only one case, the reaction of styryl diazoacetate (37) and cyclohexadiene (Table 8.3, entry 2). In addition to high levels of chemoselectivity, excellent diastereoselectivities were observed, with additional isomers undetected by 1 H NMR spectroscopy.24
8.4
FORMATION AND REACTIONS OF YLIDES
Metallocarbenes derived from diazoacetates and the appropriate transition metal are generally thought to be electrophilic in nature; that is, they are susceptible to nucleophilic addition. This reactivity manifold has been exploited with the formation of ylide species on reaction with a number of nucleophiles providing intermediates that can undergo either rearrangements (e.g., 41 ! 42, Scheme 8.8)25 or cycloadditions (e.g., 43 ! 44, Scheme 8.8)26 depending on the type of ylide formed (Scheme 8.8). In the case of silver-derived carbenes, this area has only recently begun to attract attention, and therefore there are only limited examples reported in the literature. 8.4.1
C--Hal Addition–Rearrangement Reactions
In the course of our investigation of the cyclopropanation of olefins, we observed the formation of a by-product, ethyl-2,3-dichloropropanoate (26, Scheme 8.4), when the reaction was conducted in CH2Cl2, resulting from the net insertion of the carbene
235
FORMATION AND REACTIONS OF YLIDES
TABLE 8.1. Yields and Isomer Ratios of the Silver-Catalyzed B€ uchner Reaction of Benzene Derivatives with EDA Substrate
Product
Yield (%) a, b
Isomer Ratio (2 : 3 : 4)
74 75c
NA NA
CO2Et
28a 30a
CH3
CO2Et
CH3
28b
30b
OCH3
CO2Et
64 d 62 c, e
40 28c
Cl
28d
CO2Me
28e
1 : 1.8 : 2.4
1 : 0 : 19
OCH3 30c CO2Et
49 f
1 : 8.5 : 18
14 f
0:1:1
35 g
NA
Cl
30d CO2Et
CO2Me
30e
CO2Et
28f 30f a
The yields are the average of at least two runs and refer to isolated material and, unless noted, were conducted in CH2Cl2. b In the reactions conducted in CH2Cl2, there 15–25% of the C--Cl insertion product 26 was observed in the 1 H NMR spectrum of the crude reaction mixture. c These reactions were conducted in the neat arene. d The C--H insertion product 31 ( 4%) was observed in 1 H NMR spectrum of crude reaction mixture. e The C--H insertion product 31 ( 5%) was observed in 1 H NMR spectrum of crude reaction mixture. f The purified material was contaminated with 5% of the C--Cl insertion product. g The C--H insertion product 32 ( 28%) was observed in 1 H NMR spectrum of crude reaction mixture.
236
SILVER CARBENOIDS
H Br
OEt
OEt
O
O
Br
N
N
N
N
32
Br
N Br
N
Ag Br THF
Br 31
Br
B
Br
TpBr3Ag(THF) (33)
Figure 8.3. C--H insertion products and the Perez catalyst.
R1
H
AgSbF6 (10 mol%)
N2 +
R2
R3
CO2Me
Ph
34
CH2Cl2, reflux
Ph
CO2Me
R1 R2
35
H R3 36
Scheme 8.6
Ph R1 R2
H R3
AgSbF6 (10 mol%)
N2 + CO2Me
Ph
CH2Cl2, reflux
37
34
CO2Me R1 R2
H R3 38
Scheme 8.7 TABLE 8.2. Silver-Catalyzed Cyclopropanation with 35 a Entry
R1
R2
1 2
H H
--(CH2)4---(CH¼CHCH2CH2)--
3
H
4 5 6 7 8 9
H Ph Me H Ph H
a
R3
Yield (%) 88 79 80
H Ph H Me H Ph
Approximately 5–10 equiv of the olefin was employed.
Ph H Ph Ph Ph Ph
96 82 80 86 84 54
237
FORMATION AND REACTIONS OF YLIDES
TABLE 8.3. Silver-Catalyzed Cyclopropanation with 37 a Entry
R1
R2
R3
1 2
H H
--(CH2)4---(CH¼CHCH2CH2)--
3
H
4 5 6 7
H Ph Me Ph
Yield (%) 43 67 b 57
H Ph H H
Ph H Ph Ph
82 56 c 65 34
a
Approximately 5–10 equiv of the olefin was employed. Includes 15% of insertion product. c Reaction conducted at room temperature. b
X RO2C
+
X
X
[2,3] rearrangement
RO2C
41
RO2C M 39
RO2C
+
42
M
40
Z
X
X = NR2, OR, SR, Hal
R2
R1O2C
+
Y
X R2
[3+2] cycloaddition
43
R1O2C
X R2
Y Z 44
Scheme 8.8
fragment into the C--Cl bond. The participation of halogenated solvents with metallocarbenes is rare, but has been observed previously.27,28 Similarly, the reaction of free carbenes with haloalkanes and allyl halides has also been described previously in the literature, but these processes are quite unselective.29–34 To further explore the formation of 26, the reaction was repeated in CH2Cl2 only, and in this case the insertion product was obtained in 26% yield (Scheme 8.9; Table 8.4, entry 1).35 On the basis of this result, an examination of the scope and
N2
CO2Et 18
22 R–X 44–50
CO2Et X R 26, 51–53
Scheme 8.9
238
SILVER CARBENOIDS
TABLE 8.4. Products and Yields of the Carbene Insertion into Haloalkanes Entry
Substrate Product (a ¼ Cl, b ¼ Br)
Yield (X ¼ Cl) (%) a, b
Yield (X ¼ Br) (%) a, b
26
65
60
29 c, d
62
48 c, e
39
78
70
—
63 f, g
60 f, h
73
—
CO2Et
1
X
CH2X2
CH2X 26a–b
44a-b
CO2Et
2
X
CHX3
CHX2 51a–b
45a-b
CO2Et
3
X
CX4
CX3 52a–b
46a-b
X
X
4
CO2Et 53a–b
47a–b
5
Cl
Cl
CO2Et 53a
48a X
X
6
CO2Et
53a–b
49a–b
7
Cl
Cl
CO2Et
50a
53a
a
These isolated yields are based on the average of at least two experiments and on the amount of EDA used. The material balance is accounted for by dimerization of EDA (fumarate and maleate). c Reaction performed in CH2Cl2. d In addition, 26a was formed ( 9%). e In addition, 26a was formed ( 3%). f Yield determined by GC. g In addition, cyclohexene was formed ( 47%, by GC). h In addition, cyclohexene was formed ( 57%, by GC). b
limitations of this reaction was conducted. It was found that similar products were obtained with a number of polychloromethane and polybromomethane derivatives (Table 8.4, entries 1–3). However, attempts to extend this addition–rearrangement sequence to other haloalkanes was not successful in the expected sense, providing
239
FORMATION AND REACTIONS OF YLIDES
instead the a-haloacetate derivative 53a-b via a net 1,1-hydrohalogenation/elimination process and the olefin (Table 8.4, entries 4–7).35 In the case of the reaction with halocyclohexanes, the formation of cyclohexene was established and quantified using GC analysis, which revealed that an equivalent amount (within experimental error) of olefin was formed. On the basis of these results, a mechanism (Scheme 8.10) involving the intermediacy of a silver–carbene 54 was proposed in which the insertion product arises from the formation of the halonium ylide 55, followed by a 1,2 shift (55 ! 26, or 51 or 52). Alternatively, if the substrate and thus the halonium ylide 56 contain a b-hydrogen, this could be removed by an intramolecular deprotonation with concomitant loss of halide resulting in formation of the olefin 57 and the a-haloacetate 53. At this stage, no independent evidence has been obtained to support this pathway; thus this mechanism is purely speculative (see text below). Indeed, although the pathway has been depicted as involving metal-free intermediates, it is quite likely that this is not the case, but this awaits independent experimental verification. CO2Et
N2
22
CO2Et
LnAg 54
18 X = Cl, Br
R
X
R CO2Et 53
57
X R
55
X 47–50 CO2Et
X H
R
CO2Et
R X 44–46
X CO2Et
56
R 26, 51–52
Scheme 8.10
The interception of metallocarbenes (Cu or Rh) with allylic substituted amines, ethers, thioethers, and selenides and their subsequent [2,3] rearrangement is well documented.36 On the other hand, fewer reports exist of the corresponding process with allylic halides.37–39 Very recently it has been reported by our lab that the AgTp complex 22 does, in fact, catalyze this reaction in excellent efficiencies, although the diastereoselectivities are very low at this point (Table 8.5, Schemes 8.11 and 8.12).15 In this case, allylic halides 58–62 were reacted with ethyl- or tert-butyl diazoacetate (BDA) in the presence of 22 (Scheme 8.11), leading to the formation of homoallylic haloacetates 64a,b–68a,b (Table 8.5, entries 1–5). Propargylic substrates 69–70 provided similar rearrangement products, although in this case allenes 72b–73a,b were obtained (Scheme 8.12). A mechanism involving the formation and rearrangement of the halonium ylide (63 or 71) was proposed, again via the intermediacy of a silver carbene.15 Similarly to the 1,2 shift described above, no detailed mechanistic studies have been performed, and so this pathway should be considered a working hypothesis, pending additional experimental investigation.
240
SILVER CARBENOIDS
TABLE 8.5. Products and Yields from the Rearrangement a R/Yield (%) Entry
Substrate
Product CO2R2
Br
1
Br
58
75
65
86 (1 : 1) b
96(1 : 1) b
59
84
57
89
70 (1 : 1) b
80 (1 : 1) b
—
66
71
74
64
CO2R2
Cl
2
t-Bu
Et
Cl
59 65
CO2R2
Cl
3
Cl
60
66 CO2R2
Br
4
Br
61
67
5
CO2R2
Br
Br
Br
62
Br 68
H
Br H
6
•
69
Br H
2
CO2R
72
H
Br
7
H
•
Br
70
CO2R
2
73 a
The reactions are conducted in neat halide with the exception of entry 5, in which the substrate is dissolved in CH2Cl2. b Syn/anti ratio, determined by 1 H NMR spectroscopy.
241
FORMATION AND REACTIONS OF YLIDES
X
CO2R2
CO2R2 N2
R1
22
X
[2,3]
X
rearrangement
R1
R1 64a,b–68a,b
63
58–62
CO2R2
Halonium ylide
X = Cl, Br
Scheme 8.11
X
CO2R2
CO2R2 N2
[2,3]
X
rearrangement
22
R1
R1
69–70
X
CO2R2
R1
•
72b–73a,b
71
Halonium ylide
X = Cl, Br
Scheme 8.12
Most of the substrates evaluated in this reaction were primary halides; however, one example of a secondary substrate 74 was reported to successfully engage in this rearrangement (Scheme 8.13, 74 ! 75).15 CO2Bu-t
Cl
N2
Cl
22, 77%
74
CO2Bu-t 75
Scheme 8.13
Of note in this chemistry is that in both the C-Hal insertion and C-Hal rearrangement reactions, new carbon–carbon bonds are formed as a result of the participation of the halogen, but the halogen is retained in the product. This is quite unusual as in most cases, metal-catalyzed reactions involving halogens result in the consumption of the halogen for the ensuing bond formation. This process provides substrates with multiple functional groups that can then be used in future manipulations. 8.4.2
C--S Addition–Rearrangement Reactions
Very recently, Davies and coworkers have reported a silver-catalyzed Doyle–Kirmse reaction of allylic and propargylic sulfides with ethyl diazophenylacetate (35). A selection of examples was investigated providing the rearranged products, either homoallylic thioethers (76 ! 77, Scheme 8.14a) or allenyl thioethers (78 ! 79, Scheme 8.14b) in good yields, and with good overall scope.40
242
SILVER CARBENOIDS
Ph
S
EtO2C
CO2Et N2
S
(a) AgOTf, CH2Cl2 35ºC, 83%
Br 76
Ph
35
Br 77
Ph
S
CO2Et N2
EtO2C
Ph
35
S
(b) AgOTf, CH2Cl2 35ºC, 71%
78
79
Scheme 8.14
8.5
C--H INSERTION
Several silver complexes have been investigated in C--H insertion reactions of carbenes. One of the earliest reports by Sulikowski and Burgess describes an intramolecular C--H insertion of a chiral aryl diazoacetate 80 as part of a study on the total synthesis of mitomycin family of natural products (Scheme 8.15).41 Among several catalysts that were screened was AgSbF6 in the presence of bis(isoxazoline)type ligands, which led to the formation of the desired insertion product 81 in moderate yield and modest diastereoselectivity. CO2R*
CO2R*
N2 H
N
H
AgSbF6, L* THF, then DDQ
O
R* =
O
L* =
O
O N
N Ph
O
N Ph
H
H O 80
81 (44%, dr = 2.7 : 1)
Scheme 8.15
More recently, both our group and Perezs group have demonstrated that Ag complexes 22 and 33 containing scorpionate ligands are effective for C--H insertion with diazoacetates (Scheme 8.16).17,20,42 Complex 22 efficiently effects insertion of R2
H
CO2R1 N2 18, 35, 37 22, r.t.
R2 CO2R1
18: R1 = Et, R2 = H 35: R1 = Me, R2 = Ph 37: R1 = Me, R2 = (E )–PhCH=CH–
Scheme 8.16
N--H INSERTION
243
EDA into cycloalkanes 82 and 83 (Table 8.6, entries 1 and 2), but in the case of oxygencontaining substrates 84–88, the C--H insertion yields were rather modest (Table 8.6, entries 3–5), or insertion did not occur at all (Table 8.6, entries 6 and 7).42 Interestingly, some of the same substrates were evaluated with the corresponding CuTpðCF3 Þ2 complex, and in this case the oxygen-containing substrates appear to react more efficiently.19 This reaction has been extended to include donor–acceptor carbenes, and it was found to provide the expected products in most cases.43 The Tpð3;5-CF3 Þ2 Ag complex 22 and the related silver complex 33 were examined in C--H insertion reactions with several acyclic alkanes.17,20,42 The chemical yields were excellent, and interestingly, both complexes exhibited relatively high proportions of C--H insertion at the primary carbon (Table 8.7). Perez and coworkers speculated that this increase selectivity for insertion at primary sites was due to increased electrophilic character at the metal center rather than steric factors.20 The Perez lab has reported a new set of silver complexes containing fluorinated indazolylborates (e.g., 108; Fig. 8.4).44 These complexes perform similar C--H insertion reactions to the pyrazolylborates, but at very low catalyst loadings. Perhaps not unexpectedly, the selectivities closely mirrored those obtained for the corresponding fluorinated pyrazolyl complexes.
8.6
N--H INSERTION
Both copper and rhodium complexes have been shown to catalyze carbene insertion into N--H bonds, providing a convenient method for amino acid synthesis.45 Accordingly, the corresponding silver-catalyzed process has attracted attention. Jørgensen and coworkers reported the reaction of aniline with substituted diazoacetate derivatives and several simple silver salts admixed with bis(isoxazoline) ligands (Scheme 8.17).46 Generally the yields were quite moderate, as were the enantioselectivities (Table 8.8). Parallel investigations with copper-based catalysts showed improved chemical yields, but the enantioselectivities were still poor. The authors suggested that the silver-mediated reactions proceeded via a Lewis acid–catalyzed pathway, rather than the expected carbene transfer mechanism, which was thought to be operating in the case of the copper catalysts.46 Hu and coworkers have examined N--H (and O--H and S--H) insertions in the presence of silver salts as well as with copper or rhodium catalysts with styryl diazoacetates (Schemes 8.18 and 8.19, Tables 8.9 and 8.10).47 Two possible products (114/117 and 115/118) were obtained that are derived from either direct insertion or insertion with net transposition (Schemes 8.18 and 8.19). Silver and copper salts tended to favor transposition (Table 8.9, entries 2–5; Table 8.10, entries 2–9), whereas rhodium favored direct insertion (Table 8.9, entry 1; Table 8.10, entry 1). The selectivity differences between the two products were again rationalized in terms of two mechanistic pathways. In the case of rhodium-based catalysts, it was proposed that the reaction occurs via a metallocarbene, whereas with copper and silver catalysts the reaction was interpreted as proceeding by Lewis acid activation.
244
4
3
2
1
Entry
85
O
84
O
83
82
Substrate
O
O
92
91
90
89
CO 2Et
CO 2Et
CO2Et
CO2Et
Products
41
24
88
88
Yield (%)
O
CO2Me
Ph
CO2Me
—
98
97
Ph
96
Ph
CO2Me
Products
—
91 (3 : 2)
76
87
Yield (%)
TABLE 8.6. C--H Insertion into Cyclic Substrates of Diazoacetates Catalyzed by AgTp Complex 22
O
Ph
Ph
—
104
Ph
CO2Me
103
CO2Me
102
CO2Et
Products
—
0
77
84
Yield (%)
245
7
6
5
88
O
O
87
O
O
86
O
O
O
O
O
95
94
93
O
CO2Et
CO 2Et
CO2Me
0
0
5
O
O
O
O
CO2Me
Ph
CO2Me
CO2Me
Ph
100
99
101
O
Ph
0
0
71 (3 : 2)
O
O
O
O
Ph
107
Ph
CO2Me
Ph
CO2Me
106
105
O
MeO2C
0
0
88 (3 : 2)
246
22 33 a 108
22 33 a 108
22 33 a 108
Catalyst
Substrate
80 60 75
41 29 36
CO2Et
59 (both) 42 (both) 56 (both)
CO2Et
CO2Et
Primary Site (%)
EtO2C
27 37 26
CO2Et
47 57 48
CO 2Et
CO 2Et
12 14 5
Secondary Site (%)
TABLE 8.7. C--H Insertions into Acyclic Alkanes Catalyzed by AgTp Complexes 22, 33, and 108
20 40 25
CO 2Et
14 21 18
CO2Et
Tertiary Site (%)
247
a
50 (both) 66 (both)
81
CO2Et
EtO2C
CO2Et
CO 2Et
48
CO2Et
18
46
CO2Et
22 20
These values have been normalized to account for C--H insertion products only.
55 a
33 a 108
33
a
33 a
33 a 108
EtO2C
EtO2C
CO2Et
6
CO2Et
39 24
33
CO2Et
17
56 45
CO 2Et
21 35
CO2Et
13
21 10
CO2Et
11
CO2Et
CO2Et
37
CO2Et
248
SILVER CARBENOIDS
TABLE 8.8. Silver(I)-Catalyzed N--H Insertion Entry
Ag Salt
1
AgSbF5
Ligand O
O N
N
AgOTf
O
O N
N
AgOTf
O
O N
N
AgClO4
O
O N
N
AgSbF5
O
O N
N
N
N
N
N
4
20
Me
5
48
Ph
33
13
Ph
49
9
Ph
58
9
Ph
58
()
O
O
AgOTf
Me
Ph
Ph
8
48
O
O
AgOTf
5
Bu-t
t-Bu
7
Me
O
O
AgOTf
25
Bu-t
t-Bu
6
8
Bu-t
t-Bu
5
Me
Bu-t
t-Bu
4
ee (%)
Bu-t
t-Bu
3
112 Yield (%)
Bu-t
t-Bu
2
R1
N
N
249
N--H INSERTION
F
F
F
F F
F
HB N N
F F
N NN N
CF3
F F F
F CF3 CF3
Ag(OCMe2) 108
Figure 8.4. Polyfluorinated indazolyl silver complex (108), reported in 2008.
Ph NH2
Ag(I)L*X (10 mol%)
N2
+
R1
CH2Cl2, r.t.
CO2Et
R1 Ph
110: R 1 = Ph
109
CO2Et NH 112
111: R 1 = Me
Scheme 8.17
catalyst (10 mol%)
N2 Ph NH2
+
CO2Me
Ph
109
NHPh
NHPh CO2Me
Ph
CH2Cl2, reflux
37
CO2Me
Ph
114
115
Scheme 8.18
catalyst (10 mol%)
N2 Bn XH +
116
Ph
CO2Me
XBn
XBn
37
CO2Me
CO2Me Ph
Ph
CH2Cl2, reflux
117
118
Scheme 8.19
TABLE 8.9. Metal-Catalyzed N--H Insertion Entry 1 2 3 4 5 a
Yield [114 þ 115(%)]
Catalyst Rh2(OAc)4 Cu(OTf)2 AgBF4 AgClO4 AgOTf
a
In this case 1 mol% catalyst was used.
79 72 67 64 66
114 : 115 <2 : 98 85 : 15 73 : 27 67 : 33 65 : 35
250
SILVER CARBENOIDS
TABLE 8.10. Metal-Catalyzed O--H and S--H Insertions Entry 1 2 3 4 5 6 7 8 9 a
X
Catalyst Rh2(OAc)4 Cu(OTf)2 AgBF4 AgClO4 AgOTf Cu(OTf)2 AgBF4 AgClO4 AgOTf
O O O O O S S S S
Yield [117 þ 118 (%)]
117: 118
78 53 67 71 63 68 69 71 59
<2 : 98 72 : 28 >98 : 2 >98 : 2 >98 : 2 78 : 22 69 : 31 68 : 32 72 : 28
a
In this case 1 mol% catalyst was used.
8.7
RING EXPANSION REACTIONS
It was demonstrated previously that cyclopropyl carbenes, generated by the thermal decomposition of the corresponding diazo compound, provide cyclobutenes, but this process can be nonselective and inefficient.48,49 Tang and coworkers have reported a transition-metal-mediated variant in which significant improvements in the yields and selectivities were observed.50 Several transition metal complexes were evaluated, from which AgOTf emerged as the catalyst of choice (Scheme 8.20). As can be seen in Table 8.11, a broad range of substrates (119–128) were investigated, and it was observed that this reaction is stereospecific (entries 5 and 6, Table 8.11) and regioselective (entries 7–10, Table 8.11). N2
R
R
AgOTf, CH2Cl2
OR O
r.t.
124–128
OR O 129–138
Scheme 8.20
In addition to examining the scope of the reaction, the authors observed some interesting selectivity issues with respect to the migrating bond, noting both a catalyst dependence (Scheme 8.21a) and a substrate dependence (Scheme 8.21b), which, in principle, provides substantial flexibility in the utility of the reaction.
8.8
INTERMEDIACY OF SILVER CARBENES
At this point, although there is no independent experimental evidence for the formation of silver carbenes in the reactions described above, and in some cases
251
INTERMEDIACY OF SILVER CARBENES
TABLE 8.11. Scope of Silver-Catalyzed Cyclopropyldiazoacetate Ring Expansion Entry
Cyclopropane
Cyclobutane
Yield (%)
Ratio
N2 O
O O
1
Ph
O
Ph
91
—
77
—
72
—
90
—
71
Single isomer
92
Single isomer
73
10 : 1
129
119 BnO
BnO N2 CO2Et
2
OEt O
120
130 N2 CO2Et
3
CO2Et
121
131 N2 CO2Et
4
OEt O
122
Pr
5
132 N2 CO2Et
Pr
Pr
O
123 Pr
6
OEt
Pr
133 N2 CO2Et
Pr
Pr OEt
Pr O
124
134 N2 OEt
CO2Et
7 125
O
135
(Continued)
252
SILVER CARBENOIDS
TABLE 8.11 Entry
(Continued) Cyclopropane
MeO2C
Cyclobutane
N2
CO2Et
10
Single isomer
OEt
77
Single isomer
O 137
N2
CO2Et
EtO2C
CO2Et
Me
87
CO2Et
EtO2C
127
EtO2C
OEt
136
CO2Et
9
10 : 1
O
126
CO2Et
70
OEt
CO2Et
EtO2C
Ratio
MeO2C
N2
8
Yield (%)
Me O
128
138
Ph
Ph
N2
(a)
Cat. = Cu(MeCN)4PF6;140 : 141 = 5.00
cat., CH 2Cl2 r.t.
OEt
OEt
OEt
Ph
O
O
O 139
140
Cat. = AgOTf; 140 : 141 = 0.50 Cat. = Rh2(OAc)4;140 : 141 = 0.14
141
Ar
Ar
N2
(b)
r.t.
OEt O 142
Cat. = Cu(MeCN)4PF6; 143 : 144 = >20 : 1
cat., CH 2Cl2
OEt
Ar O
Ar =
143
O
OEt Cat. = AgOTf; 143 : 144 = >20 : 1 Cat. = Rh2(OAc)4; 143 : 144 = 0.5
144
OMe
Scheme 8.21
alternative mechanistic pathways have been posited in which the silver complex behaves as a Lewis acid, the overall behavior of the intermediate in many of the other transformations discussed is consistent with the formation of a silver carbene. A DFT study by Perez and coworkers on the mechanism of Ag-catalyzed C--H insertion of
253
MISCELLANEOUS REACTIONS INVOLVING SILVER CARBENOIDS
diazoesters has appeared detailing and supporting the role of silver carbenes as intermediates in this chemistry.51
8.9 MISCELLANEOUS REACTIONS INVOLVING SILVER CARBENOIDS All the examples described above involved the reaction of diazoacetate derivatives with silver salts to initiate the formation of a putative silver carbene; however, other pathways exist. For example, Porcel and Echavarren have reported an intramolecular cyclization of an allylstannane to a pendent alkyne (Scheme 8.22) that involves the intermediacy of a silver carbene.52 As can be seen in Table 8.12, the reactions proceeded in moderate to excellent yield, providing the dienylstannane, while in some cases, reductive destannylation occurred. Several asymmetric reactions were reported with substrate (E)-145d, leading to the formation of the expected adduct in reasonable enantioselectivities (ee ¼ 73–78%) in a preliminary screen with a number of different ligands. Mechanistically, most alkyne cyclizations of this type have been interpreted in terms of the catalyst serving as a Lewis acid, activating the p bond to nucleophilic attack,53–57 and while this pathway is feasible, the authors preferred one involving the intermediacy of a silver carbene.52 Control studies ruled out a pathway involving
R1 E
SnBu3 [AgOTf(PPh 3)]3 (3 mol%)
R2
E
145
PhMe, 70ºC
E
R1
E
R2
SnBu3
E
R1
E
R2
+
146
147
O O SnBu3
148
Scheme 8.22
TABLE 8.12. Silver-Catalyzed Cyclocarbostannylation
1 2 3 4 5 6 7 8 a
Entry
Substrate
(E)-145a (E)-145b (Z)-145c (E)-145d (Z)-145d (E)-145e (E)-145f (E)-145g
E ¼ CO2Me, R ¼ H, R ¼ Me E ¼ CO2Me, R1 ¼ R2 ¼ H E ¼ SO2Ph, R1 ¼ R2 ¼ H E ¼ SO2Ph, R1 ¼ H, R2 ¼ Me E ¼ SO2Ph, R1 ¼ H, R2 ¼ Me E ¼ CH2OAc, R1 ¼ R2 ¼ H E ¼ CH2OTBDPS, R1 ¼ R2 ¼ H E ¼ CH2OH, R1 ¼ R2 ¼ H 1
2
Time (h)
Yield (%)
Yield (%)
0.5 5.0 3.5 0.5 0.5 2.5 2.0 0.2 a
146a (87) 146b (71) 146c (72) 146d (90) 146d (91) 146e (69) 146f (93) 146g (31)
147a (3) 147b (6) 147b (9)
147e (12) 147g (11)
Compound 148 was obtained in 30% yield in addition to the expected products 146g and 147g.
254
SILVER CARBENOIDS
transmetallation, along with the participation of silver acetylides; therefore, the pathway depicted in Scheme 8.23 was offered as a possibility. Porcel and Echavarren proposed that the silver complex selectively activates the alkyne to form the cyclopropyl carbene–silver(I) complex 149. Fragmentation leads to the formation of alkenylsilver complex 150, and reaction with Bu3SnOTf then leads to the formation of the final cyclization product and regeneration of the silver catalyst.
SnBu3 Z [AgOTf(L)]
SnBu3 R 145
151 R Bu3SnOTf
+(L)Ag
– OTf
AgL
SnBu3 R H 149
150
R
Bu3SnOTf
Scheme 8.23
8.10
SUMMARY
It is clear from the discussion above that the utility of silver complexes as carbene transfer catalysts has come a long way, but much remains to be done. There is a far from complete picture with respect to the precise mechanisms operating in many of these reactions, which hinders the design of more efficient second-generation catalysts. This is particularly relevant in cases where little or no precedent exists for a reaction, for example, the C-Hal insertion reactions reported by us.15,35 A second issue that demands attention concerns the development of asymmetric variants of many of these processes, which will also rely on a more complete mechanistic understanding of the design of appropriate supporting ligands.
REFERENCES
255
ACKNOWLEDGMENTS Our own endeavors in this area have been possible because of the hard work and valuable intellectual contributions of a number of coworkers, whose names appear in the references. Financial support has generously been provided by The Robert A. Welch Foundation (Y-1362). REFERENCES 1. Issue number 8 of Chemical Reviews (Chem. Rev. 2008, 108) was devoted to the utility of coinage metals in synthesis. 2. Kirmse, W., Eur. J. Org. Chem. 2002, 2193–2256. 3. Li, Z.; He, C., Eur. J. Org. Chem. 2006, 4313–4322. 4. Dias, H. V. R.; Lovely, C. J., Chem. Rev. 2008, 108, 3223. 5. Zhang, Z.; Wang, J., Tetrahedron 2008, 64, 6577–6605. 6. Arduengo, A. J.; Dias, H. V. R.; Calabrese, J. C.; Davidson, F., Organometallics 1993, 12, 3405–3409. 7. Ramnial, T.; Abernethy, C. D.; Spicer, M. D.; McKenzie, I. D.; Gay, I. D.; Clyburne, J. A. C., Inorg. Chem. 2003, 42, 1391–1393. 8. Ahrens, S.; Zeller, A.; Taige, M.; Strassner, T., Organometallics 2006, 25, 5409–5415. 9. Julian, R. R.; May, J. A.; Stoltz, B. M.; Beauchamp, J. L., J. Am. Chem. Soc. 2003, 125, 4478–4486. 10. Sudrik, S. G.; Maddanimath, T.; Chaki, N. K.; Chavan, S. P.; Chavan, S. P.; Sonawane, H. R.; Vijayamohanan, K., Org. Lett. 2003, 5, 2355–2358. 11. Sudrik, S. G.; Sharma, J.; Chavan, V. B.; Chaki, N. K.; Sonawane, H. R.; Vijayamohanan, K. P., Org. Lett. 2006, 8, 1089–1092. 12. Sudrik, S. G.; Chaki, N. K.; Chavan, V. P.; Chavan, S. P.; Chavan, S. P.; Sonawane, H. R.; Vijayamohanan, K., Chem. Eur. J. 2006, 12, 859–864. 13. We use the term carbene transfer to represent the net chemical transformation and not necessarily to imply that this is a reactive intermediate (either a free carbene or a metallocarbene). 14. Juhl, K.; Hazell, K. A.; Jorgensen, K. A., J. Chem. Soc. Perkin Trans. 1999, 1, 2293–2297. 15. Krishnamoorthy, P.; Browning, R. G.; Singh, S.; Sivappa, R.; Lovely, C. J.; Dias, H. V. R., Chem. Commun. 2007, 731–732. 16. Dias, H. V. R.; Lu, H. -L.; Kim, H. -J.; Polach, S. A.; Goh, T. K. H. H.; Browning, R. G.; Lovely, C. J., Organometallics 2002, 21, 1466–1473 and the references cited therein. 17. Lovely, C. J.; Browning, R. G.; Badarinarayana, V.; Dias, H. V. R., Tetrahedron Lett. 2005, 46, 2453–2455. 18. Dias, H. V. R.; Polach, S. A., Inorg. Chem. 2000, 39, 4676–4677. 19. Browning, R. G.,Ph.D. dissertation, Univ. Texas, Arlington, 2003. 20. Urbano, J.; Belderrain, T. R.; Nicasio, M. C.; Trofimenko, S.; Diaz-Requejo, M. M.; Perez, P. J., Organometallics 2005, 24, 1528–1532. 21. Morilla, M. E.; Dıaz-Requejo, M. M.; Belderrain, T. R.; Nicasio, M. C.; Trofimenko, S.; Perez, P. J., Organometallics 2004, 23, 293–295.
256
SILVER CARBENOIDS
22. Dıaz-Requejo, M. M.; Caballero, A.; Belderraın, T. R.; Nicasio, M. C.; Trofimenko, S.; Perez, P. J., J. Am. Chem. Soc. 2002, 124, 978–983. 23. Dıaz-Requejo, M. M.; Belderraın, T. R.; Trofimenko, S.; Perez, P. J., J. Am. Chem. Soc. 2001, 123, 3167–3168. 24. Thompson, J. L.; Davies, H. M. L., J. Am. Chem. Soc. 2007, 129, 6090–6091. 25. Sweeney, J. B., Chem. Soc. Rev. 2009, 38, 1027–1038. 26. Padwa, A., Chem. Soc. Rev., 2009, 3072–3081, DOI: 10.1039/b816701j. 27. Pirrung, M. C.; Zhang, J.; McPhail, A. T., J. Org. Chem. 1991, 56, 6269–6271. 28. Pirrung, M. C.; Zhang, J.; Lackey, J.; Sternbach, D. D.; Brown, F., J. Org. Chem. 1995, 60, 2112–2125. 29. Migita, T.; Ando, W.; Kondo, S.; Matsuyama, H.; Kosugi, M., Nippon Kagaku Zasshi 1970, 91, 374–377. 30. Urry, W. H.; Bilow, N., J. Am. Chem. Soc. 1964, 86, 1815–1819. 31. Urry, W. H.; Eiszner, J. R.; Wilt, J. W., J. Am. Chem. Soc. 1957, 79, 918–922. 32. Urry, W. H.; Wilt, J. H., J. Am. Chem. Soc. 1954, 76, 2594–2595. 33. Urry, W. H.; Eiszner, J. R., J. Am. Chem. Soc. 1952, 74, 5822–5826. 34. Urry, W. H.; Eiszner, J. R., J. Am. Chem. Soc. 1951, 73, 2977. 35. Dias, H. V. R.; Browning, R. G.; Polach, S. A.; Diyabalanage, H. V. K.; Lovely, C. J., J. Am. Chem. Soc. 2003, 125, 9270–9271. 36. Hodgson, D. M.; Pierard, F. Y. T. M.; Stupple, P. A., Chem. Soc. Rev. 2003, 30, 50–61. 37. Simonneaux, G.; Galardon, E.; Paul-Roth, C.; Gulea, M. M. S., J. Organomet. Chem. 2001, 617–618, 360–363. 38. Doyle, M. P.; Forbes, D. C.; Vasbinder, M. M.; Peterson, C. S., J. Am. Chem. Soc. 1998, 120, 7653–7654. 39. Doyle, M. P.; Tamblyn, W. H.; Bagheri, V., J. Org. Chem. 1981, 46, 5094–5102. 40. Davies, P. W.; Albrecht, S. J. -C.; Assanelli, G., Org. Biomol. Chem. 2009, 7, 1276–1279. 41. Burgess, K.; Lim, H. -J.; Porte, A. M.; Sulikowski, G. A., Angew. Chem. Int. Ed. Engl. 1996, 35, 220–222. 42. Dias, H. V. R.; Browning, R. G.; Richey, S. A.; Lovely, C. J., Organometallics 2004, 23, 1200–1204. 43. Lovely, C. J.; Flores, J. A.; Meng, X.; Dias, H. V. R., Synlett 2009, 129. 44. Despagnet-Ayoub, E.; Jacob, K.; Vendier, L.; Etienne, M.; Alvarez, C.; Diaz-Requejo, M. M.; Perez, P. J., Organometallics 2008, 27, 4779–4787. 45. Moody, C. J., Angew. Chem. Int. Ed. 2007, 46, 9148–9150. 46. Bachmann, S.; Fielenbach, D.; Jorgensen, K. A., Org. Biomol. Chem. 2004, 2, 3044–3049. 47. Yue, Y.; Wang, Y.; Hu, W., Tetrahedron Lett. 2007, 48, 3975–3977. 48. Gallucci, R. R.; Jones, M., J. Am. Chem. Soc. 1976, 98, 7704–7711. 49. Friedman, L.; Shechter, H., J. Am. Chem. Soc. 1960, 82, 1002–1003. 50. Xu, H.; Zhang, W.; Shu, D.; Werness, J. B.; Tang, W., Angew. Chem. Int. Ed. 2008, 47, 8933–8936. 51. Braga, A. A. C.; Maseras, F.; Urbano, J.; Caballero, A.; Diaz-Requejo, M. M.; Perez, P. J., Organometallics 2006, 25, 5292–5300. 52. Porcel, S.; Echavarren, A. M., Angew. Chem. Int. Ed. 2007, 46, 2672–2676.
REFERENCES
257
53. Zhao, J.; Hughes, C. O.; Toste, F. D., J. Am. Chem. Soc. 2006, 128, 7436–7437. 54. Sun, J.; Kozmin, S. A., Angew. Chem. Int. Ed. 2006, 45, 4991–4993. 55. Harrison, T. J.; Kozak, J. A.; Corbella-Pane, M.; Dake, G. R., J. Org. Chem. 2006, 71, 4525–4529. 56. Sweis, R. F.; Schramm, M. P.; Kozmin, S. A., J. Am. Chem. Soc. 2004, 126, 7442–7443. 57. Harrison, T. J.; Dake, G. R., Org. Lett. 2004, 6, 5023–5026.
9 ALDOL AND RELATED PROCESSES MASANORI KAWASAKI Medicinal Chemistry Research Laboratories, Pharmaceutical Research Division, Takeda Pharmaceutical Company Limited, Osaka, Japan
HISASHI YAMAMOTO Department of Chemistry, University of Chicago, Chicago, Illinois
9.1 Introduction 9.2 Allylation Reaction Using Allyltributyltin 9.3 Allylation Reaction Using Allylsilanes 9.4 Aldol Reaction Using Tin Enolates 9.5 Aldol Reaction Using Silyl enol Ethers 9.6 Mannich Reaction 9.7 Nitrosoaldol Reaction 9.8 Aldol Reaction with Azodicarboxylate 9.9 Conclusion References
9.1
INTRODUCTION
Carbon–carbon bond-forming reactions are essential reactions in organic synthesis. Among them, the allylation reaction of the carbonyl group, which leads to homoallylic alcohols, is one of the most important transformations, since the resulting homoallylic alcohols are useful building blocks in synthetic organic chemistry. For this transformation, allylic organometallic reagents can be used with or without catalysts. Since the 1990s, the enantioselective version of this reaction was developed using chiral allylic organometallic reagents or chiral catalysts. The reactions of chiral allylic borane or allylic titanium as allylic reagents with carbonyls afford homoallylic alcohols with high enantioselectivity and diastereoselectivity.1 Because of its high Silver in Organic Chemistry Edited by Michael Harmata Copyright 2010 John Wiley & Sons, Inc.
259
260
ALDOL AND RELATED PROCESSES
selectivity, this reaction has been applied to the synthesis of various natural and unnatural products. Although excellent results have been obtained through this approach, a stoichiometric amount of the chiral source is necessary. In the past few decades, the catalytic enantioselective version of this reaction has been developed to resolve that issue. In this chapter, we would like to focus on allylation reactions of carbonyls using silver catalysts.
9.2
ALLYLATION REACTION USING ALLYLTRIBUTYLTIN
In 1996, Yamamoto and Yanagisawa reported the allylation reaction of aldehydes with allytributyltin in the presence of a chiral silver catalyst.2 They found that the combination of silver and a phosphine ligand accelerates the allylation reaction between aldehydes and allyltributyltin. After this discovery, they screened several chiral phosphine ligands and found that chiral silver–diphosphine catalysts can effect the reaction in an enantioselective fashion (Table 9.1).2 For example, when benzaldehyde and allyltributyltin were mixed in the presence of 5 mol% of AgOTf and (S)2,20 -bis(diphenylphosphino)-1,10 -binaphthyl (BINAP), the corresponding homoallyl alcohol was obtained with 96% ee and 88% yield (Table 9.1). Generally, the reaction with aromatic aldehydes afforded the corresponding homoallyl alcohols in excellent
TABLE 9.1. Catalyst Screening for the Allylation Reaction with Aldehydes and Allyltributyltin SnBu3
Entry 1 2 3 4 5 6 7 a
OH
complex
+
R–CHO
∗
THF, -20°C
R
Complex a
Yield (%)
ee (%)
(S)-BINAP–AgOCOCF3 (S)-BINAP–AgClO4 (S)-BINAP–AgNO3 (S)-BINAP–AgOTf (R,R)-Chiraphos–AgOTf (S,S)-Me-Duphos–AgOTf (S,S)-Et-Duphos–AgOTf
47 1 26 88 97 4 13
40 26 53 96 2 48 3
Structures ligands: R
PPh2 PPh2
PPh2
P
PPh2
R P
R
R (S)-BINAP
(R,R)-Chiraphos
R = Me: (S,S)-Me-Duphos R = Et : (S,S)-Et-Duphos
261
ALLYLATION REACTION USING ALLYLTRIBUTYLTIN
TABLE 9.2. Allylation Reaction with Various Aldehydes
Entry
OH
complex
SnBu3
RCHO
+
RCHO
∗
THF, -20°C
R
Yield (%)
ee (%)
88
96 (S)
83
88 (S)
89
97
94
93
72
93
85
97
59
97
95
96
47
88
CHO
1
CHO
2
CHO
3
4
5
O
CHO CHO
n-C3H7 Me
6
CHO
7
MeO
8
Br
CHO
CHO CHO
9
enantioselectivity. On the other hand, when aliphatic aldehydes were used as substrates, the enantioselectivity was slightly decreased (Table 9.2). In addition, this silver–BINAP catalyst has been applied to the crotylation reaction between aldehydes and 2-butenylstannanes (Scheme 9.1).3 The addition of (E)-2butenylstannane to benzaldehyde in the presence of 20 mol% AgOTf/BINAP gave the corresponding adduct with 94% ee with a high anti : syn ratio. In contrast to other
262
ALDOL AND RELATED PROCESSES
SnBu3
+
PhCHO
OH
(R)-BINAP–AgOTf THF, -20°C to r.t.
Ph anti
E : Z ratio of 2-butenylstannane 95 : 5 2 : 98 53 : 47
Yield (%) 56 72 45
OH +
Ph syn
anti (% ee) : syn (% ee) 85 (94) : 15 (64) 85 (91) : 15 (50) 85 (94) : 15 (57)
Scheme 9.1. Crotylation with benzaldehyde.
Lewis acid-catalyzed allylation reactions, the syn : anti ratio observed here did not depend on the E : Z ratio of the starting stannane.4 Although the reaction mechanism was not clear, this result would suggest that the reaction took place via transmetallation from stannane to silver to generate an allylsilver species. This silver–BINAP catalyst can be applied for the reaction of 2,4-pentadienylstannane and aldehydes.5 Generally, the reaction of 2,4-pentadienylstannane and aldehydes in the presence of a Lewis acid catalyst afforded the corresponding e adduct. On the other hand, the reaction using the silver catalyst gave the corresponding g adduct predominantly (Table 9.3). These results suggested that the reaction catalyzed by silver could proceed through a six-membered cyclic transition state (Fig. 9.1). In 2000, Loh and coworkers reported the enantioselective allylation reaction in the presence of AgNO3–Tol–BINAP in an aqueous medium (Scheme 9.2).6 They surveyed the reaction media and found that a 1 : 9 mixture of water : ethanol was optimal. The reaction of allyltributyltin and aldehyde in the presence of AgNO3– Tol–BINAP in water and ethanol gave the corresponding homoallyl alcohols with moderate to good enantioselectivities. Cesarotti and coworkers have developed an allylation reaction using allyltributyltin in the presence of an AgOTf–BITIANP complex.7 In one example, the reaction was conducted with 4-bromobenzaldehyde and allyltributyltin in the presence of the catalyst to give the adduct with 78% ee (Scheme 9.3). Shi and Wang reported that the catalyst generated from AgOTf and a chiral diphenylthiophosphoramide ligand, which was prepared from chiral 1,10 -binaphthyl2,20 -diamine, could promote the allylation reaction of allyltributyltin and aromatic aldehydes (Table 9.4).8 Thus, the reaction of allyltributyltin and aldehyde in the presence of AgOTf and chiral diphenylthiophosphoramide gave the corresponding adducts with up to 98% ee. More recently, Benaglia and coworkers reported that the allylation reaction of a-iminoesters proceeded to give the homoallyl alcohols with moderate enantioselectivities (Scheme 9.4).9 The chiral diimine that has a chiral 1,2-cyclohexyldiamine framework was used as the ligand. The reaction of a-iminoesters and allyltributyltin in the presence of AgOTf and diimine ligand proceeded to afford chiral amino acid derivatives with moderate enantioselectivities.
263
ALLYLATION REACTION USING ALLYLTRIBUTYLTIN
TABLE 9.3. Dienylation of Aldehydes with Organostannanes OH SnR13
Entry
SnR13
1
SnBu3
2
SnMe3
3
SnMe3
BINAP–AgOTf
+
RCHO
R
THF, -20°C
RCHO
Yield (%)
ee (%)
61
90
68
89
57
90
41
87
62
89
73
58
68
58
52
71
<1
—
CHO Me CHO
4
SnMe3
5
SnMe3
MeO
CHO
O
6
SnBu3
7
SnMe3
8
SnBu3
9
SnBu3
CHO CHO
CHO
O
M
ε-attack
R
OH
H R
O 1
SnR
H H acyclic TS
3
LA
ε-adduct
+ RCHO
OH
H
γ-attack
R
O
M
cyclic TS
Figure 9.1. Proposed transition states.
R
γ-adduct
264
ALDOL AND RELATED PROCESSES
P(Tol) 2 P(Tol) 2
SnBu3
+
RCHO
OH
AgNO3 H2O : EtOH (1 : 9), -40°C
R 70–100% yield 30–81% ee
Scheme 9.2. Allylation reaction in aqueous media.
9.3
ALLYLATION REACTION USING ALLYLSILANES
Allylsilanes are more attractive allylation reagents than allylstannanes from an environmental perspective. However, the use of allylsilanes for organic synthesis is limited since the reactivity of allylsilanes is typically low compared to allyltributyltin. In 1999, Yamamoto and coworkers investigated the allylation reaction using allylsilanes and found that BINAP–AgF catalyzes enantioselective Sakurai– Hosomi-type reaction using allyltrimethoxysilanes (Table 9.5).10 The reaction was conducted with aromatic aldehydes and allyltrimethoxysilanes in the presence of 6 mol% BINAP and 10 mol% AgF to afford the corresponding adducts with high enantioselectivities. The use of a slight excess of AgF was essential for this reaction, because when slight excess of AgF with respect to BINAP was used, the catalytically active 1 : 1 complex of AgF and BINAP was formed predominantly. It was also established that this catalytic system could be used for asymmetric crotylation reactions (Scheme 9.5). The high anti selectivity was observed regardless of the geometry of the crotylsilane. Although this catalytic system was effective for
OH CHO +
SnBu3
(+)-BITINAP–AgOTf THF, -20°C
R R=H R = Br
R R = H: 73%, 70% ee R = Br: 55%, 78% ee
S PPh2 PPh2 S BITIANP
Scheme 9.3. Allylation using BITIANP ligand.
265
ALLYLATION REACTION USING ALLYLSILANES
TABLE 9.4. Allylation Using a Thiophosphoramide Ligand
NHEt S NHPPh2
SnBu3
+
RCHO
OH
AgOTf THF, -20°C
Entry
RCHO
R
Yield (%)
ee (%)
75
98
74
96
56
80
80
68
1 Me
CHO
Cl
CHO
2
3 MeO
CHO
4
CHO
S N N S OMe SnBu3
+
N EtO2C
OMe
AgOTf (ligand:AgOTf = 2 : 1)
HN
THF, 0°C
EtO2C 100%, 67% ee (100%, 71% ee at -40oC)
Scheme 9.4. Allylation using a diimine ligand.
266
ALDOL AND RELATED PROCESSES
TABLE 9.5. Allylation of Aldehydes with Allyltrimethoxysilane Si(OMe)3 +
a
MeOH, -20°C
R
RCHO
Yield (%)
ee (%)a
PhCHO (E)-PhCH¼CHCHO 2-Furyl-CHO 1-Naphthyl-CHO 4-Me-C6H4CHO 4-Br-C6H4CHO
80 93 70 81 67 90
94 (R) 78 (R) 83 (R) 92 (R) 93 (R) 93
Entry 1 2 3 4 5 6
OH
(R)-p-Tol–BINAP–AgF
RCHO
Configuration.
various aromatic aldehydes, aliphatic aldehydes did not react in the presence of this catalyst. To overcome this limitation, Yamamoto and coworkers investigated the reaction conditions and found that the use of KF and 18-crown-6 in the presence of AgOTf and BINAP was effective in promoting the allylation reaction (Table 9.6).11 Thus, the reaction was conducted with aliphatic aldehydes and allytrimethoxysilane in the presence of catalyst to give the corresponding adducts with satisfactory yields and high enantioselectivities. Moreover, this catalytic system can be applied to the reaction between aromatic aldehydes and allyltrimethoxysilane. The allylation reaction between ketones and allylsilanes was achieved in 2005. Yamamoto and Wadamoto developed the asymmetric allylation reaction in the presence of AgF-Difluorphos (Scheme 9.6).12 The reaction of ketones and allyltrimethoxysilane in the presence of AgF and Difluorophos afforded the corresponding tertiary homoallylic alcohols with high enantioselectivities. Additionally, a,b-unsaturated ketones could be used as substrates, and this catalytic system could be applied for the asymmetric crotylation reaction to obtain anti adducts preferentially (Schemes 9.7 and 9.8). When a,b-unsaturated ketones were used as substrates, 1,2-addition products were obtained exclusively. As described before, the anti adducts were obtained predominately, regardless of the geometry of crotyltrimethoxysilane.
Si(OMe)3
+
PhCHO
OH
(R)-BINAP–AgF MeOH, -20°C to r.t.
Ph anti
E : Z ratio 83 : 17 1< : 99 45 : 55
OH +
Yield (%) 77 82 99
Scheme 9.5. Crotylation of benzaldehyde.
Ph syn
anti (% ee) : syn (% ee) 92 (96) : 8 (62) 94 (94) : 6 (60) 93 (94) : 7 (60)
267
ALLYLATION REACTION USING ALLYLSILANES
TABLE 9.6. Allylation of Various Aldehydes Si(OMe)3
+
(R)-BINAP–AgOTf KF, 18-crown-6
RCHO
OH
THF, -20°C
Entry 1 2 3 4 5 6 7 8 9 a
R
RCHO
Yield (%)
PhCHO (E)-PhCH¼CHCHO 2-Furyl-CHO 1-Naphthyl-CHO 4-Me-C6H4CHO 4-Br-C6H4CHO 2-Me-C6H4CHO c-C6H11CHO PhCH2CH2CHO
91 81 57 95 61 95 82 62 76
ee (%)a 95 87 95 92 95 96 97 93 86
(R) (R) (R) (R) (R) (R) (R) (R) (R)
Configuration.
More recently, the catalyst that was prepared from AgOTf and Tol–BINAP was applied to the synthesis of 2,3-dihydrobenzofurans. The reaction of 2,3-dihydrobenzoxasilepine and an aromatic aldehyde was carried out in the presence of the silver catalyst, KF, and 18-crown-6 to give trans-2,3-disubstituted 2,3-dihydrobenzofuran
O Si(OMe)3
O R
Me
+
1
R
O
F F
O
O
PPh2 PPh2
O (R)-Difluorphos AgF, MeOH
R2 OH
2
R
R = 3-CF3-C6H4: 92%, 90% ee R = 4-Me-C6H4: 63%, 78% ee R = 2-Me-C6H4: 62%, 92% ee R = 1-naphthyl: 98%, 95% ee R = PhCH2Ch2: 42%, 65% ee O
F F
R1
THF, -78°C
O
n
O
n = 1: 89%, 91% ee n = 2: 63%, 93% ee n = 3: 78%, 92% ee
O X X = Cl: 97%, 96% ee X = I: 92%, 96% ee
O 74%, 92% ee
I 92%, 94% ee
Scheme 9.6. Allylation of ketones.
268
ALDOL AND RELATED PROCESSES
O
AgF, MeOH
Si(OMe)3
+
Ph
HO
(R)-Difluorphos THF, -78°C
Ph 96%, 84% ee
AgF, MeOH
O Si(OMe)3
+
Me
(R)-Difluorphos THF, -78°C
HO Me 56%, 95% ee
Scheme 9.7. Allylation of a,b-unsaturated ketones.
in a diastereoselective manner (Scheme 9.9).13 The acyl group at the ortho position of aromatic ring was essential for a successful reaction.
9.4
ALDOL REACTION USING TIN ENOLATES
The aldol reaction is one of the most useful and classical reactions in organic synthesis, since the resulting b-hydroxycarbonyl compounds are highly useful building blocks for pharmaceuticals and natural products.14 The asymmetric version of this reaction has been developed to a significant extent. We focus here on the asymmetric classical aldol reaction using a silver catalyst and related reactions.15 In 1997, Yamamoto, Yanagisawa, and others reported the asymmetric reaction catalyzed by a BINAP–AgOTf catalyst (Table 9.7).16 The reaction was conducted with tributyltin enolate and aldehydes in the presence of the BINAP–AgOTf catalyst to afford the corresponding adduct with high enantioselectivity. The observed
O
Si(OMe)3
+ Ph
AgF, MeOH (R)-Difluorphos THF, -78°C to -40°C
Me
Me OH Ph Me
syn : anti = 90 : 10, 60%, 95% ee (from E isomer) syn : anti = 90 : 10. 95%, 93% ee (from Z isomer)
Si(OMe)3
O + Ph
Me
n
AgF, MeOH (R)-Difluorphos THF, -78°C to -40°C
Me OH Ph
n = 1: syn : anti = 96 : 4, 72%, 98% ee n = 2: syn : anti = 96 : 4. 74%, 99% ee n = 3: syn : anti = >99 : 1, 97%, 98% ee
Scheme 9.8. Crotylation of ketones.
269
ALDOL REACTION USING TIN ENOLATES
MeO
CHO
Si O
+
AgOTf, (+)-Tol–BINAP KF, 18-crown-6
O
OPiv
O MeO
THF
O
O
O
OH 84% ee
F O SiMe2F
Si O
H
O
O
O
O
O O
O SiMe2F
O
Ag P
O
Ag P
O
O P
O
O
P
O O
Ag P
O
P
Scheme 9.9. Synthesis of a dihydrobenzofuran.
TABLE 9.7. Aldol Reaction with Tin Enolates
R
OSnBu3 R3 1
+
RCHO
Tin Enolate
OH
R1
THF, -20°C
R R2 R3
R2
Entry
O
(R)-BINAP–AgOTf
RCHO
Yield (%)
anti : syn
ee (%)
1
OSnBu3
PhCHO
92
89 : 11
92
2 3
OSnBu3
PhCHO PhCHO
94 95
92 : 8 93 : 7
93 94
4
OSnBu3
PhCHO
90
85 : 15
96
5 6
OSnBu3
PhCHO PhCH2CH2CHO
81 77
<1 : 99 <1 : 99
95 95
7
OSnBu3
PhCHO
98
<1 : 99
91
270
ALDOL AND RELATED PROCESSES
H
R2
R
P P R1 Ag SnBu3 O O
H
H
R
P P R1 Ag SnBu3 O O
R3
H
E
anti
Z
syn
Figure 9.2. Proposed transition state.
anti : syn ratio depended on the E : Z ratio of enolate, and this result suggested the cyclic transition state (Fig. 9.2). Since the use of stoichiometric amounts of organostannane is unfavorable for the environment, they have established reaction conditions for a process that is catalytic in organostannane. Thus, the reaction with an alkenyl trichloroacetate as the precursor of the tin enolate in the presence of the catalytic amount of Me3SnOMe and the appropriate aldehydes afforded the corresponding aldol adducts with high enantioand diastereoselectivities (Table 9.8).17 However, when the aliphatic aldehydes were used as substrates, the reaction did not take place at all. The proposed catalytic cycle is shown in Figure 9.3. It was also demonstrated that this catalytic system could be applied using a b-lactone compound as an alkenyl trichloroacetate equivalent to TABLE 9.8. Aldol Reaction Using Catalytic Amounts of Me3SnOMe O O
CCl3
+
(R)-p-Tol–BINAP–AgOTf cat. Me3SnOMe MeOH
O
O
OH
RCHO
R
THF, -20°C to r.t.
+
1 2 3
RCHO PhCHO 4-Me-C6H4CHO CHO O
R syn
anti
Entry
OH
Yield (%)
anti : syn
ee (%)
86 66 49
94 : 6 94 : 6 80 : 20
96 95 77
74 76 61 29 <1 <1
96 : 4 78 : 22 76 : 24 84 : 16 — —
92 90 83 79 — —
O
4 5 6 7 8 9
1-Naphthyl (E)-PhCH¼CHCHO (E)-CH3(CH2)2CH¼CHCHO PhCH2CH2CHO n-C5H11CHO i-PrCHO
271
ALDOL REACTION USING SILYL ENOL ETHERS
O R1
BINAP-AgOTf
MeOH
OSnMe3 R
R3 R2
O
OH
1
RCHO
R
R R3 R2
OSnMe3 R3 R1 R2
Me3SnOMe
O
MeOCOCCl3 O
CCl3 R3
R1 R2
Figure 9.3. Proposed catalytic cycle.
generate a b-ketoester with high enantioselectivity (Scheme 9.10). The alkenyl trichloroacetate reacted with Me3SnOMe to form a tin enolate, which reacted with the aldehyde in the presence of the chiral silver catalyst in a highly enantioselective manner. The resulting tin oxide was protonated by MeOH to generate the product and regenerate Me3SnOMe.
9.5
ALDOL REACTION USING SILYL ENOL ETHERS
Because these asymmetric aldol reactions are ideal methods for constructing bhydroxy carbonyl compounds in optically active form, the development of an asymmetric aldol reaction without the use of an organostannane would be advantageous. Yamagishi and coworkers have reported the Mukaiyama aldol reaction using trimethylsilyl enol ethers in the presence of the BINAP–AgPF6 complex to afford the adducts with moderate enantioselectivities (Table 9.9).18 They have also assigned
O O
+
PhCHO
(R)-p-Tol–BINAP–AgOTf cat. Me3SnOMe MeOH THF, -20°C to r.t.
O
O
Ph
O 84% ee
Scheme 9.10. Aldol reaction of diketene.
OH
272
ALDOL AND RELATED PROCESSES
TABLE 9.9. Mukaiyama Aldol Reaction with Various Aldehydes OSiMe3 R3 R1 R2
Entry
+
O
(S)-BINAP–AgPF6
RCHO
OH
1
R
R
DMF, -25°C
R2 R3
RCHO
1 2 3 4 5
PhCHO 1-Naphthyl-CHO
6
PhCHO
Silyl enol ether OSiMe3 Ph
2-Napththyl-CHO Cyclohexyl-CHO
Yield (%)
ee (%)
100 36 82 100 83
69 80 54 58 47
0
0
OSiMe3 MeO
structures to three silver species using 31 P NMR studies and X-ray analysis. The 2 : 1 complex of BINAP and AgPF6 was found to be a major complex in DMF and could not catalyze Mukaiyama aldol reactions (Fig. 9.4). The 1 : 1 complex of BINAP and AgPF6 and the 1 : 2 complex of BINAP and AgPF6 were found to be minor complexes. The 31 P NMR spectra suggested that an equilibrium existed between these two complexes and AgPF6. In addition, they have demonstrated that the 1 : 1
P
sol.
PF6
P
Ag
sol.
2PF6
P Ag
Ag P
P
sol.
P
1:1 AgPF 6-BINAP complex
minor active species
Ag
sol.
2:1 AgPF 6-BINAP complex
P
P
1:2 AgPF 6-BINAP complex
major inactive species
P : BINAP P
Figure 9.4. Structures of silver–BINAP complexes.
PF6
273
ALDOL REACTION USING SILYL ENOL ETHERS
complex coordinates an aldehyde and catalyzes the reaction via a cyclic transition state. Independently, Yamamoto, Yanagisawa, and others reported the asymmetric aldol reaction using trimethoxysilyl enol ethers.19 The reaction was conducted with aldehydes and trimethoxysilyl enol ethers in the presence of Tol–BINAP–AgF to give the corresponding adducts with high enantioselectivities and diastereoselectivities. They obtained syn-aldol adducts as major products even when silyl enol ethers derived from cyclic ketones were used. Moreover, when a,b-unsaturated aldehydes were employed as substrates, 1,2 adducts were obtained exclusively (Table 9.10). From an NMR study and correlation between the E : Z ratio of the enol ethers and diastereoselectiviy, they proposed a cyclic transition state (Fig. 9.5). Thus, the reaction of E enol ethers proceeded via a boat form, whereas the reaction of Z enol ethers took place via a chair form. In addition, the combination of KF and 18-crown-6 in the presence of BINAP–AgOTf, which was effective in accelerating the allylation reaction using allyltrimethoxysilane, could be used for the aldol reaction (Scheme 9.11).20 Thus, the reaction with the silyl enol ether derived from cyclohexanone and benzaldehyde in the TABLE 9.10. Aldol Reaction Using Trimethoxysilyl Enol Ethers OSi(OMe)3 R3 + R 2 R 1
Entry
Silyl enol ether
O
(R)-p-Tol–BINAP–AgF
RCHO
OH
1
R
MeOH, -78°C
R R2 R3
RCHO
Yield (%)
syn : anti
ee (%)
78 86 87 68 81 <1 <1 18
84 : 16 75 : 25 76 : 24 27 : 73 81 : 19 — — 75 : 25
87 92 90 76 68 — — 52
1 2 3 4 5 6 7 8
OSi(OMe)3
PhCHO 4-MeO–C6H4CHO 4-Br-C6H4CHO 1-naphthyl-CHO (E)-PhCH¼CHCHO PhCH2CH2CHO PhCOMe PhCHO
9
OSi(OMe)3
PhCHO
67
81 : 19
78
PhCHO 4-MeO–C6H4CHO 1-naphthyl-CHO PhCH2CH2CHO
84 76 63 <1
>99 : 1 >99 : 1 94 : 6 —
97 96 95 —
10 11 12 13
OSi(OMe)3
OSi(OMe)3 t-Bu
274
ALDOL AND RELATED PROCESSES
H
H
P
H O O
R R2
P
Ag
R1
OH O F
Si(OMe)3
1
R
R3
R1 R2
syn isomer
R
P P Ag
O
H
O
F Si(OMe)3
R3
Figure 9.5. Proposed transition state.
presence of catalyst proceeded to afford an adduct with high enantioselectivity and diastereoselectivity. Remarkably, the anti adduct was obtained as the major product, in contrast to the BINAP–AgF catalyst system. This catalytic system could be applied to the enantioselective hydroxymethylation of silyl enol ethers with aqueous formalin (Scheme 9.12).21 Doyle and coworkers have successfully applied the catalyst system of BINAP–AgOTf, KF, and 18-crown-6 for Mukaiyama aldol reaction of a-diazo silyl enol ether (Scheme 9.13).22 Hoveyda and coworkers reported that the silver complex generated from AgF2 and an amino acid–based ligand could be used for the asymmetric aldol reaction with silyl enol ethers and a-ketoesters (Table 9.11).23 O
OH
(R)-Tol–BINAP–AgF
Ph
MeOH–acetone, -78°C
OSi(OMe)3 +
anti : syn = 11 : 89 93% ee
PhCHO (R)-BINAP–AgOTf KF, 18-crown-6
O
OH Ph
THF, -20°C
anti : syn = 91 : 9 93% ee
O
OH
(R)-Tol–BINAP–AgF MeOH, -78°C
Ph
anti : syn = 1 : >99 97% ee
OSi(OMe)3 t-Bu
t-Bu
+
PhCHO (R)-BINAP–AgOTf KF, 18-crown-6 THF, -20°C
O t-Bu
OH Ph
anti : syn = 1 : >99 90% ee
Scheme 9.11. Diastereoselectivity of aldol reaction with a different catalyst system.
275
ALDOL REACTION USING SILYL ENOL ETHERS
OSi(OMe)3
O (R)-BINAP–AgOTf, KF
+ aq. HCHO
OH
THF, H2O, -40°C, 6 h
31%, 57% ee
Scheme 9.12. Hydroxymethylation of a trimethoxysilyl enol ether.
OTMS CO2Me
Ar
N2
OH O
(R)-BINAP–AgOTf KF, 18-crown-6
O +
CO2Me
Ar
THF, -20°C
H
N2 Ar = Ph: 4-MeO-C6H4: 4-Me-C6H4: 4-Cl-C6H4: β-styryl
80%, 91% ee 82%, 87% ee 80%, 88% ee 78%, 92% ee 74%, 88% ee
Scheme 9.13. Aldol reaction of an a-diazo trimethylsilyl enol ether. TABLE 9.11. Aldol Reaction of a-Ketoesters
H N
N N
O
O N H R
O OEt
G O
OTMS
AgF2
R
THF
+
O HO OEt
G O
Entry 1 2 3 4 5 6 7 8 9 10 11 12
G CH2CH2Ph CH2CH2CO2Me CH2i-Pr i-Pr i-Pr i-Pr Cy Cy Cyclopropyl CH2¼CH(Me) Ph 2-Thienyl
R Ph Ph Ph Ph t-Bu Me Ph Me Ph Ph Ph Ph
Yield (%)
ee (%)
92 95 95 93 61 >98 98 97 90 98 93 95
86 92 87 95 92 88 95 90 96 90 60 72
276
ALDOL AND RELATED PROCESSES
N O
Ts
Ts OTMS
H
+
(R)-BINAP–AgPF6 THF, -80°C
R
OEt
NH2 O
NH O O
O
Ph
R OH
OEt R = aromatic, t-Bu
61–90% ee
Scheme 9.14. Mannich reaction of a-iminoester with various enol ethers.
9.6
MANNICH REACTION
The Mannich reaction is a useful reaction for the synthesis of b-aminocarbonyl compounds including b-amino acids, b-lactams, and related compounds.24 Because the synthesis of chiral compounds is required for biological study, asymmetric versions of this reaction are in high demand. Several approaches to achieve asymmetric Mannich reactions have been reported. In 1998, Lectka and coworkers reported the asymmetric Mannich reaction of a-iminoesters catalyzed by a BINAP–Ag(I) complex (Scheme 9.14).25 Moreover, Hoveyda and coworkers have demonstrated that Mannich reactions between silyl enol ethers and aldimines are promoted by the chiral complex that was generated from AgOAc and an iso-Leu-derived phosphine ligand.26 When the reaction was conducted with trimethylsilyl enol ether and aldimine in the presence of AgOAc, iso-Leu-derived phosphine ligand and 2-PrOH, the b-aminoketone was obtained with high enantioselectivity (Schemes 9.15 and 9.16). This catalyst system can be applied not only to a,b-unsaturated imines but also to aromatic and aliphatic imines to give the corresponding adducts with high enantioselectivity. Furthermore, this method is applicable to a combination of silyl keteneacetals and alkynylimines (Scheme 9.17).27 In 2006, Hoveyda and coworkers developed an asymmetric Mannich reaction of silyloxyfurans and aldimines using a similar catalyst system (Scheme 9.18).28 The diastereo- and enantioselective reaction between silyloxyfurans and aldimines in the presence of catalyst gave g-butenolides, which are useful building blocks for organic synthesis. They also reported a mechanistic study of this reaction.28b
H N
N PPh2
MeO OTMS N
+ R'
R
O
OMe
OMe BocHN
AgOAc, 2-PrOH THF
NH
O
R R' 76–96% ee
R' = Ph or Me R = aromatic, heteroaromatic alkenyl, alkynyl
Scheme 9.15. Mannich reaction of various aldimines.
R R=
O Ph Ph Cl
277
NITROSOALDOL REACTION
H N
N O
PPh2
MeO
OTMS +
OMe
AgOAc, 2-PrOH
RCHO +
NH
THF
Me
H2N
OMe
O R'
R
R = n-C10H21 Cy i-Bu
92–94% ee
Scheme 9.16. A one-pot, three-component reaction.
H N
N PPh2
MeO OTMS
OMe
OMe
AgOAc, 2-PrOH
+
N
O
NH
THF, -60°C
OPh
O OPh
R
R 84–94% ee
R = TMS, aromatic, alkyl, alkenyl
Scheme 9.17. Mannich reaction of alkynyl aldimines with a silylketeneacetal.
9.7
NITROSOALDOL REACTION
In 2002, Yamamoto and Momiyama reported an unusual aldol-like reaction with silyl enol ethers and nitrosobenzene in the presence of a catalytic amount of TESOTf (nitrosoaldol reaction).29 Usually, nucleophiles react with nitrosobenzene without Lewis acid to give the N adduct predominantly. In contrast, they reported that the reaction of silyl enol ethers and nitrosobenzene catalyzed by TESOTf afforded the R2 N PPh2
MeO
R
OMe
OMe
AgOAc, 2-PrOH
+ O
R = aromatic
O
1
R N
H N
OTMS
R1 = H or Me
NH
THF, -78°C
R R2 = (s)-Bu, t-Bu, Me
R1
Ot-Bu O 79–>98% ee >98% de
Scheme 9.18. Vinylogous Mannich reaction with silyloxyfurans.
278
ALDOL AND RELATED PROCESSES
corresponding O adducts exclusively. They have extended this reaction to a regio- and enantioselective version.30 The reaction was conducted with nitrosobenzene and tin enolates in the presence of the BINAP–AgOTf catalyst to give corresponding N adducts or O adducts selectively with high enantioselectivity (Tables 9.12 and 9.13). It is noteworthy that the regioselectivity (N vs. O) can be controlled by changing the ratio of BINAP to silver salt. For example, when the reaction was conducted with trimethyltin enol ether and nitrosobenzene in the presence of a 1 : 1 mixture of a Tol–BINAP and AgOTf in THF, the O adduct was obtained as major product with high enantioselectivity (Table 9.12). The O adducts were converted into chiral alcohols with a catalytic amount of CuSO4 without the loss of enantioselectivity (Scheme 9.19).
TABLE 9.12. O-Nitrosoaldol Reaction Using Tin Enolates and Nitrosobenzene (Tol)2 P AgOTf P
Ph
O N
+
1
R
OSnR3 R2
O
(Tol)2 1
R
THF, -78°C
R3
O
O
+
NHPh R2 R3 O adduct
OH N Ph R 2 R R3 N adduct 1
>99 : 1
Entry
Enolate
1
2
3
4
Yield/ee (%)
Entry
OSnMe3
88/99
5
OSnBu3 Ph
97/88
6
OSnMe3
93/92
94/87
7
OSnMe3
95/82
92/90
8
OSnMe3
OSnMe3
Enolate OSnMe3
OSnMe3
Yield/ee (%) 96/85
92/94
279
NITROSOALDOL REACTION
O
O O
NHPh
OH
cat. CuSO4 MeOH, 0°C
Ph N N Ph
Scheme 9.19. Cleavage of the N–O bond.
In contrast, the reaction of tributyltin enol ethers and nitrosobenzene in the presence of a 1 : 2 mixture of BINAP and AgOTf in ethylene glycol diethyl ether afforded the N adduct predominantly with high enantioselectivity (Table 9.13). Momiyama and Yamamoto have determined the structures of silver–BINAP complex by an X-ray analysis and a 31 P NMR study30b. After their discovery, several other groups reported nitrosoaldol reactions using different catalysts including acid catalysts and organocatalysts.31 Although the asymmetric nitrosoaldol reaction was achieved with high enantioselectivity, the use of a stoichiometric amount of organostannane was unacceptable. In TABLE 9.13. N-Nitrosoaldol Reaction Using Tin Enolates and Nitrosobenzene
Ph
O N
+
R1
Ph Ph P AgOTf P AgOTf Ph Ph
OSnBu3 R2 R3
O R1
EtOCH2CH2OEt, -78°C
OH N Ph R2 R3
O +
N adduct
Entry
Enolate
Yield (%)
N:O
R1
O NHPh R2 R3 O adduct
ee (%)
1 2 3
OSnBu3 n=1 2 3 n
90 95 96
97 : 3 96 : 1 >99 : 1
86 >99 97
4
OSnBu3 Ph
94
>99 : 1
77
5
OSnBu3
97
>99 : 1
98
280
ALDOL AND RELATED PROCESSES
2008, Yamamoto et al. reported the asymmetric O-nitrosoaldol reaction using silyl enol ethers in the presence of the silver catalyst.32 In order to achieve this reaction, they developed a novel combination of silver and a chiral phosphite derived from BINOL. The disilanyl enol ether was used to ensure high yield and enantioselectivity. The reaction was conducted with disilanyl enol ether and nitrosobenzene in the presence of AgBF4 and the chiral phosphite ligand in THF to produce the O adduct with high regio- and enantioselectivity (Table 9.14). In addition, a chiral silyl enol ether could be used as a substrate. The reaction was conducted with chiral silyl enol TABLE 9.14. O-Nitrosoaldol Reaction Using Silyl Enol Ethers and Nitrosobenzene Ph O P OPh O OSiMe2TMS R3 PhNO + R1 R2
Entry
THF, MeOH, -78°C
OSiMe2TMS
2
R=H R = Me R = OCH2CH2O
3 R
R
O
AgBF4, CsF
Silyl enol ether
1
4
Ph
OSiMe2TMS
O
R1
NHPh
2
R R3
Yield (%)
ee (%)
85
95
72
98
85
96
66
97
41
90
37
64
84
92
99
79
80
76
O
5
OSiMe2TMS
6 n
7
8 9
n=1 n=3
OSiMe2TMS
OSiMe2TMS Ar Ar = Ph Ar = 2-naphthyl
281
CONCLUSION
OSiMe2TMS PhNO
ligand
THF, MeOH, -78°C 91%
Ph
Ph NHPh
O P OPh
Ph
O
dr = >99 : 1 99% ee
99% ee
OSiMe2TMS PhNO
O
AgBF4, CsF
+
ligand
O
ligand AgBF4, CsF
+
THF, MeOH, -78°C 70%
Ph
Ph
O O
NHPh
Ph dr = 91 : 9 99% ee
99% ee
Scheme 9.20. Diastereoselective O-nitrosoaldol reaction.
ether in the presence of the catalyst to give O adduct with high diastereoselectivity. Remarkably, the stereochemical outcome of nitrosoaldol reaction could be controlled by the catalyst regardless of the configuration of the stereocenter in silyl enol ether (Scheme 9.20). 9.8
ALDOL REACTION WITH AZODICARBOXYLATE
Azodicarboxylates are attractive aminating reagents, since a-aminocarbonyl compounds are not only useful building blocks in organic synthesis, but also essential tools for biological studies.33 Hence, the development of general and efficient methods for the preparation of these compounds is an important issue in organic synthesis. Kobayashi and coworkers have developed an amination reaction using a silyl enol ether and an azodicarboxylate (Scheme 9.21).34 The reaction was conducted with trimethylsilyl enol ether and azodicarboxylate in the presence of the AgClO4–BINAP complex to afford the corresponding adducts with high enantioselectivities. 9.9
CONCLUSION
The allylation reaction of carbonyl compounds is a classical but excellent method for constructing highly functionalized organic molecules. More recently, allylation OTMS R3 R1 R2
+ BnO2C
N N
CO2Bn
O
1. (R)-BINAP, AgClO4, -45°C 2. HF, THF
R1
HN N
CO2Bn
CO2Bn R2 R3 ≤ 86% ee
Scheme 9.21. Amination reaction using azodicarboxylate.
282
ALDOL AND RELATED PROCESSES
reactions have been rapidly extended to asymmetric versions by a combination of various catalysts and reagents. Those catalysts can also be applied for the crotylation reaction in a highly diastereo- and enantioselective manner. In addition to the allylation reaction, we have also described the aldol reaction. The aldol reaction is one of the most powerful methods for preparing 1,3-oxygenated organic compounds from carbonyl compounds, and the aldol reaction has been applied to the synthesis of numerous natural products and drugs. This reaction has also been broadened to useful variants that use nitroso compounds or azodicarboxylates instead of carbonyl compounds. The study of these reactions is just beginning, and improvements are to be expected in the near future. Although many catalysts have been developed for these reactions, most of them are Lewis acid catalysts that are typically sensitive to moisture. The silver catalysts we mentioned above might become a preferable choice since silver catalysts can be used in water and many silver salts are commercially available and inexpensive. The chemistry of silver catalysts in this context can be expected to expand significantly in the future. REFERENCES 1. (a) Yamamoto, Y.; Asao, N., Chem. Rev. 1993, 93, 2207–2293; (b) Roush, W. R., in Methoden der Organischen Chemie (Houben-Weyl), Helmchen, G.; Hoffmann, R. W.; Mulzer, J.; Shaumann, E., eds., Thieme, Stuttgart, 1996, Vol. 3 E21, pp. 1410–1486; (c) Bubnov, Y., in Science of Synthesis, Kaufmann, D. E.; Matteson, D. S., eds., Thieme, Stuttgart, 2004, Vol. 6, pp. 945–1072; (d) Ramachandran, P. V., Aldrichim. Acta 2002, 35, 23–35; (e) Brown, H. C.; Ramachandran, P. V., J. Organomet. Chem. 1995, 500, 1–19; (f) Ishihara, K., in Lewis Acids in Organic Synthesis, Yamamoto, H., ed., Wiley-VCH, Weinheim, 2000, Vol. 1, pp. 176–179; (g) Mikami, K.; Terada, M., in Lewis Acids in Organic Synthesis, Wiley-VCH, Weinheim, 2000, Vol. 2, pp. 800–805. 2. (a) Yanagisawa, A.; Nakashima, H.; Ishiba, A.; Yamamoto, H., J. Am. Chem. Soc. 1996, 118, 4723–4724; (b) Yanagisawa, A.; Nakashima, H.; Nakatsuka, Y.; Ishiba, A.; Yamamoto, H., Bull. Chem. Soc. Jpn. 2001, 74, 1129–1137. 3. Yanagisawa, A.; Ishiba, A.; Nakashima, H.; Yamamoto, H., Synlett 1997, 88–90. 4. (a) Yamamoto, Y.; Yatagai, H.; Naruta, Y.; Maruyama, K., J. Am. Chem. Soc. 1980, 102, 7107–7109; (b) Yamamoto, Y.; Yatagai, H.; Ishihara, Y.; Maeda, N.; Maruyama, K., Tetrahedron 1984, 40, 2239–2246. 5. Yanagisawa, A.; Nakatsuka, H.; Nakashima, H.; Yamamoto, H., Synlett 1997, 933–934. 6. Loh, T.-P.; Zhou, J.-R., Tetrahedron Lett. 2000, 41, 5261–5264. 7. Cesarotti, E.; Araneo, S.; Rimoldi, I.; Tassi, S., J. Mol. Catal. A. 2003, 204–205, 221–226. 8. Wang, C.-J.; Shi, M., Eur. J. Org. Chem. 2003, 2823–2828. 9. Colombo, F.; Annunziata, R.; Benaglia, M., Tetrahedron Lett. 2007, 40, 2687–2690. 10. Yanagisawa, A.; Kageyama, H.; Nakatsuka, Y.; Asakawa, K.; Matsumoto, Y.; Yamamoto, H., Angew. Chem. Int. Ed. 1999, 38, 3701–3703. 11. Wadamoto, M.; Ozasa, N.; Yanagisawa, A.; Yamamoto, H., J. Org. Chem. 68, 2003, 5593–5601.
REFERENCES
283
12. Wadamoto, M.; Yamamoto, H., J. Am. Chem. Soc. 2005, 127, 14556–14557. 13. Gonzalez, L. J-.; Munoz, S. G-.; Corral, M. A-.; Dorado, M. M-.; Garcia, I. R., Chem. Eur. J. 2006, 12, 8762–8769. 14. Mahrwald, R., ed., Modern Aldol Reactions; Wiley-VCH, Weinheim, 2004, Vols.1, 2. 15. (a) Gennari, C., in Comprehensive Organic Synthesis, Trost, B. M.; Fleming, I.; Heathcock, C. H., eds., Pergamon Press, Oxford, 1991, Vol. 2, p. 629; (b) Hollis, T. K.; Bosnich, B., J. Am. Chem. Soc. 1995, 117, 4570–4581; (c) Braun, M., in Methoden der Organischen Chemie (Houben-Weyl), Helmchen, G.; Hoffmann, R. W.; Mulzer, J.; Shaumann, E., eds., Thieme, Stuttgart, 1995, Vol. E21, pp. 1410–1486; (d) Nelson, S. G., Tetrahedron: Asymmetry 1998, 9, 357–389; (e) Gr€oger, H.; Vogl, E. M.; Shibasaki, M., Chem. Eur. J. 1998, 4, 1137–1141; (f) Mahwald, R., Chem. Rev. 1999, 99, 1095–1120; (g) Carreira, E. M., in Comprehensive Asymmetric Catalysis, Jacobsen, E. N.; Pfaltz, A.; Yamamoto, H.;eds., Springer, Heidelberg, 1999, Vol. 3, p. 997; (h) Arya, P.; Qin, H., Tetrahedron 2000, 56, 917–947; (i) Machajewski, T. D.; Wong, C. -H., Angew. Chem. Int. Ed., 2000, 39, 1353–1357; (j) Carreira, E. M., in Modern Carbonyl Chemistry, Otera, J., ed., Wiley-VCH, Weinheim, 2000, Chapter 8, p. 227. 16. Yanagisawa, A.; Matsumoto, Y.; Nakashima, H.; Asakawa, K.; Yamamoto, H., J. Am. Chem. Soc. 1997, 119, 9319–9320. 17. (a) Yanagisawa, A.; Matsumoto, Y.; Asakawa, K.; Yamamoto, H., J. Am. Chem. Soc. 1999, 121, 892–893; (b) Yanagisawa, A.; Matsumoto, Y.; Asakawa, K.; Yamamoto, H., Tetrahedron 2002, 58, 8331–8339. 18. Ohkouchi, M.; Masui, D.; Yamaguchi, M.; Yamagishi, T., J. Mol. Catal. A Chem. 2001, 170, 1–15. 19. (a) A. Yanagisawa, A.; Y. Nakatsuka, Y.; A. Asakawa, A.; H. Kageyama, H.; H. Yamamoto, H., Synlett 2001, 69–72; (b) Yanagisawa, A.; Nakatsuka, Y.; Asakawa, K.; Wadamoto, M.; Kageyama, H.; Yamamoto, H., Bull. Chem. Soc. Jpn. 2001, 74, 1477–1484. 20. Wadamoto, M.; Ozasa, N.; Yanagisawa, A.; Yamamoto, H., J. Org. Chem. 2003, 68, 5593–5601. 21. Ozasa, N.; Wadamoto, M.; Ishihara, K.; Yamamoto, H., Synlett 2003, 2219–2221. 22. Kundu, K.; Doyle, M. P., Tetrahedron: Asymmetry 2006, 17, 574–577. 23. Akullian, L. C.; Snapper, M. L.; Hoveyda, A. H., J. Am. Chem. Soc. 2006, 128, 6532–6533. 24. (a) Kobayashi, S.; Ishitani, H., Chem. Rev. 1999, 99, 1069–1094; (b) Co´rdova, A., Acc. Chem. Res. 2004, 37, 102–112; (c) Friestad, G. K.; Mathies, A. K., Tetrahedron 63, 2007, 2541–2569. 25. (a) Ferraris, D.; Young, B.; Cox, C.; Dudding, T.; Drury, W. J. III; Ryzhkov, L.; Taggi, A. E.; Lectka, T., J. Am. Chem. Soc. 2002, 124, 67–77; (b) Ferraris, D.; Young, B.; Dudding, T.; Lectka, T., J. Am. Chem. Soc. 1998, 120, 4548–4549. 26. Josephsohn, N. S.; Snapper, M. L.; Hoveyda, A. H., J. Am. Chem. Soc. 2004, 126, 3734–3735. 27. Josephsohn, N. S.; Carswell, E. L.; Snapper, M. L.; Hoveyda, A. H., Org. Lett. 2005, 7, 2711–2713. 28. (a) Carswell, E. L.; Snapper, M. L.; Hoveyda, A. H., Angew. Chem. Int. Ed. 2006, 45, 7230–7233; (b) Wieland, L. C.; Vieira, E. M.; Snapper, M. L.; Hoveyda, A. H., J. Am. Chem. Soc. 2009, 131, 570–576. 29. Momiyama, N.; Yamamoto, H., Angew. Chem. Int. Ed. 2002, 41, 2986–2988.
284
ALDOL AND RELATED PROCESSES
30. (a) Momiyama, N.; Yamamoto, H., J. Am. Chem. Soc. 2003, 125, 6038–6039; (b) Momiyama, N.; Yamamoto, H., J. Am. Chem. Soc. 2004, 126, 5360–5361. 31. (a) Yamamoto, H.; Momiyama, N., Chem. Commun. 2005, 3514–3525; (b) Janey, J. M., Angew. Chem. Int. Ed. 2005, 44, 4292–4300; (c) Yamamoto, H.; Kawasaki, M., Bull. Chem. Soc. Jpn. 2007, 80, 595–607. 32. Kawasaki, M.; Li, P.; Yamamoto, H., Angew. Chem. Int. Ed. 2008, 47, 3795–3797. 33. (a) Greck, C.; Drouillat, B.; Thomassingny, C., Eur. J. Org. Chem. 2004, 1377–1385; (b) Erdik, E., Tetrahedron 2004, 60, 8747–8782. 34. Yamashita, Y.; Ishitani, H.; Kobayashi, S., Can. J. Chem. 2000, 78, 666–672.
10 COUPLING REACTIONS PROMOTED BY SILVER JEAN-MARC WEIBEL, AURELIEN BLANC,
AND
PATRICK PALE
Laboratory for Organic Synthesis and Reactivity Institute of Chemistry, University of Strasbourg, France
10.1 10.2 10.3 10.4 10.5
Introduction sp3–sp3 Coupling Reactions Promoted by Silver Salts sp3–sp2 Coupling Reactions Promoted by Silver Salts sp3–sp Coupling Reactions Promoted by Silver Salts sp2–sp2 Coupling Reactions Promoted by Silver Salts 10.5.1 Homocoupling of Vinyl- or Arylsilver Species 10.5.2 Organosilver Species as Nucleophilic Reagents 10.5.3 Silver as a Lewis Acid Reagent 10.5.4 Silver as a Halogen Scavenger 10.5.4.1 Silver in Pd-Catalyzed Couplings 10.5.4.2 Silver in Pd(II)-Promoted Electrophilic Substitution of Arenes (C--H Activation) 10.5.4.3 Silver as Reagent for Decarboxylative Coupling 10.5.4.1 Silver as a Halogen Scavenger 2 10.6 sp –sp Coupling Reactions Promoted by Silver Salts 10.6.1 Organosilver Species as Nucleophilic Reagents 10.6.2 Organosilver Species in Transmetallations 10.6.3 Silver as a Lewis Acid Reagent 10.6.4 Organosilver Species as Intermediates in Catalyzed Enyne or Arylyne Synthesis 10.7 sp–sp Coupling Reactions Promoted by Silver Salts 10.8 Conclusion References
Silver in Organic Chemistry Edited by Michael Harmata Copyright 2010 John Wiley & Sons, Inc.
285
286
10.1
COUPLING REACTIONS PROMOTED BY SILVER
INTRODUCTION
Carbon–carbon bond formation is clearly a key transformation in organic synthesis. Since the early times of organic chemistry, numerous methods have been developed to establish C--C bonds. Among them, those using organometallic compounds constitute the vast majority. Originating near the end of the nineteenth century with Frankland and organozincs,1 and continuing into the twenthieth century with Grignard and his reagents, methods based on stoichiometric amounts of organometallics established themselves as key tools in organic synthesis.2 Looking for milder conditions and for modulation in reactivity, chemists explored other metals, especially transition metals, usually through transmetalation. During this quest, the peculiar behavior of group VIII metals, especially palladium, was noticed; and catalytic reactions were developed, leading to a burst in this chemistry in the early 1980s.3 More recently, other metals of this group, especially coinage metals,4 also proved useful in carbon–carbon bond formation, either as such or in combination with other metals, mainly palladium. Although among the first organometallics ever produced,5 organosilver compounds were applied to organic chemistry only sparingly.6 However, these reagents exhibit an extreme mildness and a very low basicity, rendering them very useful in various transformations.6–8 Indeed, since the end of the twentieth century, silver salts have increasingly garnered interest as catalysts or cocatalysts in various reactions in which C--C bonds are created.7a,8 In these reactions, preformed organosilver species can be directly used as reagents, and organosilvers are often produced as transient intermediates in catalyzed reactions, but silver could also play different roles, such as a Lewis acid or halide scavenger. In this chapter, the Ag-promoted formation of C--C bonds is described, with the presentation organized according to the hybridization of the carbons involved in the bond formation. 10.2
sp3–sp3 COUPLING REACTIONS PROMOTED BY SILVER SALTS
One of the earliest, if not the first, uses of silver in C--C bond formation was the homocoupling reaction originally described by Gardner et al. and further investigated by Kharash and Reinmuth and by Kochi et al.9–11 In this reaction, Grignard reagents in THF solution reacted with stoichiometric amounts of various silver salts leading to homocoupling of alkyl groups (Scheme 10.1). Interestingly, this reaction could be performed with catalytic amounts of silver provided that the nitrate counterion was present. The latter could be obtained from silver nitrate or by addition of lithium nitrate to silver bromide. Mixtures of alkanes were obtained starting from two different organomagnesiums, suggesting radical formation. R–MgX
+
AgY
2 R–Ag
R–Ag R–R +
Scheme 10.1
+
MgXY 2 Ag
sp3–sp3 COUPLING REACTIONS PROMOTED BY SILVER SALTS
MgBr
O O
287 O
Br
Br 1.2 equiv
O
AgOTs 0.01 equiv THF, r.t.
O O 80%
Scheme 10.2
However, their nonstatistical distribution as well as retention of stereochemistry if alkenyl organometallics were used (see Scheme 10.14),12 suggested that free radicals were not intermediates in such reactions. Such Wurtz-type couplings therefore probably proceeded by transmetallation and then recombination (Scheme 10.1). More recently, this reaction was slightly modified using the more soluble silver tosylate as catalyst (1 mol%) and 1,2-dibromoethane as reoxidant (Scheme 10.2).13 In a more interesting way, nonsymmetric couplings have been achieved much more recently.14 Tertiary and secondary alkyl halides readily reacted with benzyl or allyl Grignard reagents in the presence of silver nitrate, giving the corresponding benzyl or allyl derivatives in high yields (Scheme 10.3). In competitive experiments, tertiary alkyl halides proved more reactive than the secondary ones. Electronic effects on benzyl Grignard reagents did not modify reactivity. Experimental evidence suggested radicals as intermediates in this reaction. Instead of starting from Grignard reagents, Johnson15 and then Brown et al.16–18 showed that alkylboranes, readily accessible through hydroboration of the corresponding alkenes, also reacted in the presence of silver nitrate or oxide and produced alkanes resulting from homocoupling. The presence of hydroxide brought a very rapid reaction at room temperature or below. In such reactions, alkylboranes are probably transmetallated to silver and the thus-formed alkylsilver reacts to give the corresponding homocoupling product (Scheme 10.4). More functionalized boranes have also been coupled in the presence of silver oxide, but the reaction proved more efficient when chromous chloride was added as catalyst (Scheme 10.5).19
MgBr
R1
R2
AgNO3 0.1 equiv
X
Et2O, r.t.
R2
R1 66–88%
X= Cl, Br
Scheme 10.3
BH3
Ag2O 1.5 equiv
B 3
NaOH 3 equiv < 20°C
3
Scheme 10.4
Ag
20°C
71%
288
COUPLING REACTIONS PROMOTED BY SILVER
OMe
OH B OH
MeO
Ag2O 3 equiv
MeO
OMe
CrCl2 0.05 equiv THF, 65°C
MeO
MeO
72%
Scheme 10.5 Ag2O 2 equiv
Ph
KOH 2 equiv
OH + Br B OH
Ph
PdCl2(dppf) 0.03 equiv dioxane, 80°C
71%
Scheme 10.6
As before, nonsymmetric couplings could also be realized from boronic acids, but using a palladium catalyst instead of chromium(II) chloride. Silver oxide was used as an activator for the Pd-catalyzed coupling of cyclopropylboronic acids to allyl bromides (Scheme 10.6).20 In a related reaction, Caddick et al. expanded this Suzuki-type coupling to other alkyl–alkyl couplings.21 Although the yields were still modest, the reaction seemed compatible with various functional groups and required only mild conditions (Table 10.1).
TABLE 10.1. Suzuki–Miyaura Cross-Coupling of Alkyl Bromides R1
Br
n
+
BBN
R2
R1
AgOTf 0.04 equiv
R2
n
Pd(dba)2 0.04 equiv IPr.HCl 0.08 equiv THF, 40°C tBuOK
iPr
Cl N
iPr
iPr
IPr.HCl =
n 10
R
1
CH3
R
2
Product
Yield (%)
10
n-Hex
iPr
N
56
5
10
10
4 4
CH3
CH2Ph-pOMe
COOEt
n-Bu
CN
n-Bu
46 MeO 4
EtOOC
53
4 4
NC 4
52
sp3–sp2 COUPLING REACTIONS PROMOTED BY SILVER SALTS
10.3
289
sp3–sp2 COUPLING REACTIONS PROMOTED BY SILVER SALTS
No coupling reactions connecting sp2 and sp3 carbons directly promoted by silver salts seem to have been reported. However, silver salts have been used as activators or cocatalysts in a few palladium-catalyzed couplings. The Caddick protocol set up for alkyl–alkyl coupling (see Section 10.2) could also be used for the coupling of a vinylborane with an alkyl bromide (Scheme 10.7). The yield was, however, modest.17 The synthesis of medicinally relevant 4-substituted 2(5H)-furanones has been achieved by a coupling reaction of the triflate derived from b-tetronic acid with alkyland cyclopropylboronic acids (Scheme 10.8).22 This reaction was mediated by bis(acetonitrile)dichloropalladium and silver oxide. Interestingly, optically active cyclopropyl boronic acids were not racemized during this coupling reaction. In 2006, Yu et al. combined pyridinyl-directed C--H activation and C--C bond formation with alkylboronic acids (see Section 10.5.4.2).23 The success of this transformation relied on the combination of palladium acetate (10 mol%), benzoquinone (1 equiv), and silver oxide or carbonate (0.5 equiv) in a protic solvent, but an excess of boronic acid (3 equiv) was required (Scheme 10.9). Interestingly, in this reaction silver oxide played a dual role as promoter for the transmetallation step and as cooxidant with benzoquinone. BBN
Ph
AgOTf 0.04 equiv
+ 10
Br
Ph
Pd(dba)2 0.04 equiv i -Pr.HCl 0.08 equiv THF, 40°C t-BuOK
10
37%
Scheme 10.7 TfO Ph
B(OH)2
Ag2O
+ O
O
O
Ph
PdCl2(MeCN)2
O
THF, 70°C
66% ee 91%
Scheme 10.8
HO N H
+
B R HO
Ag2O 1 equiv benzoquinone 0.5 equiv
N Pd(OAc)2 0.1 equiv t-amyl alcohol
R
air, 100°C, 6 h
R = Me 67% R = nBu 75% R = cPr 52%
Scheme 10.9
290
COUPLING REACTIONS PROMOTED BY SILVER
R
R HO +
Ag2O 1.0 equiv benzoquinone
B R1
HO
N
Pd(OAc)2, L, THF, 60°C
R
N R1 R
CO2H L=
HN
O
43–96% ee 54–95%
-(–)-menthyl
O
Scheme 10.10
In 2008, the same group developed an asymmetric version of this reaction (Scheme 10.10).24 Run under similar conditions, but with more silver oxide (1 equiv) and thus less reoxidant (benzoquinone 0.5 equiv) and in the presence of catalytic amounts of chiral ligand (20 mol%), the best enantiomeric excesses and yields were obtained with menthyl-L-leucine ester as the chiral ligand. 10.4
sp3–sp COUPLING REACTIONS PROMOTED BY SILVER SALTS
The title reactions refer to alkylation reactions using alkynyl silvers as reagents. It seems that no catalytic version has so far been reported, and only a few examples of stoichiometric reactions are known.8 As any organometallic derivatives of terminal alkynes, silver acetylides readily reacted with alkyl halides, leading to the corresponding disubstituted acetylenes. For example, methyl iodide gave the methylacetylene derivative (Scheme 10.11, top).25 Surprisingly, silver arylacetylides react with adamantyl bromide or iodide, although refluxing in N-methylmorpholine (NMM) was required (Scheme 10.11, bottom).26 Some evidence suggested an SRN1 mechanism for this alkylation. Silver acetylides also reacted with epoxides yielding propargylic alcohols.27 As for related alkylations with silver acetylides (see Section 10.6.1), this reaction required zirconocene dichloride and catalytic amount of silver triflate. This method proved useful for both electron-rich and electron-deficient alkynes and compatible with various acid- and base-sensitive functional groups (Table 10.2).
MeI
Ag MeOOC
8
I +
MeOOC
Ag
13
Me
NMM reflux 62%
Cl
Cl
Scheme 10.11
sp2–sp2 COUPLING REACTIONS PROMOTED BY SILVER SALTS
291
TABLE 10.2. Epoxide Opening with Silver Acetylides
R1
Ag
+
OH
R3 R2
R2
AgOTf 0.2 equiv Cp2ZrCl2 1.2 equiv
O
R3
CH2Cl2, r.t.
R1
Epoxide
Product
7
Yield (%)
OH 7
O
CH2OTHP
R1
82 OTHP OH
O
n-Bu
53 nBu OH
O
COOMe
54 COOMe
OH
TBSO
n-Bu
TBSO
O
52
n-Bu OH
TBSO
COOMe
TBSO
46
O COOMe
Interestingly, 1,2-disubstituted epoxides led to a,a-disubstituted propargyl alcohols, indicating that 1,2 shifts and therefore epoxide-to-aldehyde rearrangements occurred during the reactions (Scheme 10.12). Mechanistic studies by NMR validated this rearrangement and also supported the transmetallation–abstraction mechanism as already proposed (Scheme 10.13).8
10.5
sp2–sp2 COUPLING REACTIONS PROMOTED BY SILVER SALTS
Such couplings are the most prominent C--C bond-forming reactions in which silver salts have been involved. In such reactions, silver could play a large variety of roles, OH Ph
Ph O
+ Ag
n
Bu
AgOTf 0.2 equiv Cp2ZrCl2 1.2 equiv CH2Cl2, r.t.
Scheme 10.12
Ph Ph
nBu
292
COUPLING REACTIONS PROMOTED BY SILVER O
n-Bu
Ph Ph
n-Bu
Ag
Ph
Ph OH
n-Bu
Cp2ZrCl
n-Bu
Cp2Zr
n-Bu
nBu
AgOTf Cp2Zr
Ph
Cp2Zr
Ph
O
O
Ph
Ph
Scheme 10.13
inducing homocoupling or acting as a Lewis acid or as a halide scavenger. Moreover, preformed organosilver species, depending on their stability, can be directly used as reagent for the formation of C--C bonds. 10.5.1
Homocoupling of Vinyl- or Arylsilver Species
Vinylsilvers could be obtained by transmetallation from the corresponding vinyllithiums.12 More stable than alkylsilvers (see Section 10.2 and Ref. 28), they nevertheless underwent thermal decomposition leading to homocoupling products (Scheme 10.14). This method was applied to 2- or 4-bromopyridine derivatives and used to produce the corresponding bipyridines, useful as intermediates for herbicide manufacture. Because arylsilvers are more stable, heating was required in this case (Scheme 10.15).29
R R
Ag(PBu3)I
Li
-78°C
R r.t.
Et2O
R
Ag
95%
Scheme 10.14
Mg N
Br
EtI, I2 Et2O, r.t.
AgBr
N
MgBr
PhCl, 60°C
N 60%
Scheme 10.15
N
sp2–sp2 COUPLING REACTIONS PROMOTED BY SILVER SALTS AgF, MeCN
K2 R
SiF5
R
or AgNO3, H2O–Et2O
R 47–72%
r.t.
R
293
Ag
Scheme 10.16
Interestingly, (E)-alkenylpentafluorosilicates, readily available from alkynes by hydrosilylation, could also be dimerized to the corresponding symmetric (E,E)-1,3dienes in good yields on treatment with silver fluoride in acetonitrile or with silver nitrate in water/ether (Scheme 10.16).30 This reaction suggested that pentafluorosilicates could be transmetallated to silver, leading to a vinylsilver that dimerized. Boranes are also a source of organosilvers through transmetallation (see Section 10.2) and thus, have been coupled in the presence of silver oxide, but the reaction proved more efficient when chromous chloride was added as catalyst.19 Arylboranes led to biaryls, and vinylboranes gave the corresponding dienes, but noticeably, with neither isomerization of the double bond nor cyclization (Scheme 10.17). Interestingly, these results confirmed other reports (see Section 10.2) suggesting the absence of free radicals as intermediates in the homocoupling of organosilver compounds. 10.5.2
Organosilver Species as Nucleophilic Reagents
Arylargentates are stable species. Their chemistry31 is so far not well developed, but they have been used as nucleophilic reagents, mainly in transmetallation processes.32,33 With (h3-allyl)palladium complexes, new complexes were formed, and, on addition of benzoquinone, reductive elimination occurred. In this reaction, the aryl group was transferred to the allyl system, providing allylbenzene derivatives (Scheme 10.18).
n-C5H11 n-C5H11 Ag2O 3 equiv
88%
CrCl2 0.05 equiv
B(OH)2
THF, 65°C
n-C5H11
B(OH)2
Ag2O 3 equiv CrCl2 0.05 equiv THF, 65°C
Scheme 10.17
95%
294
COUPLING REACTIONS PROMOTED BY SILVER
F R
Ag
Cl Pd
2
Nn-Bu 4
+
X
F
F
R Pd
CH2Cl2
F
81%
F
Nn-Bu 4 F
2
Cl
O
X = Cl, Br R or
=
Cl
R
C6F5
Cl
2
O
CDCl3 97%
F
Cl F
F
Cl
Scheme 10.18
10.5.3
Silver as a Lewis Acid Reagent
With its d10 electronic configuration and its f orbitals available, silver(I) ion can act as a Lewis acid,34 and relativistic effects reinforce such behavior.35 Moreover, silver salts are able to act as either or both a s-Lewis acid or a p-Lewis acid, with a slight preference for s coordination over p coordination, as revealed by calculations. This preference is higher for nitrogen than for oxygen Lewis bases (Scheme 10.19).36 On this basis, silver-catalyzed aldolization and allylation reactions have been developed, expanding the technology for creating a C--C bond using silver (see Chapter 9). Asymmetric versions have also been developed.7d,37 In related reactions, Snapper and Hoveyda et al. reported an asymmetric version of Mannich-type reactions catalyzed by silver acetate.38,39 Propargyl imines proved to be susceptible to nucleophilic addition of silylenol ethers derived from acetate esters in the
R
X
AgCl
R X
Ag+ Cl-
and/or
∆ Hf (kcal/mol) -22.6
Ph
-24.4
Ph Ph Ph
O
-26.4
NH
-39.6
R
X
Ag+ Cl-
∆Hf (kcal/mol) -25.2
Cy
-34.7
Cy Cy Cy
O
-26.0
NH
-40.4
Scheme 10.19 Computed heats of formation (B3LYP/SDD, kcal/mol) of various substrates with AgCl (from Ref. 36).
sp2–sp2 COUPLING REACTIONS PROMOTED BY SILVER SALTS
295
TABLE 10.3. Asymmetric Mannich Reactions of Enol Ethers with Alkynyl Imines MeO N + R
PPh2
R
O
OMe
Yield (%)
ee (%)
91 84
94 92
89
90
67
84
Ph SiMe3
O
H N
N
L*=
Ar OPh
L* 0.05 equiv i-PrOH, THF
OPh
R
HN
AgOAc 0.05 equiv
OTMS
SiMe3
OSiMe3
presence of silver acetate and a ligand derived from an amino acid (Table 10.3). In this reaction,chiralb-aminoesterswere obtainedin highyieldandgoodenantioselectivity. It is worth noting that propargyl imines are ideal substrates for both s and p coordination with silver ion, the dual Lewis acidities of silver acting in a cooperative way. This methodology was applied to the asymmetric synthesis of ()-sedamine (Scheme 10.20). An imine produced in situ from o-anisidine and methyl-5-oxopentanoate was alkylated with an enol silane in the presence of silver acetate as catalyst, OMe
OTMS + Ph
N CO2Me
OMe
AgOAc 0.05 equiv
NH O
L* 0.05 equiv THF, 4°C, 16 h 56%
MeO2C
Ph ee > 98%
NH
OH Ph
(–)-sedamine
Scheme 10.20
296
COUPLING REACTIONS PROMOTED BY SILVER
MeO MeO +
N Ar
H MeO2C N
AgOAc 0.011 equiv
O
OTMS
L* 0.01 equiv
Ar
(acidic workup)
CO2Me
O
72–95% H N
N
L*=
O
ee > 87%
O
PPh2
OMe
Scheme 10.21
together with a chiral phosphine ligand. This transformation gave a b-aminoketone in 56% yield with enantiomeric excess higher than 98%.39 Under similar conditions, the same authors were able to control two stereogenic centers in an asymmetric vinylogous Mannich reaction. Indeed, treatment of imines derived from aryl a-ketoesters with siloxyfuran under related conditions gave functionalized g-butenolides with high diastereo- and enantioselectivities (Scheme 10.21).40 Li et al. reported that b-diketones reacted with styrenes on treatment with silver triflate. The product structure suggested that such reactions proceeded through nucleophilic addition of the enol derived from b-diketones to the silver p complex of styrene (Scheme 10.22).41 Very recently, an intramolecular version with alkynyl b-diketones or b-diketoesters has been reported (Scheme 10.23).42 With such substrates, the reaction required a mixture of silver tetrafluoroborate and copper triflate as catalysts. Here, also, nucleophilic addition of a ketoenol to a p complex seemed responsible for an exo-dig cyclization, but depending on substituents, exocyclic or endocyclic alkenes were obtained. Similarly, b-diketones have been shown to regioselectively add to furans in the presence of silver triflate (Scheme 10.24).43
O R1
O R2
R1
AgOTf 0.1 equiv
+ R3
O
R3
DCE, 80°C
O R2 61–79%
R2 O
Ag+
H O
Ar R1
Scheme 10.22
sp2–sp2 COUPLING REACTIONS PROMOTED BY SILVER SALTS
O
297
O
R1
O CuOTf , AgBF4 0.1 equiv
R2
O R2
R1
DCE, 100°C
40–95%
Scheme 10.23
R2
R1
AgOTf 0.1 equiv
+ O
R1
DCE, 80°C
O
R2
O
O R1 = Me, R2 = H 60% R1 = Me, R2 = Ph 64% R1 = Ph, R2 = H 70%
O
Scheme 10.24
10.5.4
Silver as a Halogen Scavenger
One of the most prominent roles of silver in the formation of carbon–carbon bonds is due to another typical property of silver: its halophilicity. Silver ions exhibit strong Lewis acid–base interaction with halides, and silver halides are often insoluble materials, especially in organic solvents (Table 10.4). Silver salts have thus been used to abstract halogen(s) in organic or organometallic molecules and in metal complexes. In metal-catalyzed coupling reactions, the main use of silver is to generate a metal intermediate that has an unsaturated coordination sphere by halide abstraction. This resulting metal complex is thus able to coordinate other species, including unsaturated organic compounds. Depending on the nature of the ligand, various reactions can then occur (Scheme 10.25). A vast number of examples have been now described, and it is not possible to mention all of them here. Only a few are reported, emphasizing the probable role of silver.
TABLE 10.4. Molar Solubility Product (Ksp) of Common Silver Salts in Water at 25 C Silver Salts AgCl AgBr AgI AgCN AgOAc Ag3PO4 Ag2CO3 Source: Reference 44.
Ksp 1.77 1010 5.35 1013 8.52 1017 5.97 1017 1.94 103 8.89 1017 8.46 1012
298
COUPLING REACTIONS PROMOTED BY SILVER
Ag+ Z LnM
X
LnM
LnM
Z reactions
AgX
Scheme 10.25
10.5.4.1
Silver in Pd-Catalyzed Couplings
Such processes are well known in palladium chemistry and quite important, because of the role of palladium-catalyzed C--C bond formation in organic chemistry.3 Abstraction of chloride from a vinylpalladium complex by silver acetate has been reported,45 as well as halide abstraction by silver triflate from arylpalladium halides.46 More interestingly in the context of creating C--C bonds, silver perchlorate was able to promote the reaction of (h1-aryl)palladium bromide with dienes. Silver-promoted bromide abstraction led to the formation of (h1-h2-enyl)palladium complexes, which evolved further through the regioselective formation of a C--C bond between the aryl group and the former diene. Reactions with nonconjugated dienes suggested that the reaction proceeds via carbometallation of the less crowded double bond. Isomerization and b elimination led to a (h3-allyl)palladium complex (Scheme 10.26).47 In a similar but more complex reaction, dienes carrying a b-diketo side chain have been alkylated with various vinyl bromides and cyclized to the corresponding alkylated cyclopropyl alkenes in the presence of silver carbonate (Scheme 10.27). With a relative cis stereochemistry for the newly created C--C bonds, the mechanism is F iPr
F
Br Pd
F
F
F F 2
+
R1
F F
Br Pd
+ R
F
iPr
F
CH2Cl2
F
F
AgClO4
F
50%
Pd
Br
2 2 AgClO4 CH2Cl2
F
R1
R1 R
F
Pd PhF5
Br Pd
F5Ph Br
2
R
Scheme 10.26
COMe COMe
+
Ag2CO3 1 equiv
R1
Br R2
K2CO3 3 equiv Pd(PPh3)4 0.02 equiv
Scheme 10.27
R1 R2 COMe COMe
sp2–sp2 COUPLING REACTIONS PROMOTED BY SILVER SALTS
OH C6H13
I
+
Pd(OAc)2
C5H11
AgOAc DMF
299 C5H11
C6H13 80%
OH
Scheme 10.28
not obvious and probably different from those of the related reactions mentioned above; a palladium enolate has been proposed for this reaction.48 In the latter reactions, at least one step corresponded to a carbometallation reaction. Such carbometallations also arose in Heck reactions, and indeed, several conditions set up for such Heck reactions relied on silver salts.49 In these reactions, silver salts only seemed to abstract halide from intermediate palladium species; their role could, however, be more complex. The first example was described in the mid-1980s independently by Jeffery, Hallberg, and others in their work on Heck coupling. The former demonstrated the beneficial role of silver salts in the formation of dienols from allylic alcohols and vinyl iodides, as well as the role of counterion or phase transfer conditions (Scheme 10.28).50 The latter found that silver salts accelerated the reaction between aryl iodide and allyl or vinyl silanes, reinforcing regioselectivity and avoiding desilylation (Scheme 10.29).51 Such methodology was successfully applied to the synthesis of isopanepoxydone, a secondary metabolite isolated from the basidiomycete Panus conchatus (Scheme 10.30).52 The natural polyenic lactone ()-callystatin A, isolated in 1997 from the marine sponge Callyspongia truncata and exhibiting remarkable growth inhibition of cancer cells (IC50 10–20 pg/mL), was also prepared using this Heck-like reaction under Jeffery conditions (Scheme 10.31).53 As part of studies pertaining to the total synthesis of the complex hexacyclic alkaloid gelsemine, Overman and Abelman applied the Heck reaction to the cyclization of N-(2-halophenyl)enamides.54 They found that the addition of silver salt suppressed alkene isomerization, a frequent problem in Heck reactions, and dramatically accelerated the reaction (Scheme 10.32). These results led other investigators to propose that the Heck reaction proceeds via an ionic pathway or a neutral variant, depending on the starting materials and on the Pd(OAc)2, P(oTol)3
SiMe3
I +
32%
120°C SiMe3 55°C Pd(OAc)2, P(oTol)3 AgNO3
Scheme 10.29
SiMe3
55%
300
COUPLING REACTIONS PROMOTED BY SILVER
OAc
OAc Pd(OAc)2, Ag2CO3
Br +
O
OH
89%
OTBS
OH
O
DMF, 90°C
OTBS
OH OH
O
Isopanepoxydone O
Scheme 10.30 OTBS
OTBS I
Pd(OAc)2, AgOAc
+
DMF, r.t., 70%
OH
OH
OH O
O (–)-Callystatin A
O
Et
Scheme 10.31
conditions,49,55 although an alternative pathway that involves a PdII/PdIV redox system has been proposed.49,56 The ionic pathway would be rendered favorable when silver salts are used, or when starting from vinyl or aryl triflates.57 Controlling the coordination of palladium intermediates with silver salts was the key to developing highly diastereo- and enantioselective Heck reactions.58 For example, the presence of silver phosphate in the key cyclization step during the synthesis of various alkaloids reversed the stereochemistry at the newly created spiro center (Scheme 10.33).59 O
Me N
O
Me N
Pd(OAc)2 PPh3 NEt3 MeCN, reflux
O N Me
I
MeCN, r.t. NEt3 PPh3 Pd(OAc)2 AgNO3
Scheme 10.32
O
1:1
70%
Me N
sp2–sp2 COUPLING REACTIONS PROMOTED BY SILVER SALTS
301
O
R N
O NR Br O H
NR
NEt3
PG
N
Br Pd
Br
L
N PG
Pd2(dba)3
Br
PG N Pd2(dba)3
Br
Pd
NR
NR O
Ag3PO4 PG
O PG
N
N Br
Br
Scheme 10.33 R
PdOAc2(R-BINAP) 0.1 equiv Ag2CO3 2 equiv NMP, 60°C
I
R
H 55–74% 33–46% ee
R = COOMe, CH2OTBS, CH2OAc
Scheme 10.34
The Shibasaki cyclization of meso-cyclohexa-1,4-dienes in the presence of a chiral palladium complex and silver carbonate in 1-methyl-2-pyrrolidinone is probably the first example of an enantioselective Heck reaction (Scheme 10.34).60 The enantiomeric excess could be slightly improved by replacing the chiral phosphine ligand BINAP with the corresponding less coordinating arsine ligand, as well as replacing silver carbonate with silver phosphate.61,62 With such possibilities for regio-and stereocontrol, Pd–Ag-catalyzed Heck reactions have been and still are widely applied to total synthesis. However, just a few more examples are provided here. Brown et al. described a very convergent enantioselective synthesis of ()-galanthamine (Scheme 10.35), an Amaryllidaceae
BocMeN I
OH
BocMeN
OH
OH
Pd(OAc)2 AgCO3 ,dppf
O
O OMe
N
58%
OMe
Scheme 10.35
O OMe (–)-Galanthamine
302
COUPLING REACTIONS PROMOTED BY SILVER
OMe OHC
MeO N Boc
O Br
CHO
Pd(OAc)2
O
AgCO3, dppf
82%
TBSO
Br
N Boc
TBSO
MeO
MeO
MeO Pd(OAc)2
O
O
O
AgCO3, dppf
N Boc
H 44%
TBSO
NMe
N Boc
TBSO
HO (+)-Codeine
Scheme 10.36
alkaloid that is in clinical use for the symptomatic treatment of Alzheimers disease. Its preparation has been realized starting from isovanillin.63 The Heck reaction offered a convenient and ingenious way for the formation of three of the four rings of galanthamine. Hudlicky et al. described an elegant strategy based on two intramolecular Heck coupling reactions for the enantioselective synthesis of codeine from 2-phenylbromoethane (Scheme 10.36).64 Palladium/silver-catalyzed Heck reactions have usually involved vinyl or aryl halides and alkenes, but these reaction conditions were also extended to allenes. Indeed, Zenner and Larock65 showed that simple alkyl allenes readily reacted with aryl and vinyl iodide derivatives in the presence of palladium acetate or chloride and silver phosphate. Moreover, the reaction could be rendered asymmetric using chiral ligands; the best one was a bisoxazolidine derivative (Scheme 10.37). More recently, substituted allenyl sulfones were also engaged in Pd/Ag-catalyzed Heck-type reactions. Treated with palladium tetrakis(triphenylphosphine) and a mixture of silver and potassium carbonate in DMF at 80 C, such allenyl sulfones
H N
Ts
+
Ts N
Pd(OAc)2, L* 0.05 equiv
n-C8H17
•
n-C 8H17
Ag3PO4 0.4 equiv DMF, 100°C
I
O
L* =
62–90% ≤ 86% ee
O N
Ph
Scheme 10.37
N Ph
sp2–sp2 COUPLING REACTIONS PROMOTED BY SILVER SALTS
PhSO2
•
R1 R2
+
Ph
Pd(PPh3)4, Ag2CO3 0.05 equiv
I
303 PhSO2
R1 R2
•
K2CO3 4 equiv DMF, 80°C
Ph 62–76%
Scheme 10.38
gave more substituted allenyl sulfones (Scheme 10.38).66 The regioselectivity is very peculiar compared to similar processes already described in the literature; the roles of the sulfonyl group and the silver salt remain unclear. With such benefits offered by silver salts in these Heck-type reactions, chemists looked for similar improvements in other palladium-catalyzed couplings and related reactions. Stille reactions have been improved by the addition of silver salts. For example, 5-iodouracil could be coupled with stannylated pyridine, forming 5-(2-pyridinyl) uracil in good yield (Scheme 10.39).67 Under similar conditions, addition of silver oxide in dimethylacetamide allowed the production of hindered 1,2-substituted binaphthyl derivatives in high yields (Scheme 10.40).68 These compounds were useful precursors for the preparation of asymmetric reagents exhibiting axial chirality. Silver salts are also common activators in numerous Suzuki coupling reactions. The earliest example could probably be found in the Kishis palytoxin synthesis. Silver oxide as well as thallium hydroxide provided dramatic rate enhancements in the couplings of vinylboronic acids (Scheme 10.41).69 Both thallium and silver ions are most probably abstracting halide in palladium intermediates, but silver is clearly the most efficient. Moreover, with the right counterion, the silver salt also acts as a mild base and activator. O I + N
N H
Snn-Bu3
O NH O
NH
Ag2O PdCl2(PPh3)2 DMF
O N
NH 68%
Scheme 10.39
iPrO
O
I
SnMe3
MeO
MeO +
Ag2O
COOiPr
Pd(PPh3)4 DMA, 150°C
81%
Scheme 10.40
304
COUPLING REACTIONS PROMOTED BY SILVER I
TBSO
OTBS OTBS
TBSO
OAc OAc
OAc OAc
+ O
(HO)2B
TBSO
Pd(PPh3)4 0.25 equiv
OAc
O
THF, r.t. additive
OAc
OAc
OTBS
TBSO
OH
OTBS
OAc OH
KOH
2h
86%
TlOH
30 min
94%
Ag2O
5 min
92%
Scheme 10.41
In both Stille and Suzuki couplings, the role of silver salts may be not only as halide scavengers but also as promoters in the transmetallation step. In both coupling types, rate enhancements due to silver could be observed (see, e.g., Scheme 10.41). Indeed, the role that silver could play in Stille reactions could be similar to the one observed with copper.70 In the Suzuki reaction, the boronic acid species must be activated in order to transfer its organic moiety in the transmetallation step, and an oxygen atom intervenes in this process.71,72 The silver ion, and especially its basic counterion, could act in this capacity. For example, silver oxide was used to good effect in the synthesis of ( þ )-fostriecin to promote the Suzuki cross-coupling, finalizing the assembly of the carbon backbone (Scheme 10.42).73 Interestingly, Hiyama couplings, similar to Stille and Suzuki couplings but involving silanes instead of stannanes or borates, required stoichiometric amounts of silver salts.74 However, the silanes must be activated to obtain effective transformations, and usually, vinyl- or arylsilanols are used in such couplings. Among the silver salts, the oxide proved again to be the most active promoter (Scheme 10.43).75
HO O
OSitBuPh2 Br
O OTBS +
SitBuPh2 Ag2O Pd(PPh3)4, THF, 65°C, 8 h
O
B
O
OH O O
OSitBuPh2
O
OSitBuPh2
OTBS
2-O PO 3
O
74%
OH
O OH (+)-Fostriecin
Scheme 10.42
OH
sp2–sp2 COUPLING REACTIONS PROMOTED BY SILVER SALTS
Si
OH
305 Ph
Pd(PPh3)4 0.05 equiv
+ Ph–I
Ag salt 1 equiv
MeO
MeO
THF, 60°C
CuO AgNO3 AgOTf AgBF4 Ag2O
0% 16% 21% 23% 80%
Scheme 10.43
O Ag X
Ag
R Pd
OH C Si
Scheme 10.44
The free hydroxyl group at the silicon center played a key role in this reaction, since aryltrimethylsilane and aryldimethylsilyloxysilane did not react while silanediols or triols76 and polyalkenylsiloxanes did.77 As before, a silver ion and its counterion could act together in the transmetallation step, such as depicted in Scheme 10.44. Such a hypothesis seemed supported by the discovery in 2006 that 2-pyridinyl, but not 3-pyridinyl, allylsilanes reacted under similar conditions and transferred the pyridine moiety to various aryl iodides (Scheme 10.45).78 In this coupling, the nitrogen atom of the pyridinyl moiety probably played a role similar to that of the hydroxyl group in the silanol-based Hiyama couplings. In an analogous way, 2-trimethylsilylpyridine has been engaged in coupling with aryl iodides in the presence of a stoichiometric amount of silver oxide and a catalytic amount of tetrabutylammonium fluoride and palladium tetrakistriphenylphosphine.79 Only the iodides substituted by electron-withdrawing groups gave good 2-arylpyridines yields (Scheme 10.46). Interestingly, the concomitant formation of 2,20 dipyridine suggested the intermediate formation of 2-pyridinyl silver (see Scheme 10.15, Section 10.5.1). I
N
Si
Pd(PPh3)4 0.05 equiv
+ R
Ag2O 1.5 equiv THF, 60°C
Scheme 10.45
R N 59–93%
306
COUPLING REACTIONS PROMOTED BY SILVER
I + N
R
Pd(PPh3)4, Ag2O
R
TBAF
SiMe3
DMF, 90°C
N R = p-NO2 75% R = p-OMe 45%
Scheme 10.46 AgNO3 1 equiv
Cl2Pd
+
2
AcOH, air, 20 h
99%
Scheme 10.47
10.5.4.2 Silver in PdII-Promoted Electrophilic Substitution of Arenes (C--H Activation) During their work on the arylation of aromatic compounds by substitution, Fujiwara, et al. observed biaryl formation when aromatic compounds were placed in the presence of olefin–palladium complexes and silver nitrate.80 Developing this reaction as a method for biphenyl synthesis, these authors showed that the more stable the olefin–palladium complex was, the lower the yield. Ethylene dichloropalladium proved to be the best choice, when used with silver nitrate. However, the reaction required stoichiometric amounts of both “catalysts” (Scheme 10.47). Benzene derivatives substituted by electron-donating or -withdrawing groups reacted as well, but a mixture of regioisomers was produced, except for nitrobenzene, which only gave m,m0 -dinitrobiphenyl. Similarly, the vinylation of benzene was achieved on treatment with silver acetate and vinylpalladium(II) complexes (Scheme 10.48).45 In these reactions, silver ion abstracted chloride from the starting palladium species, rendering the coordination of benzene possible. As a Lewis acid due to its oxidation state, PdII then initiated electrophilic addition to the aromatic system, leading to C--C bond formation after reductive elimination (Scheme 10.49). Surprisingly not used in organic synthesis for decades, this methodology is now emerging with the synthesis of biaryls and of functionalized arenes, heterocyclic or not, as the goal.81
Cl +
H PPh3
Cl Pd Ph3P Cl
Cl
AgOAc AcOH, reflux 1.5 h
80%
Scheme 10.48
Cl
sp2–sp2 COUPLING REACTIONS PROMOTED BY SILVER SALTS
H
Ag+
PdL2
X PdL2 PhH
R
307
R
PdL2
PdL2
R
R
R
Scheme 10.49 OMe OMe
PdCl2(PPh3)2 0.03 equiv
I +
X S
AgF 1 equiv
EtO2C
X=N X=C
DMSO
X
S
46% 68%
CO2Et
Scheme 10.50
In 2005 and 2006, Mori et al. reported the arylation of heterocycles with aryl iodides using palladium dichloride complexes as catalysts and silver nitrate or fluoride as activators (Scheme 10.50). The best coupling conditions were obtained with a combination of silver nitrate and potassium fluoride.82 With organic thin-film transistors, light-emitting devices, and dye-sensitized organic solar cells as the ultimate targets; the same group more recently reported a synthesis of poly(arylated)thiophenes through a series of coupling reactions and a Pd–Ag-promoted arylation (Scheme 10.51).83 Applications of this reaction are not limited to advanced materials, but can be applied to natural product synthesis. Indeed, indoles have quite recently (in 2008) been arylated in the presence of palladium acetate and silver oxide (Scheme 10.52).84 In the preceding reactions, the arylation was regioselective with an outcome similar to electrophilic aromatic substitution. However, with simple benzene derivatives, mixtures of biaryl derivatives have been obtained (Scheme 10.53).85 The role of silver trifluoroacetate in these arylations was crucial and, as proposed by the authors, this silver salt could enhance the reactivity and reoxidize Pd species. R
R PdII cat., Ar–I
H
AgNO3–KF, DMSO, 100°C
S R
26–86%
Scheme 10.51
S Ar
R
308
COUPLING REACTIONS PROMOTED BY SILVER Ag2O
+ N
OH
I
Pd(OAc)2
N
DMF, r.t.
OH 87%
Scheme 10.52 NO2
O 2N AgOOCCF3
+ tBu
O 2N +
tBu
Pd(OAc)2 CF3CH2OH, 120°C
I
t Bu
50%
55:45
Scheme 10.53 Ar N + Ar–I
AgOAc, TFA
R
O
N
Pd(OAc)2
R
O
34–95%
Scheme 10.54
Although discussed in terms of a Pd0/PdII versus PdII/PdIV mechanism, the lack of regioselectivity suggests an electrophilic aromatic substitution. Other examples, as well as competitive experiments, also supported such a mechanism (see text below). Suchanarylationreactioncanneverthelessberenderedregioselectiveifcoordinating atoms are present in the starting molecule at the appropriate position, directing the C--H activation at this position. Thus, 2-arylbenzoxazoles have been regioselectively arylated in the presence of palladium acetate and silver acetate (Scheme 10.54).86 Various directing groups could be used in such directed arylations; even carboxylates seemed effective at controlling C--H activation of benzoic acid derivatives (Scheme 10.55).87 Interestingly, direct coupling of aromatic systems has been achieved starting from nonactivated arenes. Such a process is based on direct C--H activation of both partners. Although mixture of compounds could be obtained, good regioselectivity was R1
COOH +
COOH
I R2
AgOAc 1.3 equiv Pd(OAc)2 0.05 equiv
R1
R2
AcOH, 100°C
R1 = OiPr, R2 = p-Cl 69% R1 = Me, R2 = m-CF3 59% R1 = Br, R2 = m-Cl 53%
Scheme 10.55
sp2–sp2 COUPLING REACTIONS PROMOTED BY SILVER SALTS
R
309
R AgOAc
+
Pd(OOCCF3)2
N PG
PhH, 110°C
N PG PG = Piv 75% = Ac 61%
Scheme 10.56 N N
Ag2CO3 2 equiv
+
benzoquinone 0.5 equiv Pd(OAc)2 0.1 equiv DMSO, 130°C
69%
Scheme 10.57
observed by using silver acetate; 2-arylindoles were favored in this case (Scheme 10.56).88 Interestingly, the reversed regioselectivity, that is, formation of 3-arylindoles, was achieved in the presence of copper acetate as cocatalyst. As in the preceding arylation, such direct reactions of nonactivated arenes could be completely regioselective with a coordinating substituent. Benzoquinoline has been used for such purposes (Scheme 10.57).89 Interestingly, such directed arylations have just been rendered asymmetric by using chiral amino acids as ligands on prochiral diarylpyridinylmethane derivatives (see Scheme 10.10, Section 10.3).24 Screening revealed that rigid ligands such as 1,1cyclopropylaminocarboxylates or leucine derivatives N-substituted with bulky carbamates were the most effective in inducing high enantiodiscrimination. 10.5.4.3
Silver as Reagent for Decarboxylative Coupling
In the presence of palladium(II) and silver(I) salts, arene carboxylates could be converted to aryl palladium species, which were engaged in Heck coupling reactions.90 Since the more electrophilic palladium trifluoroacetate proved to be the best catalyst, decarboxylation probably occurred by aromatic electrophilic substitution O MeO MeO
OH + R
Ph
Ag2CO3 Pd(OOCCF3)2 DMSO–DMF, 120°C
Ph
MeO MeO
R
R = OMe 91% R = NO 2 85%
Scheme 10.58
310
COUPLING REACTIONS PROMOTED BY SILVER
O
O OH
O
Pd(OOCCF3)2
OMe
N
MeO
Ag2CO3
+
DMSO–DMF, 80°C
MeO
N
OMe
63%
Scheme 10.59
O O
N
N
Ag2CO3
COOH
dioxane, reflux
N PG
Pd(OOCCF3)2
N PG
N LiAlH4
DMSO–DMF, 80°C
98%
N Cryptosanginolentine
PG = SO2Ph 45% PG = Bn 20%
Scheme 10.60
MeO
OMe
O OH +
I+
Ag2CO3 PdCl2, L
OMe
2
DMSO, 150°C
OMe
Scheme 10.61
of palladium(II) at the ipso position (Scheme 10.58). Silver carbonate dramatically improved this decarboxylation reaction, and DMSO was also critical for the efficiency of the process. This reaction has been extended to other heteroarene carboxylates and to cycloalkenones,leadingtoconjugatedarylorheteroarylenonesingoodyields(Scheme10.59).91 This reaction has also been applied to the synthesis of indoloquinoline alkaloids (Scheme 10.60).92 Similarly, biaryls, including sterically hindered ones, have more recently been obtained by decarboxylative coupling of arenecarboxylic acids and diaryliodonium triflates.93 The PdCl2/DPEphos system in the presence of silver carbonate in DMSO was found to be the most efficient, giving yields ranging from 37% to 85% (Scheme 10.61).
10.6
sp2–sp COUPLING REACTIONS PROMOTED BY SILVER SALTS
The title processes are well documented as stoichiometric, as well as catalytic, procedures. In the former, silver acetylides are used or produced in situ from terminal
sp2–sp COUPLING REACTIONS PROMOTED BY SILVER SALTS O Ag AcO
Br
+ OAc
MeO
reflux 85%
O
O
O
O
O
AcO
CCl4
O
311
AcO
MeO
O
O
O O
Cl
OR
(+)-Methynolide
Cl
Scheme 10.62
alkynes.8 In the latter, the formation of a Csp2 --Csp bond is usually performed in association with palladium catalysis.3 10.6.1
Organosilver Species as Nucleophilic Reagents
Silver acetylides, easily produced either from the corresponding alkynes94 or silylated alkynes,95 were used as nucleophilic reagents toward activated carbonyl groups or iminiums. Silver acetylides exhibit an extreme mildness and a very low basicity, so that theyarecompatiblewithawiderange offunctionalgroups andcan be producedand used in protic solvents and even in water, rendering them green reagents in green processes. Since such reactivity has recently been reviewed,8 only a few examples are reported here. Acyl halides or epoxides can be alkynylated with preformed silver acetylides. Such processes have been applied to the total synthesis of the antibiotic macrolide ( þ )-methynolide,96 and in the total synthesis of the antitumor agent FR901464 isolated from a Pseudomonas species (Scheme 10.62).97 In the latter, the alkynylation proved to be effective only in the presence of dichlorozirconocene and catalytic amounts of silver triflate, suggesting a complex mechanism in which Ag–Zr transmetallation and chloride abstraction by the added silver triflate occurred (Scheme 10.63).8 Nevertheless, this reaction proved quite general.98
MeO O
OH
H
Cp2ZrCl2 1.3 equiv
Ag + O
AgOTf 0.2 equiv
O
MeO
AcO
84% dr = 6 : 1
O
O
O O
CH2Cl2, r.t.
O
OH
HO
N H
O FR901464
Scheme 10.63
312
COUPLING REACTIONS PROMOTED BY SILVER
TABLE 10.5. Aldehyde Alkynylation Catalyzed by Ag(I) in Water R1
R2–CHO
+
Cy3PAgCl 0.1 equiv i-Pr2EtN 0.2 equiv H2O, r.t.–100°C
Alkyne
Aldehyde
OH R2 R1
Conditions
Yield (%)
CHO Ph
60 C, 12 h
90
90 C, 12 h
93
70 C, 2 days
81
80 C, 6 h
93
90 C, 1.5 days
77
Br MeO
CHO
Ph
OMe CHO Ph
CHO Ph
CF3
n -Hex
CHO CF3
Interestingly, the first catalytic version of such nucleophilic additions of alkynyl silvers to aldehydes has been described.99 Indeed, silver chloride in the presence of tricyclohexylphosphine and mild bases such as ethyldiisopropylamine catalyzed the addition of terminal alkynes to aldehydes in good to high yields (Table 10.5). The reaction proved to be almost insensitive to electronic effects; however, alkylacetylenes were less reactive than arylacetylenes. The solvent had a dramatic effect on the reaction course. Control experiments with preformed phenylethynylsilver showed that both phosphine and water activated the silver acetylide. Similarly, terminal alkynes added to a-iminoesters derived from ethylglyoxylate in the presence of silver salts.100 In this case, the reaction worked best in apolar solvents (Scheme 10.64). In a closely related reaction, Li et al.101 devised a silver-catalyzed three-component coupling between aldehydes, alkynes, and amines. This transformation was really efficient in water and in an ionic liquid, although both required overnight heating at 100 C (Table 10.6). In this reaction, a silver acetylide that was sufficiently nucleophilic to add to the iminium ion that formed by condensation of the amine on the aldehyde was certainly produced.8
sp2–sp COUPLING REACTIONS PROMOTED BY SILVER SALTS
313 OMe
MeO R
AgNO3 or AgOTf 0.1 equiv
N
+
HN
hexane, r.t.
H
COOEt
COOEt R
Scheme 10.64
Supported versions of such silver-catalyzed, three-component couplings have been recently reported. Silver oxide on multiwall carbon nanotubes or on alumina,102 as well as silver-doped zeolites,103 proved to be efficient and reusable catalysts for such coupling reactions. In the former, water was used as solvent, while in the latter, no solvent was required, making it a truly green process. Within the same green context, and in an effort to utilize carbon dioxide in organic chemistry, Anastas et al. devised a similar reaction in which the silver acetylide intermediate was trapped by carbon dioxide and the resulting carboxylate was trapped again with benzyl bromide derivatives. The thus-formed ester reacted
TABLE 10.6. Propargylic Amine Formation from Aldehydes and Alkynes n n R1
O
+ R2
AgI 0.015–0.03 equiv
+ H
N H
H2O, 100°C
N R2 R1
Alkyne
Aldehyde
Amine
CHO
Ph
N H
CHO
N H
Ph
Yields (%) 95
70
CHO Ph
N H Ph
Ph
CHO
96
85 N H
314
RPh
COUPLING REACTIONS PROMOTED BY SILVER
Br + CO 2
+ Ar
O
R
AgI 0.05–0.1 equiv
R O
O
K2CO3 DMAc 100°C
Ar
Ar O
R
O
O
R
O
O Ar
Scheme 10.65
intramolecularly in a Diels–Alder-type reaction, possibly also catalyzed by silver iodide. This reaction offered a new and concise synthesis of arylnaphthalene lactones, analogs of natural lignans (Scheme 10.65).104 10.6.2
Organosilver Species in Transmetallations
Combining the mild nucleophilicity of alkynylsilvers with the halogenophilicity of silver ions suggests their use in transmetallation reactions. This idea led to only a few reactions with alkynylsilver species, although such processes certainly occurred in various catalyzed reactions (see Sections 10.4, 10.5.1, 10.5.2, 10.5.4.1, 10.6.4, and Ref. 8). For mechanistic investigations on Pd/Ag-catalyzed enyne or arylyne synthesis, Pale et al. mixed various alkynylsilvers with vinyl triflates or aryl iodides in the presence of palladium salts or complexes (Scheme 10.66).105 This reaction clearly demonstrated that organosilvers can be transmetalated to organopalladium species, even if no halide was present, as demonstrated when starting from vinyl triflates. In the nucleophilic opening of oxiranes by silver acetylides (see Section 10.4), Koide et al. provided good evidence for transmetallation of silver acetylides to dichlorozirconocene leading to alkynyl zirconium species.98
OTf
t-Bu
t-Bu Ag
Pd(PPh3)4
n-Bu
5 min
DMF, r.t. O
n-Bu
1h
98% O
OTf
n-Bu 78%
Scheme 10.66
sp2–sp COUPLING REACTIONS PROMOTED BY SILVER SALTS
OTIPS
EWG
+
AgNTf2 0.05 equiv
315
TIPSO
EWG
CH2Cl2, r.t.
Ag+ OTIPS
Scheme 10.67
10.6.3
Silver as a Lewis Acid Reagent
Several cycloadditions are known to be promoted by silver salts (see Chapter 2), driven by the Lewis acidity of silver ions. Probably more formal than true cycloadditions, such reactions forming C--C bonds are, indeed, often initiated by the Lewis acidity of silver ions, as revealed by the NMR observation of silver p coordination to siloxyalkynes (Scheme 10.67).106 Similarly, g- or d-alkynylaldehydes proved to be more p-coordinated than s-Ocoordinated in the presence of silver hexafluoroantimonate in solution, as shown again by NMR.107 It is worth noting that such experimental observations disagree with theoretical calculations (see Section 10.5.3 and Ref. 36). Interestingly, w-alkynyl– aldehydes reacted intramolecularly, yielding a product formally resulting from a [2 þ 2] cycloaddition and a retro[2 þ 2] cycloaddition (Scheme 10.68). This reaction could also be applied to intermolecular coupling of alkynes to aldehydes. The fact that tetrafluoroboric acid, as well as trifluoroborane–etherate, gave the same products, suggested that the active species is nevertheless the s-coordinated species. The intramolecular carbostannylation reported in 2007 by Echavarren et al. and related reactions are also initiated by the p-Lewis acidity of silver ions.108 Silver salts catalyzed the cyclization of alkynylated allylstannanes to the corresponding 1stannylmethylene-2-vinyl five-membered rings (Scheme 10.69). Interestingly, the more electrophilic the salt is, the more rapid the reaction is; the hexafluroroantimonate
O
R1
AgSbF6 0.1 equiv
R2 X R3
CH2Cl2
O
R1
R2 X R3
Ag+
R3
R2– X O
R1
R1
R1 R2–X
R3
R2– X O
Ag+
Scheme 10.68
R3
O
316
COUPLING REACTIONS PROMOTED BY SILVER
SnBu3
Ag(PPh3)OTf 0.03 equiv
Z SnBu3
Z
Tol., 70°C
Ag+
TfOSnBu3
Ag
-OTf
Z
Z SnBu3
Scheme 10.69
affords the product in one minute. However, more consistent results were achieved with the triphenylphosphine complex of silver triflate. A similar process is probably responsible for the cyclization of N-sulfonyl and NBoc alkynylated enamines. For such reactions, silver triflate was also the best catalyst, leading in high yields to azahydrindane systems, although as a mixture of isomers (Scheme 10.70).109 Although not described as such, this reactivity of silver as a Lewis acid in C--C bond formation via enamines was already known and actually described in the synthesis of complex indole alkaloids. A N-sulfonyldienamine embedded within a polycyclic indole ring system added to the trimethylsilylated propargyl arm of this system, leading in high yield to a new six-membered ring (Scheme 10.71).110 10.6.4 Organosilver Species as Intermediates in Catalyzed Enyne or Arylyne Synthesis Alkynylsilvers are often intermediates in various coupling reactions catalyzed by both silver and palladium salts, leading to enynes or arylynes.
AgOTf 0.04 equiv
R N PG
CH2Cl2, 40°C
Ag+ R N PG
Scheme 10.70
R N PG
sp2–sp COUPLING REACTIONS PROMOTED BY SILVER SALTS
O
N
PhS N
Ag
O SiMe3
317 O
N
AgOAc 0.1 equiv
PhS
AcOH/H2SO4 r.t.
SiMe3 N
SO2Ar
H2O
N
PhS N
SO2Ar
80%
Scheme 10.71 R
OTf +
R
t-Bu
Pd(PPh3)4 0.1 equiv DIPEA 1.2 equiv AgI 0.2 equiv DMF
t -Bu 60–88%
Scheme 10.72
In designing a convergent and versatile synthesis toward antitumoral dienediynes, Pale et al. had to develop alternatives to the classical Sonogashira coupling reaction.111–113 Most silver salts proved effective cocatalysts in such couplings, but the halides were the most efficient, giving coupling products very cleanly and in good yields, even at room temperature and with the most sensitive compounds (Scheme 10.72). Moreover, under these conditions, the formation of homocoupling products, common side products resulting from Glaser-type reactions when using CuI catalysts,114–116 was totally suppressed. Further investigations including 109Ag NMR confirmed the in situ formation of alkynyl silver through p complexation and proton abstraction.117 They also revealed that the best palladium catalyst, namely, palladium tetrakis(triphenylphosphine), played a key role, in that a phosphine liberated by decoordination ended up on the alkynyl silver, stabilizing it and rendering it more soluble.118 The resulting alkynyl silver then entered the palladium catalytic cycle at the transmetallation step (See section 10.6.2).105 To avoid a deprotection step, as well as to solve synthetic problems, Pale et al. modified their Pd–Ag catalytic conditions to directly couple 1-trialkylsilyl-1-alkynes, first with AgF as cocatalyst, expecting that silylated alkynes would be directly converted into silver acetylides with this catalyst (Scheme 10.73).119
OTf t-Bu
R3Si
C 4H 9
C 4H 9
Pd(PPh3)4 0.1–0.5 equiv AgF 1 equiv DMF
Scheme 10.73
tBu
30–95%
SiMe3
318
COUPLING REACTIONS PROMOTED BY SILVER R = SiMe3 R
OTf t -Bu
n-Bu
t-Bu
nBu
87%
R = Sit-BuMe2 84% R = SiPh2t-Bu 75%
Pd(PPh3)4 0.1 equiv
AgI 0.2 equiv DMF, r.t. n-Bu4NF.3H2O 1.5 equiv R
OH t-Bu
O
OH
R = SiMe3
78%
R = Sit-BuMe2 67% R = SiPh2t-Bu 66%
O
Scheme 10.74
Tetrabutylammonium fluoride trihydrate eventually proved to be the best activator for this goal, and various 1-trialkylsilyl-1-alkynes could be coupled under such conditions (Scheme 10.74).120,121 Moreover, alkynes protected with various silyl groups could be distinguished and selectively coupled depending on conditions. 1-Trimethylsilyl-1-alkynes were selectively coupled in the presence of other silylated acetylenic moieties using Pd (PPh3)4–AgCl as the catalyst in the presence of a slight excess of methanol and potassium carbonate in DMF at room temperature.122 Successive coupling reactions could thus be achieved, using differentially protected polyynes,123 even in a one-pot process (Scheme 10.75).124 TMS
t-Bu
OTf
t-Bu
K2CO3 4 equiv MeOH Pd(PPh3)4 0.1 equiv AgCl 0.2 equiv DMF, r.t.
TIPS
86%
TIPS O
TfO AgCl 0.2 equiv
n-Bu4NF.3H2O 1.2 equiv Pd(PPh3)4 0.1 equiv DMF, r.t. 66%
O
t-Bu
Scheme 10.75
sp2–sp COUPLING REACTIONS PROMOTED BY SILVER SALTS
319
TABLE 10.7. Pd/Ag Coupling Reaction with Triisopropylsilylethynyl Phosphonate AgCl 0.2 equiv Pd(PPh3)4 0.1 equiv
O Ar
I + TIPS
P OEt OEt
Ar
n-Bu 4NF.3H2O 1.5 equiv DMF, 50°C
Aryl Halide
O P OEt OEt
Product
PhI
Yield (%)
O P OEt OEt
Ph
95
MeO
I
MeO
O P OEt OEt
80
O 2N
I
O2N
O P OEt OEt
60
N
N
O P OEt OEt
N N
I
66
Such direct couplings of silylated alkynes have been applied to the synthesis of a variety of aryl- and heteroarylethynyl phosphonates, available only in poor yields with classical routes (Table 10.7).125 The same method allowed the preparation of aryl- and heteroaryl propolates directly from trimethylsilylpropiolates (Scheme 10.76).125 These methods have also been applied to the development of a novel route to conjugated diynes, directed toward natural polyacetylenic products. Substituted
O O
I
O
COOEt
O AgCl 0.2 equiv Pd(PPh3)4 0.1 equiv TMS
COOEt
66%
n-Bu4NF.3H2O 1.5 equiv DMF, 50°C N
N
I
Scheme 10.76
N COOEt
N 60%
320
COUPLING REACTIONS PROMOTED BY SILVER
TMS
1. MeLi, LiBr
TMS
2. n-PrI, HMPA -78°C to r.t.
60%
OH
(S)-CBS reagent
n-Bu
BH3.THF
TMS
O AgCl K2CO3, MeOH n-Bu Pd(PPh3)4 DMF, r.t.
I 69%
O nBu
THF, r.t.
n-Pr
(S)-Dehydroxyvirol A
nPr
Scheme 10.77
conjugated diynes were obtained through sequential coupling of 1,4-bis(trimethylsilyl)butadiyne (Scheme 10.77).126–128 Silver salts having a basic counterion also proved successful in various enyne or arylyne syntheses by coupling. Conditions based on stoichiometric amounts of silver oxide have been developed by Mori et al. for the synthesis of arylated alkynes from terminal alkynes and aryl iodides. Under such conditions, neither silylated alkynes nor aryl bromides or triflates did not undergo coupling (Scheme 10.78).129,130 With silver carbonate as cocatalyst and with a palladium(II) complex as catalyst, enynyl phosphonates have been successfully obtained (Scheme 10.79).131 In 2007, Wu et al. showed that arylboronic acids and the corresponding boronates reacted with terminal alkynes in the presence of a palladacycle as catalyst and silver oxide (Scheme 10.80).132 As in the Mori reaction using basic silver salts, stoichiometric amounts of silver ion were required for high efficiency. Under similar conditions, but with the addition of chlorides, Nagasaka et al. were able to directly engage 1-trimethylsilyl-1-alkynes in coupling reactions with aryl
I + R1
R2
Pd(PPh3)4 0.05 equiv Ag2O 1 equiv THF, 60°C
R1 = H, R2 = SiMe3 R1 = OMe, R2 = SiMe3 R1 = COMe, R2 = SiMe3 R1 = COMe, R2 = t-Bu R1 = H, R2 = Ph
Scheme 10.78
R1
79% 60% 99% 87% 60%
R2
sp2–sp COUPLING REACTIONS PROMOTED BY SILVER SALTS
O + Ph P EtO OEt
I
321 O P EtO OEt
Ag2CO3 1.5 equiv PdCl2(PPh3)2 0.05 equiv
Ph
MeCN, r.t.
81%
Scheme 10.79
B(OH)2 +
R2
R1
Pd cat. 0.01 equiv Ag2O 0.5 equiv KOAc 0.75 equiv
R1
R2
CH2Cl2, r.t.
Pd cat. =
N Tol Pd PPh 3 Cl
Fe
R1 = H, R2 = Ph R1 = H, R2 = Phn-C5H11 R1 = OMe, R2 = Ph R1 = H, R2 = COOEt R1 = Me, R2 = n-C5H11
95% 88% 81% 80% 81%
Scheme 10.80
iodides in the presence of silver carbonate. As for the Mori reaction, only iodides (not bromides or triflates) were converted to the expected arylynes. Stoichiometric amounts of silver ion were also required (Scheme 10.81).133 Under essentially the same conditions, Jeong et al. were able to prepare disubstituted trifluoromethylated enediynes, either directly through a dicoupling reaction from bis(trimethylsilyl)enediynes or through successive couplings from mono(trimethyl silyl)enediynes (Scheme 10.82).134 The first direct coupling of terminal alkynes with aryl iodides or bromides without palladium was reported by Wang and Li in 2006.135 Silver iodide and triphenylphosphane in polar solvents proved to be the best catalyst combination, while potassium carbonate proved to be the better base, giving diarylacetylenes in high yields (Table 10.8).
O
I NHR SiMe3
R=
Pd(OAc)2/4 PPh3 0.05 equiv
+
RHN
n-Bu 4NCl 1 equiv
X
Ag2CO3 0.5 equiv THF, 60°C
OMe
Scheme 10.81
X
X = OMe X = Cl
81% 89%
O
322
COUPLING REACTIONS PROMOTED BY SILVER
TABLE 10.8. Silver-Catalyzed Palladium-Free Coupling Reactions AgI 0.1 equiv PPh3 0.3 equiv
X +
Ar
R
K2CO3 2 equiv
R
Ar
DMF, 100°C
Arylhalide
Ar
Product
PhI
p-MePh
PhI
p-FPh
Yields (%) 96
F
98
MeO
I
Ph
MeO
92
O2N
I
Ph
O2N
99
Ph N NC
Br
CF3
NC
Ph
Ph
92
TMS
CF3 +
R
99
N
Br
Ar
I
Ar
Ag2CO3 0.5 equiv
n-Bu 4NBr 1 equiv PdCl2(PPh3)2 0.1 equiv
R
Ph
toluene, r.t.
30–71%
Scheme 10.82
10.7
sp–sp COUPLING REACTIONS PROMOTED BY SILVER SALTS
There are very few reports of sp–sp coupling reactions promoted by silver salts in the literature, and it seems that only a single example has been reported so far. This work described the homocoupling of aliphatic acetylenes on a solid support by using a catalytic system combining silver tosylate, cupric chloride, and tetramethylethylenediamine (TMEDA) (Scheme 10.83).136 All other methods for this coupling reaction proved to be unsuccessful. Control experiments showed that the coupling did not occur with copper or silver salts alone or when cupric tosylate was associated with silver tosylate. These results suggested that copper(II) chloride was required in
323
CONCLUSION
O O
Pri Si O Pri
Ph O OMe CuCl2, AgOTs (0.1 equiv each) 90%
TMEDA, DBU, O2
O
O
O
Pri Si O Pri
O
Ph
O
Ph
O
Si iPr iPr
O
OMe
OMe
Scheme 10.83
addition to the complex formed from silver tosylate and TMEDA. The authors proposed that silver ion activated the alkyne by complexation, facilitating the formation of a dinuclear copper(II) acetylide complex, which directly collapsed to the homocoupling product, releasing copper(I).
10.8
CONCLUSION
Although neglected in organic chemistry for a century, organosilver species and silver salts have been the subject of increasing interest since the end of the twentieth century and clearly even more so in the past decade. Silver salts have been implied in various C--C bond-forming reactions, and, as shown in this chapter, almost all kinds of coupling have been documented, in either stoichiometric or catalytic processes. These Ag-promoted coupling reactions are based on very different methods and mechanisms, depending on the way that silver is involved in the reaction. On one hand, organosilver species can now be produced from a wide variety of other reagents through transmetallation (organolithium and magnesium, organoboranes and borates, organosilicates) or directly from alkynes for alkynylsilvers. Organosilver species can be produced in situ and used as nucleophilic reagents or in homocoupling reactions. They can also be transmetallated to other metals, such as zirconium and palladium, and thus can be involved in other reactions, such as crosscouplings and alkylations. On the other hand, silver salts can act as a mild Lewis acid, promoting various reactions such as allylations, aldolizations, cycloadditions, and cyclizations. Silver salts can also be used as halide scavengers, acting as cocatalysts in cross-coupling reactions catalyzed by other metals, especially palladium. In the latter context, the exact role of silver salts is far from clear and may be more complex than just halide
324
COUPLING REACTIONS PROMOTED BY SILVER
abstraction, possibly with a role in transmetallation. In reactions with alkynes, it has been established that silver salts catalyze the in situ formation of alkynylsilvers as intermediates. Organosilver species and silver salts thus exhibit a wide variety of behaviors, and further work will probably unravel new roles in organic chemistry, especially for the formation of C--C bonds.
REFERENCES 1. Frankland, E., Liebigs Ann. Chem. 1848, 71, 171–213. 2. Grignard, V., Compt. Rend. Acad. Sci. 1900, 130, 1322–1325. 3. (a) De Meijere, A.; Diederich, F., Metal–Catalyzed Cross–Coupling Reactions, WileyVCH, Weinheim, 2004; (b) Tsuji, J., Palladium Reagents and Catalysts, Wiley, Hoboken, NJ, 2004. 4. Lipshutz, B. H.; Yamamoto, Y. Coinage metals in organic synthesis, Chem. Rev. 2008, 108, 2793–3442. 5. 3-Ethoxyprop-1-ynyl silver and phenylethynyl silver were reported as early as 1865 and 1870; see (a) Berend, M.; Liebermann, C., Liebigs Ann. Chem. 1865, 135, 259–290; (b) Glaser, C., Liebigs Ann. Chem. 1870, 154, 157–171; (c) Chevastelon, R., Compt. Rend. Acad. Sci. 1897, 124, 1364. 6. Patai, S.; Rappoport, Z., Organic Derivatives of Gold and Silver, Wiley-VCH, Weinheim, 2003. ´ lvarez7. (a) Weibel, J.-M.; Blanc, A.; Pale, P., Chem. Rev. 2008, 108, 3149–3173; (b) A Corral, M.; Mun˜oz-Dorado, M.; Rodrıguez-Garcıa, I., Chem. Rev. 2008, 108, 3174–3198; (c) Yamamoto, Y., Chem. Rev. 2008. 108, 3199–3222; (d) Naodovic, M.; Hisashi Yamamoto, H., Chem. Rev. 2008, 108, 3132–3148. 8. Halbes-Letinois, U.; Weibel, J.-M.; Pale, P., Chem. Soc. Rev. 2007, 36, 759–769. 9. (a) Gardner, J. H.; Bergstrom, P., J. Am. Chem. Soc. 1929, 51, 3375–3377; (b) Gardner, J. H.; Joseph, L., J. Am. Chem. Soc. 1939, 61, 2551–2552; (c) Biclkey, E. A.; Gardner, J. H., J. Org. Chem. 1940, 5, 126–132. 10. Kharash, M. S.; Reinmuth, O., Grignard Reagents of Nonmetallic Substances, PrenticeHall, New-York, 1954. 11. (a) Tamura, M.; Kochi, J., J. Am. Chem. Soc. 1971, 93, 1483–1485; (b) Kochi, J., J. Organomet. Chem. 2002, 653, 11–19. 12. Whitesides, G. M.; Casey, C. P., J. Am. Chem. Soc. 1966, 88, 4541–4543. 13. Nagano, T.; Hayashi, T., Chem. Lett. 2005, 34, 1152–1153. 14. Someya, H.; Ohmiya, H.; Yorimitsu, H.; Oshima, K., Org. Lett. 2008, 10, 969–971. 15. Snyder, H. R.; Kuck, J. A.; Johnson, J. R., J. Am. Chem. Soc. 1938, 60, 105–111. 16. Brown, H. C.; Verbrugge, C.; Snyder, C. H., J. Am. Chem. Soc. 1961, 83, 1001. 17. Brown, H. C.; Hebert, N.; Snyder, C. H., J. Am. Chem. Soc. 1961, 83, 1001–1002. 18. Brown, H. C.; Snyder, C. H., J. Am. Chem. Soc. 1961, 83, 1002–1003. 19. Falck, J. R.; Mohapatra, S.; Bondela, M.; Venkataraman, S. K., Tetrahedron Lett. 2002, 43, 8149–8151.
REFERENCES
325
20. Chen, H.; Deng, M.-Z., J. Org. Chem. 2000, 65, 4444–4446. 21. Arentsen, K.; Caddick, S.; Georey, F.; Cloke, N.; Herring, A. P.; Hitchcock, P. B., Tetrahedron Lett. 2004, 45, 3511–3515. 22. Yao, M.-L.; Deng, M.-Z., J. Org. Chem. 2000, 65, 5034–5036. 23. Chen, X.; Goodhue, C. E.; Yu, J.-Q., J. Am. Chem. Soc. 2006, 128, 12634–12635. 24. Shi, B.-F.; Maugel, N.; Zhang, Y.-H.; Yu, J.-Q., Angew. Chem. Int. Ed. 2008, 47, 4882–4886. 25. Isabelle, M. E.; Leitch, L. C., Can. J. Chem. 1958, 36, 440–448. 26. (a) Pouwer, R. H.; Williams, C. M.; Raine, A. L.; Harper, J. B., Org. Lett. 2005, 7, 1323–1325; (b) Pouwer, R. H.; Harper, J. B.; Vyakaranam, K.; Michl, J.; Williams, C. M.; Jessen, C. H.; Bernhardt, P. V., Eur. J. Org. Chem. 2007, 241–248. 27. Albert, B. J.; Koide, K., J. Org. Chem. 2008, 73, 1093–1098. 28. (a) Noltes, J. G.; Van Koten, G.,in Comprehensive Organometallic Chemistry I., Wilkinson, G.; Abel, E. W.; Stone, F. G. A.,eds., Pergamon Press, Oxford, 1982, Vol. 2, pp. 709–763; (b) Van Koten, G.; James, S. L.; Jastrzebski, J. T. B. H., in Comprehensive Organometallic Chemistry II, Abel, E. W.; Stone, F. G. A.; Wilkinson, G. eds., Pergamon Press, Oxford, 1996, Vol. 3, pp. 57–133. 29. Lang, G. H.,US Patent 3,962,268. (1976). 30. Tamao, K.; Matsumoto, H.; Kakui, T.; Kumada, M., Tetrahedron Lett. 1979, 13, 1137–1140. 31. (a) Uso´n, R.; Laguna, A.; Abad, J. A., J. Organomet. Chem. 1983, 246 341–345; (b) Kronenburg, C. M. P.; Jastrzebski, J. T. B. H.; Boersma, J.; Lutz, M.; Spek, A. L.; van Koten, G., J. Am. Chem. Soc. 2002, 124, 11675–11683. 32. Albeniz, A. C.; Espinet, P.; Lo´pez-Cimas, O.; Martın-Ruiz, B., Chem. Eur. J. 2004, 11, 242–252. 33. Albeniz, A. C.; Espinet, P.; Martın-Ruiz, B., Chem. Eur. J. 2001, 7, 2481–2489. 34. Bagus, P. S.; Lee, Y. S.; Pitzer, K. S., Chem. Phys. Lett. 1975, 33, 408–411. 35. (a) Pyykk€o, P., Angew. Chem. Int. Ed. 2004, 43, 4412–4456; (b) Gorin, D. J.; Toste, F. D., Nature 2007, 446, 395–403. 36. Yamamoto, Y., J. Org. Chem. 2007, 72, 7817–7831. 37. Yanasigawa, A.; Arai, T., Chem. Commun. 2008, 1165–1172. 38. Josephsohn, N. S.; Carswell, E. L.; Snapper, M. L.; Hoveyda, A. H., Org. Lett. 2005, 7, 2711–2713. 39. Josephsohn, N. S.; Snapper, M. L.; Hoveyda, A. H., J. Am. Chem. Soc 2004, 126, 3734–3735. 40. Wieland, L. C.; Vieira, E. M.; Snapper, M. L.; Hoveyda, A. H., J. Am. Chem. Soc 2008, 131, 570–576. 41. Yao, X.; Li, C.-J., J. Org. Chem. 2005, 70, 5752–5755. 42. Deng, C.-L.; Zou, T.; Wang, Z.-Q.; Song, R.-J.; Li, J.-H., J. Org. Chem. 2009, 74, 412–414. 43. Arcadi, A.; Alfonsi, M.; Marinelli, F., J. Organomet. Chem. 2007, 692, 5322–5326. 44. Lide, D. R., ed., CRC Handbook of Chemistry and Physics, Chemical Rubber Co., 1999. 45. Moritani, I.; Fujiwara, Y.; Danno, S., J. Organomet. Chem. 1971, 27, 279–282.
326
COUPLING REACTIONS PROMOTED BY SILVER
46. (a) Grove, D. M.; van Koten, G.; Louwen, J. N.; Noltes, J. G.; Spek, A. L., J. Am. Chem. Soc. 1982, 104 6609–6616; (b) Rimml, H.; Venanzi, L. M., J. Organomet. Chem. 1984, 260, C52–C54. 47. Albeniz, A. C.; Espinet, P.; Lin, Y.-S., Organometallics 1995, 14, 2977–2986. 48. Yeh, M.-C. P.; Tsao, W.-C.; Wang, Y.-J.; Pai, H.-F., Organometallics 2007, 26, 4271–4277. 49. (a) Larhed, M.; Hallberg, A., Intermolecular Heck reaction: Scope, mechanism, and other fundamental aspects of the intermolecular Heck reaction, in Handbook of Organopalladium Chemistry for Organic Synthesis, Negishi, E. I. ed., Wiley, 2003; (b) Whitcombe, N. J.; Hii, K. K.; Gibson, S. E., Tetrahedron 2001, 57, 7449–7476. 50. (a) Jeffery, T., J. Chem. Soc. Chem. Commun., 1984, 1287–1289; (b) Jeffery, T., Tetrahedron Lett. 1985, 26, 2667–2670; (c) Jeffery, T., Tetrahedron Lett. 1990, 31, 6641–6644; (d) Jeffery, T., J. Chem. Soc. Chem. Commun. 1991, 324–325. 51. (a) Karabelas, K.; Westerlund, C.; Hallberg, A., J. Org. Chem. 1985, 50, 3896–3900; (b) Karabelas, K.; Hallberg, A., J. Org. Chem. 1986, 51, 5286–5290; (c) Karabelas, K.; Hallberg, A., J. Org. Chem. 1988, 53, 4909–4914. 52. Shotwell, J. B.; Hu, S.; Medina, E.; Abe, M.; Cole, R.; Crews, C. M.; Wood, J. L., Tetrahedron Lett. 2000, 41, 9639–9643. 53. Kalesse, M.; Chary, K. P.; Quitschalle, M.; Burzlaff, A.; Kasper, C.; Scheper, T., Chem. Eur. J. 2003, 9, 1129–1136. 54. Abelman, M. M.; Oh, T.; Overman, L. E., J. Org. Chem. 1987, 52, 4130–4133. 55. Cabri, W.; Candiani, I., Acc. Chem. Res. 1995, 2, 2–7. 56. Sundermann, A.; Uzan, O.; Martin, J. M. L., Chem. Eur. J. 2001, 7, 1703–1711. 57. (a) Jutand, A., Appl. Organomet. Chem. 2004, 18, 574–582; (b) Jutand, A., Pure Appl. Chem. 2004, 76, 565–576. 58. Ashimori, A.; Overman, L. E., J. Org. Chem. 1992, 57, 4571–4572. 59. Madin, A.; Overman, L. E., Tetrahedron Lett. 1992, 33, 4859–4862. 60. Sato, Y.; Sodeoka, M.; Shibasaki, M., J. Org. Chem. 1989, 54, 4738–4739. 61. Sato, Y.; Honda, T.; Shibasaki, M., Tetrahedron Lett. 1992, 33, 2593–2596. 62. Sato, Y.; Watanabe, S.; Shibasaki, M., Tetrahedron Lett. 1992, 33, 2589–2592. 63. Satcharoen, V.; McLean, N. J.; Kemp, S. C.; Camp, N. P.; Brown, R. C. D., Org. Lett. 2007, 9, 1867–1869. 64. Omori, A. T.; Finn, K. J.; Leisch, H.; Carroll, R. J.; Hudlicky, T., Synlett 2007, 2859–2862. 65. Zenner, J. M.; Larock, R. C., J. Org. Chem. 1999, 64, 7312–7322. 66. Fu, C.; Ma, S., Org. Lett. 2005, 7, 1605–1607. 67. Malm, J.; Bj€ork, P.; Gronowitz, S.; H€ornfeldt, A.-B., Tetrahedron Lett. 1992, 33, 2199–2202. 68. Fukui, H.; Fukushi, Y.; Tahara, S., Tetrahedron Lett. 2003, 44, 4063–4065. 69. Uenishi, Y.; Beau, J.-M.; Armstrong, R. W.; Kishi, Y., J. Am. Chem. Soc. 1987, 109, 4756–4758. 70. Espinet, P.; Echavarren, A. M., Angew. Chem. Int. Ed. 2004, 43, 4704–4734. 71. Ridgway, B. H.; Woerpel, K. A., J. Org. Chem. 1998, 63, 458–460. 72. Matos, K.; Soderquist, J. A., J. Org. Chem. 1998, 63, 461–470. 73. Reddy, Y. K.; Falck, J. R., Org. Lett. 2002, 4, 969–971.
REFERENCES
327
74. For a review of related coupling without silver activators, see Handy, C. J., Manoso, A. S., McElroy, W. T., Seganish, W. M., DeShong, P., Tetrahedron 2005, 61, 12201–12225. 75. Hirabayashi, K.; Kawashima, J.; Nishibara, Y.; Mori, A.; Hiyama, T., Org. Lett. 1999, 1, 299–301. 76. Hirabayashi, K.; Mori, A.; Kawashima, J.; Suguro, M.; Nishibara, Y.; Hiyama, T., J. Org. Chem. 2000, 65, 5342–5349. 77. Suguro, M.; Mori, A., Synlett 2001, 845–847. 78. Nokami, T.; Tomida, Y.; Kamei, T.; Itami, K.; Yoshida, J.-I., Org. Lett. 2006, 8, 729–731. 79. Napier, S.; Marcuccio, S. M.; Tye, H.; Whittaker, M., Tetrahedron Lett. 2008, 49, 6314–6315. 80. Fujiwara, Y.; Moritani, I.; Ikegami, K.; Tanaka, R.; Teranishi, S., Bull. Chem. Soc. Jpn. 1970, 43, 863–867. 81. For a review on biaryl synthesis, see Bringmann, G.; Walter, R.; Weirich, R.. Angew. Chem. Int. Ed. 1990, 29, 977–991. 82. (a) Kobayashi, K.; Sugie, A.; Masabumi, T.; Masui, K.; Mori, A., Org. Lett. 2005, 7, 5083–5085; (b) Kobayashi, K.; Ahmed, M. S. M.; Mori, A., Tetrahedron 2006, 62, 9548–9553. 83. Arai, N.; Miyaoku, T.; Teruya, S.; Mori, A., Tetrahedron Lett. 2008, 49, 1000–1003. 84. Lebrasseur, N.; Larrosa, I., J. Am. Chem. Soc. 2008, 130, 2926–2927. 85. Lu, W.; Qin, C., J. Org. Chem. 2008, 73, 7424–7427. 86. Yang, F.; Wu, Y.; Zhu, Z.; Zhang, J.; Li, Y., Tetrahedron 2008, 64, 6782–6787. 87. Chiong, H. A.; Pham, Q.-N.; Daugulis, O., J. Am. Chem. Soc. 2007, 129, 9879–9884. 88. Stuart, D. R.; Villemure, E.; Fagnou, K., J. Am. Chem. Soc. 2007, 129, 12072–12073. 89. Hull, K. L.; Sanford, M. S., J. Am. Chem. Soc. 2007, 129, 11904–11905. 90. Myers, A. G.; Tanaka, D.; Mannion, M. R., J. Am. Chem. Soc. 2002, 124, 11250–11251. 91. Tanaka, D.; Myers, A. G., Org. Lett. 2004, 6, 433–436. 92. Miki, Y.; Kuromatsu, M.; Miyatake, H.; Hamamoto, H., Tetrahedron Lett. 2007, 48, 9093–9095. 93. Becht, J.-M.; Le Drian, C., Org. Lett. 2008, 10, 3161–3164. 94. Davies, R. B.; Scheiber, D. H., J. Am. Chem. Soc. 1956, 78, 1675–1678. 95. Viterisi, A.; Orsini, A.; Weibel, J.-M.; Pale, P., Tetrahedron Lett. 2006, 47, 2779–2781. 96. Inanaga, J.; Katsuki, T.; Takimoto, S.; Ouchida, S.; Inoue, K.; Nakano, A.; Okukado, N.; Yamaguchi, M., Chem. Lett. 1979, 1021–1024. 97. Albert, B. J.; Sivaramakrishnan, A.; Naka, T.; Czaicki, N. L.; Koide, K., J. Am. Chem. Soc. 2007, 129, 2648–2659. 98. (a) Naka, T.; Koide, K., Tetrahedron Lett. 2003, 44, 443–445; (b) Shahi, S. P.; Koide, K., Angew. Chem. Int. Ed. 2004, 43, 2525–2527. 99. Yao, X.; Li, C.-J., Org. Lett. 2005, 7, 4395–4398. 100. Ji, J.-X.; Au-Yeung, T. L. T.; Wu, J.; Chiu, W. Y.; Chan, A. S. C., Adv. Synth. Catal. 2004, 346, 42–44. 101. (a) Wei, C.; Li, Z.; Li, C.-J., Org. Lett. 2003, 5, 4473–4475; (b) Li, Z.; Wei, C.; Chen, L.; Varma, R. S.; Li, C.-J., Tetrahedron Lett. 2004, 45, 2443–2447. 102. Zhou, Y.; He, T.; Wang, Z., Arkivoc 2008, xiii, 80–90.
328
COUPLING REACTIONS PROMOTED BY SILVER
103. 104. 105. 106. 107. 108. 109. 110. 111. 112. 113. 114. 115. 116. 117. 118. 119. 120. 121. 122. 123. 124. 125. 126.
Maggi, R.; Bello, A.; Oro, C.; Sartori, G.; Soldi, L., Tetrahedron 2008, 64, 1435–1439. Eghbali, N.; Eddy, J.; Anastas, P., J. Org. Chem. 2008, 73, 6932–6935. Bertus, P.; Dillinger, S.; Pale, P., Org. Lett. 2001, 3, 1661–1664. Sweis, R. F.; Schramm, M. P.; Kozmin, S. A., J. Am. Chem. Soc. 2004, 126, 7442–7443. Rhee, J. U.; Krische, M., Org. Lett. 2005, 7, 2493–2495. Porcel, S.; Echavarren, A. M., Angew. Chem. Int. Ed. 2007, 46, 2672–2676. Harrison, T. J.; Dake, G. R., Org. Lett. 2004, 6, 5023–5026. Magnus, P.; Schultz, J., Tetrahedron Lett. 1986, 27, 655–658. Bertus, P.; Pale, P., Tetrahedron Lett. 1996, 37, 2019–2022. Bertus, P.; Pale, P., Tetrahedron Lett. 1997, 38, 8193–8196. Bertus, P.; Pale, P., J. Organomet. Chem. 1998, 567, 173–180. Glaser, C., Ber. Dtsch. Chem. Ges. 1869, 2, 422. Hay, A., J. Org. Chem. 1960, 25, 1275–1276. Siemsen, P.; Livingston, R. C.; Diederich, F., Angew. Chem. Int. Ed. 2000, 39, 2632–2657. Halbes-Letinois, U.; Pale, P.; Berger, S., J. Org. Chem. 2005, 70, 9185–9190. Halbes-Letinois, U.; Pale, P.; Berger, S., Magn. Reson. Chem. 2004, 42, 831–834. Halbes, U.Ph.D. thesis, Univ. Louis Pasteur, 2003. Bertus, P.; Halbes, U.; Pale, P., Eur. J. Org. Chem. 2001, 4391–4393. Halbes, U.; Bertus, P.; Pale, P., Tetrahedron Lett. 2001, 42, 8641–8644. Halbes, U.; Pale, P., Tetrahedron Lett. 2002, 43, 2039–2042. Halbes-Letinois, U.; Vasiliev, A.; Pale, P., Eur. J. Org. Chem. 2005, 2828–2834. Halbes, U.; Pale, P., J. Organomet. Chem. 2003, 687, 420–424. Lecercle, D.; Mothes, C.; Taran, F., Synth. Commun. 2007, 37, 1301–1311. Fiandanese, V.; Bottalico, D.; Marchese, G.; Punzi, A., Tetrahedron Lett. 2003, 44, 9087–9090. Fiandanese, V.; Bottalico, D.; Marchese, G.; Punzi, A., Tetrahedron 2006, 62, 5126–5132. Fiandanese, V.; Bottalico, D.; Cardellicchio, C.; Marchese, G.; Punzi, A., Tetrahedron 2005, 61, 4551–4556. Mori, A.; Kawashima, J.; Shimada, T.; Suguro, M.; Hirabayashi, K.; Nishibara, Y., Org. Lett. 2000, 2, 2935–2937. Ahmed, M. S. M.; Sekiguchi, A.; Shimada, T.; Kawashima, J.; Mori, A., Bull. Chem. Soc. Jpn. 2005, 78, 327–330. (a) Huang, X.; Zhang, C.; Lu, X., Synthesis 1995, 769–771; (b) Peng, A.-Y.; Ding, Y.-X., Org. Lett. 2005, 7, 3299–3301. Yang, F.; Wu, Y.-J., Eur. J. Org. Chem. 2007, 3476–3479. Koseki, Y.; Omino, K.; Anzai, S.; Nagasaka, T., Tetrahedron Lett. 2000, 41, 2377–2380. Jeon, H. H.; Son, J. B.; Choi, J. H.; Jeong, I. H., Tetrahedron Lett. 2007, 48, 627–631. Li, P.-H.; Wang, L., Synlett 2006, 2261–2265. Liao, Y.; Fathi, R.; Yang, Z., Org. Lett. 2003, 5, 909–912.
127. 128. 129. 130. 131. 132. 133. 134. 135. 136.
11 SUPRAMOLECULAR CHEMISTRY OF SILVER WEI-YIN SUN AND ZHENG-SHUAI BAI Coordination Chemistry Institute, State Key Laboratory of Coordination Chemistry, School of Chemistry and Chemical Engineering, Nanjing National Laboratory of Microstructures, Nanjing University, China
JIN-QUAN YU Department of Chemistry, The Scripps Research Institute, La Jolla, California
11.1 Introduction 11.2 Cage-Like Complexes 11.3 Tube-Like Compounds 11.4 Polycatenanes with Silver(I) 11.5 Polyrotaxanes with Silver(I) 11.6 Silver(I) Coordination Polymers with Specific Topology 11.7 Conclusion References
11.1
INTRODUCTION
Since the late 1980s, there has been tremendous progress and great achievements in supramolecular chemistry; especially after Lehn, Pedersen, and Cram won the Nobel Prize in Chemistry in 1987, which can be considered a milestone in the development of the field.1 The results reported to date demonstrate that the simultaneous assembly of multiple components (e.g., organic ligands and metal salts) is a very powerful method for the construction of compounds with specific structures and properties.2 It is evident that there are two important aspects of so-called metal-directed or metalmediated assembly: (1) owing to their diverse properties (structure, flexibility, coordination ability, etc.), organic ligands can significantly influence the formation Silver in Organic Chemistry Edited by Michael Harmata Copyright 2010 John Wiley & Sons, Inc.
329
330
SUPRAMOLECULAR CHEMISTRY OF SILVER
and structure of the resulting complex; and (2) different metal centers can have different coordination numbers, geometries, and intrinsic properties, so the choice of metal salt is critical. Accordingly, metal complexes with a variety of structures and properties can be achieved through the assembly of rationally designed organic ligands with particular metal salts. As a part of this larger theme, this chapter focuses on the supramolecular chemistry of silver(I). In the area of silver(I) coordination chemistry, it is known that several different coordination numbers and geometries exist.3 In addition to well-known two- (linear), three- (triangular or T-shaped), and four- (tetrahedral or square planar) coordinate silver(I) complexes, there are five-, six-, or even seven- and eight-coordinate structures (Fig. 11.1).4–14 Furthermore, in complexes containing multiple metal centers, there can be argentophilic silver(I)–silver(I) interactions with strength on the order of hydrogen bonds when the metal– metal distance is shorter than the Ag–Ag van der Waals contact distance of 3.44 A.15,16 In combining the variety in coordination environment with the possibility for strong metal–metal interactions, a diverse range of silver(I) complexes with specific topologies and structures can be obtained.17 This chapter describes the structures and properties of supramolecular silver(I) complexes with specific topologies such as cages, tubes, catenanes and polycatenanes, rotaxanes and polyrotaxanes, and multidimensional frameworks, as synthesized by reactions of various silver(I) salts with predesigned organic ligands. 11.2
CAGE-LIKE COMPLEXES
In contrast to “planar” molecules, such as M2L2 macrocycles, M3L3 triangles, M4L4 squares (Fig. 11.2), isolated individual cages (also referred as cones, capsules, boxes,
Figure 11.1. Schematic representation of typical silver(I) (central atom) coordination environments with different coordination numbers and geometries: (a) linear two-coordinate; (b) triangular three-coordinate; (c) T-shaped three-coordinate; (d) square planar four-coordinate; (e) tetrahedral four-coordinate; (f) trigonal bipyramidal five-coordinate; (g) tetragonal pyramidal five-coordinate; (h) octahedral six-coordinate; (i) trigonal prism six-coordinate; (j) sevencoordinate; (k) tetragonal prism eight-coordinate.
331
CAGE-LIKE COMPLEXES
: metal atom;
: ligand
Figure 11.2. Schematic representation of “planar” M2L2 macrocycles, M3L3 triangles and M4L4 squares, formed through metal–ligand coordination.
prisms, cylinders, or containers depending on shape) have discrete three-dimensional (3D) hollow structures and normally have inner cavities with defined shapes and sizes. Accordingly, these cavities can perform a variety of functions, including hosting ions or uncharged molecules, stabilizing unstable species and regulating or promoting specific reactions, such as the [2 þ 2] intermolecular photodimerization of olefins.18 With this breadth of utility, such cage-like compounds have attracted great interest among chemists. Studies have shown that these metal-containing cage-like compounds can be constructed efficiently through the assembly of structurally defined building blocks.19 However, it should be emphasized that cage-like complexes are isolated molecular compounds, containing metal–ligand coordination interactions that are enclosed within a defined space. This space contains a finite number of metal centers and ligands, in contrast to the repeat units of coordination polymers, which form networks that extended infinitely. To achieve this structural motif in the case of a metal atom that has more than two coordination positions, partial coordination of the metal by chelating ligands such as ethylenediamine (en), 2,20 -bipyridine (bipy), or diethylenetriamine may be required to prevent the formation of infinite polymeric compounds. For example, multicomponent assembly of cis-protected, square planar metal complexes of the type [M(en)](NO3)2 or [M(bipy)](NO3)2 [M ¼ Pd(II) or Pt(II)] with multidentate organic ligands leads to the formation of cage-like complexes with nanometer-scale inner cavities.20 In contrast to palladium(II) and platinum(II), as mentioned above, silver(I) can have a linear geometry with two coordination sites, and as a result usually no protection via additional chelating ligands is needed for the construction of silver(I) cage-like compounds. The organic ligands are also important in the construction of silver(I) cage-like complexes. The most popular ligands are compounds with nitrogen donor groups (e. g., pyridyl, imidazolyl, pyrazolyl). Such ligands can be divided into two types according to their relative conformational variability.21 One type is the “flexible” ligand, which can adopt different conformations and coordination modes according to the geometric requirement of the given metal ion. Such flexible ligands can form metal complexes with a variety of structures, making it difficult to predict the structure of the resulting complexes. The other class consists of “rigid” ligands, which have a lower degree of conformational variation when they coordinate with metal ions. Both flexible and rigid ligands have been used to build silver(I) cages. Flexible imidazole-containing tripodal ligands [e.g., 1,3,5-tris(imidazol-1-ylmethyl)benzene and 1,3,5-tris(imidazol-1-ylmethyl)-2,4,6-trimethylbenzene (1)]
332
SUPRAMOLECULAR CHEMISTRY OF SILVER
Figure 11.3. An assembly reaction and the resulting crystal structure of the [Ag3(1)2]3 þ cylindrical cage with the inclusion of PF6. Hydrogen atoms and anions outside the cage have been omitted for clarity.
(Fig. 11.3) have been shown to be versatile components in the construction of metal complexes with specific topologies and structures when reacted with different metal salts.21–24 Reactions of ligand 1 with various silver(I) salts have been systematically investigated, and cage-like complexes have been successfully isolated.12,25–28 For example, when ligand 1 was reacted with silver(I) hexafluorophosphate [AgPF6] in a solution of methanol and acetonitrile at room temperature, a cage complex [Ag3(1)2](PF6)3 resembling a cylinder was obtained containing one hexafluorophosphate anion in its inner cavity (Fig. 11.3).25 Analysis of the resulting crystal structure revealed that the ligand 1 adopts a cis,cis,cis conformation, with two such ligands orienting themselves in a face-to-face fashion. The ligands, in turn, are linked together by 3 two-coordinate Ag(I) atoms through the Ag–N coordination. The height of the cage, namely, the distance between the two central benzene ring planes, is approximately 1.1 nm and the Ag(I)–Ag(I) separation is about 7 A, implying that the cage has an inner cavity of sufficient size for hexafluorophosphate anion inclusion. It is worth noting that this anion inclusion by the [Ag3(1)2]3 þ cylindrical cage is also an anion recognition process. As illustrated in Figure 11.4, the same [Ag3(1)2]3 þ
2
+ 3 nothing anions solvent water
[Ag3(1)2]3+
Anion: ClO4-, BF4-, NO3-, N3-, CF3SO3-, PF6- or SbF6-
Figure 11.4. Schematic drawing for the assembly reactions of flexible ligand 1 and silver(I) salts with different counteranions to generate [Ag3(1)2]3 þ cylindrical cages with different species included in the inner cavity of the cage.
CAGE-LIKE COMPLEXES
333
cage-like framework was obtained by reactions of ligand 1 and silver(I) salts with different counteranions; however, the specific species included in the inner cavity of the cage depends on the corresponding anions.12,25–28 The cage can be empty (i.e., with nothing included), or it can contain anions, neutral solvent molecules, or water molecules as exhibited in Figure 11.5. Only anions with suitable size and shape can be included inside the cage. In addition to simple anions such as ClO 4 , BF4 , PF6 , and 2 SbF6 , complex anions such as [Ag(NO3)3] can also be included (Fig. 11.5c). Rigid ligands have also been used to build silver(I) cage-like complexes.29,30 The reaction between the disk-shaped rigid tripodal ligand 1,3,5-tribenzimidazolyl-2,4,6trimethylbenzene (2) (Fig. 11.6) and silver(I) methanesulfonate (AgO3SCH3) or silver(I) trifluoromethanesulfonate (AgOTf) was studied by 1 H and 19 F NMR and electrospray ionization–time-of-flight (ESI-TOF) mass spectroscopy.29 The results revealed that two types of cage-like complexes were formed depending on the metalto-ligand ratio: [Ag3(2)2]3 þ with a triangular shape and [Ag4(2)4]4 þ with a tetrahedral shape (Fig. 11.6). The cages have different cavities of unique shape and size. Only the cavity of the [Ag4(2)4]4 þ cage has the ability to include a guest methanesulfonate or triflate anion, as confirmed by NMR spectroscopy as well as analysis of the X-ray crystal structure (Fig. 11.7). Furthermore, there exists a dynamic
Figure 11.5. Crystal structures of [Ag3(1)2]3 þ cylindrical cages with different species included in the inner cavity of the cage: (a) empty; (b) ClO4 anion; (c) [Ag(NO3)3]2 anion; (d) one methanol and one water molecule (in this cage two of three Ag are three-coordinate); (e) two water molecules. Hydrogen atoms and anions outside the cage have been omitted for clarity.
334
SUPRAMOLECULAR CHEMISTRY OF SILVER
CH 3 O3
HN
N N H
N
N
NH
q 5e
S Ag
[Ag3(2)2]3+
1. 1.0
eq
Ag
SO
3 CH 3
2
[Ag4(2)4]4+
Figure 11.6. Schematic representation of rigid tripodal ligand 2 and the corresponding assembly reaction with silver(I) methanesulfonate to give [Ag3(2)2]3 þ and [Ag4(2)4]4 þ cages.
equilibrium between the [Ag3(2)2]3 þ and [Ag4(2)4]4 þ cages, which can be controlled by the inclusion and release of the guest anion. When the rigid tridentate ligand 1,3,5-tri-3-pyridyl-2,4,6-tri-p-tolyl-benzene (3) was reacted with silver(I) triflate or silver(I) hexafluorophosphate, cage-like
Figure 11.7. Crystal structure of the [Ag4(2)4]4 þ cage with the inclusion of one methanesulfonate anion inside the cage. Hydrogen atoms and anions outside the cage have been omitted for clarity.
335
CAGE-LIKE COMPLEXES
CF 3 SO 3
N
1.5
g qA
e
[Ag6(3)4]6+
N 1.0
eq
Ag
SO
3 CF 3
N 3
[Ag4(3)4]4+ Figure 11.8. Schematic representation of rigid tripodal ligand 3 and the assembly reaction with silver(I) triflate to give [Ag6(3)4]6 þ and [Ag4(3)4]4 þ cages.
complexes [Ag6(3)4]6 þ and [Ag4(3)4]4 þ with octahedral and tetrahedral shapes, respectively, were formed as illustrated in Figure 11.8.30 The two cages have different shapes and volumes and possess unique internal cavities and thus have different guest inclusion properties.30 For example, the [Ag4(3)4]4 þ cage has a high affinity toward a guest adamantane molecule, as evidenced by its X-ray crystal structure (Fig. 11.9), while the [Ag6(3)4]6 þ cage has no interaction with adamantane. Interestingly, interconversion between these two cages can be achieved by changing the metal-to-ligand ratio, thereby effecting the guest molecules inclusion or release. For instance, the encapsulated guest molecule of adamantane in the [Ag4(3)4]4 þ cage was released on the addition of 2 equiv of silver salt, promoting an accompanying structural conversion of [Ag4(3)4]4 þ to [Ag6(3)4]6 þ . Similarly, the [Ag4(3)4]4 þ cage can be recovered from [Ag6(3)4]6 þ by addition of 2 equiv of [2,2,2]-cryptand to trap silver(I). The interconversion between [Ag4(3)4]4 þ and [Ag6(3)4]6 þ is quantitative and fast.30 In a 2008 study of silver cages using 1,3,5-tris(2-oxazolinyl)benzene (4) as the connecting ligand, assembly reactions of 4 with various silver(I) salts were carried out and silver(I) complexes with different structures were obtained.31 When ligand 4 was reacted with silver(I) perchlorate in a solution of acetonitrile and chloroform at ambient temperature, a cage-like complex [Ag3(4)2(CH3CN)3](ClO4)3 was formed, as shown in Figure 11.10. This structure constitutes a small cage, as indicated by the centroid-to-centroid distance of 3.64 A between two central benzene ring planes and the Ag--Ag distance of 7.04 A. Therefore, there is no true cavity inside the cage, and no inclusion properties exist. This particular complex crystallizes in an acentric space group R3c, and is second-harmonic generation (SHG) active, with a response of about 0.5 times of that of urea.31 Several other silver(I) cage-like complexes have been reported.32–35 3,8,13-Tris (4-pyridylmethylamino)-2,7,12-trimethoxy-10,15-dihydro-5H-tribenzo[a,d,g]
336
SUPRAMOLECULAR CHEMISTRY OF SILVER
Figure 11.9. Crystal structure of the [Ag4(3)4]4 þ cage with the inclusion of one guest adamantane molecule inside the cage. 2,4,6-Tri-p-tolyl groups, hydrogen atoms, and anions have been omitted for clarity.
cyclononene (5, Fig. 11.11), a derivative of cyclotriveratrylene (CTV), was reacted with silver(I) tetrafluoroborate (AgBF4) in acetonitrile solution leading to the formation of a stellated tetrahedral cage-like complex [Ag4(5)4(CH3CN)4] (BF4)47CH3CN2.8H2O with encapsulation of acetonitrile molecules in the inner cavity of the cage.32 In addition to the N-donor ligands 1–5 mentioned above, assembly reactions of thiacalix[4]arene-p-sulfonate (TCAS) with silver(I) and terbium(III) salts were studied by electrospray ionization mass spectrometry under
Figure 11.10. The assembly reaction and resulting crystal structure of [Ag3(4)2(CH3CN)3]3 þ . Hydrogen atoms and perchlorate anions have been omitted for clarity.
TUBE-LIKE COMPOUNDS
337
Figure 11.11. Ligands 5 and 6 used in the construction of silver(I) cage-like complexes.
varying pH, resulting in cage-like complexes with different compositions and photoluminescent properties.33 Triphosphine 6 (Fig. 11.11) has also been used to react with silver(I) salts, and the resulting cage-like complexes have been reported.34
11.3
TUBE-LIKE COMPOUNDS
Carbon-based nanotubes (CNTs) have attracted a great deal of attention in the fields of chemistry, physics, and materials science and have been extensively studied since their initial discovery in 1991.36 Much of the excitement in this area of research stems from their unique structures, fundamental electronic and physical properties, and potential applications.37 Inspired by these carbon nanotube studies, coordination chemists have undertaken significant efforts toward the construction of tube-like complexes through metal–ligand coordination.38–41 Examples of silver(I) tubular complexes obtained via assembly reactions of predesigned organic ligands with silver(I) salts are described here.42–44 A pyridyl- and sulfanyl-containing flexible ligand 2,4,6-tris[(4-pyridyl)methylsulfanyl]-1,3,5-triazine (7, Fig. 11.12) with multiple coordination sites has been reported,45 and a nanometer-scale, tube-like silver(I) complex, [Ag7(7)4](ClO4)2 (NO3)5(DMF)2, was isolated following an assembly reaction of ligand 7 with silver nitrate (AgNO3) and silver perchlorate (AgClO4) salts in N,N-dimethylformamide (DMF) solution.42 As revealed by X-ray crystallographic analysis, shown in Figure 11.13a, two molecules of 7 with a cis,cis,cis conformation are linked together by 3 two-coordinate silver(I) atoms to form a [Ag3(7)2]3þ unit. In turn, two of these units are further linked by two Ag(I) atoms through Ag–N (the nitrogen of triazine, not of pyridine in 7) and Ag–S coordination to generate one [Ag7(7)4]7 þ tube with dimensions of 1.34 0.96 0.89 nm (Figs. 11.13b and 11.13c). There are two DMF
338
SUPRAMOLECULAR CHEMISTRY OF SILVER
N
N
N O
S N S
O N
N
N S N
N
O N
N
O
N N 8
7
cis -conformation
N trans -conformation
Figure 11.12. Multidentate organic ligands 7 and 8 (with cis and trans conformations) used in the construction of silver(I) tube-like complexes.
molecules and two perchlorate anions in the large cavity of the tube (Fig. 11.13b). Furthermore, an infinite one-dimensional (1D) chain was formed by [Ag7(7)4]7þ tubes sharing the Ag(I) atoms, as exhibited in Figure 11.13d. The Ag(I) is fourcoordinate with a square planar geometry and an N2S2 donor set. This is a unique 1D chain structure with numerous nanosized tubes linked by silver(I) atoms.42
Figure 11.13. (a) Crystal structure of two [Ag3(7)2]3þ units linked by two Ag(I) atoms through Ag–N and Ag–S coordination to give a [Ag7(7)4]7þ unit; (b) [Ag7(7)4]7þ tube with inclusion of two DMF molecules and two perchlorate anions; (c) schematic drawing of the tube; (d) [Ag7(7)4]7þ tubes linked by four-coordinate silver(I) atoms to form an infinite 1D chain. The hydrogen atoms, solvent molecules, and anions have been omitted for clarity.
POLYCATENANES WITH SILVER(I)
339
Another silver(I) tube-like complex, [Ag(8)]BF40.5(C6H6), was obtained through the combination of ethyleneglycol ether–bridging tetradentate ligand 8 (Fig. 11.12) with silver(I) tetrafluoroborate.43 The flexible ligand 8 can have both cis and trans conformations (Fig. 11.12) through the rotation of the central C--C single bond. Interestingly, these different conformations can be fine-tuned by adjusting the reaction temperature.43 When ligand 8 was reacted with AgBF4 in a solution of dichloromethane and benzene at 30 C, complex [Ag(8)]BF40.5C6H6 with a 1D tube structure was obtained in which the ligand 8 adopted a cis conformation (Figs. 11.14a and 11.14b). However, when the same reaction was carried out at 0 C, the ligand 8 was found to adopt a trans conformation, and complex [Ag2(8)(H2O)](BF4)2 with a two-dimensional (2D) porous structure was formed (Fig. 11.14c).43 It is immediately obvious that the 1D tube formed from ligand 8 (Figs. 11.14a and 11.14b) is different from the complex obtained from ligand 7 (Fig. 11.13d), as the [Ag7(7)4]7þ tube has a defined length while the [Ag(8)] þ structure is an infinite 1D chain tube with internal dimensions of approximately 1.2 0.82 nm.
11.4
POLYCATENANES WITH SILVER(I)
Catenanes and polycatenanes are an interesting class of compounds with interlocking units, as schematically represented in Figure 11.15. These structures have attracted a great deal of attention from chemists due to their topological novelty, which offers the potential for possible application in molecular switches, machines, and devices.46 A number of catenanes and polycatenanes have been prepared by assembly reactions of metal salts with rationally predesigned organic ligands;47–53 and two kinds of silver(I) polycatenanes with bis(2-methylimidazol-1-yl)methane (9) and 1,5-(3-oxapentane) diyl-bis(isonicotinate) (10) (Fig. 11.16) have been reported.51–53 When ligand 9 was reacted with AgNO3 and AgBF4, complexes with the formulas [Ag4(9)4](NO3)43H2O and [Ag4(9)4](BF4)4H2O were obtained, respectively. The results of structural analysis showed that they have the same cationic framework structure. As exhibited in Figure 11.17a, it is clear that four twocoordinate silver(I) atoms link four molecules of 9 through Ag–N coordination to form M4L4 macrocycles, which interlock with each other to generate an infinite 1D polycatenane.51,52 This is an example of transformation of individual macrocycles to an infinite 1D chain through catenation. It is worth noting that the formation of such interlocking polycatenanes can be influenced by subtle variation of the ligand and counteranion. For example, a complex with a simple 1D zigzag chain structure was isolated by reaction of the AgNO3 with bis(imidazol-1-yl)methane.51 On the other hand, a discrete complex with a macrometallocyclic structure of [Ag6(9)6] (SO3CF3)6 was produced by reaction of ligand 9 with AgOTf as shown in Figure 11.17b.52 Another example is the catenation leading to the formation of a 2D network from 1D chains (Fig. 11.18).53 One molecule of ligand 10 links to two silver(I) atoms, and one metal atom, in turn, coordinates with four molecules of 10 to give an infinite 1D
340
SUPRAMOLECULAR CHEMISTRY OF SILVER
Figure 11.14. Top (a) and side (b) views of an infinite 1D [Ag(8)]BF40.5C6H6 tube obtained at 30 C (hydrogen atoms, anions, and solvent molecules have been omitted for clarity); (c) 2D network structure of [Ag2(8)(H2O)](BF4)2 isolated at 0 C (hydrogen atoms have been omitted for clarity).
341
POLYCATENANES WITH SILVER(I)
(a)
(b)
Figure 11.15. Schematic drawing of [2]-catenane (a) and infinite 1D [n]-polycatenane (b).
N N
N
N 9
N O
N
O
O
O O
10
Figure 11.16. Structures of organic ligands 9 and 10 used for construction of polycatenaned silver(I) complexes.
Figure 11.17. (a) Crystal structure of the infinite 1D polycatenane of [Ag4(9)4].4 þ Top (b) and side (c) views of the macrometallocyclic structure of [Ag6(9)6].6 þ Hydrogen atoms, anions, and solvent molecules have been omitted for clarity.
342
SUPRAMOLECULAR CHEMISTRY OF SILVER
Figure 11.18. Infinite 1D chain (a) and 2D polycatenaned (b) structures of [Ag(10)2] þ . Hydrogen atoms and anions have been omitted for clarity.
chain (Fig. 11.18a). Such 1D chains interlock with one another to generate 2D polycatenane structures (Fig. 11.18b).
11.5
POLYROTAXANES WITH SILVER(I)
Rotaxanes and polyrotaxanes are another type of compound consisting of string-like components threaded through ring-like components without covalent bonding (Fig. 11.19) and have received attention for a long time.46 Reported ring parts of
343
POLYROTAXANES WITH SILVER(I)
Stopper
Stopper
(a)
(b)
Figure 11.19. Schematic drawing of a rotaxane (a) and an infinite 1D polyrotaxane (b).
rotaxanes and polyrotaxanes include crown ethers, cyclophanes, cyclodextrins, calixarenes, and cucurbiturils as well as the metal complexes themselves.46,54–57 Cucurbit[6]uril (CB[6]), with its six glycoluril units linked by 12 methylene bridges (Fig. 11.20a), has been found to be an efficient building block in the preparation of rotaxanes and polyrotaxanes.54 Organic compounds with diamine units are well suited to serve as the string part of rotaxanes, since the protonated diamine moiety can incorporate CB[6] via N--H--O hydrogen-bonding interactions. For example, N,N0 -bis(3-pyridylmethyl)-1,6-hexanediamine (11) and N,N0 -bis(3pyridylmethyl)-1,4-benzenedimethylamine (12) (Fig. 11.20b) were reacted with CB[6] followed by AgNO3 under acidic conditions to yield polyrotaxanes; the effects of the flexibility and length of the ligands on the formation and structure of the resulting polyrotaxanes have also been explored.56 The protonated ligand (H212) was included in the inside cavity of CB[6] to give pseudorotaxanes, which are further connected by silver(I) through Ag–N coordination, and an infinite 1D polyrotaxane was ultimately formed (Fig. 11.21).56
Figure 11.20. (a) Top (left) and side (right) views of the host cucurbit[6]uril; (b) organic ligands 11, 12, and 13 used for construction of silver(I) polyrotaxanes.
344
SUPRAMOLECULAR CHEMISTRY OF SILVER
Figure 11.21. Formation of the infinite 1D polyrotaxane of [Ag(H212)(CB[6])](NO3)310H2O and the corresponding crystal structure.
Figure 11.22. A 2D polyrotaxane of [Ag2(13)3](NO3)2.
SILVER(I) COORDINATION POLYMERS WITH SPECIFIC TOPOLOGY
345
Polyrotaxanes can also be formed through interlocking coordination polymers, a typical example of which has been reported by Robson and coworkers.57 1,4-Bis (imidazol-1-yl-methyl)benzene (13) (Fig. 11.20b) coordinates with silver(I) to give 1D polymeric chains that are further knitted together to provide a 2D polyrotaxane (Fig. 11.22).
11.6 SILVER(I) COORDINATION POLYMERS WITH SPECIFIC TOPOLOGY In addition to the infinite tubes, polycatenanes, and polyrotaxanes, all of which are polymeric complexes, a great number of silver(I) coordination polymers with 1D, 2D and 3D framework structures have been reported more recently. Selected examples and their specific topologies of some of these are listed in Table 11.1. Schematic structures of the corresponding organic ligands are shown in Figure 11.23.9,58–75 An infinite 1D complex {[Ag(14)(15)]}n with a single-stranded helical chain structure was obtained through the reaction of 6-amino-1-naphthalenesulfonic acid (H-14) and 2-methylpyrazine (15) with Ag2CO3.58 Furthermore, when tetra- and hexaethyleneglycol-spaced ligands, 1,11-(3,6,9-trioxaundecane)diyl-bis(nicotinate) (16) and 1,17-(3,6,9,12,15-pentaoxaheptadecane)diyl-bis(nicotinate) (17), were reacted with silver(I) salts, double-stranded helical complexes (as exhibited in Fig. 11.24) were formed.59 Various 2D networks with common (4,4) and (6,3) topologies have been reported. The silver(I) complex {[Ag(18)(NO3)]}n was formed from 1,3-bis(2-pyridylaminomethyl)benzene (18) and AgNO3, and similarly, {[Ag(19)2/3]3[Ag(19)]2(CF3SO3)5}n was isolated from the reaction of 1,3,5-tris(4-((2-ethyl-benzimidazol-1-yl)methyl) phenyl)benzene (19) with AgO3SCF3. These two complexes have (4,4) and (6,3) topologies, respectively.60,61 When the imidazole-containing tripodal ligand 1,3-bis(1-imidazolyl)-5-(imidazol-1-ylmethyl)benzene (20) was reacted with [Ag(NH3)2] ClO4 in an aqueous methanol solution, the twofold interpenetrated complex {[Ag(20)]ClO4}n with 4.82 topology was generated (Fig. 11.25).62 To date, a number of different topologies have been reported for silver(I) complexes with 3D structures, as listed in Table 11.1.9,63–75 The reaction of a silver– ammonia solution with 3,30 ,5,50 -tetramethyl-4,40 -bipyrazole (H2-21) provides a polymeric complex {[Ag2(21)]}n with a fourfold interpenetrated (10,3)-a framework structure.63 The complex {[Ag4(22)2](NO3)4MeCN2H2O}n with interpenetrating (10,3)-b nets was prepared from the reaction of N0 ,N0 -bis[1-(pyrazine-2-yl)ethylidene]benzildihydrazone (22) with AgNO3.64 Another type of (10,3) topology is (10,3)-d, which has been observed in the complex {[Co(23)3AgBF4]}n [23 ¼ 5-(4pyridyl)-4,6-dipyrrinato].9 In this case, the [Co(23)3] acts as building block and is further connected by Ag(I) to give a 3D framework with twofold interpenetration, as illustrated in Figure 11.26. The reaction of 1H-1,2,4-triazole-1-acetic acid (H-24) with AgNO3 in the presence of sodium hydroxide under hydrothermal conditions gave {[Ag(24)]}n. This complex
346
{[Ag(14)(15)]}n {[Ag(16)](BF4)}n/{[Ag(17)](PF6)}n {[Ag(18)(NO3)]}n {[Ag(19)2/3]3[Ag(19)]2(CF3SO3)5}n {[Ag(20)]ClO4}n {[Ag2(21)]}n {[Ag4(22)2](NO3)4MeCN2H2O}n {[Co(23)3AgBF4]}n {[Ag(24)]}n {[Ag6(25)4Cl](OH)6H2O}n {[Ag(26)] CH3OH}n {[Ag(27)3](NO3)}n {[Ag(28)](NO3)(CH3OH)}n {[Ag2(29)](ClO4)2}n {[Ag2(30)](ClO4)2}n {[Ag(31)](ClO4)4.5H2O}n {[Ag(32)2](SbF6)}n {Fe(33)2[Ag(CN)2]22/3H2O}n {Fe(34)[Ag(CN)2][Ag2(CN)3]}n {[Ag(35)](BF4)1/2(OH)1/2(CH3CN)1/4}n {[Ag(35)](BF4)1.5H2O}n
Complex 3, 2, 4, 2, 3, 2, 4, 4, 4, 2, 5, 6, 4, 4, 4, 6, 4, 2, 2, 4, 4,
trigonal linear tetrahedral linear; 3, trigonal trigonal linear tetrahedral trigonal pyramidal tetrahedral linear tetragonal pyramidal octahedral tetrahedral tetrahedral tetrahedral trigonal prism tetrahedral linear linear square planar tetrahedral
Coordination Number and Geometry of Ag(I)
TABLE 11.1. Selected Silver(I) Coordination Polymers with Specific Topology
1D single-stranded helix 1D double-stranded helix 2D (4,4) network 2D (6,3) network 2D (4.82) network 3D (10,3)-a or SrSi2 3D (10,3)-b or ThSi2 3D (10,3)-d 3D diamond 3D 4.142 (dia-f) 3D SrAl2 (sra) 3D a-Po (pcu) 3D ð8;34 Þ-b 3D pyrite (pyr) 3D rutile (rtl) 3D (8,3)-b 3D coesite (coe) 3D NbO 3D (6,6) 3D lvt 3D PtS
Topology
58 59 60 61 62 63 64 9 65 66 67 68 69 70 71 71 72 73 74 75 75
Reference
347
SILVER(I) COORDINATION POLYMERS WITH SPECIFIC TOPOLOGY O
O
SO3-
O
O
O
O
O
O
O
N
N
O
O
O
O
O
O
O
N NH2
14
N
N
N
16
15
17
N
N H N
H N
N
N
18 N N N
19
N
N
N
N N
N N N
N
N N
N N
N N
N
20
21
N N
N
22
N COO-
N N
N N N
HN
23
H2N COO-
24
N N
25
N
N N
N N
N
26
Figure 11.23. Structures of organic ligands 14–35 used for construction of the silver(I) coordination polymers listed in Table 11.1.
has a 3D structure with a typical diamond topology in which the silver(I) has a tetrahedral coordination geometry and serves as a four-connector.65 The polymeric complex {[Ag6(25)4Cl](OH)6H2O}n was obtained by reaction of an ammonia solution of AgCl and 3-amino-1,2,4-triazole (H-25). The 3D framework structure was confirmed by X-ray crystallographic analysis exhibiting 4.142 (dia-f) topology
348
SUPRAMOLECULAR CHEMISTRY OF SILVER N
H2N
N
NH2
27
N
N
28
N N
S
S
S
S
N
N
S
S
S
S
N N
N
N
N
29
30
N
N
N N N
N
N
N
N N
N
N
31
N O
32
N
N
33
34
N
N
35
Figure 11.23. (Continued).
(Fig. 11.27) by considering ligand 25 as three-connected nodes and two-coordinated Ag(I) as linkers.66 Coordination polymers {[Ag(26)]CH3OH}n and {[Ag(27)3] (NO3)}n with SrAl2 (sra) and a-Po (pcu) topologies, respectively, were readily obtained by reacting the corresponding ligands bis(1,2,4-triazol-1-yl)acetic acid (H-26) and 4,40 -diaminodiphenylmethane (27) with silver(I) salts, as reported
Figure 11.24. Space-filling model of the double-stranded helix {[Ag(16)](BF4)}n.
SILVER(I) COORDINATION POLYMERS WITH SPECIFIC TOPOLOGY
349
Figure 11.25. Parallel interpenetration of two 4.82 nets of {[Ag(20)]ClO4}n, crystal structure (a) and schematic representation (b).
by Du and Ciani.67,68 The topology of ð8;34 Þ-b was predicted by Wells in the late 1970s, and a silver(I) complex {[Ag(28)](NO3)(CH3OH)}n with such a topology was synthesized by using 3,30 ,5,50 -tetrakis(4-pyridyl)bimesityl (28) as a ligand.69 Hanton and coworkers reported silver(I) complexes with different topologies induced by the ligands 1,2,4,5-tetrakis(3-pyridylmethylsulfanylmethyl)benzene
Figure 11.26. Crystal structure of {[Co(23)3AgBF4]}n with (10,3)-d topology and twofold interpenetration.
350
SUPRAMOLECULAR CHEMISTRY OF SILVER
Figure 11.27. Three-dimensional 4.142 (dia-f) net observed in the crystal structure of {[Ag6(25)4Cl](OH)6H2O}n.
(29) and 1,2,4,5-tetrakis(4-pyridylmethylsulfanylmethyl)benzene (30).70 The complexes {[Ag2(29)](ClO4)2}n and {[Ag2(30)](ClO4)2}n have 3D structures with pyrite (pyr) and rutile (rtl) topologies, as shown in Figure 11.28. The compound 1,4,5,8,9,12-hexaazatriphenylene (31) was reacted with different silver(I) salts, and complexes with (10,3)-a topology were formed; however, {[Ag(31)](ClO4) 4.5H2O}n showed a quite different (8,3)-b topology.71 The combination of 2ethylpyrazine (32) and AgSbF6 leads to the formation of complex {[Ag(32)2](SbF6)}n with coesite (coe) topology.72 The complex [Ag(CN)2] was used as a building block to react with Fe(BF4)26H2O/3-cyanopyridine (33) and Fe(BF4)26H2O/pyrimidine (34) leading to the formation of complexes {Fe(33)2[Ag(CN)2]22/3H2O}n and {Fe(34)[Ag(CN)2] [Ag2(CN)3]}n with NbO and (6,6) topologies, respectively.73,74 The reaction of 2,5-bis(pyrazine)-1,3,4-oxadiazole (35) with AgBF4 provided topological isomers of {[Ag(35)](BF4)1/2(OH)1/2(CH3CN)1/4}n and {[Ag(35)](BF4) 1.5H2O}n depending on the corresponding guest and solvent medium.75 Both complexes have 3D framework structures but have different topologies: lvt and PtS, respectively (Fig. 11.29).
11.7
CONCLUSION
More recent studies have demonstrated that silver(I) is a versatile metal because of its variable coordination number and geometry. As a consequence of these inherent properties, a wide variety of structures, ranging from individual cage-like molecular complexes to infinite 1D, 2D, and 3D architectures with specific topologies, have been
CONCLUSION
351
Figure 11.28. {[Ag2(29)](ClO4)2}n (a) and {[Ag2(30)](ClO4)2}n (b) with pyrite (pyr) and rutile (rtl) topologies, respectively.
352
SUPRAMOLECULAR CHEMISTRY OF SILVER
Figure 11.29. Topological isomers of {[Ag(35)](BF4)1/2(OH)1/2(CH3CN)1/4}n (a) and {[Ag (35)](BF4)1.5H2O}n (b) with lvt and PtS topologies, respectively.
obtained, as described in this chapter. Despite this progress, the challenge of predicting the exact structure of the assembled products remains because of the complicated assembly processes that can be influenced by factors such as the chosen counteranion and the reaction conditions.76–78 Thus, further studies that systematically explore the influence of such factors are needed. In addition to their structural diversity, silver(I) complexes also have interesting properties and applications in organic chemistry,26,30,31,63,79,80 which are explored in other chapters in this book. ACKNOWLEDGMENTS We are grateful to the National Natural Science Foundation of China for financial support (Grant 20828002).
REFERENCES 1. Lehn, J. M., Angew. Chem. Int. Ed. 1988, 27, 89–112. 2. Lehn, J. M., Supramolecular Chemistry: Concepts and Perspectives, VCH, Weinheim, 1995. 3. Young, A. G.; Hanton, L. R., Coord. Chem. Rev. 2008, 252, 1346–1386. 4. For an example of two-coordinated linear silver(I), see Fitchett, C.M.; Steel, P.J., Dalton Trans. 2006, 4886–4888. 5. For an example of three-coordinated triangular silver(I), see Chiang, L. M.; Yeh, C. W.; Chan, Z. K.; Wang, K. M.; Chou, Y. C.; Chen, J. D.; Wang, J. C.; Lai, J. Y., Cryst. Growth Des. 2008, 8, 470–477.
REFERENCES
353
6. For an example of three-coordinated T-shaped silver(I), see Ganguly, S.; Chattopadhyay, S.; Sinha, C.; Chakravorty, A., Inorg. Chem. 2000, 39, 2954–2956. 7. For an example of four-coordinated square planar silver(I), see Reger, D. L.; Gardinier, J. R.; Smith, M. D., Inorg. Chem. 2004, 43, 3825–3832. 8. For an example of four-coordinated tetrahedral silver(I), see Argent, S. P.; Adams, H.; RiisJohannessen, T.; Jeffery, J. C.; Harding, L. P.; Clegg, W.; Harringtond, R. W.; Ward, M. D., Dalton Trans. 2006, 4996–5013. 9. For an example of five-coordinated trigonal bipyramidal silver(I), see Halper, S. R.; Do, L.; Stork, J. R.; Cohen, S. M., J. Am. Chem. Soc. 2006, 128, 15255–15268. 10. For an example of five-coordinated tetragonal pyramidal silver(I), see Habata, Y.; Noto, K.; Osaka, F., Inorg. Chem. 2007, 46, 6529–6534. 11. For an example of six-coordinated octahedral silver(I), see Niu, C. Y.; Wu, B. L.; Zheng, X. F.; Zhang, H. Y.; Li, Z. J.; Hou, H. W., Dalton Trans. 2007, 5710–5713. 12. For an example of six-coordinated trigonal prism silver(I), see Fan, J.; Zhu, H. F.; Okamura, T. A.; Sun, W. Y.; Tang, W. X.; Ueyama, N., Chem. Eur. J. 2003, 9, 4724–4731. 13. For an example of seven-coordinated silver(I), see Prince, P. D.; Steed, J. W., Supramol. Chem. 1998, 10, 155–158. 14. For an example of eight-coordinated tetragonal prism silver(I), see Jones, P. G.; Gries, T.; Grutzmacher, H.; Roesky, H. W.; Schimkowiak, J.; Sheldrick, G. M., Angew. Chem. Int. Ed. 1984, 23, 376–376. 15. Katz, M. J.; Sakai, K.; Leznoff, D. B., Chem. Soc. Rev. 2008, 37, 1884–1895. 16. Xu, J.; Yuan, Q.; Bai, Z. S.; Su, Z.; Sun, W. Y., Inorg. Chem. Commun. 2009, 12, 58–61. 17. Steel, P. J.; Fitchett, C. M., Coord. Chem. Rev. 2008, 252, 990–1006. 18. Sun, W. Y.; Yoshizawa, M.; Kusukawa, T.; Fujita, M., Curr. Opin. Chem. Biol. 2002, 6, 757–764. 19. Fujita, M.; Umemoto, K.; Yoshizawa, M.; Fujita, N.; Kusukawa, T.; Biradha, K., Chem. Commun. 2001, 509–518. 20. Fujita, M.; Tominaga, M.; Hori, A.; Therrien, B., Acc. Chem. Res. 2005, 38, 371–380. 21. Sun, W. Y. in New Development in Organometallic Chemistry Research, Cato, M. A., ed., Nova Science Publishers, New York, 2006, Chapter 2, pp. 37–61. 22. Xu, G. C.; Ding, Y. J.; Okamura, T. A.; Huang, Y. Q.; Liu, G. X.; Sun, W. Y.; Ueyama, N., Cryst. Eng. Comm. 2008, 10, 1052–1062. 23. Xu, G. C.; Ding, Y. J.; Huang, Y. Q.; Liu, G. X.; Sun, W. Y., Micropor. Mesopor. Mater. 2008, 113, 511–522. 24. Wang, X. F.; Lv, Y.; Okamura, T. A.; Kawaguchi, H.; Wu, G.; Sun, W. Y.; Ueyama, N., Cryst. Growth Des. 2007, 7, 1125–1133. 25. Liu, H. K.; Huang, X.; Lu, T.; Wang, X.; Sun, W. Y.; Kang, B. S., Dalton Trans. 2008, 3178–3188. 26. Sun, W. Y.; Fan, J.; Okamura, T. A.; Xie, J.; Yu, K. B.; Ueyama, N., Chem. Eur. J. 2001, 7, 2557–2562. 27. Sun, W. Y.; Fan, J.; Yu, K. B., Acta Chim. Sinca. 2001, 59, 1102–1105. 28. Fan, J.; Sun, W. Y.; Okamura, T. A.; Xie, J.; Tang, W. X.; Ueyama, N., New J. Chem. 2002, 26, 199–201. 29. Hiraoka, S.; Yi, T.; Shiro, M.; Shionoya, M., J. Am. Chem. Soc. 2002, 124, 14510–14511.
354
SUPRAMOLECULAR CHEMISTRY OF SILVER
30. Hiraoka, S.; Harano, K.; Shiro, M.; Shionoya, M., Angew. Chem. Int. Ed. 2005, 44, 2727–2731. 31. Huang, Y. Q.; Shen, Z. L.; Okamura, T. A.; Wang, Y.; Wang, X. F.; Sun, W. Y.; Yu, J. Q.; Ueyama, N., Dalton Trans. 2008, 204–213. 32. Sumby, C. J.; Hardie, M. J., Angew. Chem. Int. Ed. 2005, 44, 6395–6399. 33. Iki, N.; Ohta, M.; Horiuchi, T.; Hoshino, H., Chem. Asian J. 2008, 3, 849–853. 34. Zhang, J.; Xu, X.; James, S. L., Chem. Commun. 2006, 4218–4220. 35. Willans, C. E.; Anderson, K. M.; Junk, P. C.; Barbour, L. J.; Steed, J. W., Chem. Commun. 2007, 3634–3636. 36. Iijima, S., Nature 1991, 354, 56–58. 37. Special issue on carbon nanotubes, Acc. Chem. Res. 2002, 35, 997–1113. 38. Fan, J.; Zhu, H. F.; Okamura, T. A.; Sun, W. Y.; Tang, W. X.; Ueyama, N., Inorg. Chem. 2003, 42, 158–162. 39. Harada, R.; Matsuda, Y.; Okawa, H.; Kojima, T., Angew. Chem. Int. Ed. 2004, 43, 1825–1828. 40. Tashiro, S.; Tominaga, M.; Kusukawa, T.; Kawano, M.; Sakamoto, S.; Yamaguchi, K.; Fujita, M., Angew. Chem. Int. Ed. 2003, 42, 3267–3270. 41. Yamaguchi, T.; Tashiro, S.; Tominaga, M.; Kawano, M.; Ozeki, T.; Fujita, M., Am. Chem. Soc. 2004, 126, 10818–10819. 42. Hong, M. C.; Zhao, Y. J.; Su, W. P.; Cao, R.; Fujita, M.; Zhou, Z. Y.; Chan, A. S. C., Angew. Chem. Int. Ed. 2000, 39, 2468–2470. 43. Dong, Y. B.; Jiang, Y. Y.; Li, J.; Ma, J. P.; Liu, F. L.; Tang, B.; Huang, R. Q.; Batten, S. R., J. Am. Chem. Soc. 2007, 129, 4520–4521. 44. Kleina, C.; Graf, E.; Hosseini, M. W.; Cianb, A. D.; Fischer, J., Chem. Commun. 2000, 239–240. 45. Hong, M. C.; Zhao, Y. J.; Su, W. P.; Cao, R.; Fujita, M.; Zhou, Z. Y.; Chan, A. S. C., J. Am. Chem. Soc. 2000, 122, 4819–4820. 46. Sauvage, J. P.; Dietrich-Buchecker C., eds., Molecular Catenanes, Rotaxanes, and Knots, Wiley-VCH, Weinheim, 1999. 47. Vignon, S. A.; Wong, J.; Tseng, H. R.; Stoddart, J. F., Org. Lett. 2004, 6, 1095–1098. 48. Hori, A.; Sawada, T.; Yamashita, K.; Fujita, M., Angew. Chem. Int. Ed. 2005, 44, 4896–4899. 49. Zhu, H. F.; Fan, J.; Okamura, T. A.; Sun, W. Y.; Ueyama, N., Cryst. Growth Des. 2005, 5, 289–294. 50. Fan, J.; Sun, W. Y.; Okamura, T. A.; Tang, W. X.; Ueyama, N., Inorg. Chim. Acta 2004, 357, 2385–2389. 51. Jin, C. M.; Lu, H.; Wu, L. Y.; Huang, J., Chem. Commun. 2006, 5039–5041. 52. Jin, C. M.; Wu, L. Y.; Lu, H.; Xu, Y., Cryst. Growth Des. 2008, 8, 215–218. 53. Sague, J. L.; Fromm, K. M., Cryst. Growth Des. 2006, 6, 1566–1568. 54. Kim, K., Chem. Soc. Rev. 2002, 31, 96–107. 55. Fan, J.; Sun, W. Y.; Okamura, T. A.; Zheng, Y. Q.; Sui, B.; Tang, W. X.; Ueyama, N., Cryst. Growth Des. 2004, 4, 579–584. 56. Wang, Z. B.; Zhu, H. F.; Zhao, M.; Li, Y. Z.; Okamura, T. A.; Sun, W. Y.; Chen, H. L.; Ueyama, N., Cryst. Growth Des. 2006, 6, 1420–1427.
REFERENCES
355
57. Hoskins, B. F.; Robson, R.; Slizys, D. A., J. Am. Chem. Soc. 1997, 119, 2952–2953. 58. Liu, H. Y.; Wu, H.; Ma, J. F.; Song, S. Y.; Yang, J.; Liu, Y. Y.; Su, Z. M., Inorg. Chem. 2007, 46, 7299–7311. 59. Jouaiti, A.; Hosseini, M. W.; Kyritsakas, N., Chem. Commun. 2003, 472–473. 60. Deng, Z. P.; Zhu, L. N.; Gao, S.; Huo, L. H.; Ng, S. W., Cryst. Growth Des. 2008, 8, 3277–3284. 61. Zhang, X. L.; Guo, C. P.; Yang, Q. Y.; Lu, T. B.; Tong, Y. X.; Su, C. Y., Chem. Mater. 2007, 19, 4630–4632. 62. Fan, J.; Sun, W. Y.; Okamura, T. A.; Tang, W. X.; Ueyama, N., Inorg. Chem. 2003, 42, 3168–3175. 63. Zhang, J. P.; Horike, S.; Kitagawa, S., Angew. Chem. Int. Ed. 2007, 46, 889–892. 64. Bai, Y.; Duan, C. Y.; Cai, P.; Dang, D. B.; Meng, Q. J., Dalton. Trans. 2005, 2678–2680. 65. Hu, T. L.; Du, W. P.; Hu, B. W.; Li, J. R.; Bu, X. H.; Cao, R., Cryst. Eng. Comm. 2008, 10, 1037–1043. 66. Zhang, J. P.; Lin, Y. Y.; Zhang, W. X.; Chen, X. M., J. Am. Chem. Soc. 2005, 127, 14162–14163. 67. Du, M.; Zang, Z. H.; Tang, L. F.; Wang, X. G.; Zhao, X. J.; Batten, S. R., Chem. Eur. J. 2007, 13, 2578–2586. 68. Carlucci, L.; Ciani, G.; Proserpioa, D. M.; Portab, F., Cryst. Eng. Comm. 2006, 8, 696–706. 69. Natarajan, R.; Savitha, G.; Dominiak, P.; Wozniak, K.; Moorthy, J. N., Angew. Chem. Int. Ed. 2005, 44, 2115–2119. 70. Cordes, D. B.; Hanton, L. R., Inorg. Chem. 2007, 46, 1634–1644. 71. Abrahams, B. F.; Jackson, P. A.; Robson, R., Angew. Chem. Int. Ed. 1998, 37, 2656. 72. Carlucci, L.; Ciani, G.; Proserpio, D. M.; Rizzato, S., J. Chem. Soc. Dalton Trans. 2000, 3821–3827. 73. Galet, A.; Niel, V.; Munoz, M. C.; Real, J. A., J. Am. Chem. Soc. 2003, 125, 14224–14225. 74. Niel, V.; Thompson, A. L.; Goeta, A. E.; Enachescu, C.; Hauser, A.; Galet, A.; Munoz, M. C.; Real, J. A., Chem. Eur. J. 2005, 11, 2047–2060. 75. Du, M.; Zhao, X. J.; Guoa, J. H.; Batten, S. R., Chem. Commun. 2005, 4836–4838. 76. Domasevitch, K. V.; Solntsev, P. V.; Guralskiy, I. A.; Krautscheid, H.; Rusanov, E. B.; Chernegac, A. N.; Howard, J. A. K., Dalton Trans. 2007, 3893–3905. 77. Northrop, B. H.; Yang, H. B.; Stang, P. J., Chem. Commun. 2008, 5896–5908. 78. Gimeno, N.; Vilar, R., Coord. Chem. Rev. 2006, 250, 3161–3189. 79. Yan, Z. Q.; McCracken, T.; Xia, S. J.; Maslak, V.; Gallucci, J.; Hadad, C. M.; Badji, J. D., J. Org. Chem. 2008, 73, 355–363. 80. Halbes-Letinois, U.; Weibel, J. M.; Pale, P., Chem. Soc. Rev. 2007, 36, 759–769.
12 A CRITICAL COMPARISON: COPPER, SILVER, AND GOLD A. STEPHEN K. HASHMI Institute of Organic Chemistry, Ruprecht-Karls-University, Heidelberg, Germany
12.1 Introduction 12.2 Reactions Catalyzed by Copper, Silver, or Gold 12.2.1 Aldehyde–Alkyne–Amine Coupling 12.2.2 Carbene Insertion Reactions 12.2.3 In Silico Comparison of Organocopper(I), Organosilver(I), and Organogold(I) -Ate Compounds 12.2.4 Cyclization of ortho-Alkynylbenzaldehydes 12.2.5 Allenyl Ketones: The Cycloisomerization to Furans 12.2.6 A Thiol in the Substrate: The Cyclization of a-Thioallenes 12.2.7 Furans by Propargyl Claisen Reaction 12.2.8 Tandem Cyclization/Pinacol Rearrangement 12.2.9 Furanones by Domino Heterocyclization/1,2 Shift 12.2.10 Conia-ene Reaction 12.3 Reactions Catalyzed by Silver or Gold 12.3.1 Cyclization of N-Propargylcarboxamides 12.3.2 Dake’s Pyrrole Synthesis 12.3.3 Combination with Organocatalysis 12.3.4 Vinylallenes Deliver Cyclopentadienes 12.3.5 a-Pyrones by a Cascade Reaction 12.3.6 Dihydrofurans from Propargyl Esters 12.3.7 Methylene Butyrolactones by Addition of Carboxylates to Alkynes 12.3.8 Hydroarylation of Allenes 12.3.9 Different Products by Silver and Gold Catalysts 12.4 Reactions Catalyzed by Copper or Silver 12.4.1 General Trends
Silver in Organic Chemistry Edited by Michael Harmata Copyright 2010 John Wiley & Sons, Inc.
357
358
A CRITICAL COMPARISON: COPPER, SILVER, AND GOLD
12.4.2 Pyrroles by Hydroamination 12.4.3 Copper/Silver Cocatalysis 12.4.4 Carbonylations 12.5 Conclusion References
12.1
INTRODUCTION
In organic synthesis, catalysis by silver is an important methodology. The first time that chemists hear about silver catalysts during their studies is in an introductory organic chemistry course—a technical process, in which the oxidation of ethene to oxirane is catalyzed by silver.1 This process is not possible with propene, as the allylic position will preferentially be oxidized, but the research since the late 1990s has demonstrated that certain gold catalysts are able to give good selectivities for this top target in industrial chemistry.2,3 This basic example demonstrates what will be discussed in this chapter, similarities and differences between the elements of the copper triad, but with a focus on homogeneous catalysis. All three metals have been used extensively in homogeneous catalysis of organic reactions. Early work focused on copper; thus the catalysis-related literature for this element is abundant.4 Silver had sustained continuous interest,5 but never to the extent that copper experienced. Gold is the “youngest” member in the field of catalysis, but is currently (as of 2009) catching up at an incredible rate.6 This chapter discusses the similarities and differences that one can find in the publications covering homogeneous catalysis by these metals. 12.2 12.2.1
REACTIONS CATALYZED BY COPPER, SILVER, OR GOLD Aldehyde–Alkyne–Amine Coupling
The coupling of the three components -aldehyde, alkyne and amine- has been called “A3 coupling.”7 It delivers propargyl amines, which are important synthetic intermediates and, in contrast to previously used routes, avoids the use of strong-base reagents such as BuLi, ethylmagnesium bromide or LDA, and the need for the kinds of solvents necessary for these reagents. An initial finding was the catalytic activity of CuI under solvent-free conditions in which an acceleration by microwave irradiation was reported (Scheme 12.1).8 The reaction probably proceeds through an iminium or imine species; thus there is a close relationship to the historic work of Reppe for the formation of the corresponding propargyl alcohols.
359
REACTIONS CATALYZED BY COPPER, SILVER, OR GOLD
1
R
H
R3 N R4 H
O
+
H
+
R2
CuI, microwave
R1
R3 N R4 +
R2
H 2O
Scheme 12.1. Copper-catalyzed reaction of aldehydes with alkynes and secondary amines.
Subsequent research revealed that with different soluble silver salts only low conversions between 25% and 45% could be obtained, which was due to the in situ reduction of the more noble silver. However, with AgCl, AgBr, and AgI, good catalytic activity was observed.9 For example, with as little as 0.2 mol% AgI in water at 100 C, excellent yields could be obtained. Investigation of gold catalysts showed that with AuBr3 a very clean reaction can be observed in water.10 Similar results were obtained with AuCl and AuI; it was speculated that gold(III) is reduced in situ to gold(I). The choice of AuBr3 for most of the discussion in this chapter ultimately resulted from the fact that at the time of the investigation, on normalization of gold for catalysis, this was the least expensive gold halide commercially available. The use of an ionic liquid, [Bmim][BF4], as the solvent for these conversions has been reported; in this case copper and gold catalysts tended to deliver trimers of the aliphatic aldehydes as side products.11 Only the copper-catalyzed version of this reaction was extended to an asymmetric version (known as “AA3 coupling”). Both CuOTf and the pybox ligand gave excellent enantioselectivity values of up to 99% ee for anilines as amine components. By anomalous diffraction on a derivative bearing a bromo atom, the absolute configuration of the product could be determined to be (R) for the ligand configuration shown in Scheme 12.2.12 There is no evidence in the literature that chiral gold or silver complexes have also been tested; in the context of the success of chiral gold and silver complexes in the catalytic asymmetric aldol reaction,6b this would have been an interesting option.
O N Ph R1
H
+
O H
R2
+
H N Ar H
O
N N + CuOTf toluene or water
Ph R1
H N Ar R2
Scheme 12.2. Catalytic asymmetric version of the copper-catalyzed reaction.
+
H 2O
360
A CRITICAL COMPARISON: COPPER, SILVER, AND GOLD
N
O O Ph
+ H 2O
Ph catalyst
+
+
N H
O water
Scheme 12.3. Diastereoselective A3 coupling.
This kind of comparison has been made only for the diastereoselective version of this reaction.13 Water was still used as solvent for the conversion of phenylacetylene, 2-benzyloxyheptanal, and piperidine (Scheme 12.3), but the reaction temperature was lowered to 60 C, room temperature, or even 0 C. As shown in Table 12.1, neither with copper(I) halides (entry 1) nor with silver(I) halides (entry 2) could any satisfactory results be obtained. Only gold(I) halides were still reactive enough at these temperatures. With AuCl at 60 C, the diastereomeric ratio (dr) was still low (entry 3); lowering the temperature to room temperature led to a slight increase of the dr (entry 4); and going to 0 C further increased the dr, but at the same time diminished the yield. With AuI at room temperature (entry 5) and 0 C (entry 6), only marginally better results were obtained. 12.2.2
Carbene Insertion Reactions
Cyclohexane is a saturated hydrocarbon in which no regioselectivity problems of C--H insertion can occur. The reaction of cyclohexane with ethyl diazoacetate was investigated in a thorough study by Perez et al.14 The N-heterocyclic carbene ligand IPr (IPr ¼ 1,3-bis(diisopropylphenyl)imidazol-2-yliden) was used, all three compounds IPrMCl (M ¼ Cu, Ag, Au) were inactive as catalysts in cyclohexane (Table 12.2, entry 1). Addition of the sodium BARF salt (NaB[3,5-(CF3)2C6H3]4) gave ethyl cyclohexyl acetate as the C--H insertion product of the carbene (Scheme 12.4). In this case, the best yields were obtained for the gold(I) catalyst (Table 12.2, entry 4), the copper(I) catalyst was second in this ranking (entry 2), with silver(I) affording the lowest yield (74%, entry 3). TABLE 12.1. Substrate-Induced Diastereoselectivity in A3 Coupling in Water Entry 1 2 3 4 5 6 7
Catalyst
Temperature
Time (h)
Yield (%)
dr
CuCl, CuBr, or CuI AgCl, AgBr, or AgI AuCl AuCl AuCl AuI AuI
18 18 18 18 12 18 12
— — 89 89 57 100 58
— — 64 : 36 72 : 28 75 : 25 73 : 27 76 : 24
60 C 60 C 60 C r.t. 0 C r.t. 0 C
361
REACTIONS CATALYZED BY COPPER, SILVER, OR GOLD
TABLE 12.2. Yields of the Metal-Catalyzed Carbene Insertion Entry 1 2 3 4
Catalyst
Yield (%)
IPrCuCl, IPrAgCl, or IPrAuCl IPrCuCl þ NaBAr0 4 IPrAgCl þ NaBAr0 4 IPrAuCl þ NaBAr0 4
— 80 74 84
N
M
Cl
N
O N2
O
NaBAr ′4
+
O
O
Scheme 12.4. Carbene insertions catalyzed by copper(I), silver(I), and gold(I).
12.2.3 In Silico Comparison of Organocopper(I), Organosilver(I), and Organogold(I)-Ate Compounds Organocuprate(I) reagents like R2CuLi readily react with organic electrophiles (E þ ), for example, alkyl halides or a,b-unsaturated carbonyl compounds (Scheme 12.5).15 On the other hand, there are only a few reports where one could assume a participation of organoargentate16 or organoaurate17 intermediates in organic reactions. In a computational study, Nakamura et al.17 examined these reactions. They had a close look at both steps, the association with the electrophile to form the intermediate R2CuE, and the subsequent elimination of the product R--E. The higher reactivity of the organocuprate(I) compounds is based on the high-lying 3d orbitals; the 4d orbitals of the corresponding silver(I) species and the 5d orbitals of the gold(I) species are lowlying. Owing to the relativistic effect, the 5d orbitals and the 4d orbitals are close in energy. In addition, the next step, the elimination of R--E from the intermediate, is easier for copper than for silver or gold. This results from the higher stability of the
[R–CuI–R]
_
+
E
+
E R–Cu–R
Scheme 12.5. Typical reactions of organocuprates.
R–Cu
+
R–E
362
A CRITICAL COMPARISON: COPPER, SILVER, AND GOLD
carbon–gold bond, which again is a relativistic effect18 and follows the general rule that the transition metal–carbon bond strength increases from the first-row transition metals to the heavier analogs. 12.2.4
Cyclization of ortho-Alkynylbenzaldehydes
This reaction type was intensively investigated by Yoshinori Yamamoto et al., who focused mainly on copper and gold.19 In the context of an application in the synthesis of azaphilones, the Porco group thoroughly compared the reactivity of the three metals in the conversion of a hydroxylated ortho-alkynylbenzaldehyde to the corresponding pyrylium salt (Scheme 12.6, Table 12.3).20 The catalyst screening involved control experiments without a catalyst [excluding catalytic activity of the trifluoroacetic acid (TFA), Table 12.3, entry 1]. Copper salts (entries 2 and 3) were effective; after 20 min copper(I) had converted more substrate than had copper(II). Silver(I) (entry 4) was even better. Four different gold(III) catalysts were tested; the acetate (entry 7) was the most active catalyst, followed by the chloride (entry 6) and the bromide (entry 5). In general, this type of screening had been helpful for the authors, as it delivered a highly active catalyst, but in the overall context of this chapter it is not very helpful, as systematic changes are missing. For example, keeping the counterion constant and comparing the different metals in the same oxidation state would have been helpful. Still, this nicely points out one of the difficulties of the field, namely, that in skipping from one entry to the next in the tables, two or more changes may be made simultaneously, without allowing for a proper analysis of the data at the end. 12.2.5
Allenyl Ketones: The Cycloisomerization to Furans
The cycloisomerization of allenyl ketones was initially described as being catalyzed by rhodium(I) or silver(I) by Marshall et al.21 The activity of copper, silver, and gold for this reaction was first compared in two papers published later (Scheme 12.7).22 In the case of copper and silver, only a cycloisomerization was observed (Table 12.4, entries 1 and 2); with gold, a dimer is obtained as well (entry 3). This initial work led to a beautiful synthetic methodology by Gevorgyan et al.23 They discovered that two different products are formed from bromoallenyl ketones, and again this depends not only on the catalyst but also on the solvent. This time, there are two different monomers; no dimer is observed (Scheme 12.8). A more detailed view of the selectivities (Table 12.5) reveals a selective formation of the 2,4-disubstituted furan with the copper(I) catalysts (entries 1 and 2), a low HO O OH
catalyst 1,2-dichloroethane/TFA (10:1)
HO O+
r.t.
OH
Scheme 12.6. Formation of 2-benzopyrylium salts.
_ CF3 CO 2
363
REACTIONS CATALYZED BY COPPER, SILVER, OR GOLD
TABLE 12.3. Different Catalysts for the Formation of 2-Benzopyrylium Salts Entry
Catalyst
1 2 3 4 5 6 7 a
Time (min)
Conversion (%) <1 41 74 94 48 75 100
a
— 5 mol% Cu(OTf)2 2.5 mol% [Cu(OTf)]2toluene 5 mol% AgNO3 5 mol% AuBr3 5 mol% AuCl3 5 mol% Au(OAc)3
20 20 20 20 20 20 1
Reaction performed at 40 C.
•
catalyst
1
R
MeCN, r.t.
R1
O
O
+
R1 R1
O
O
Scheme 12.7. Depending on the metal, different products can be observed.
reactivity with the silver(I) catalysts (entry 3), and a high selectivity for the 2,5-disubstitution with AuCl3 in toluene (entry 4). However in THF, the selectivity is completely inverted (entry 5). With Ph3AuCl in toluene the selectivity is lower than with AuCl3 in this solvent (entry 6), but with Et3PAuCl, the selectivity switches again (entry 7). Similar effects have been observed with the diphenylallenyl ketones (Scheme 12.9).23 In this case, only one product is formed; one of the phenyl groups has to migrate. The individual results are listed in Table 12.6. Copper(I) halides are not active for this conversion (entries 1–3); only the copper triflates deliver the product, with copper (II) affording a better yield (entries 4 and 5). All silver salts are active, and again the triflates are superior and in dichloromethane even operate at room temperature
TABLE 12.4. Cycloisomerization or Cycloisomerization/Dimerization of Allenyl Ketones (R1 ¼ Me) Entry
Catalyst
1 2 3
Conversion (%)
1 mol% CuCl 1 mol% AgNO3 1 mol% AuCl3
Ratio Monomer : Dimer
100 100 94
1:0 1:0 1:1
Br
Br
•
catalyst
O
C8 H 17
solvent, r.t.
Br
O
C8 H 17
2,5-Disubstituted
+
O
C8 H 17
2,4-Disubstituted
Scheme 12.8. Two different cycloisomers from bromoallenyl ketones.
364
A CRITICAL COMPARISON: COPPER, SILVER, AND GOLD
TABLE 12.5. Product Distribution of the Bromoallenyl Ketone Cyclization
Entry 1 2 3 4 5 6 7 a
Ratio Conversion 2,5-Disubstitution : 2,4-Disubstitution (%)
Catalyst
Solvent/Time
10 mol% CuCl 10 mol% CuI 5 mol% AgBF4 1 mol% AuCl3 1 mol% AuCl3 1 mol% Ph3AuCl 1 mol% Et3PAuCl
Toluene/1 day Toluene/1 day CH2Cl2/1 day Toluene/5 min THF/5 min Toluene/9 h Toluene/9 h
29 21 Traces 86 78
0:1 0:1 — 95 : 5 5 : 95 16 : 84 <1 : 99
a a
Conversion not determined.
Ph Ph
catalyst
•
4-Br–Ph
solvent temperature
O
Ph Ph
4-Br–Ph
O
Scheme 12.9. Only one isomer by a 1,2-phenyl shift.
(entries 6–8). Switching to gold(I) shows that AuI is not active (entry 9), while Ph3PAuOTf is active again, even in DCM at room temperature (entries 10 and 11). On the other hand, AuBr3 leads to a low yield only (entry 12). 12.2.6
A Thiol in the Substrate: The Cyclization of a-Thioallenes
In a spectacular work, Morita and Krause24 studied the cyclization of allenylthiocarbinols to dihydrothiophenes by copper, silver, and gold catalysts. The reaction was the first example of a gold-catalyzed C--S bond formation and was highly stereoTABLE 12.6. Conditions and Yields for the 1,2-Phenyl Shift in Diphenylallenyl Ketones Entry 1 2 3 4 5 6 7 8 9 10 11 12
Catalyst 5 mol% CuCl 5 mol% CuBr 5 mol% CuI 5 mol% [CuOTf]2toluene 5 mol% Cu(OTf)2 5 mol% AgPF6 5 mol% AgOTf 20 mol% AgOTf 5 mol% AuI 1 mol% Ph3PAuOTf 1 mol% Ph3PAuOTf 5 mol% AuBr3
Solvent/Temperature
Toluene/100 C Toluene/100 C Toluene/100 C Toluene/100 C Toluene/100 C Toluene/100 C Toluene/100 C DCM/r.t. Toluene/100 C Toluene/100 C DCM/r.t. Toluene/100 C
Yield (%) — — — 42 95 47 80 70 Traces 100 99 23
365
REACTIONS CATALYZED BY COPPER, SILVER, OR GOLD
•
OMe
H
catalyst CH2Cl2, 20°C
OMe S
HS dr >99 : 1
dr >99 : 1
Scheme 12.10. Transition-metal-catalyzed cyclization of an allenylthiocarbinol.
selective, the diastereomerically pure starting material delivering only one diastereomer of the product. In the case of gold(III) chloride, the formation of the disulfide derived from the starting material provided some evidence for an in situ reduction of gold(III) to gold(I). Similar evidence for a possible in situ reduction of gold(III) was obtained in related work on allenyl carbinols (C--C bond formation in the latter case).25 The test substrate for the exceptional systematic screening is shown in Scheme 12.10. The results of the study (in Table 12.7) clearly show that the copper(I) halides tested are not active (entries 1 and 2). The same is true for silver chloride and silver tetrafluoroborate (entries 3 and 4). Only gold catalysts show activity, with gold(I) as the chloride and iodide (entries 5–7) as well as the cationic (phosphane)gold(I) species (entry 9), but not as the coordinatively saturated phosphanegold(I) chloro complex. Both gold(III) chloride and bromide are active, but the formation of the disulfide mentioned above, indicated that redox processes changes the nature of the gold complex. The authors also screened different solvents. Dichloromethane was by far superior to THF, toluene, or hexane. 12.2.7
Furans by Propargyl Claisen Reaction
The Kirsch group studied the propargyl Claisen reaction, which consists of an initial [3,3]-sigmatropic reaction and a subsequent ring closure (Scheme 12.11).26 TABLE 12.7. Systematic Comparison of Catalysts Entry 1 2 3 4 5 6 7 8 9 10 11
Catalyst
Time
20 mol% CuCl 20 mol% CuI 15 mol% AgCl 15 mol% AgBF4 5 mol% AuCl 5 mol% AuI 1 mol% AuI 5 mol% Ph3PAuCl 5 mol% Ph3PAuCl/10 mol% AgBF4 5 mol% AuCl3 5 mol% AuBr3
1h 2 days 2 days 2 days 1.5 h 5 min 1.5 h 7 days 4h 3h 20 min
Yield (%) Traces — — Traces 56 88 64 Traces 52 58 56
366
A CRITICAL COMPARISON: COPPER, SILVER, AND GOLD
O
O EtO
catalyst
O
EtO
EtO +
DCM 23°C, time
O
•
O
O
Scheme 12.11. Highly substituted furan heterocycles by coinage metal catalysis.
Again, the authors tested some variations of the metal catalyst; the entries with the copper triad metals are listed in Table 12.8. Copper(I) iodide was inactive (entry 1), silver tetrafluoroborate and AuCl3 both delivered the allene without cycloisomerization within 24 h (entries 2 and 4), and potassium tetrachloroaurate was inactive (entry 3). Without activation, phosphanegold(I) chlorides were inactive (entry 5), but after removal of the chloro ligand by silver salts, excellent yields were obtained in short reaction times of 40 min (entries 6–8). 12.2.8
Tandem Cyclization/Pinacol Rearrangement
Here the substrates are 3-silyloxy-1,5-enynes; the concept was to combine a metalinduced 6-endo-trig cyclization with a ring contraction by a sigmatropic rearrangement (Scheme 12.12).27 From the cationic intermediate, elimination to the biaryl product can be observed in addition to the aldehyde. The different catalysts tested are shown in Table 12.9. The obligatory CuI (entry 1) delivers a mixture of starting material and the two products, and AgSbF6 does not convert the substrate at all (entry 2). Among the (phosphane)gold(I) complexes, only those with SbF6 counterions are active (entries 3–7), affording good yields of the aldehyde in selective reactions. 12.2.9
Furanones by Domino Heterocyclization/1,2 Shift
The easily accessible (Scheme 12.13).28
a-hydroxypropargyl
ketones
deliver
furanones
TABLE 12.8. Cationic Gold(I) Complexes are Crucial for the Rearrangement Entry 1 2 3 4 5 6 7 8
Catalyst
Time
Yield (%)
5 mol% CuI 10 mol% AgBF4 5 mol% K[AuCl4] 2 mol% AuCl3 2 mol% Ph3PAuCl 2 mol% Ph3PAuCl/2 mol% AgBF4 2 mol% Ph3PAuCl/2 mol% AgSbF6 2 mol% Ph3PAuCl/2 mol% AgOTf
24 h 24 h 24 h 24 h 24 h 40 min 40 min 40 min
— 100 (only allene) — 7 þ 47 allene — 97 83 60
367
REACTIONS CATALYZED BY COPPER, SILVER, OR GOLD
catalyst
Et3 SiO
O
+
conditions
Et3 SiO
Ph
H
Ph
+ Ph
Ph
[M]
Scheme 12.12. Kirsch’s tandem reaction.
TABLE 12.9. Silver Delivers Only Poor Results in the Tandem Sequence Yield (%) Entry 1 2 3 4 5 6 7
Catalyst 10 mol% CuI 5 mol% AgSbF6 5 mol% [(Ph3PAu)3O]BF4 10 mol% Ph3PAuCl/5 mol% AgBF4 10 mol% Ph3PAuCl/5 mol% AgSbF6 2 mol% Ph3PAuCl/1 mol% AgSbF6 10 mol% Ph3PAuCl/5 mol% AgSbF6
Conditions
Aldehyde Biaryl
i-PrOH, toluene, 100 C, 24 h i-PrOH, DCM, 23 C, 1 h i-PrOH, DCM, 23 C, 24 h i-PrOH, DCM, 23 C, 24 h H2O, DCM, 23 C, 1 h i-PrOH, DCM, 23 C, 2.5 h i-PrOH, DCM, 23 C, 10 min
21 — — — 39 81 93
15 — — — — — —
Here, while the classical CuI (Table 12.10, entry 1) and AgSbF6 (entry 2) deliver some product, the gold(I) complex that was tested does not deliver any (entry 3). Thus, for the oxidation state þ 1, silver is the optimum, but it cannot compete with AuCl3, which delivered superior yields (entries 4 and 5), with temperatures slightly above room temperature preferred.
Ph O
Ph OH
catalyst conditions
+
O [M] OH
Ph O O
Scheme 12.13. An interesting pathway to furanones.
TABLE 12.10. Gold Outperforms the Other Catalysts, but it has to Be Gold(III) Entry 1 2 3 4 5
Catalyst 5 mol% CuI 5 mol% AgSbF6 5 mol% Ph3PAuBF4 5 mol% AuCl3 5 mol% AuCl3
Conditions
Yield (%)
80 C, DMF, 2 h 23 C, DCM, 1.5 h 23 C, DCM, 1.5 h 23 C, DCM, 1.5 h 38 C, DCM, 1.5 h
22 50 — 83 95
368
A CRITICAL COMPARISON: COPPER, SILVER, AND GOLD
O
O
O OMe
O
catalyst
OMe
conditions
Scheme 12.14. Alkylidenecyclopentanes from 1,3-dicarbonyl compounds.
TABLE 12.11. A 5-Exo-Dig Cyclization Delivers Five-Membered Rings Entry 1 2 3 4 5 6 7
12.2.10
Catalyst
Conditions
Yield (%)
10 mol% AgOTf 10 mol% AuCl3 10 mol% Ph3PAuCl 10 mol% Ph3PAuOTf 1 mol%[(Ph3PAu)3O]BF4 10 mol% (CuOTf)2C6H6/10 mol% Ag2O 10 mol% (CuOTf)2C6H6/10 mol% AgBF4
DCE, r.t., 18 h DCE, r.t., 30 min DCE, 60 C, 6 h DCE, r.t., <15 min HOTf, DCE, r.t., <15 min DCE, 100 C, 22 h DCE, 100 C, 23 h
50 30 — >95 >95 55 100
Conia-ene Reaction
The Conia-ene reaction was brought to the field of gold catalysis by Toste et al.29 It provided very mild conditions for this cyclization to a functionalized five-membered ring (Scheme 12.14). In this reaction, silver(I) is successful (Table 12.11, entry 1); gold(III) leads to an unselective conversion (entry 2), but gold(I) is more active and selective. As usual, coordinatively saturated gold(I) complexes fail (entry 3), but cationic gold(I) complexes deliver very good yields in short reaction times (entries 4 and 5). In a recent (2009) publication, Jin-Heng Li reported a copper/silver-cocatalyzed version of this isomerization.30 With the same substrate, here the combination of copper(I) and silver(I) is able to compete with the gold(I) catalysts (entries 6 and 7).
12.3 12.3.1
REACTIONS CATALYZED BY SILVER OR GOLD Cyclization of N-Propargylcarboxamides
A mild conversion of a simple class of substrates, the N-propargylcarboxamides, is possible with gold catalysts. With gold(I), the alkylidene oxazoles are formed; with gold(III), the aromatic oxazoles are formed.31 The reactions readily proceed at room temperature, and there are numerous examples of this reaction (Scheme 12.15). A 2008 publication by Harmata and Huang32 revealed that for a number of propargyl amide derivatives, 5 mol% of AgSbF6 at room temperature can achieve the same conversion to alkylidene oxazoles in 89–98% yield in minutes! In the case of disubstitution at the propargylic position, the reactions proceeds in DCM; without the
369
REACTIONS CATALYZED BY SILVER OR GOLD
R1
O
catalyst conditions
N H
R1
O
[Au(III)]
R1
N
O
N
Scheme 12.15. A simple route to the oxazole ring system.
propargylic substitution, MeNO2 was used as solvent. Still, gold catalysts can reach several thousand turnovers in these reactions, while for silver there are 20 turnovers. 12.3.2
Dake’s Pyrrole Synthesis
The reaction of homopropargylic ketones with primary amines is an excellent method for the synthesis of highly substituted pyrroles. As shown for one example in Scheme 12.16, the yields often are better with the silver(I) catalyst.33
O
N
catalyst
+
NH 2
+
1,2-dichloroethane 50°C
5 mol% AgOTf, 3.5 h
78%
5 mol% AuCl, AgOTf, Ph3 P, 3.5 h
74%
H2O
Scheme 12.16. An easy pathway to anellated pyrroles.
12.3.3
Combination with Organocatalysis
The transformation of an electrophilic carbonyl group to a nucleophilic enamine group with the help of a secondary amine was also applied in coinage metal catalysis.34 The reaction often, in much the same way as do other enamine catalyses, requires quite high catalyst loadings of the secondary amine. A test substrate for catalyst optimization was the malonate shown in Scheme 12.17. In a complex system like this, several control experiments are necessary. A characteristic set is listed in Table 12.12. Without any of the catalysts (entry 1), or with only one of the catalysts (entries 2 and 3), no product is formed. Only the combination of the gold and the organocatalyst affords the desired product (entry 4). O
Catalyst A
O
Catalyst B conditions
MeO2C CO2Me
MeO2C CO2Me
Scheme 12.17. Enamine formation and activation of the triple bond lead to ring closure.
370
A CRITICAL COMPARISON: COPPER, SILVER, AND GOLD
TABLE 12.12. Selected Experiments for Optimization Entry
Catalyst A
Catalyst B
Conditions
Yield (%)
1 2 3
— 10 mol% [(Ph3PAu)3O]BF4 —
— —
120 C, toluene, 24 h 70 C, CDCl3, 6 h 70 C, CDCl3, 6 h
RSMa RSM RSM
4
10 mol% [(Ph3PAu)3O]BF4
5
10 mol% AgSbF6
6
10 mol% AgOTf
a b
HNði-PrÞb2 HNði-PrÞb2 HNði-PrÞb2 HNði-PrÞb2
70 C, CDCl3, 3 h
86
70 C, CDCl3, 24 h
16
70 C, CDCl3, 2 h
59
Recovered starting material. Catalyst loading: 20 mol%
The use of AgSbF6 instead of the gold(I) catalyst shows only a little conversion after a long reaction time (entry 5). Again, AgOTf is the more reactive silver(I) complex, affording a higher yield in significantly less time (entry 6). 12.3.4
Vinylallenes Deliver Cyclopentadienes
The electrophilic cyclization of vinylallenes delivers cyclopentadienes with interesting substitution patterns.35 The anellated cyclopentadienes shown in Scheme 12.18 can be prepared in this way. Silver and gold catalysts have been compared using this substrate. Some of the catalysts tested are listed in Table 12.13. With only a silver(I) catalyst, the substrate decomposed (entry 1), while AuCl3 at least afforded 30% of the product (entry 2). As expected, the coordinatively saturated gold(I) complex gave
•
Ph catalyst CH 2Cl 2 conditions
Ph
Scheme 12.18. Tetrasubstituted cyclopentadienes from silver and gold catalysis.
TABLE 12.13. Different Efforts to Initiate the Cyclization Entry 1 2 3 4 5 6
Catalyst 5 mol% AgSbF6 5 mol% AuCl3 5 mol% Ph3PAuCl 2 mol% Ph3PAuCl/2 mol% AgSbF6 2 mol% Ph3PAuCl/2 mol% AgSbF6 1 mol% Ph3PAuCl/1 mol% AgSbF6
Conditions
0 C, 5 min 0 C, 5 min 23 C, 180 min 0 C, 1 min 20 C, 5 min 0 C, 1 min
Yield (%) Decomposition 30 — 97 93 96
371
REACTIONS CATALYZED BY SILVER OR GOLD
O O O
catalyst
O
CH 2Cl 2 temperature
Scheme 12.19. Difficult substitution patterns at a-pyrones made possible.
no conversion even after long reaction times at room temperature (entry 3), while the cationic gold(I) complex generated in situ by chloride abstraction was very successful under a range of conditions, including lower catalyst loadings (entries 4–6). 12.3.5
a-Pyrones by a Cascade Reaction
From propiolic acid propargyl esters an elegant route to a-pyrones has been developed by Schreiber and coworkers.36 The conversion is shown in Scheme 12.19. Out of a long table with reaction conditions and catalysts tested, the entries listed in Table 12.14 are the most important ones in this context. The silver(I) salt used delivered small amounts of the desired product only at reflux temperature (entries 1 and 2). Only the gold(I) catalyst gave a high yield of the a-pyrone; a higher temperature was preferable (entries 3 and 4). 12.3.6
Dihydrofurans from Propargyl Esters
The synthesis of dihydrofurans from propargyl esters is another reaction in which both gold and silver have been employed. In early work, Shigemasa et al.37 reported the efficient chirality transfer in the cycloisomerization of monoesters of butynediols. They used 8–15 mol% of AgBF4 as catalyst in benzene at 80 C in the dark. The corresponding gold(I)-catalyzed variation was reported by Gagosz and coworkers.38 A typical reaction is shown in Scheme 12.20. The diastereomeric ratio shifts only slightly in the direction of the thermodynamically more stable trans product; the enantioselectivity remains unchanged, and overall only 2 mol% of catalyst at room temperature is needed.
TABLE 12.14. Cationic Gold(I) Complexes Dominate This Reaction Entry 1 2 3 4 a
Catalyst
Temperature
Yield (%)
5 mol% AgSbF6 5 mol% AgSbF6 5 mol% Ph3PAuCl/5 mol% AgSbF6 5 mol% Ph3PAuCl/5 mol% AgSbF6
r.t. 40 C r.t. 40 C
Trace 11 a 61 81
Decomposition of most of the starting material.
372
A CRITICAL COMPARISON: COPPER, SILVER, AND GOLD
OBz BzO
OH
2 mol% Ph3PAuNTf 2
C5H11
C5H11
CH2Cl2, r.t.
O
35 min, 99% yield
d.r. 94:6 ee >98%
d.r. 92:8 ee >98%
Scheme 12.20. Stereoselective synthesis of 2,5-dihydrofurans.
12.3.7
Methylene Butyrolactones by Addition of Carboxylates to Alkynes
The silver-catalyzed synthesis of methylene butyrolactones has been known for a long time;39 the field was revived by Geneˆt more recently.40 Scheme 12.21 shows that the silver(I) catalysis required 10 mol% of catalyst and 110 C, but then gave a very good yield of product. Scheme 12.22 displays the corresponding investigation of Genet and coworkers; the different conditions and catalysts tested are shown in Table 12.15.
O
O
HO
10 mol% Ag2CO3
C16H 33
O
C16 H33
PhH, reflux
OH
99%
OH
Scheme 12.21. Pale’s silver-catalyzed cyclization.
O HO
O CO 2Me
O
catalyst
CO2 Me
solvent, r.t., 2 h
Scheme 12.22. Geneˆt’s gold-catalyzed cyclization.
TABLE 12.15. Lactone Formation Entry 1 2 3 4 5
Catalyst 5 mol% 5 mol% 5 mol% 5 mol% 5 mol%
AgOTf AuCl3 AuCl AuCl AuCl
Solvent
Yield (%)
MeCN MeCN MeCN Toluene DCM
10 84 90 60 63
373
REACTIONS CATALYZED BY SILVER OR GOLD
Silver(I) (entry 1) is significantly less active than gold(III) (entry 2) or gold(I) (entry 3), but again the table lists two changes at a time, the metal and the counterion. Tests with two more solvents (entries 4 and 5) did not give superior results.
12.3.8
Hydroarylation of Allenes
In the context of the total synthesis of ()-rhazinilam, an intramolecular hydroarylation of an allene (Scheme 12.23) was investigated.41 The authors report that the Ag(I) complexes were completely ineffective, while AuCl3 delivered 27% of the product and a mixture of 5 mol% AuCl3/20 mol% AgOTf gave 82%. The optimum was then reached with 5 mol% Ph3PAuOTf (92% yield). In all cases, the diastereoselectivity was better than 92 : 8. MeO2C
MeO2C catalyst
•
23°C
N
N
Scheme 12.23. Anellated pyrroles by hydroarylation of allenes.
12.3.9 12.3.9.1
Different Products by Silver and Gold Catalysts The Epoxide–Alkyne Reaction
For the reaction of the epoxide shown in Scheme 12.24 with a gold catalyst, the formation of a 1,2-dicarbonyl compound was reported; while with the silver salt, the acylindene was obtained.42 It is difficult to put this observation into a wider context, as this substrate type usually gives only the acylindenes with the gold catalysts as well.43
2 mol% AgSbF6
O
O
DCM, 23°C, 0.5 h
O
86%
5 mol% Ph3PAuCl 2 mol% AgSbF6 DCM, 23°C, 0.5 h
41% O
Scheme 12.24. Divergent pathways depending on the catalyst.
374
A CRITICAL COMPARISON: COPPER, SILVER, AND GOLD
TABLE 12.16. A Remarkable Change in Selectivity Yield (%) Entry 1 2 3 4 5 6
Catalyst
Five-Membered Ring
Six-Membered Ring
5 mol% AuCl3 5 mol% Ph3PAuCl 5 mol% Ph3PAuCl/5 mol% AgSbF6 5 mol% AgSbF6 5 mol% AgOCN 5 mol% Ag2O
— — —
— 25 >95
— 40 100
>95 60 0% OMe
OMe catalyst
O N
O
MeOH, r.t.
N
OMe
O
+ OMe
N
Scheme 12.25
12.3.9.2
The Carbonyl–Alkyne Reaction
With a quite specific class of compounds, Belmont and coworkers reported a selectivity switch when changing the catalyst.44 The reaction is shown in Scheme 12.25. With gold(III), no conversion was observed (Table 12.16, entry 1); the coordinatively saturated gold(I) complex gave a low yield of the 6-endo-dig cyclization product (entry 2). The cationic gold(I) complex with a free coordination site gives an excellent yield with the same selectivity (entry 3); the same is true for AgSbF6 alone (entry 4). But this is not true for all silver(I) catalysts; with a systematic increase in the pKa value of the conjugate acid of the silver counterion, increased portions of the 5-exo-dig product were produced (entries 5 and 6). 12.4 12.4.1
REACTIONS CATALYZED BY COPPER OR SILVER General Trends
A comparison of the carbonyl frequencies of a tris(pyrazolyl borate) complexes of copper(I) [TpBr3Cu(CO), n(CO) ¼ 2105 cm1] provides good evidence for a lower electron density at silver(I) [TpBr3Cu(CO), n(CO) ¼ 2157 cm1].14a By comparison of the vibrational spectra of cationic copper(I) and silver(I) tetra (acetonitrile) complexes, the different bond strengths of the metal nitrogen bonds have been determined and the influence of this bond strength on catalysis has been shown to be strong.45 12.4.2
Pyrroles by Hydroamination
An investigation comparing different hydroamination catalysts for the conversion of enamine alkynes to pyrroles is illustrated in Scheme 12.26.46 In this test, different
OMe
375
REACTIONS CATALYZED BY COPPER OR SILVER
TABLE 12.17. Some Catalysts Tested for the Reaction in Scheme 12.26 Entry 1 2 3 4 5 6 7 8 9
Catalyst
Yield (%)
4 mol% CuO 4 mol% CuAc2 4 mol% Cu(NO3)2 4 mol% CuCl2 4 mol% Ag2O 4 mol% AgOAc 4 mol% AgNO3 4 mol% AgCl 4 mol% Zn(NO3)2
13 33 53 65 6 7 14 9 99
EtO2 C
EtO 2C
catalyst
N Bu
H
MeCN, r.t.
N Bu
Scheme 12.26. Trisubstituted pyrroles by hydroamination.
copper(II) salts were compared to different silver(I) salts. It remains unclear why neither copper(I) nor gold catalysts were tested (the paper appeared in 2008). Table 12.17 shows these results. In the end, copper(II) was more reactive than silver(I), but the most effective systems were based on zinc(II) (entry 9, Table 2.17). In another more recent publication, the effect of copper(I) or silver(I) on the intramolecular addition of a sulfonamide to an alkyne was explored. In the case of silver, the reaction temperature was lower and the reaction times were shorter, and the best results were finally obtained with ethanol as solvent, but this was then never tested for copper.47 12.4.3
Copper/Silver Cocatalysis
Another version of the Conia-ene reaction is the focus of this research. Here the reaction of g-alkynic b-ketoesters was tested.48 The screening resulted in a copper(I)/ silver(I) cocatalysis system. It remains unclear why the authors did not at least run some reactions with gold(I) catalysts, too, especially since they even cite the work on the gold-catalyzed Conia-ene reaction in the introduction (Section 12.1). 12.4.4
Carbonylations
A paper dating from 1976 investigated the reaction of olefins or alcohols with cationic copper(I) and silver(I) complexes of carbon monoxide.49 The remarkable outcome of these kinetic measurements is that there is no difference between the catalytic behavior of Cu(CO)n þ and Ag(CO)2 þ .
376
12.5
A CRITICAL COMPARISON: COPPER, SILVER, AND GOLD
CONCLUSION
Most of the investigations in the area of coinage metal catalysis only provide data that do not allow a sound comparison of the catalytic activity or selectivity of the different metal complexes. In an often random manner, not only the metal but also the counterions and/or ligands and/or solvents are changed, and perhaps the reaction temperature as well. One gets the impression that the complexes that were at hand were the ones to be tested, not the ones that scientifically make sense at that point of the investigation. Still, as a general pattern, it seems that gold catalysts often have higher activity and operate at lower temperatures and lower catalyst loadings. A vast number of reactions still proceed with one catalyst, but apparently not with the other; there are too many reactions to list here, but often some of the crucial control experiments are missing. Much more work needs to be done here. For industrial applications where several coinage metals work as catalysts, certainly the cheaper metal is preferred, even if the catalyst loading would be higher, but for sensitive molecules, where functional group tolerance is an issue, the heavier analogs might be more interesting and useful.
REFERENCES 1. Boxhoorn, G. (Shell Internationale Research Maatschappij B.V., Netherlands), Patent EP 255975, 1988, p. 8. 2. Sinha, A. K.; Seelan, S.; Okumura, M.; Akita, T.; Tsubota, S.; Haruta, M., J. Phys. Chem. B 2005, 109, 3956–3965. 3. For a detailed discussion, see the only review that covers both heterogeneous and homogeneous catalysis by gold: Hashmi, A. S. K., Hutchings, G., Angew. Chem. 2006, 118, 8064–8105; Angew. Chem. Int. Ed. 2006, 45, 7896–7936. 4. (a) Alexakis, A.; B€ackvall, J. E.; Krause, N.; Pmie, O.; Diguez, M., Chem. Rev. 2008, 108, 2793–2795; (b) Harutyunyan, S. R.; den Hartog, T.; Geurts, K.; Minnaard, A. J.; Feringa, B. L., Chem. Rev. 2008, 108, 2824–2852; (c) Shibasaki, M.; Kanai, M., Chem. Rev. 2008, 108, 2853–2873; (d) Yamada, K.-I.; Tomioka, K., Chem. Rev. 2008, 108, 2874–2886; (e) Stanley, L. M.; Sibi, M. P., Chem. Rev. 2008, 108, 2887–2902; (f) Poulsen, T. B.; Jørgensen, K. A., Chem. Rev. 2008, 108, 2903–2915; (g) Deutsch, C.; Krause, N.; Lipshutz, B. H., Chem. Rev. 2008, 108, 2916–2927; (h) Meldal, M.; Tornøe, C. W., Chem. Rev. 2008, 108, 2952–3015; (i) Evano, G.; Blanchard, N.; Toumi, M., Chem. Rev. 2008, 108, 3054–3131. 5. (a) Naodovic, M.; Yamamoto, H., Chem. Rev. 2008, 108, 3132–3148; (b) Weibel, J.-M.; Blanc, A.; Pale, P., Chem. Rev. 2008, 108, 3149–3173; (c) Alvarez-Corral, M.; Mun˜ozDorado, M.; Rodrgıuez-Garca, I., Chem. Rev. 2008, 108, 3174–3198; (d) Yamamoto, Y., Chem. Rev. 2008, 108, 3199–3222. 6. (a) Sawamura, M.; Ito, Y., Chem. Rev. 1992, 92, 857–871; (b) Sawamura, M.; Ito, Y., in Catalytic Asymmetric Synthesis, 2nd ed., Ojima, I., ed., Wiley-VCH, Weinheim, 2000, pp. 493–512; (c) Dyker, G., in Organic Synthesis Highlights V.; Schmalz, H.-G.; Wirth, T., eds., Wiley-VCH, Weinheim, 2004, pp. 48–55; (d) Hashmi, A. S. K., Gold Bull. 2003, 36, 3–9; (e) Hashmi, A. S. K., Gold Bull. 2004, 37, 51–65; (f) Arcadi, A.; Di Giuseppe, S.,
REFERENCES
7. 8.
9. 10. 11. 12. 13. 14.
15. 16.
17. 18. 19.
20. 21. 22.
377
Curr. Org. Chem. 2004, 8, 795–812; (g) Hoffmann-R€ oder, A.; Krause, N., Org. Biomol. Chem. 2005, 3, 387–391; (h) Widenhoefer, R.; Han, X., Eur. J. Org. Chem., 2006, 4555–4563; (i) Hashmi, A. S. K.; Hutchings, G., Angew. Chem. 2006, 118, 8064–8105; Angew. Chem. Int. Ed. 2006, 45, 7896–7936; (j) Hashmi, A. S. K., Chem. Rev. 2007, 107, 3180–3211; (k) F€urstner, A.; Davies, P. W., Angew. Chem. 2007, 119, 3478–3519; Angew. Chem. Int. Ed. 2007, 46, 3410–3449; (l) Gorin, D. J.; Toste, F. D., Nature 2007, 446, 395–403; (m) Shen, C. H., Tetrahedron 2008, 64, 3885–3903; (n) Skouta, R.; Li, C.-J., Tetrahedron 2008, 64, 4917–4938; (o) Muzart, J., Tetrahedron 2008, 64, 5815–5849; (p) Li, Z.; Brouwer, C.; He, C., Chem. Rev. 2008, 108, 3239–3265; (q) Arcadi, A., Chem. Rev. 2008, 108, 3266–3325; (r) Jimenez-Nu´nez, E.; Echavarren, A. M., Chem. Rev. 2008, 108, 3326–3350; (s) Gorin, D. J.; Sherry, B. D.; Toste, F. D., Chem. Rev. 2008, 108, 3351–3378; (t) Hashmi, A. S. K.; Rudolph, M., Chem. Soc. Rev. 2008, 37, 1766–1775. Wei, C.; Li, Z.; Li, C.-J., Synlett 2004, 1472–1483. (a) Kabalka, G. W.; Wang, L.; Pagni, R. M., Synlett 2001, 676–678; (b) Sharifi, A.; Farhangian, H.; Mohsenzadeh, F.; Naimi-Jamal, M. R., Monatsh. Chem. 2002, 133, 199–204; (c) Wang, L.; Li, P.-H., Chin. J. Chem. 2003, 21, 710–713. Wei, C. M.; Li, Z. G.; Li, C. J., Org. Lett. 2003, 5, 4473–4475. Wei, C. M.; Li, C. J., J. Am. Chem. Soc. 2003, 125, 9584–9585. Li, Z.; Wei, C.; Chen, L.; Varma, R. S.; Li, C.-J., Tetrahedron Lett. 2004, 45, 2443–2446. (a) Wei, C. M.; Li, C. J., J. Am. Chem. Soc. 2002, 124, 5638–5639; (b) Wei, C. M.; Mague, J. T.; Li, C. J., Proc. Natl. Acad. Sci. USA 2004, 101, 5749–5754. Huang, B.; Yao, X.; Li, C.-J., Adv. Synth. Catal. 2006, 348, 1528–1532. (a) Mar Dıaz-Requejo, M.; Perez, P. J., J. Organomet. Chem. 2005, 690, 5441–5450; (b) Fructos, M. R.; de Fremont, P.; Scott, N. M.; Nolan, S. P.; Diaz-Requejo, M. M.; Perez, P. J., Angew. Chem. 2005, 117, 5418–5422; Angew., Chem., Int., Ed. 2005, 44, 5284–5288. (a) Posner, G. H., Org. React. 1972, 19, 1–113; (b) Marshall, J. A., Chem. Rev. 1989, 89, 1503–1511. (a) James, P. F.; O’Hair, R. A. J., Org. Lett. 2004, 6, 2761–2764; (b) Pale, P.; Chuche, J., Tetrahedron Lett. 1987, 28, 6447–6448; (c) Dalla, V.; Pale, P., Tetrahedron Lett. 1994, 35, 3525–3528; (d) Dalla, V.; Pale, P., New J. Chem. 1999, 23, 803–805; (e) Dillinger, S.; Bertus, P.; Pale, P., Org. Lett. 2001, 3, 1661–1664; (f) Paul, A. M.; Bent, B. E., J. Catal. 1994, 147, 264–271. Nakanishi, W.; Yamanaka, M.; Nakamura, E., J. Am. Chem. Soc. 2005, 127, 1446–1453. (a) Tai, H. C.; Krossing, I.; Seth, M.; Deubel, D. V., Organometallics 2004, 23, 2343–2349; (b) Pyykko, P., J. Am. Chem. Soc. 1995, 117, 2067–2070. (a) Asao, N., Synlett 2006, 1645–1656; (b) Patil, N. T.; Yamamoto, Y., Arkivoc 2007, 6–19; for related work, see (c) Dyker G.; Hildebrandt, D.; Liu, J.; Merz, K., Angew. Chem. 2003, 115, 4536–4538; Angew. Chem. Int. Ed. 2003, 42, 4399–4402. Zhu, J.; Germain, A. R.; Porco, J. A., Angew. Chem. 2004, 116, 1259–1263; Angew., Chem., Int., Ed.2004, 43, 1239–1243. Marshall, J. A.; Robinson, E. D., J. Org. Chem. 1990, 55, 3450–3451. (a) Hashmi, A. S. K., Angew. Chem. 1995, 107, 1749–1751; Angew. Chem. Int. Ed. 1995, 34, 1581–1583; (b) Hashmi, A. S. K.; Schwarz, L.; Choi, J.-H.; Frost, T. M., Angew. Chem. 2000, 112, 2382–2385; Angew., Chem., Int., Ed. 2000, 39, 2285–2288.
378
A CRITICAL COMPARISON: COPPER, SILVER, AND GOLD
23. Dudnik, A. S.; Sromek, A. W.; Rubina, M.; Kim, J. T.; Kel’in, A. V.; Gevorgyan, V., J. Am. Chem. Soc. 2008, 130, 1440–1452. 24. Morita, N.; Krause, N., Angew. Chem. 2006, 118, 1930–1933; Angew., Chem., Int., Ed.2006, 45, 1897–1899. 25. Hashmi, A. S. K.; Blanco, M. C.; Fischer, D.; Bats, J. W., Eur. J. Org. Chem. 2006, 1387–1389. 26. Suhre, M. H.; Reif, M.; Kirsch, S. F., Org. Lett. 2005, 7, 3925–3927. 27. Kirsch, S. F.; Binder, J. T.; Crone, B.; Duschek, A.; Haug, T. T.; Liebert, C.; Menz, H., Angew. Chem. 2007, 119, 2360–2363; Angew. Chem. Int. Ed. 2007, 46, 2310–2313. 28. Kirsch, S. F.; Binder, J. T.; Liebert, C.; Menz, H., Angew. Chem. 2006, 118, 6010–6013; Angew. Chem. Int. Ed. 2006, 45, 5878–5880. 29. (a) Kennedy-Smith, J. J.; Staben, S. T.; Toste, F. D., J. Am. Chem. Soc. 2004, 126, 4526–4527; (b) Staben, S. T.; Kennedy-Smith, J. J.; Toste, F. D., Angew. Chem. 2004, 116, 5464–5466; Angew. Chem. Int. Ed. 2004, 43, 5350–5352. 30. Deng, C.-L.; Zou, T.; Wang, Z.-Q.; Song, R.-J.; Li, J.-H., J. Org. Chem. 2009, 74, 412–414. 31. (a) Hashmi, A. S. K.; Weyrauch, J. P.; Frey, W.; Bats, J. W., Org. Lett. 2004, 6, 4391–4394; (b) Milton, M. D.; Inada, Y.; Nishibayashi, Y.; Uemura, S., Chem. Commun. 2004, 2712–2713; (c) Dorsch, D.; Burgdorf, L. T.; Gericke, R.; Beier, N.; Mederski, W., PCT International Patent Application WO 2005123688 A2, CAN 144:69837, 2005; (d) Hashmi, A. S. K.; Rudolph, M.; Schymura, S.; Visus, J.; Frey, W., Eur. J. Org. Chem. 2006, 4905–4909. 32. Harmata, M.; Huang, C., Synlett 2008, 1399–1401. 33. Harrison, T. J.; Kozak, J. A.; Corbella-Pane, M.; Dake, G. R., J. Org. Chem. 2006, 71, 4525–4529. 34. Binder, J. T.; Crone, B.; Haug, T. T.; Menz, H.; Kirsch, S. F., Org. Lett. 2008, 10, 1025–1028. 35. Lee, J. H.; Toste, F. D., Angew. Chem. 2007, 119, 930–932; Angew. Chem. Int. Ed. 2007, 46, 912–914. 36. Luo, T.; Schreiber, S. L., Angew. Chem. 2007, 119, 8398–88401; Angew. Chem. Int. Ed. 2007, 46, 8250–8253. 37. Shigemasa, Y.; Yasui, M.; Ohrai, S.-I.; Sasaki, M.; Sashiwa, H.; Saimoto, H., J. Org. Chem. 1991, 56, 910–912. 38. Buzas, A.; Istrate, F.; Gagosz, F., Org. Lett. 2006, 8, 1957–1959. 39. (a) Dalla V.; Pale P., Tetrahedron Lett. 1994, 35, 3525–3528; for related work with allenes, see Pale, P.; Chuche, J., Tetrahedron Lett. 1987, 28, 6447–6448. 40. Genin, E.; Tullec, P. Y.; Antoniotti, S.; Brancour, C.; Gen^et, J.-P.; Michelet, V., J. Am. Chem. Soc. 2006, 128, 3112–3113. 41. Liu, Z.; Wasmuth, A. S.; Nelson, S. G., J. Am. Chem. Soc. 2006, 128, 10352–10353. 42. Lin, G.-Y.; Li, C.-W.; Hung, S.-H.; Liu, R.-S., Org. Lett. 2008, 10, 5059–5062. 43. Hashmi, A. S. K.; B€uhrle, M.; Salathe, R.; Bats, J. W., Adv. Synth. Catal. 2008, 350, 2059–2064. 44. Godet, Th.; Vaxelaire, C.; Michel, C.; Milet, A.; Belmont, P., Chem. Eur. J. 2007, 13, 5632–5641.
REFERENCES
379
45. Zhang, Y.; Sun, W.; Freund, C.; Santos, A. M.; Herdtweck, E.; Mink, J.; K€ uhn, F. E., Inorg. Chim. Acta 2006, 359, 4723–4729. 46. Prior, A. M.; Robinson, R. S., Tetrahedron Lett. 2008, 49, 411–414. 47. Barange, D. K.; Nishad, T. C.; Swamy, N. K.; Bandameedi, V.; Kumar, D.; Sreekanth, B. R.; Vyas, K.; Pal, M., J. Org. Chem. 2007, 72, 8547–8550. 48. Deng, C.-L.; Song, R.-J.; Guo, S.-M.; Wang, Z.-Q.; Li, J.-H., Org. Lett. 2007, 9, 5111–5114. 49. Souma, Y.; Sano, H., Bull. Chem. Soc. Jpn. 1976, 49, 3296–3299.
INDEX
Acetone, 46, 48, 49, 52, 53, 55, 130, 134, 137, 170, 173, 188, 230, 289, 293, 332, 335, 336, 374 Acetonitrile, 48, 49, 53 Acetophenone, 12, 188, 204 Acetoxyallenes, 96, 97 1-Acetoxy-3-methylbut-1,2-diene, 95 p-Acidic, 160 Acridine derivatives, 160 Acridine synthesis, 159 Acrylonitrile, 45 Activated heteroaromatics, nucleophilic substitution of, 24 Activation parameters, 197 Acyl halides/epoxides, 311 1,2-Acyl migration, 111 Acyloxy alkynones, 111 Acyloxyallene, 146 Adamantyl (Ad) halides, 7 Adamantyl bromide/iodide, 290 Addition–rearrangement reactions, 234 Addition–rearrangement sequence, 238 [Ag(acac)], 46 Ag–Ag van der Waals contact distance, 330
AgBF4, 71, 73, 88, 90–92, 97, 101–102, 104–105, 108, 109, 112, 126–127, 130–131, 135–136, 138, 145, 146–148, 152–153, 156–160, 192, 249, 250, 280, 297, 305, 336, 339, 350, 364–368, 371 AgBF4(Ph3P), 75 AgBr, 3, 19, 121, 132, 148, 292, 297, 359, 360 AgBr2LiBr, 5, 16 Ag-catalyzed di-tert-butylsilylene, 215 aldimines and ketimines, scope of, 215 Ag(CB11H12), 75 Ag(CB11H6Br6), 75 AgCl, 15, 20, 23, 35–36, 76, 137, 169, 294, 297, 318–320, 359–360, 365, 375 AgClO4(Ph3P), 75 AgClO4, 44–45, 74–75, 77, 87, 90–94, 96, 102–103, 106, 123, 128–130, 248–250, 298, 337 AgClO4–BINAP, 281 AgCN, 148, 297 Ag2CO3, 29, 34–35, 75, 86, 98, 297–298, 300–301, 303, 309–310, 321–322, 345
Silver in Organic Chemistry Edited by Michael Harmata Copyright 2010 John Wiley & Sons, Inc.
381
382 AgF, 8, 9, 13, 17, 50, 159, 264, 266–268, 273, 274, 293, 307, 317 AgF2, 274 AgF-Difluorphos, 266 Ag(fod), 46 AgI, 29, 30, 35, 313–314, 317–318, 322, 359–360 Ag–N coordination, 332, 343 109 Ag NMR spectra, 20, 29, 317 AgNO3, 2, 4, 11, 14, 19, 20, 29, 32, 33, 87, 93, 95, 100, 125, 145–146, 148–149, 151–153, 159, 169, 183, 260, 262, 287, 293, 299, 300, 305–307, 313, 337, 339, 343, 345, 363, 375 [Ag(NO3)3]2– anion, 333 AgNO3-catalyzed 5-endo-dig cyclization, 148 AgNO3–Tol–BINAP, 262 AgNTf2, 45, 156, 315 AgO, 159 Ag2O, 18, 30, 49, 68, 85, 150–151, 159, 287–290, 293, 303–306, 308, 320–321, 368, 374–375 AgOAc, 23, 47–49, 51–57, 61–64, 67–71, 77–78, 96, 119, 122–123, 126–127, 132, 148, 150–151, 159, 276–277, 295–297, 299, 300, 306, 308–309, 317, 375 AgOCOCF3, 33, 119, 124–125, 260 AgO2CPh, 183 AgO3SCF3, 345 AgOTf, 20, 22, 24, 25, 29, 34, 45, 58, 60, 63–65, 72, 74–78, 85, 98, 110–111, 113, 147–151, 153–159, 161, 192–194, 196–199, 208, 213–215, 217, 242, 248–250, 252–254, 260–271, 273–275, 278–279, 288–289, 291–292, 296–297, 305, 311, 313, 316, 333, 339, 364, 366, 368–370, 372–373 AgOTf/BINAP, 261 AgOTf–BITIANP complex, 262 AgOTf(Ph3P), 75 AgPF6, 156, 159, 271–272, 332, 364 AgSbF6, 77, 98, 101, 106, 109–110, 152, 154, 156, 159–161, 192, 232, 234, 236, 242, 315, 350, 366–368, 370–371, 374 Ag2SO4, 159
INDEX
AgTp complex, diazoacetates catalyzed, 244–245 AgTp–diazomalonate, 233 Ag–TpMes complex, 178 Aldehyde–alkyne–amine coupling, 358 Aldehydes, 313 alkynylation catalyzed by Ag(I) in water, 312 terminal alkynes addition, 312 Aldimines, asymmetric Mannich reaction of, 276 Aldol adducts, 270 syn-Aldol adducts, 273 Aldol reaction, 144, 282 of diketene, 271 using allylsilanes, 271 using catalytic amounts, of Me3SnOMe, 270 using tin enolates, 268 using trimethoxysilylenol ethers, 273 with azodicarboxylate, 281 with tin enolates, 269 Aldol reaction, diastereoselectivity, 274 2,3-Alkadienenitriles, 5 N-Alkenyl alkynylamides, 158 Alkenyl organometallics, 287 (E)-Alkenyl pentafluorosilicates, 293 Alkenylsilver complex, 254 Alkenylsilver compounds, 13, 14 Alkenyl trichloroacetate, 271 Alkenyl trichloroacetate equivalent, 270 Alkoxy bromocyclononenes, 93 formation of, 94 Alkoxymethylenecyclopropanes, [2 þ 2] cycloaddition of, 46 Alkoxystyrene derivatives, 90 1,2-Alkyl/aryl shift, 146 Alkylboranes, 287 Alkylcopper reagents, 5 Alkylidenecyclopentanes, from 1,3-dicarbonyl compounds, 368 Alkylidene cycloproparenes, 90 Alkylidene oxazoles, 368 Alkylmagesium halides, silver-catalyzed dimerization, 3 Alkyl migration, 110 [1,2]/[1,5]-Alkyl shift, 111 Alkylsilver compounds, 2–7 Alkylsilver intermediate, 111
INDEX
Alkyne activation, 160 Alkynes, 18–20, 22–23, 26–27, 29–33, 35–36, 45, 71, 96, 99, 110, 144, 148, 150–151, 153–158, 160–161, 189, 190, 201, 204–206, 253–254, 290, 293, 311–313, 315, 317–321, 323–324, 359, 372–375 thiourea additions, 161 p-Alkyne–silver complex, 20, 22 g-Alkynic b-ketoesters, 375 Alkynones, asymmetric reduction of, 104 O-Alkynylacylimines, 154 g-/d-Alkynylaldehydes, 315 w-Alkynylaldehydes, 315 b,g-Alkynyl a-amino acid derivatives, 22 O-Alkynylaryl aldimines, 155 2-Alkynylbenzaldehyde, 151 O-Alkynylbenzaldehydes, 149 cyclization of, 362 2-Alkynylbenzaldoximes, 156 2-Alkynylbenzenamines, 161 b-Alkynyl ketones, 157 Alkynylpalladium complex, 30 Alkynyl silvers, 290 Alkynylsilver species, 314 Alkynyl zirconium species, 314 Allenes, 5 hydroarylation of, 373 Allenic b-ketoesters, 101 Allenic oximes, 152 Allenic sulfones, 108 silver-catalyzed preparation of, 108 Allenones, 104, 110 Allenyl alcohols, 144 Allenyl aldehydes, 144 Allenyl ketones, 144, 146 cycloisomerization of, 362 dimerization of, 363 to furans, 362 Allenyllithium reagents, 15 Allenylsilver compounds, 15 Allenylsilver reagents, 5 Allenyl sulfones, 302, 303 Allenylthiocarbinols cyclization, to dihydrothiophenes, 364 transition-metal-catalyzed cyclization of, 365 Allenyl thioethers, 241 Allylation, 144
383 of aldehydes with allyltrimethoxysilane, 266 of ketones, 267 reaction between ketones and allylsilanes, 266 reaction of aldehydes, 260 reaction using allylsilanes, 264 reaction using allyltributyltin, 260 reaction with various aldehydes, 261 of a,b-unsaturated ketones, 268 using BITIANP ligand, 264 using diimine ligand, 265 using thiophosphoramide ligand, 265 of various aldehydes, 267 Allyl bromides, 9, 288 Allyl cation, trapping by indole nucleophiles, 131 Allylic acetate, 122 Allylic alcohols, 299 Allylic amines, 204, 221 Allylic borane, 259 Allylic cation, intermolecular trapping of, 122 Allylic disilane, 219, 220 Allylic ethers, 219, 221 silver-catalyzed silylene transfer, 220 Allylic organometallic reagents, 259 Allylic silane, 220 Allylic titanium, 259 Allylic transposition, 221 (h3-Allyl)palladium complexes, 293 2-Allylphenol, 106 Allylsilane, 219 Allylsilver species, 262 Allylstannane, 253 Allyltributyltin, 262 Allyltrimethoxysilanes, 264, 266 (S)-Alpine Borane , 104 Alzheimer’s disease, 302 Amaryllidaceae alkaloids, 120, 301 Amidation, 173 Aminating reagents, 281 b-Amino acids, 84, 276 1,2-Aminoalcohol, 204 1-Aminobenzimidazolium salts, 25 b-Aminoesters, 295 b-Aminoketone, 276, 296 6-Amino-1-naphthalenesulfonic acid, 345 3-Amino-1,2,4-triazole, 347
384 Ammonia, 19 silver oxide, 84 to styrene, 172 Anellated pyrroles, 370 by hydroarylation of allenes, 373 Anion metathesis, 143 O-Anisidine, 295 Anomalous diffraction, 359 Anthonomus grandis grandis, 97 anti Markovnikov selectivity, 188 anti,syn-selective aldol reaction, 195 Antitumor agent FR901464, 311 Antitumoral dienediynes, 317 Areneazoethynes, 30 Arene trapping, 136 Argentocarbenium ion, 91 Arenediazonium chlorides, 30 Argento carbonium ion, 89 Argento cation, 89 Argento cationic intermediates, 88, 89 Argentocyclopropyl cation, 90 Argentodiazoesters, 17 Arndt–Eistert homologation, 84, 85, 230 Arndt–Eistert rearrangement, 84, 86 Aromatic aldehydes, 266 Aromatic electrophilic substitution, 309 Aromatic oxazoles, 368 ArSCl, 153 Arylaldimines, pka of, 51 a-Aryl alkanoic acids, 108 Arylargentates, 293 2-Arylbenzoxazoles, 308 Aryl boronic acids, palladium catalyzed coupling of, 30 Aryl bromides, 320 Arylcopper compounds, 11 Aryl iminoesters catalytic asymmetric cycloaddition, 65 intramolecular catalytic asymmetric cyclization, 63 2-Arylindoles, 309 Aryl iodides, 35, 314 (h1-Aryl)palladium bromide, 298 Aryl palladium species, 309 N-Arylsilaaziridines, 217 N-Arylsilaziridines, dearomatization reactions of, 218 Arylsilver compounds, 10 decomposition temperatures, 11
INDEX
Arylsilvers, 292 trans-4-Aryl-/trans-4-alkenyl-3-bromo-4,6dimethyl-3,4-dihydro-2-pyrones, 109 ()-Ascofuranone, 103 A1,3 steric interactions, 119 Asymmetric aldol reaction, 271, 273 Asymmetric crotylation reactions, 264 Asymmetric [3 þ 2] cycloadditions, 50 Asymmetric Mannich reactions, of enol ethers with alkynyl imines, 295 Asymmetric vinylogous Mannich reaction, 296 titanium silylenoid intermediate [Cp2Ti¼SiHR], 184 AuBr3, 46, 359, 363–365 AuCl, 161, 359, 360, 365, 369, 372 AuCl3, 146, 151, 363–368, 370, 372–374 AuClPPh3, 106, 158–159, 161, 186, 363–368, 370–371, 373–374 AuI, 156, 359, 360, 364–365 Au(IMes)Otf, 156 AuPPh3BF4, 158 Au(PPh3)OTf, 157 AuPPh3SbF6, 159, 161 Axial chirality, 303 7-Azabenzonorbornadiene, syn invertomer of, 87 1-Azabicyclo[3.2.1]heptane, 86 2-Azabicyclo[2.2.1]heptane derivatives, 86 1-Azabicyclo[3.2.0]hept-3-ene derivative, 87 2-Azabicyclo[2.1.1]hexanes, 88 2-Azabicyclo[3.3.1]non-7-enes, 128 2-Azabicyclo[2.2.2]octane derivatives, 86 6-Azabicyclo[3.2.1]oct-3-enes, 128 Aza-Diels–Alder cycloaddition, 75 Aza-Diels–Alder reaction, 77 Azaenolate, 53 Azasilacyclopentadienes, 205 allylic amines, formation of, 205 Azasilacyclopentenes, 215, 217 Azasilalactone, 212 Aziridination, 232 Aziridine, 171 Aziridinium intermediate, 88 Aziridinium ion, 94 Azodicarboxylates, 282 Azomethine cycloaddition, 49, 62 Azomethine [3 þ 2] cycloadditions, 66
INDEX
Azomethine ylides, 50, 58, 218 alkenes, cycloadditions of, 46 [3 þ 2] cycloadditions of, 47 stereoselective cycloadditions, 51 BARF, 360 Bathophenanthroline–silver complex, 175 Benzaldehyde, 47, 218 crotylation of, 266 dearomatization reactions of, 218 Benzene, 10 in CH2Cl2, 234 Benzocycloctatriene, 44 Benzocyclopropene1,3-butadienenaphtho [b]-cyclopropene, 71 Benzohydrindenone, 137, 138 Benzoic acid derivatives, C–H activation of, 308 Benzopyrylium salts, 149 2-Benzopyrylium salts formation, 362 catalysts for, 363 Benzopyrylium synthesis, 149 Benzoquinoline, 309 Benzoquinone, 289, 290 Benzoyl chloride, 24 2-Benzyloxyheptanal, 360 Benzyne, 44 BF4, 333 Biaryls, 310 Bicyclic pyrrolidines, 169 Bicyclo[6.4.0]dodecanone derivative, 125 BINAP, 58, 60, 77, 78, 264, 266, 272, 278, 279, 301 BINAP–AgF, 264, 274 catalyst system, 274 BINAP–Ag(I) complex, 276 BINAP–AgOTf, 273, 274, 278 BINAP–AgOTf catalyst, 268 BINAP–AgPF6, 271 1,10 -Binaphthyl-2,20 -diamine, 262 BINOL, 280 Biocides, 143 Biphenyl, 10, 44 synthesis, 306 2,20 -Bipyridine, 331 Bipyridines, 292 Bis(alkylmagnesium halides), silvermediated ring-closing reaction, 4, 5
385 (S)-2,20 -Bis(diphenylphosphino)-1,10 binaphthyl (BINAP), 260 Bishomocubanes, 93 Bis(isoxazoline) ligands, 243 Bis(isoxazoline)-type ligands, 242 Bis(2-methylimidazol-1-yl)methane, 339 2,5-Bis(pyrazine)-1,3,4-oxadiazole, 350 1,3-Bis(2-pyridylaminomethyl) benzene, 345 N,N0 -Bis(3-pyridylmethyl)-1, 6-hexanediamine, 343 1,4-Bis(silylated butadiyne), 320 Bis(1,2,4-triazol-1-yl)acetic acid, 348 Bis(trimethylsilyl)enediynes, 321 [Bmim][BF4], 359 s-Bond migration, 113 Boranes, 4 Boronic acids, 288 Boxes, 330 Bridged annulenes, 89 Bromination, 17 Bromine, 15, 31 Bromoallenyl ketone cyclization, 364 Bromoallyl cation, 93 2-Bromoallyl cation, 126 a-Bromoaryl ketones, 108 4-Bromobenzaldehyde, 262 Bromoform, 94 2/4-Bromopyridine derivatives, 292 N-Bromosuccinimide, 31 Bromovinyl tetrahydrofurans, 94 Brønsted acids, 131 catalysis, 148 1,3-Brook rearrangement, 206 4-tBu-pyridine, 174 B€uchner reaction, 233 BuLi, 358 Bullatenone, 103 t-BuMe2SiOTf, 198 Bu4NCl, 35 t-Bu2SiCl2, 191 Bu3SnOTf, 254 1,3-Butadienes, 71, 88, 89 n-Butane, 5 (2R,3R)-Butane-2,3-diol, 131 n-Butanol, 122 Butatrienes, 6 cis-But-2-ene, 15 cis-2-Butene, 192, 195
386 trans-2-Butene, 192 But-3-ene-2-one, 74 Butenolide, 124 formation from gem-dihalocyclopropanecarboxylic acids, 125 g-Butenolides, 276, 296 2-Butenylstannanes, 261 t Bu3tpy, 175 silver complex, 174 Butyl acrylate, 61, 65 tert-Butyl allenylsilver, 15 tert-Butylcarbonyloxybutynols, 104 tert-Butyl diazoacetate, 107 tert-Butylisocyanide, 217 Butylsilver, 5 Butynediols, 371 1,4-Butynediols, 102 (1-Butynyl)silver, 19 Butyraldehyde, 190 Cage complex, 332 Cage-like complexes, 331, 335 Calixarenes, 343 Callyspongia truncata, 299 ()-Callystatin A, 299 Campholenaldehyde, 95 Capsules, 330 Carbamates, 174 Carbamate-protected imines silaaziridines, decomposition of, 216 Carbenes, 207 transfer reactions, 107 p bonds, 232 Carbenoid, 111, 146, 207, 229–255 Carbocation intermediate, 93, 160 Carbometallations, 298, 299 Carbon-based nanotubes, 337 Carbon dioxide, 15, 313 Carbon disulfide, 15 Carbon monoxide, 375 Carbonolide B, 85 Carbon–silver bond, radically mediated cleavage, 2 Carbonyl–alkyne reaction, 374 Carbonylations, 375 Carbonyl compounds, insertion of, 190 Carbonyl group, allylation reaction of, 259 Carborate anion, 27 Carbostannylation, 315
INDEX
Cascade electrocyclization/[3,3]sigmatropic rearrangement, 210 Catalyst screening, 260 Catalytic asymmetric [3 þ 2] cycloadditions, 58 Catalytic asymmetric version of the coppercatalyzed reaction, 359 Catalytic version of nucleophilic addition, 312 Catenanes, 339 Cationic aza-Cope rearrangement, 99 Cationic cyclizations, 129 Cationic gold(I) complexes, 157 Cationic intermediates, 117, 120 nucleophilic trapping of, 121–132 Cation–olefin cyclizations, 129 CC, C¼C ene-yne cycloisomerization, 157–160 CC, C¼N, cycloisomerization, 153–157 CC, C¼O, cycloisomerization, 148–152 C¼C¼C, C¼N, cycloisomerization, 152–153 C¼C¼C, C¼O, cycloisomerization, 144–148 Cesium/potassium fluoride, 8 C–H activation, 306 Chair-like transition state, 209, 212 C-Hal insertion, 241, 254 C-Hal rearrangement reactions, 241 C–H bond activation, 144 CH2Cl2, 101, 103, 105–106, 109–110, 112, 123, 134–136, 147, 149, 152–154, 156–158, 160, 169, 198, 207, 233–238, 240, 242, 248–250, 252, 291, 294, 298, 311, 315–316, 321, 364, 372 Chemoselectivity, 201, 217, 234 Chemoswitch, 23 C–H insertions, 234, 236, 242, 244–245 into acyclic alkanes catalyzed by AgTp complexes, 246 Chiral auxiliary, 131 Chiral carbamates, 128 Chiral catalysts, 259 Chirality transfer, 104, 108 Chiral phosphite, 280 Chiral silver catalyst, 260 Chiral silver–diphosphine catalysts, 260 N-Chloramines, 86 Chloramine-T, 169–170
INDEX
Chloride abstraction, 371 Chlorine, 31 a-Chloroacrylic esters, 122 Chloroalkyne, 32 N-Chloro-2-azabicyclo[2.2.1]heptane, 87 a-Chlorocyclopentenone, 134, 135 1-Chloro-2,2-diphenylethene, 15 1-Chloro-2,2-diphenylvinylsilver, 15 2-Chlorofuro[2,3-b]pyrazines, 149 N-Chloroisoquinuclidines, 86 3-Chloro-3-methylbut-1-yne, 95 (S)-2-Chloropropiophenone dimethyl acetal, 108 2-Chloro-3-methyl hex-4-enoates, 107 N-chlorosuccinimide, 32, 123 N-chlorosulfonyl isocyanate, 199 Chromium(II) chloride, 288 Chromous chloride, 293 Cinchona-derived ligands, 65 14 C-labeled amino acid, 84 Claisen rearrangement, 106 ClO4 anion, 333 Cobalt chloride–ephedrine complex, 58 C–O bonds, silver-mediated di-tertbutylsilylene insertion, 219 Codeine, 302 (cod)PtCl2, 189 Coesite (coe) topology, 350 Coinage metal catalysis, 369 Collision-activated dissociation, 230 p Complex, 96 p Complexation, 317 Computed heats of formation (B3LYP/SDD, kcal/mol), of various substrates with AgCl, 294 Cones, 330 Conia-ene reaction, 368, 375 Conrotatory ring closure, 132 Containers, 331 Coordination polymers, 331 Copper acetate, 309 Copper catalysts, 243 Copper-catalyzed reaction of aldehydes with alkynes and secondary amines, 359 ring expansion, 202 Copper complexes, 243 Copper-mediated alkyne formation, 206 Copper-mediated ring expansion, 207
387 Copper/silver salts, 106 Copper triflate, 296 Copper(I) halides, 360 Copper(I) iodide, 28 Copper(II) bromide, 202 Copper(II) chloride, 322 Copper(II) triflate, 210 Crinine, 100 type alkaloids maritinamine, 126 Cross-conjugated dienone, 132 Crossover experiments, 206, 220 Crotonaldehyde, 190 Crotylation, with benzaldehyde, 262 Crotyl chloride, 107 Crotylsilane, 264 Crotyl tiglate, 209 Crotyltrimethoxysilane, 266 18-crown-6, 266, 267, 273, 274 Crown ethers, 343 [2,2,2]-Cryptand, 335 Crystal structure, 332 C–S addition–rearrangement reactions, 241 C–Si bond, oxidation of, 209 Cubanes, 92 rearrangements of, 92 CuCl, 49, 156, 360–361, 363–365 CuCl2, 323, 375 CuCN, 49 Cucurbiturils, 343 CuI, 29, 149, 194, 196, 202–207, 358–360, 364–368 Cupric chloride, 30, 322 CuTp(CF3)2, 243 1-Cyanoisoquinolines, 156 N,O-Cyclic acetal, 217 Cyclic nitrones, 152, 153 Cyclic oxonium intermediate, 111 Cyclic silanes, 191 Cyclic transition state, 270, 273 Cyclization cascades, 102 Cycloadditions, 144, 315 [2 þ 1] cycloaddition, 207 [2 þ 2] cycloadditions, 44, 45 [3 þ 2] cycloadditions, 46 [3 þ 3] cycloaddition, 156 [4 þ 2] cycloaddition, 74, 202 Cyclodextrins, 343 1,3-Cycloheptadiene, 89 trans-Cycloheptene, 74
388 Cyclohepta-1,3,5-trienes, 89 Cyclohexene, 197, 239 Cyclohexene silacyclopropane, 192, 197, 198, 201, 221 Cyclohexenone, 45 1,2-Cyclohexyldiamine, 262 Cycloisomerization, 104, 366 Cycloisomerization/dimerization reactions, 146 Cycloisomerization reaction, to produce furan rings, 144, 145 E-Cyclooctene, 93 trans-Cyclooctene, 74 Cyclopentadienes, 74, 370 2-Cyclopentenone, formation of, 161 3-Cyclopentenone, 161 Cyclophanes, 343 Cyclopropanation, 138, 233 with styrene, 233 Cycloproparenes, 90 Cyclopropenes, 90 derivatives, silver(I)-promoted rearrangement of, 91 1,1-Cyclopropylaminocarboxylates, 309 Cyclopropylboronic acids, 289 Pd-catalyzed coupling, 288 Cyclopropyl carbenes, 250 Cyclopropylcarbinyl cation, 89 Cyclotriveratrylene, 336 Cylinder, 331, 332 Cy3PAgCl, 23–24, 312 (Cy2P)2Pt¼SiMes2, 198 Dake’s pyrrole synthesis, 369 Danishefsky’s diene, 74, 76, 77 DBU, 32, 51, 52, 70, 156, 162 DDQ oxidation, 103 Dearomatization, mechanisms, 218 Decamethylsilicocene, 186 Decarboxylative coupling, 310 Deethynylation, 32 Dehydrochlorination, 138 Dehydrolinalol, 96 Dehydropyrroles, 66 (S)-1-Dehydroxyvirol A, 36 Desilylation, 32, 33 Deuterium labeling, 45 Dewar benzenes, 91 DFT study, 252
INDEX
4.142 (dia-f) topology, 347 cis-1,2-Dialkenylcyclopropanols, 125 Dialkylcyclopropene carboxylates, 90 4,40 -Diaminodiphenylmethane, 348 Diamond topology, 347 Diaryliodonium triflates, 310 Diarylpyridinylmethane, 309 Diaryltetrayne, 113 Diastereoselectivity, 47–48, 51, 53, 55–56, 58, 65, 67, 70, 127–128, 131, 193, 195, 202, 206, 208–212, 217, 232, 234, 239, 242, 259, 268, 270, 273–274, 281, 360, 373 Diazoalkanes, 18, 26 a-Diazocarbonyl compounds, Wolff rearrangement of, 183 Diazoesters, Ag-catalyzed C–H insertion of, 253 a-Diazoketones, 84, 85 a-Diazosilyl enol ether, Mukaiyama aldol reaction, 274 a-Diazotrimethoxysilyl enol ether, aldol reaction, 275 Dibromoaziridinium ion, 94 Dibromocarbene, 122 Dibromocyclopropanation reaction, 133 Dibromocyclopropanes, 93, 94 1,2-Dibromoethane, 3 as reoxidant, 287 Dichlorocyclopropanation, 133 Dichloromethane, 32–33, 53, 108, 110–111, 124, 134–135, 339, 363, 365 3,5-Dichloropyrazine-2(1H)-ones, 148 Dichlorozirconocene, 311, 314 Diels–Alder reaction, 53, 74, 97, 158, 314 cis-2,4-Dien-1-al, 161 Dienylstannane, 253 Diethylacetylene dicarboxylate, 212 Diethylenetriamine, 331 Diethyl fumarate, 232 Diethyl phosphite, 156 HPO(OEt)2, 151 Difluorphos, 77–78, 266–268 Difuroperylene, 113 1,1-Dihalo-2-(silyloxy)-2vinylcyclopropanes, 133 synthesis, 134 Dihydrobenzofuran, 106 synthesis of, 269
INDEX
2,3-Dihydrobenzofurans, 267 2,3-Dihydrobenzoxasilepine, 267 Dihydroborated internal acetylenes, 5 Dihydrofurans, from propargyl esters, 371 2,5-Dihydrofurans, 102 enolacetate preparation, 103 stereoselective synthesis of, 372 1,5-dihydro-2H-pyrrol-2-ones, 148 Dihydroisoquinoline, 154, 156 synthesis, 155 Dihydroisoquinoline core, access, 154 1,2-Dihydroisoquinolines, 155, 156 2,5-Dihydropyrans, 104 Diimine ligand, 262 Diisopropylethylamine, 20, 29 Diisopropyl fumarate, 65 b-Diketones, 296 Dimerisation, 17 Dimesitylsilane, 185 Dimethoxybenzene, 27 Dimethyl acetamide, 303 Dimethylacetylenedicarboxylate (DMAD), 154 Dimethyl argentate, 3 Dimethyl carbonate, 187 1,3-Dimethylcyclobutadiene, 45 Dimethylcyclopropane-1,1-dicarboxylate, 156 Dimethyl diazomalonate, 230 Dimethyldichlorosilane, 186 N,N-Dimethylformamide (DMF), 20, 25–26, 28–30, 35–36, 85, 122, 149, 191, 272, 299, 300, 302–303, 306, 308–310, 314, 317–320, 322, 337–338, 367 Dimethyl fumarate, 56, 66 Dimethyl maleate, 56, 58, 60, 61, 62, 66 cis-Dimethylsilacyclopropane, 190, 192, 195 trans-Dimethylsilacyclopropane, 192 Dinitrobenzene, 27 m,m0 -Dinitrobiphenyl, 306 Dinuclear copper(II) acetylide complex, 323 DIOP, 59–60 Dioxosilacyclopentane, 212 Diphenylacetylene, 15 Diphenylallenyl ketones, 363 1,4-Diphenylbutadiene, 14 Diphenyl-/diethyl-/dimethylsilane, 185 4,7-Diphenyl-1,10-phenanthroline, 174 Diphenylsilane, 188
389 Diphenylthiophosphoramide ligand, 262 1,3-Dipolar adducts, 153 Dipolar cycloaddition, 26, 48, 154, 213–214 Dipolarophile, 212 1,3-Dipolarophiles, 152 2,20 -Dipyridine, 305 2,20 -Dipyridyl, 14 Dirhodium carboxylate catalyst, 168 Disilanylenol ether, 280 Disilver diazomethane, 18 Disilver nitrene transfer catalyst structure, 170 Disrotatory ring opening, 137 trans-2,3-Disubstituted 2,3dihydrobenzofuran, 267 1,2-Disubstituted naphthalenes, 110 trans-2,6-Disubstituted piperidines, 153 1,2-/1,3-Disubstituted oxasilacyclopentanes, formation of, 194 2,6-Di-tert-butylhydroxytoluene (BHT), 178 2,6-Di-tertbutyl-4-methylpyridine (DTBMP), 154 Di-tert-butylsilacyclopropanes, 190 Di-tert-butylsilylene, 190, 192, 198 to olefins, silver-mediated transfer, 192 Di-tert-butylsilylene transfer, 192, 193, 195, 204, 214 g-Dithiolactones, 15 Diynes, 30 d3-Methyl iodide, 27 DMF, see Dimethyl formamide DMSO, 32, 52, 205, 307, 309–310 Domino reactions, 118 Double-stranded helix, 348 Duphos, 58–59, 260 Dye-sensitized organic solar cells, 307 Dynamic equilibrium, 334 EDA, 243 Electrocyclic processes, 117 Electrocyclic ring opening, 44–45, 118, 120–121, 123–124, 126–127, 130–131, 134, 139 Electrocyclization, 132, 209 4p electrocyclization, 133–134, 137–138, 218 6p-electrocyclization, 105, 208, 210, 215 Ireland–Claisen rearrangement cascade reaction, 209 Electrocylic ring opening, 139
390 2p-Electron disrotatory ring opening, 133 2p-Electron ring opening, 118 Electrophilic aromatic substitution, 138, 307, 308 Electrospray ionization–time-of-flight (ESI-TOF) mass spectroscopy, 333 Electrospray mass spectrometry, 336 Elemental sulfur, 9 b Elimination, 206 syn Elimination, 53 Enamine catalyses, 369 Enamines, 316 Enaminoester, 106 Enantioselective Sakurai–Hosomi reaction, 264 Enantioselectivity, 280 5-Endo-dig cyclization, 148 6-Endo-dig cyclization, 374 6-Endo-dig product, 150 6-Endo-trig cyclization, 366 Enesulfonamides cyclization, 157, 158 Enol ethers, 157, 273 Enol silane, 295 Enol triflates, 28 Enoxysilanes, 207 Ent-erythramine, 128 Enthalpic control, 198 Entropic control, 198 Enynenitriles, trans addition, 5 Enynes, 35 Enynylphosphonates, 320 Enynylsulfinates, 6 ( þ )-Epi-acetomycin, 210, 221 Epibatidine, 50 3-Epi-erythramine, 128 Epi-maritinamine, 126 10 -Epi-stegobinone, 195, 221 retrosynthesis of, 195 Epoxide–alkyne reaction, 373 Epoxide opening, with silver acetylides, 291 Epoxide-to-aldehyde rearrangement, 291 Epoxyalkynes, 28 Erythrina alkaloids, 126 Ethane, 5 Ethanol, 14 Ethene, 188 Ethylcyclohexyl acetate, 360 Ethyl diazoacetate, 232 Ethyldiazophenyl acetate, 241
INDEX
Ethyl 2,3-dichloropropanoate, 237 Ethyldiisopropylamine, 312 Ethylene, 5 3,30 -Ethylenebicyclopropenyls, 91 Ethylenediamine, 331 Ethylene–dichloro-palladium, 306 Ethylglyoxylate, 312 Ethyl iodide, 27 Ethylmagnesium bromide, 358 Ethyl-tert-butyldiazoacetate, 239 17a-Ethynylandrost-5-ene-3b, 17b-diol, 96 4-trans-Ethynyl-cis-2,6,6-trimethylbicyclo [3.1.1]heptan-4-ol, 31 Et3N, 34, 52, 54–55, 65, 85, 149 Et3PAuCl, 363–364 (Et3P)2Pt(H)–Si(St-Bu)2(OTf), 197 Et3SiH, 184 Exo-dig cyclization, 296 5-Exo-dig cyclization, 106, 368, 374 5-Exo-dig product, 150 6-Exo-iodo/bromo-2-azabicyclo[2.2.0] hexanes, 88 Feringa’s phosphoramidite ligand, 65 Ferrocenyl group, 160 Fesulphos, 59, 65 Fischerindole, 130 Fischer-type metal carbene complexes, 186 3-Fluoroacrylate, 67 Fmoc-protected amino acids, 84 19 F NMR, 333 F orbitals, 84 Formalin, 274 ( þ )-Fostriecin, 304 FR901464, 24 Fragmentation, 32 Free radicals, 287 Free silylene transfer, product-determining step of, 198 Friedel–Crafts reaction, 131, 132 Fuming HNO3, 10 Functionalized furans, access to, 147 Furan heterocycles, by coinage metal catalysis, 366 Furanones, by domino heterocyclization/1,2 shift, 366 2(5H)-Furanones, 289 Furanuronic acid, 85 2-(5H)-Furanylidenamines, 148
INDEX
Furo[2,3-b]pyrazine synthesis, 149 Furodihydrofuranones, 151 Furo[2,3-d]pyrimidines, 148 access, 149 Furopyrrolidinones, 151, 152 synthesis, 151 Furoquinoline, 150 ()-Galanthamine, 301, 302 GC analysis, 239 Geiparvarin, 103 Gelsemine, 299 Gem-dibromocyclopropane, 93–94, 120–124, 129–131, 133 silver(I)-mediated ring opening of, 124 Gem-dibromospiropentanes, 131 Gem-dichlorocyclopropane, 122, 125, 126, 128, 133, 134–136, 138 Gem-dihalocyclopropanes, 133 Gem-dimethylsilacyclopropane, 212 Geneˆt’s gold-catalyzed cyclization, 372 Glaser-type reaction, 317 Glycine, 47 Gold(III) catalysis, 145 Gold(III) chloride, 365 Gibberellic acid, access to central core of, 160 Green chemistry, 143 Green reagents, 311 Grignard reagents, 3, 286, 287 Halide scavenger, 292 a-Haloacetates, silver-catalyzed preparation of, 107 Haloalkanes, 31 carbene insertion, 238 Haloallyl cations, 125, 129–133 Halocyclopropane, 118, 120, 139 electrocyclic ring opening of, 123 ring-opening reactions of, 118 silver(I)-mediated electrocyclic ring opening, 131 silver(I)-mediated ring opening, 128, 129 silver(I)-promoted ring opening, 138 substituent effects on the ring opening of, 120 a-Halogenated aryl ketones, 108 Halogenoamines, 86 a-Haloimines, silver-induced reactions of, 86 a-Haloketones, 86
391 Halonium ylide, 239, 241 N-(2-Halophenyl)enamides Heck reaction, cyclization of, 299 Halophilicity, 297 a-Halo-g-unsaturated esters, 107 Hammett correlation, 198 Hammett equation, 198 Hapalindole, 130 Heck coupling, 309 Heck reaction, 299–302 b-H elimination, 89 Heptafluoro-2-nitrosopropane, 9 Herbicide manufacture, 292 Heteroaryl propiolates, 319 Hetero-Diels–Alder reactions, 77 Heterogeneous silver catalysis, 172 1,4,5,8,9,12-Hexaazatriphenylene, 350 Hexafluoroantimonate, 101, 108 Hexafluoroisopropanol (HFP), 124 Hexafluorophosphate anion inclusion, 332 Hexahydroazepines, 153 Hexamethylsilirane, 186 1-Hexene, 188 n-Hexene silacyclopropane, 192 n-Hexylsilane, 188 1-Hexyne, 20, 29 Histrionicotoxin, 126 Hiyama couplings, 304 1 H NMR, 333 H2O2, 204 Homoallylic ethers, 199 Homoallylic haloacetates, 239 Homoallylic thioethers, 241 Homocoupling, 286, 287, 292, 293, 317, 322 Homocubane isomerization, to norsnoutane, 92 Homocubanes, 92 Homogeneous catalysts, 229 1H-1,2,4-triazole-1-acetic acid, 345 H€unig’s base, 58, 62 Hydroamination, 167 Hydroborated bisalkenes, 4 silver-mediated ring closing of, 6 Hydroboration, 4, 204, 287 Hydrogenation, 202 Hydrogen bonds, 63 Hydrogen peroxide, 119 Hydrosilation, 188 Hydrosilylation, 184
392 a-Hydroxy acids, 211, 221 synthesis, from a-ketoesters, 211 a-Hydroxyalkynones, rearrangement to 3 (2H)-furanones, 110 b-Hydroxycarbonyl compounds, 268 2-Hydroxy-2-ethynyladamantane, 31 I2, 15–16, 153, 292 ICl, 153 a-Imidoylcarbenium intermediates, 86 Imine tautomer, 203 Iminium intermediate, 100 Iminium salts, 21, 22 Iminoalkynes, electrophilic cyclization of, 153 a-Iminoesters, 212, 262 Iminoiodanes, 168, 172 as nitrene precursors, 169 Indazolylborates, 243 Indene products, 131 3-1H-Indenyl ketones, 160 Indole alkaloids, 316 Indolizidinones, 70 Indolizines, 157 Indoloquinoline alkaloids, 310 Infinite 1D polycatenane, 339 Infinite 1D polyrotaxane, 343 Infinite tubes, 345 Insertion product, 239 Integrin inhibitors, 56 Intermolecular amination, with phenanthroline ligands, 174 Intermolecular competition reactions, 213 Intermolecular dimerization, 5 Intermolecular trapping, 124 Interrupted Nazarov reactions, 135–138 Intramolecular amidation, 168, 173 C–H bonds, 174 by silver-terpyridine catalyst, 175 intermolecular amination of, 168 Intramolecular deprotonation, 239 Intramolecular Heck coupling reactions, 302 Intramolecular rearrangement mechanism, 206 Invertomers, 87 Iodine, 9, 15, 31 1-Iodoadamantane, alkynylation, 27 Iodobenzene, 15 Iodoform, 31
INDEX
1-Iodo-2-phenylacetylene, 31 cis-1-Iodopropene, 15 N-Iodosuccinimide, 31, 33 5-Iodouracil, 303 Ionic liquid, 23, 93, 312, 359 Ireland–Claisen reaction, diastereoselectivity, 211 Ireland–Claisen rearrangement, 209 Iridium silylenoids, 188 Isobutene, 14 Isobutenyl radical, 14 Isobut-1-enylsilver, 13, 14 Isobut-1-enyltriethyllead, 14 Isobutyraldehyde, 217 Isocyanate anion, 121 Isocyanides, 66, 214 Isokinetic temperature, 198 Isopanepoxydone, 299 Isopropyl alcohol, 124 Isoquinoline-N-oxide structures, formation of, 156 Isoquinolines, 94, 153 Isothiocyanate group, 162 Isovanillin, 302 Julandine, 122 Kallolide B, 145 Karachine alkaloid structure, 154 Ketene, 85 a-Ketoesters, 211–212, 274, 296 aldol reaction, 275 b-Ketoesters, 101, 104, 154–155, 296, 375 KF, 266–267, 269, 273–275, 307 KOH, 4, 54, 288, 304 b-Lactams, 276 Lactone formation, 372 Lamellarin alkaloid family, 154 Lamellarins, 71 Late transition metal silylenoid complexes, 184 synthesis of, 185 LDA, 358 Lennoxamine, 35 s-Lewis acidity, 144 p-Lewis acidity, 144 Lewis acids, 131, 253 p-Lewis acid, 294
393
INDEX
s-Lewis acid, 294 activation, 243 catalyst, 262 Li(0), 191 Ligand, 60 Ligand exchange, 198 Light-emitting devices, 307 Light-induced silicon–silicon bond cleavage, 184 Light-induced 1,2-silyl shift, 185 LiHMDS, 53–54, 127–128 Lithium, 49, 191 lithium acetate, 47 zinc, 49 Lithium aluminum hydride, 206–207, 209, 310 Lithium bromide, 5, 48, 49 Lithium hexamethyldisilazide, 128 Lithium nitrate, 286 Lithium octylthiolate, 56 Lithium perchlorate, 55 Lithium silylenoid, 191 lvt topologies, 350, 352 g-Lycorane, total synthesis, 126 Macrolide antibiotic, 85 Maleate, 232 Maleic anhydride, 44, 45, 47 Mannich reaction, 276 alkynyl aldimines with silyl ketene acetal, 277 a-iminoester with enolethers, 276 Mannich-type reactions, 294 Masked allylic amines, 217 Mass spectrometry, 154, 230, 233 Me2AgMgBr, 6 ()-Menthol, 128 Menthyl acrylate, 51, 52 Menthyl-L-leucine ester, as chiral ligand, 290 Me3SiCF3, 8 Mesitylsilane, 11, 185 Me3SnOMe, 270, 271 Meso-cyclohexa-1,4-dienes, Shibasaki cyclization, 301 Meso-tetra-p-tolylporphyrin, 186 Metal carbenoids, 207 Metal-catalyzed N–H insertion, 243, 249 Metal-catalyzed O–H insertions, 250 Metallic silver, 2, 3, 12, 14
Metallocarbene, 234, 239, 243 Metal salt, effect of, 192 Metal silylenoids/silylmetal complexes, 184 reactivity and attributes of, 184–190 N-ortho-methoxyaniline-substituted imines, 77 5-Methoxyindane, 124 N-Methoxypyridinium iodide, 25 a-Methylacrolein, 158 Methyl acrylate, 45, 49, 51, 56, 58, 65, 68, 204, 214 Methylation, 17 Methyl-bicyclo[2.2.2]octane, 27 (S)-2-Methyl-1-butanol, 174 Methyl carbamate, 204 Methyl 2-cyanocinnamate, 74 3-Methylcyclohex-2-enol, 6 Methylene butyrolactones, carboxylates addition to alkynes, 372 Methylenesilaaziridines, 214 Methyl iodide, 7, 15, 27, 290 N-Methylmaleimide, 64 N-Methylmorpholine, 27, 290 Methyl-5-oxopentanoate, 295 Methyl propiolate, 47 2-Methylpyrazine, 345 1-Methyl-2-pyrrolidinone, 301 Methylsilver, 3 Methyl S-para-tolylsulfinylacrylate, 53 Methyl vinyl ketone, 133 ( þ )-Methynolide, 311 1,2-Migration mechanism, 112, 146 Migratory aptitudes, 146–147 Mitomycin, 242 Molar solubility product (Ksp), of silver salts, in water, 297 Montmorillite K10, 74 MOP, 60 Mukaiyama aldol reaction, 206–207, 271–272, 274 Munchnone, 67 NaOH, 204 Naphthalene-1,4-dione, 74 Naphthalene products, 131 Naphtho[b]cyclopropene, 73 Natural polyacetylenic products, 319 Natural products, 282 Nazarov cyclization, 133–135
394 NbO topologies, 350 Neocarzinostatin chromophor, 28 N–H bonds, carbene insertion, 243 N–H insertion, 243 NiCl2(dppe), 49 NiCl2(PPh3)2, 49 Ni(CO)4, 186 Nitrene transfer process, 173 Nitrene transfer reaction, 168 Nitric acid, 33 Nitrile insertion, 204 Nitroalkenes, 55 Nitrobenzene, 306 Nitrogen, 294 Nitroolefins, 67 3-Nitropyrrolidines, 55 Nitrosoaldol reaction, 277, 279, 281 N-nitrosoaldol reaction using tin enolates, 279 O-nitrosoaldol reaction using silylenol ethers, 280 using tin enolates, 278 Nitrosobenzene, 277–280 Nitroso compounds, 282 Nitrosyl chloride, 9 N–O bond, cleavage of, 279 Norcaradiene, 233 Norcarene derivative, 90 NsNH2, 175 Nucleophilic additions, 312 (OC)5Mo¼SiH2, 187 17 O labeling, 146, 147 Olefin–palladium complexes, 306 Olefins aziridination of, 168, 171, 174 [2 þ 2] intermolecular photodimerization of, 331 One-pot, three-component reaction, 277 Orbital symmetry rules, 118 Organic thin-film transistors, 307 Organocatalysis, 369 Organocatalysts, 279 Organocopper(I) compounds, in silico comparison of, 361 Organogold(I)-ate compounds, in silico comparison of, 361 Organosilver(I) compounds, in silico comparison of, 361
INDEX
Organosilver species, 93, 286, 292, 323–324 as nucleophilic reagents, 311 in enyne/arylyne synthesis, 316 in transmetallations, 314 Organostannane, 271 aldehydes, dienylation of, 263 Orthoester, 122 [Os(TTP)]2, 186 6p-Oxa electrocyclization, 105 1,5-(3-Oxapentane) diyl-bis(isonicotinate), 339 Oxasilacyclohexenes, platinum-catalyzed formation of, 189 Oxasilacyclopentenes, 190, 193–195, 201, 202, 206, 207, 209, 210, 212 Mukaiyama aldol functionalization of, 207 platinum-catalyzed formation of, 189 Oxasilacyclopropane, 208 Oxazines, 148 Oxazole synthesis, 151 Oxazoline, 217 Oxetane intermediate, 161 Oxidation, 202, 204 Oxidative condensation, 30 Oxidative cyclization, 152 Oxidative homocoupling, 3 Oxirane, ethane, oxidation of, 358 Oxirenium/dioxolenium pathways, 146 Oxonium intermediate, 110 Oxonium ion, 27 Oxonium ylide [1,2]-Stevens rearrangement of, 221 transition-metal-catalyzed [2,3]sigmatropic rearrangement of, 219 3-Oxo-N-propargylamides, 151 Oxophilic, 160 Oxyallyl cationic intermediate, 138 Oxy-Cope rearrangement, 98–99, 125 Oxygen Lewis bases, 294 Paclitaxel, 123 Pale’s silver-catalyzed cyclization, 372 Palladacycle, 190, 320 Palladium, 189 Palladium acetate, 289, 302, 308 Palladium bis(acetonitrile dichloride), 289 Palladium catalysis, 153, 311 Palladium catalyst, 288, 317
INDEX
Palladium-catalyzed C–C bond formation, 298 Palladium-catalyzed coupling, 15, 289 Palladium-catalyzed di-tert-butylsilylene transfer, 190 Palladium-catalyzed reactions, 303 Palladium chloride, 302 Palladium dichloride complexes, 307 Palladium(II) BP complex, 175 catalysis, 145 ipso position, aromatic electrophilic substitution, 310 Palladium silylenoid, 190 Palladium tetrakistriphenylphosphine, 302, 305, 317 Palladium trifluoroacetate, 309 Palytoxin synthesis, 303 Panus conchatus, 299 p Participation, 87 Pd–Ag-catalyzed Heck reactions, 301 Pd/Ag coupling reaction, with triisopropylsilylethynylphosphonate, 319 PdCl2/DPEphos system, 310 PdCl2(MeCN)2, 146, 289 PdII/PdIV redox system, 300 Pd(OAc)2, 34–35, 46, 289–302, 308–309, 321 Pd0/PdII vs. PdII/PdIV mechanism, 308 Pd(PPh3)4, 16, 28, 35–36, 217, 298, 303–306, 314, 317–320 Pd (PPh3)4–AgCl, 318 Pearson’s hard–soft acid–base theory, 217 Penicillin N, biosynthesis of, 84 PennPhos, 58–59 Pentadienyl cation, 135 cationic species, 133 2,4-Pentadienyl-stannane, 262 Pentamethyldisilane (HSiMe2SiMe3), 189 1,17-(3,6,9,12,15-Pentaoxaheptadecane) diyl-bis(nicotinate), 345 cis-/trans-2 Pentene, 178 Perchlorophenylacetylene, 31 Perez catalyst, 236 Perfluoroalkenylsilver compounds, 16, 17 Perfluoroalkyl iodides, 9 Perfluoroalkylsilver compounds, 7–9 Perfluoroalkyl thiosilver, 9 Perfluoroalkynes, 16
395 Perfluoroallenes, 16 Perfluoroarylsilver compounds, 12, 13 Perfluorobiphenyl, 12 Perfluoroisopropylsilver, 9, 16 Perfluorophenyl bromide, 12 Perfluorophenyllithium, 12 Perfluorophenylsilver, 12, 16 Perfluoro-tert-butylmethyl sulfide, 9 Pericyclic reactions, 207 Peroxyl silver acetylides, 21 (PhAg)2AgNO3, 10 PhBr¼NNs, 168 Phenanthrolines, 176 Phenanthroquinolizidine alkaloids, 122, 129 Phenylacetylene, 360 (E)-1-Phenylbut-2-en-1-one, 74 Phenylaziridines, 172 2-Phenylbromoethane, 302 Phenylethynylsilver, 19, 312 Phenylimines, 76 N-Phenylmaleimide, 56, 62, 64 Phenyl isocyanates, 55 azomethine ylides, reaction of, 55 Phenylmagnesium iodide, 11 (S)-2-Phenylpropionic acid, 108 1,2-Phenyl shift, 161 in diphenylallenyl ketones, 364 Phenylsilane, 188 Phenylsilver, 10–12 Phenyl vinyl sulfone, 50, 56, 58 Pheromone, 195 PhI¼NNs, 168 PhI¼NSO2Ph, 172 PhI¼NTs, 172, 178 PhI(OAc)2, 173, 175 Phosphatydyloxy allenone, 101 Phosphatyloxy, 111 Phosphatyloxy alkynone, 101, 112 Phosphatyloxy allene, 112 Phosphine ligand, effect on rate of silylene transfer, 196 Phosphonate-substituted pyrrolidines, 67 Photoluminescent properties, 337 PHOX type ligand, 63 (Ph3P)2AgOTf, 196–199 Ph3PAuCl, 186, 363–368, 370–371, 374 Ph3PAuOTf, 364, 373, 386 Ph3PAu-SiCP2Cl, 198 (Ph3P)2PdCl2, 190
396 (Ph3P)2Pt (C2H4), 184 (Ph3P)3RhCl, 184 PhSiH3, 184 Ph2Sn¼S, 74 Pictet–Spengler cyclization, 71 PINAP, 59–61 Piperidine, 360 derivatives, 99 2-Piperidinemethanol, 122, 123 Platinum cyclobutene intermediate, 187 Platinum(II) catalysis, 158 Platinum silylenoid, 187 31 P NMR, 272, 279 PNPCl, 8 Poly(arylated thiophenes), 307 Polycatenanes, 339, 345 with silver(I), 339 Polyfluorinated indazolyl silver complex, 248 Polymerizations, 5 Polypropionates, 221 Polyproprionate synthons, 195 Polyrotaxanes, 345 1D polyrotaxane, 344 2D polyrotaxane, 344 with silver(I), 342 Polysilanes, 184 Polyynes, 318 a-Po (pcu) topologies, 348 Porphyrin ruthenium nitrenoid, 179 Potassium acetylide, 19 Potassium carbonate, 29, 35, 36, 218, 302, 321 Potassium cyanide, 32 Potassium fluoride, 8, 307 Potassium nitrate, 99 (PPh3)Ag(CB11H12), 75 (PPh3)Ag(CB11H6Br6), 75 Prisms, 331 Propargyl acetates 1,2 migration, 104 [3,3]-sigmatropic rearrangement of, 104 Propargyl bromides, 107 Propargyl–Claisen rearrangement, 101, 104–106, 162, 365 N-Propargylcarboxamides, cyclization of, 368 Propargyl esters, silver-mediated rearrangement of, 113 Propargylic alcohols, 23 Propargylic amine formation, 313
INDEX
Propargylic amines, 21–23 Propargylic sulfinates, 108 Propargyl imines, 295 cis-1-Propenylsilver, 14 Propiolic acid propargyl esters, 371 Propionitrile, 12 Propynylsilver, 19 Protodesilylation, 33, 138, 202, 217 Protolysis, 17 Pseudopterolide diterpene, 145 P(t-Bu)2(2-biphenyl), 189 PtS topologies, 350, 352 ‘‘Push–pull’’ mechanism, 120 Pyranoquinoline, 150 Pyrans, 94 Pyrazoles, 25 Pyrazolylborates, 243 ligands, 177 Pyridine, 8–9, 24–25, 33, 100, 156, 169–170, 174, 293, 303, 305, 337 Pyridinium intermediate, 155 2-Pyridinyl, 305 3-Pyridinyl allylsilanes, 305 2-Pyridinyl silver, 305 2-Pyridones, 158 Pyridone synthesis, 159 Pyridyliminoalkynes, 153 Pyrite (pyr), 351 topologies, 350 a-Pyrones by cascade reaction, 371 silver-assisted preparation of, 109 Pyrroles, 106 by hydroamination, 374 scaffold, access, 157 Pyrrolidines, 48, 50, 53, 55–56, 68, 70, 99, 153, 169 Pyrrolidinone, 152 D1-Pyrrolines, 66 trans-2,5-Pyrrolidines, 153 Pyrrolizidine alkaloids, 153 Pyrroloisoquinoline scaffold, 154 Pyrroloquinoxalines, 157 Pyrrolothiazoles, 157 Pyrylium intermediate, 150 Pyrylium salt, 362 Quinap, 59–61 Quinolines, 94
INDEX
Quinoxaline, 156 Racemic kallolide B, synthesis of, 145 Radicals as intermediates, 287 pathway, 180 Rearrangement reactions cyclization cascade, 103 [1,2] rearrangements, 219 [2,3] rearrangement, 239 Reductive destannylation, 253 Regioselectivity, 55, 151,190, 191, 194, 203, 207, 217, 278, 299, 303, 308–309, 360 Relativistic effects, 84, 294, 361, 362 Retro[2 þ 2] cycloaddition, 315 Rh2(OAc)4, 234, 252 ()-Rhazinilam, 373 Rhodium-based catalysts, 243 Rhodium complexes, 243 Ring expansion reactions, 250 RNA polymerase inhibitors, 54 Rotaxane, 343 Ruthenium porphyrin catalyst, 168 Ruthenium silylenoid, 185 Rutile (rtl) topologies, 350, 351 Saturation kinetic behavior, 197 SbF6 counterion, 333, 366 Scandium triflate–catalyzed Mukaiyama aldol process, 206 Schlenk filtration, 193 Schrock carbene, 187 Scorpionate ligands, 242 SEAr process, 130 ()-Sedamine, asymmetric synthesis of, 295 Segphos, 77–78 Selective insertion reactions, 217 Sesquiterpene synthesis, 97 1,2 Shift, 111, 145, 239, 291 S–H insertions, 250 Sigmatropic rearrangements, 83–113 [1,2]-rearrangements, 108–113 [2,3]-rearrangements, 107–108, 219, 221 [3,3]-rearrangements, 95–97, 106, 111–113, 209, 210, 212, 219, 365– 366 29 Si {1H} NMR, 197 Silaaziridines, 214, 215 accessing method, 214 Silacarbonyl ylide, 207, 208, 212
397 Silacyclopentenes, functionalization of, 203 Silacyclopropane, 191, 200, 201, 204, 208 copper-catalyzed ring expansion, 206 in situ functionalization of, 194, 202 intramolecular rearrangement of, 205 relative reactivities of, 196 synthesis, 191, 205 synthetic utility of, 191 Silalactone, 209 Silane, 188 Silanol-based Hiyama couplings, 305 Silaoxazolidinone, 215 Sila–Wittig reaction, 187 [2,3]-Sila–Wittig rearrangement, 219 Silaziridines, synthetic utility of, 217 Silole, 189, 200 Siloxanes, 207 Silver, 49 as halogen scavenger, 297 as Lewis acid reagent, 294, 315 as reagent for decarboxylative coupling, 309 p complex, 29 p coordination to siloxyalkynes, 315 desilylation/coupling catalytic, 35 in Pd-catalyzed couplings, 298 in PdII-promoted electrophilic substitution of arenes, 306 intermolecular C–H amination, bathophenanthroline catalysis, 176 promoted cascade reaction, 211 promoted cationic aza-Cope rearrangement, 99 promoted 1,3-dipolarcycloaddition reactions, 212 promoted rearrangement, 87 Silver acetate, 22, 47–49, 51–56, 58, 60–63, 65, 66, 68, 70, 77, 122, 124, 126, 131, 294, 295, 298, 306, 308, 309 triethylamine-catalyzed reaction, 68 triphenylphosphine, 56 Silver acetylides, 18, 20, 24, 25, 27, 28, 30, 32, 33, 290, 311, 314, 312 addition to acid halides and carbon dioxide, 21 addition to aldehyde, 23 reactivity, 20 synthesis of, 19 to acid halides, 21
398 Silver azomethine complex, 63 Silver benzoate, 84, 85 Silver–BINAP catalyst, 261, 262 Silver BINAP complex, 64, 272, 279 structures of, 272 Silver bromide, 3, 5, 19 Silver carbene, 230, 239, 253 complex, 233 intermediacy of, 253 species, 230 Silver carbenoids, reactions, 253 Silver carbonate, 32, 49, 289, 298, 301, 302, 310, 320, 321 Silver catalysts, 232, 243, 280 Silver-catalyzed, 104, 313 aldehyde insertion, 217 aldolization, 294 B€uchner reaction, 235 carbene transfer, 230 cyclization, 102, 111 cycloadditions, 46, 53 [3 þ 2] cycloaddition, 67 cyclocarbostannylation, 253 cycloisomerization, 144 cyclopropanation, 233, 234, 236, 237 cyclopropyldiazoacetate ring expansion, scope of, 251 Doyle–Kirmse reaction, 241 free palladium coupling reactions, 322 oxy-Cope rearrangement, 99 propargyl Claisen rearrangement, 105 rearrangements, 93 ring expansion, 94 ring-opening cycloaddition, 71 ring rearrangement, 89 [3,3]-sigmatropic rearrangements, 102 silacarbonyl ylide cycloaddition, scope of, 213 three-component coupling, 23 of aldehydes/alkynes/amines, 312 Silver-catalyzed di-tert-butylsilylene transfer, 193, 219, 222 to aldehydes, 212 to an enolizable ester, 208 to di-/mono-substituted alkynes, 201 to disubstituted olefins, 193 to imines, 215 to a,b-unsaturated carbonyl compounds, 208
INDEX
Silver-catalyzed silylene transfer, 192, 195, 221 kinetic studies, 197 mechanism of, 197 product-determining step of, 198 Silver chloride, 11, 15, 169, 312, 365 Silver diacetate, 3 Silver(diazomethyl)phosphoryl compounds, 18 Silver-2,9-dimethyl-phenanthroline complex, 176 Silver fluoride, 8, 12, 16, 50, 65, 293 promoted desilylation, 50 Silver halides, 297 Silver hexafluoroantimonate (AgSbF6), 97, 104, 109, 315 Silver hexafluorophosphate, 110 Silver-impregnated solids, 74 Silver iodide, 29, 321 Silver ions, 108, 113 p-Lewis acidity, 315 Silver isocyanate, 121, 123 Silver-mediated aza-Cope rearrangements, 100 Silver-mediated carbene transfer reactions, 184 Silver-mediated di-tert-butylsilylene transfer, 200 catalytic cycle, 199 to acetylenes, 200 to carbonyl compounds, 207 to imines, 214 Silver-mediated [1,2]-sigmatropic rearrangements, 109 Silver metal, 10, 14 Silver nanoclusters, 84, 231 Silver nitrate (AgNO3), 2, 4, 5, 8, 10, 11, 14, 19, 31, 32, 49, 84, 86, 94, 96, 98, 99, 108, 169, 286, 287, 293, 306, 307, 337 Silver O-benzoyltartrate, 51 Silver oxide, 17, 18, 32, 49, 84, 182, 287–290, 293, 303–305, 307, 313, 320 Silver perchlorate (AgClO4), 64, 87, 89, 91–93, 96, 102, 129, 298, 337 Silver phenylacetylide, 24, 31 Silver phosphate, 23, 201, 204, 300–302 Silver phosphine complexes, 74, 198 Silver–pyrazolylborate-catalyzed aliphatic C–H amination, 179
INDEX
Silver salts, 84, 94, 97, 98, 101, 102, 110, 113, 144, 229, 289 molar solubility product (Ksp), water, 297 Silver silylenoid intermediate, 208 Silver-substituted diazomethyl compounds, 17 Silver tartrate, 51 Silver–terpyridine catalyst, 171, 173 Silver tetrafluoroborate (AgBF4), 96, 101, 108, 111, 112, 126, 128, 130, 134, 135, 137, 138, 296, 365, 366 Silver tosylate (AgOTs), 212, 322 Silver triflate (AgOSO3CF3), 20, 29, 58, 62, 71, 98, 108, 110, 113, 192, 195, 201, 214–215, 290, 296, 298, 311, 316 perchlorate, 106 silaaziridines, 215 triphenylphosphine complex, 316 Silver trifluoroacetate, 8, 12, 16, 84, 96, 119, 124, 201, 208, 219, 307 Silver trifluoromethanesulfonate, 4 Silver trinitromethane, 7 Silver tris(pyrazoyl)borate, 107 Silver(I) assisted ring-opening reactions, 120 catalysis, 145 catalyzed N–H insertion, 248–249 coordination polymers, 345, 346 perchlorate, 335 promoted electrocyclic ring opening, 126 promoted Nazarov reaction, 132 salts as Lewis acids, 118 tube-like complex, 339 tubular complexes, 337 Silver(I) carborane triphenylphosphine complex, 74 Silver(I) cation halophilic character of, 117 halophilicity of, 120 Silver(I) halides, 360 Silver(I) hexafluorophosphate (AgPF6), 332, 334 Silver(I) hydride, 6 Silver(I) methanesulfonate (AgO3SCH3), 333 Silver(I)–silver(I) (AgPF6), interactions, 330 Silver(I) tetrafluoroborate (AgBF4), 336, 339 Silver(I) triflate, 334
399 Silver(I) trifluoromethanesulfonate (AgOTf), 333 Silyl anion, 200 Silylation, 17 b-Silyl elimination, 198 Silylene, 189 transfer, 200 Silylenoid–catalyzed atom transfer reactions, 189 Silylenol ethers, 154, 159, 274, 277, 280, 281, 294 enantioselective hydroxymethylation, 274 Silylgold complex, 186 Silylmethyl allylic silanes, 200 formation mechanism, 199, 200 Silyloxyalkynes, 206 2-Silyloxycyclopentenyl cations, 136, 137 Silyloxyfurans, asymmetric Mannich reaction of, 276 1,2-Silyl shift, 184 Silylsilver intermediate electronic nature of, 198 spectroscopic attributes of, 197 Single-electron transfer (SET), 50 Single-stranded helical chain structure, 345 Six-membered cyclic transition state, 262 Skeletal rearrangement, 113 SN2 reaction, 94 SN20 reaction, 94, 95 Sodium acetate, 152 Sodium borohydride, 156 Sodium hydroxide, 345 Sodium tetraphenylborate, 185 Sodium thiosulfate, 84 Solvolysis reactions, 121, 122 Sonogashira coupling reaction, 29, 317 Spirocyclic furanones, 110 Spirocyclization, 127 Spiropyrrolidines, 68 sp–sp coupling reactions, 322 sp2–sp coupling reactions, 310 sp2–sp2 coupling reactions, 291 sp3–sp coupling reactions, 290 sp3–sp2 coupling reactions, 289 sp3–sp3 coupling reactions, 286 Square planar metal complexes, 331 SRN1 mechanism, 290 Stabilized haloallyl cation, 122 Stable silylenes, 186
400 p Stacking, 128 Stellated tetrahedral cage-like complex, 336 Stereoelectronic effect, 93, 118 Stereoelectronic factors, 118 Stereoselectivity, 49 Stereospecific, 192 Steric interactions, 202 [1,2]-Stevens rearrangement, 219, 221 Stille couplings, 304 Stille reactions, 303 Styrenes, 188, 197, 234, 296 silver p complex, 296 Styryl diazoacetates, 234, 243 Styrylsilver, 14 Substituted furo[2,3-b]pyrazines, silver(I)mediated synthesis of, 148 3-Substituted 2-pyrones, 109 Substrate-induced diastereoselectivity, in A3 coupling in water, 360 Sulfamate esters, 174 Sulfide imination, 172 Sulfilimines, 172 Sulfoxide imination, 172, 173 Sulfoximines, 172, 173 Suzuki coupling reactions, 288, 303, 304 Suzuki–Miyaura cross-coupling, 288 Tandem addition–cyclization reaction, 161 Tandem cyclization/pinacol rearrangement, 366 Tandem cycloisomerization–cycloaddition reaction, 158 Tandem Michael addition–Mannich reaction, 154 Tandem/sequential one-pot procedures, 71 Tartrate, 49 TBAF3H2O, 34–35, 306 TEMPO, 27 Terminal alkynes palladium–catalyzed coupling of, 30 coupling with aryl iodides/bromides without palladium, 321 TESOTf, 277 Tetrabutylammonium fluoride, 305, 318 Tetradentate ligand, 339 Tetrafluoroborate, 91, 102 3,30 ,5,50 -Tetrakis(4-pyridyl)bimesityl, 349 1,2,4,5-Tetrakis(3-pyridylmethylsulfanylmethyl)benzene, 349
INDEX
1,2,4,5-Tetrakis(4-pyridylmethylsulfanylmethyl)benzene, 350 Tetrakis(triphenylphosphine)palladium, 35, 302, 305, 317 3,30 ,5,50 -Tetramethyl-4,40 -bipyrazole, 345 Tetramethylethylenediamine (TMEDA), 322 Tetraphenylbutatriene, 15 Tetrasubstituted cyclopentadienes, 370 Tetrasubstituted furans, silver-catalyzed preparation of, 105 b-Tetronic acid, 289 TFA, 50, 71, 148–149, 308, 362 Thallium hydroxide, 303 Thermodynamic driving force, 218 Thiacalix[4]arene-p-sulfonate (TCAS), 336 Thiazole, 157 a-Thioallenes, cyclization of, 364 Thiourea intermediates, 162 Three-component coupling, 313 Tin enolates, 271, 278 TIPS alkynylethers, 45 TIPSNTf2, 45 TMEDA, 322–323 TMSCN, 156 TMSOTf, 154 Tol–BINAP, 77, 267, 278 (R)-Tol-BINAP, 232 Tol–BINAP–AgF, 273 Toluene-d8, 197 para-Toluenesulfenic acid, 53 Topisomerase I inhibitors, 71 Topological isomers, 350, 352 (4,4) Topologies, 345 (6,3) Topologies, 345 (6,6) Topologies, 350 4.82 Topology, 345 (10,3) Topology, 345, 349 (10,3)-a Topology, 350 (8,3)-b Topology, 350 (10,3)-d Topology, 349 Tosyl allenones, 111 N-Tosyl aziridines, 232 Tp*, 178 Tp*,Br, 178 TpBr3, 178 Tp(3;5-CF3)2Ag complex, 243 TpMs, 178 Transition metal complexes
401
INDEX
catalyzed insertion of di-tert-butylsilylene into C–O bonds, 220 silylenoid complexes, generation of, 186 catalyzed hydrosilation reactions, 187 reactivity of, 187 synthesis of silylenes, 184 Transmetallation, 15, 24, 30, 262, 287, 289, 291, 292, 293, 304, 314, 317 organosilver species, 314 1-Trialkylsilyl-1-alkynes, 317, 318 Tributylphosphinesilver iodide, 4 Tributyltin enolate, 268 Tributyltin enol ethers, 279 Tricyclic indoles, 123 Tricyclo[4.1.0.02,7]heptanes, 89 Tricyclohexylphosphine, 312 Triethylamine, 14, 35, 48, 49, 51, 54–56, 64, 68, 84, 85 Trifluoroacetic acid (TFA), 131, 148, 362 Trifluoroborane–etherate, 315 2,2,2-Trifluoroethanol (TFE), 124 Trifluoromethylsilver, 8 Trifluoromethylsulfenyl chloride, 9 2,4,6-Triisopropylphenylsilane, 185 Triisopropylsilyl (TIPS) group, 134 Trimethoxysilyl enol ethers, 273 hydroxymethylation of, 275 Trimethyl(perfluorophenyl)silane, 12 1-Trimethylsilyl-1-alkynes, 20, 32, 320 deprotection of, 32, 33 Trimethylsilyl chloride, 32 Trimethylsilyl diazomethane, 232 Trimethylsilyl enol ethers, 271, 281 1-Trimethylsilyl perfluoroalkenes, 16 Trimethylsilylpropiolates, 319 Trimethylsilyl-protected alkyne, 35 2-Trimethylsilylpyridine, 305 Trimethylsilyl triflate, 185 Trimethyltin enol ether, 278 Trinitromethyl iodide, 7 1,11-(3,6,9-Trioxaundecane)diyl-bis (nicotinate), 345 Triphenylethyllead, 10 Triphenylphosphane, 321 Triphenylphosphine, 20, 29, 56, 74, 75, 190, 195, 197, 198, 316 1,3,5-Tris(4-((2-ethyl-benzimidazol-1-yl) methyl) phenyl)benzene, 345 Tripodal ligands, 331
1,3-bis-(1-imidazolyl)-5-(imidazol-1ylmethyl)benzene, 345 1,3,5-tribenzimidazolyl-2,4,6trimethylbenzene, 333 1,3,5-Tri-3-pyridyl-2,4,6-tri-p-tolylbenzene, 334 1,3,5-Tris(imidazol-1-yl-methyl) benzene, 331 1,3,5-Tris(imidazol-1-ylmethyl)-2,4,6trimethylbenzene, 331 1,3,5-Tris(2-oxazolinyl)benzene, 335 Tris(pyrazolyl borate), 374 Tris(pyrazolyl)borato complex, 233 3,8,13-Tris(pyridylmethylamino)-2,7, 12-trimethoxy-10,15-dihydro-5Htribenzo[a,d,g]cyclonone, 335 2,4,6-Tris(tert-butyl)pyrimidine (TTBP), 111, 148 Trisubstituted pyrroles by hydroamination, 375 4,40 ,400 -Tri-tert-butyl-2,20 :60 ,200 -terpyridine (tBu3tpy), 170 Triton X-100, 77 Trost ligand, 58–60 TsN–NaBr, 168 TsN–NaCl, 168 1D tube, 339 Tube-like compounds, 337 Tungsten silylene complexes, 184 Turnover frequency, 75 Two-dimensional (2D) porous structure, 339 a,b-Unsaturated imines, 276 silaaziridines, decomposition of, 216 a,b-Unsaturated ketones, 266 r Value, 198 Vinylallenes, 370 Vinyl-/arylsilver species, homocoupling of, 292 Vinylboronic acids, 303 Vinyl iodides, 299 Vinyllithium reagents, 14 Vinylogous Mannich reaction with silyloxyfurans, 277 Vinylpalladium complex, 298 Vinylpalladium(II) complexes, 306 Vinylsilanes, 192, 204, 207
402
INDEX
Vinylsilvers, 14, 91, 99, 292–293 Vinyl triflates, 35, 36, 314
Ylides, formation and reactions of, 234 Ynones, 21
Wang resin, 71, 77 Wolff rearrangement, 84, 85, 229–231 Wurtz-type couplings, 287
Zeolites, 49 Zinc acetate, 47 Zinc bromide, 48, 193 Zinc iodide, 193 Zirconium acetylides, 24 Zirconium catalyzed nucleophilic addition, 24
X-ray analysis, 272 X-ray crystallography, 137 X-ray crystal structure, 333