<<

Selective : Synthetic Methods and Catalysts

Selective Glycosylations

Synthetic Methods and Catalysts

Edited by Clay S. Bennett Editor All books published by Wiley-VCH are carefully produced. Nevertheless, authors, editors, and Prof. Clay S. Bennett publisher do not warrant the information contained Tufts University in these books, including this book, to be free of Department of Chemistry errors. Readers are advised to keep in mind that 62 Talbot Ave. statements, data, illustrations, procedural details or Medford, MA 02155 other items may inadvertently be inaccurate. United States

Cover Library of Congress Card No.: applied for

British Library Cataloguing-in-Publication Data The cover image was kindly provided by the Editor A catalogue record for this book is available from the British Library.

Bibliographic information published by the Deutsche Nationalbibliothek The Deutsche Nationalbibliothek lists this publication in the Deutsche Nationalbibliografie; detailed bibliographic data are available on the Internet at http://dnb.d-nb.de.

© 2017 Wiley-VCH Verlag GmbH & Co. KGaA, Boschstr. 12, 69469 Weinheim, Germany

All rights reserved (including those of translation into other languages). No part of this book may be reproduced in any form – by photoprinting, microfilm, or any other means – nor transmitted or translated into a machine language without written permission from the publishers. Registered names, trademarks, etc. used in this book, even when not specifically marked as such, are not to be considered unprotected by law. Print ISBN: 978-3-527-33987-7 ePDF ISBN: 978-3-527-69622-2 ePub ISBN: 978-3-527-69624-6 Mobi ISBN: 978-3-527-69625-3 oBook ISBN: 978-3-527-69623-9 Cover Design Schulz Grafik-Design, Fußgönheim, Germany Typesetting SPi Global, Chennai, India Printed on acid-free paper v

Contents

List of Contributors xi Preface xv

Part I Introduction 1

1 Stereoselective Glycosylations – Additions to Ions 3 Bas Hagen, Stefan van der Vorm, Thomas Hansen, Gijs A. van der Marel, and Jeroen D.C. Codée 1.1 Introduction 3 1.2 Stability, Reactivity, and Conformational Behavior of Glycosyl Oxocarbenium Ions 4 1.3 Computational Studies 10 1.4 Observation of Glycosyl Oxocarbenium Ions by NMR Spectroscopy 14 1.5 Oxocarbenium Ion(-like) Intermediates as Product-Forming Intermediates in Reactions 15 1.6 Conclusion 24 References 26

2 Application of Armed, Disarmed, Superarmed, and Superdisarmed Building Blocks in Stereocontrolled Glycosylation and Expeditious Synthesis 29 Mithila D. Bandara, Jagodige P. Yasomanee, and Alexei V. Demchenko 2.1 Introduction: Chemical Synthesis of and 29 2.2 Fraser-Reid’s Armed–Disarmed Strategy for Oligosaccharide Synthesis 31 2.3 Many Reactivity Levels Exist between the Armed and Disarmed Building Blocks 33 2.4 Modes for Enhancing the Reactivity: Superarmed Building Blocks 35 2.5 Modes for Decreasing the Reactivity: Superdisarmed Building Blocks 38 2.6 Application of Armed and Disarmed Building Blocks in Stereocontrolled Glycosylation 41 2.7 Application of Armed/Superarmed and Disarmed Building Blocks in Chemoselective Oligosaccharide Synthesis 46 2.8 Conclusions and Outlook 54 References 55 vi Contents

3 Solvent Effect on Glycosylation 59 KwoK-Kong Tony Mong, Toshiki Nokami, Nhut Thi Thanh Tran, and Pham Be Nhi 3.1 Introduction 59 3.2 General Properties of Solvents Used in Glycosylation 60 3.3 Polar and Noncoordinating Solvents in Glycosylation 62 3.4 Weakly Polar and Noncoordinating Solvents in Glycosylation 63 3.5 Polar and Coordinating Solvents in Glycosylation 64 3.6 Weakly Polar and Coordinating Solvents in Glycosylation 68 3.7 Solvent Effect of Ionic Liquid on Glycosylation 71 3.8 Solvent Effect on Electrochemical Glycosylation 73 3.9 Molecular Dynamics Simulations Studies on Solvent Effect 73 3.10 Conclusions 74 References 75

Part II Stereocontrolled Approaches to Glycan Synthesis 79

4 Intramolecular Aglycon Delivery toward 1,2-cis Selective Glycosylation 81 Akihiro Ishiwata and Yukishige Ito 4.1 Introduction 81 4.2 Ketal Type Tethers 82 4.3 Silicon Tethers 82 4.4 2-Iodoalkylidene Acetals as Tether 84 4.5 Benzylidene Acetals as Tether 86 4.6 IAD through Hemiaminal 93 4.7 Conclusions 93 References 94

5 Chiral Auxiliaries in Stereoselective Glycosylation Reactions 97 Robin Brabham and Martin A. Fascione 5.1 Introduction 97 5.2 Neighboring Group Participation of O-2 Chiral Auxiliaries 97 5.3 Neighboring Group Participation of O-2 Achiral Auxiliaries 103 5.4 Preconfigured Chiral Auxiliaries 106 5.5 Conclusion 111 References 112

6 Glycosylation with Glycosyl Sulfonates 115 Luis Bohé and David Crich 6.1 Introduction 115 6.2 Formation of Glycosyl Sulfonates 115 6.3 Evidence for Glycosyl Sulfonates 118 6.4 Location of the Glycosyl Sulfonates in the General Glycosylation Mechanism 119 6.5 Applications in O- Synthesis 123 6.6 Applications in S-Glycoside Synthesis 128 6.7 Applications in C-Glycoside Synthesis 128 Contents vii

6.8 Polymer-Supported Glycosylation with Sulfonates 129 6.9 Conclusion 130 References 130

Part III Catalytic Activation of Glycosides 135

7 Stereoselective C-Glycosylation from Glycal Scaffolds 137 Kim Le Mai Hoang, Wei-Lin Leng, Yu-Jia Tan, and Xue-Wei Liu 7.1 Introduction 137 7.2 Classification of C-Glycosylation Reactions 138 7.3 Ferrier-Type Rearrangement 138 7.4 Pd-Catalyzed Heck-Type 140 7.5 Tsuji–Trost-Type C-Glycosylation 145 7.6 Sigmatropic Rearrangement 147 7.7 NHC-Catalyzed C-Glycosylations 149 7.8 Conclusion 151 References 151

8 Brønsted- and Lewis-Acid-Catalyzed Glycosylation 155 David Benito-Alifonso and M. Carmen Galan 8.1 Introduction 155 8.2 Chiral Brønsted Acids 155 8.3 Achiral Brønsted Acids 159 8.3.1 Homogeneous Brønsted Acid Catalysis 159 8.3.2 Heterogeneous Brønsted Acid Catalysis 161 8.4 Lewis-Acid-Catalyzed Glycosylations 161 8.4.1 Synthesis of O-Glycosides 161 8.4.2 Conformationally Constraint Glycosyl Donors 163 8.4.3 Synthesis of O-Glycoside Mimics 164 8.5 Metals as Lewis Acids 165 8.5.1 Gold 165 8.5.2 Cobalt 167 8.5.3 Nickel 168 8.5.4 Iron 168 8.6 Synthesis of C-Glycosides 169 8.7 Conclusions and Outlook 170 References 170

9 Nickel-Catalyzed Stereoselective Formation of 1,2-cis-2-Aminoglycosides 173 Eric T. Sletten, Ravi S. Loka, Alisa E. R. Fairweather, and Hien M. Nguyen 9.1 Introduction 173 9.2 Biological Importance of 1,2-cis-Aminoglycosides 173 9.3 Use of Nonparticipatory Groups to Form 1,2-cis-Aminoglycosides 175 9.4 Nickel-Catalyzed Formation of 1,2-cis-Aminoglycosides 178 9.5 C(2)-N-Substituted Benzylidene Glycosyl Trichloroacetimidate Donors 180 9.5.1 Comparison to Previous Methodologies 182 viii Contents

9.5.2 Expansion of Substrate Scope 183 9.6 Studies of C(2)-N-Substituted Benzylideneamino Glycosyl N-Phenyl Trifluoroacetimidate Donors 187 9.6.1 Synthetic Advantage of N-Phenyltrifluoroacetimidate Donors 189 9.7 1,2-cis-Amino Glycosylation of Thioglycoside Acceptors 190 9.8 Application to the Synthesis of Biologically Active Glycans 194 9.8.1 Mycothiol 196 9.8.2 GPI Anchor 197 9.8.3 O- Component of Gram-Negative Bacteria S. enterica and P. rustigianii 200 9.8.4 TN Antigen 201 9.8.5 Heparin 204 9.9 Conclusion 206 References 207

10 Photochemical Glycosylation 211 Justin Ragains 10.1 Introduction 211 10.2 Photochemistry Basics 212 10.3 Photosensitized O-Glycosylation with Chalcogenoglycoside Donors 214 10.4 Photochemical O-Glycosylation with Other Donors 223 10.5 Photosensitized C-Glycosylation 224 10.6 Conclusions 228 References 229

Part IV Regioselective Functionalization of Monosaccharides 231

11 Regioselective Glycosylation Methods 233 Mark S. Taylor 11.1 Introduction 233 11.2 Substrate Control: “Intrinsic” Differences in OH Group Reactivity of Glycosyl Acceptors 234 11.3 Substrate Control: Modulation of Acceptor OH Group Reactivity by Variation of Protective Groups 239 11.4 Substrate Control: /Acceptor Matching in Regioselective Glycosylation 241 11.5 Reagent-Controlled, Regioselective Glycosylation 243 11.6 Enzyme-Catalyzed Regioselective Glycosylation 246 11.7 Synthetic Catalysts for Regioselective Glycosylation 248 11.8 Summary and Outlook 250 References 252

12 Regioselective, One-Pot Functionalization of 255 Suvarn S. Kulkarni 12.1 Introduction 255 Contents ix

12.2 Regioselective, Sequential Protection/Functionalization of Polyols 256 12.3 Regioselective, One-Pot Protection of via TMS Protection of Polyols 262 12.4 Orthogonally Protected d-Glycosamine and Bacterial Rare Building Block via Sequential, One-Pot Nucleophilic Displacements of O-Triflates 268 12.5 Summary and Outlook 273 References 273

Part V Stereoselective Synthesis of Deoxy Sugars, Furanosides, and Glycoconjugate Sugars 277

13 Selective Glycosylations with Deoxy Sugars 279 Clay S. Bennett 13.1 Introduction 279 13.2 Challenges in 2-Deoxy-Sugar Synthesis 281 13.3 Protecting Group Strategies 283 13.4 Addition to Glycals 284 13.5 Additions to Glycosyl Halides 285 13.6 Latent Glycosyl Halides 287 13.7 Reagent-Controlled Approaches 289 13.8 Umpolung Reactivity 291 13.9 Conclusion 293 References 293

14 Selective Glycosylations with Furanosides 297 Carola Gallo-Rodriguez and Gustavo A. Kashiwagi 14.1 Introduction 297 14.2 Construction of the Furanose Template 299 14.3 Stereoselective Glycosylation with Furanoside Donors 300 14.3.1 1,2-trans Furanosides 300 14.3.2 1,2-cis Furanosides 303 14.3.2.1 Flexible Donors 303 14.3.2.2 Conformationally Restricted Donors 307 14.3.2.3 Indirect Methods 314 14.3.2.4 Oxidation–Reduction 316 14.4 Reactivity Tuning of Furanosides for Oligosaccharide Synthesis 319 14.5 Conclusion 321 References 321

15 De novo Asymmetric Synthesis of Carbohydrate Natural Products 327 Pei Shi and George A. O’Doherty 15.1 Introduction 327 15.2 Danishefsky Hetero-Diels–Alder Approach 328 15.3 MacMillan Proline Aldol Approach 330 x Contents

15.4 The O’Doherty Approaches 334 15.4.1 O’Doherty Iterative Dihydroxylation Approach 334 15.4.2 O’Doherty Achmatowicz Approach 335 15.4.3 De novo Use of the Achmatowicz Approach to Pyranose 336 15.4.4 De novo Access to Monosaccharide Natural Products 337 15.4.5 De novo Access to Oligosaccharide Natural Products 341 15.5 Conclusion 347 References 347

16 Chemical Synthesis of Sialosides 353 Yu-Hsuan Lih and Chung-Yi Wu 16.1 Introduction 353 16.2 Chemical Synthesis of Sialosides 354 16.3 Conclusions 366 References 367

Index 371 xi

List of Contributors

Mithila D. Bandara Robin Brabham University of Missouri - St. Louis York Structural Biology Laboratory Department of Chemistry and Department of Chemistry Biochemistry University of York One University Blvd. Heslington, York St. Louis, MO 63121 YO10 5DD USA UK

David Benito-Alifonso Jeroen D.C. Codée University of Bristol Leiden University School of Chemistry Leiden Institute of Chemistry Cantock’s Close PO Box 9502 Bristol BS8 1TS 2300 RA Leiden UK The Netherlands

Clay S. Bennett David Crich Tufts University Wayne State University Department of Chemistry Department of Chemistry 62 Talbot Ave. 5101 Cass Ave. Medford, MA 02155 Detroit, MI 48202 USA USA

Luis Bohé Alexei V. Demchenko Institut de Chimie des Substances University of Missouri - St. Louis Naturelles Department of Chemistry and CNRS-ICSN UPR2301 Biochemistry Université Paris-Sud One University Blvd. Avenue de la Terrasse St. Louis, MO 63121 91198 Gif-sur-Yvette USA France Alisa E. R. Fairweather University of Iowa Department of Chemistry E331 Chemistry Building Iowa City, IA 52245 USA xii List of Contributors

Martin A. Fascione Akihiro Ishiwata York Structural Biology Laboratory RIKEN Department of Chemistry Synthetic Cellular Chemistry Laboratory University of York 2‐1 Hirosawa Wako Heslington, York Saitama 351‐0198 YO10 5DD Japan UK Yukishige Ito M. Carmen Galan RIKEN University of Bristol Synthetic Cellular Chemistry Laboratory School of Chemistry 2‐1 Hirosawa Wako Cantock’s Close Saitama 351‐0198 Bristol BS8 1TS Japan UK Gustavo A. Kashiwagi Carola Gallo-Rodriguez Universidad de Buenos Aires Universidad de Buenos Aires CIHIDECAR Departamento de Química CIHIDECAR Departamento de Química Orgánica Orgánica Ciudad Universitaria, Pabellón II Ciudad Universitaria, Pabellón II 1428 Buenos Aires 1428 Buenos Aires Argentina Argentina Suvarn S. Kulkarni Bas Hagen Indian Institute of Technology Bombay Leiden University Department of Chemistry Leiden Institute of Chemistry Powai PO Box 9502 Mumbai 400076 2300 RA Leiden India The Netherlands Wei-Lin Leng Thomas Hansen Nanyang Technological University Leiden University Division of Chemistry and Biological Leiden Institute of Chemistry Chemistry PO Box 9502 School of Physical and Mathematical 2300 RA Leiden Sciences The Netherlands SPMS-CBC-02-01 21 Nanyang Link, 637371 Kim Le Mai Hoang Singapore Nanyang Technological University Division of Chemistry and Biological Yu-Hsuan Lih Chemistry Genomics Research Center, Academia School of Physical and Mathematical Sinica Sciences 128 Academia Road, Section 2, Nankang SPMS-CBC-02-01 Taipei 115 21 Nanyang Link, 637371 Taiwan Singapore List of Contributors xiii

Xue-Wei Liu Toshiki Nokami Nanyang Technological University Tottori University Division of Chemistry and Biological Department of Chemistry and Chemistry Biotechnology School of Physical and Mathematical 4-101 Koyama-Minami Sciences Tottori city SPMS-CBC-02-01 680-8552 Tottori 21 Nanyang Link, 637371 Japan Singapore George A. O’Doherty Ravi S. Loka Northeastern University University of Iowa Department of Chemistry and Chemical Department of Chemistry Biology E331 Chemistry Building 360 Huntington Ave. Iowa City, IA 52245 Boston, MA 02115 USA USA

Gijs van der Marel Justin Ragains Leiden University Louisiana State University Leiden Institute of Chemistry Department of Chemistry PO Box 9502 232 Choppin Hall 2300 RA Leiden Baton Rouge, LA 70803 The Netherlands USA

KwoK-Kong Tony Mong Pei Shi National Chiao Tung University Corden Pharma Applied Chemistry Department 1-B Gill Street 1001 Ta Hsueh Road Woburn, MA 01801 Hsinchu 300 USA Taiwan ROC Eric T. Sletten Hien M. Nguyen University of Iowa University of Iowa Department of Chemistry Department of Chemistry E331 Chemistry Building E331 Chemistry Building Iowa City, IA 52245 Iowa City, IA 52245 USA USA Yu-Jia Tan Pham Be Nhi Nanyang Technological University National Chiao Tung University Division of Chemistry and Biological Applied Chemistry Department Chemistry 1001 Ta Hsueh Road School of Physical and Mathematical Hsinchu 300 Sciences Taiwan ROC SPMS-CBC-02-01 21 Nanyang Link, 637371 Singapore xiv List of Contributors

Mark S. Taylor Chung-Yi Wu University of Toronto Genomics Research Center, Academia Department of Chemistry Sinica 80 St. George Street 128 Academia Road, Section 2, Nankang Toronto, ON M5S 3H6 Taipei 115 Canada Taiwan

Nhut Thi Thanh Tran Jagodige P. Yasomanee National Chiao Tung University University of Missouri - St. Louis Applied Chemistry Department Department of Chemistry and 1001 Ta Hsueh Road Biochemistry Hsinchu 300 One University Blvd. Taiwan ROC St. Louis, MO 63121 USA Stefan van der Vorm University of Leiden Leiden Institute of Chemistry PO Box 9502 2300 RA Leiden The Netherlands xv

Preface

The past decade has seen increased recognition of the important roles oligosaccharides play in an array of biological process including (but not limited to) protein folding, pathogen invasion, cell adhesion, and immune response. As a consequence, the field of glycoscience is undergoing rapid growth, with an ever‐increasing number of investiga- tors turning their attention to it. Despite all of this, the field is still in its infancy, espe- cially compared to other areas of biology such as genomics and proteomics. There are several reasons for this; chief among them is the fact that glycoscientists do not enjoy ready access to homogeneous material for study, an advantage that was critical for the advances in other areas of biomedical research. This is because, unlike the other major classes of biopolymers, cells produce carbohydrates as heterogeneous mixtures, which are often intractable. As a consequence, organic synthesis (including chemoenzymatic synthesis) remains the only avenue for the production of pure oligosaccharides for bio- medical evaluation. The synthesis of most oligosaccharides is a nontrivial undertaking, however, owing to issues of regiochemistry and . Thus, while has been known for over a century, the construction of a new oligosac- charide can still be a research project in and of itself. Among the challenges facing the chemist who wishes to synthesize oligosaccharides, one of the most significant is controlling selectivity in the glycosylation reaction. Typical glycosylation reactions proceed through a mechanism somewhere along the SN1–SN2 continuum, which renders controlling selectivity in the reaction immensely difficult. While several elegant solutions to this problem have been devised, a general approach to controlling selectivity in glycosylation reactions with a broad range of substrates remains to be developed. This has prompted calls for the development of new approaches to glycosylation from numerous sectors. Before embarking on developing a new approach to glycosylation, however, it is first necessary to understand the advances that currently constitute the state of the art. The purpose of this volume is to describe the principles of chemical glycosylation. Rather than break down the text into chapters focusing on activating different classes of leaving groups, the focus is instead largely on mechanistic aspects that are responsible for selectivity. Furthermore, technologies for automated and one‐pot synthesis have been extensively reviewed elsewhere and will only be covered when relevant. This volume is organized into five parts. The first part deals with an introduction to the basic principles or . In Chapter 1, Codeé et al. outline the factors responsible for controlling additions to oxocarbenium cations. Next, Demchenko xvi Preface

and coworkers describe the roles protecting groups play in both attenuating glycan reactivity and controlling stereoselectivity in glycosylations in Chapter 2. This part con- cludes with Chapter 3 from Mong, Nokami, and coworkers, which details the roles solvents play in controlling the stereochemical outcome of glycosylation. Part II describes ways in which electrophilic glycosyl donors can be modified to undergo selective reactions. In Chapter 4, Ishiwata and Ito provide a detailed introduc- tion to the use of Intramolecular Aglycone Delivery (IAD) for the stereoselective syn- thesis of cis‐1,2‐glycans. This is followed by a discussion of the use of chiral auxiliaries in oligosaccharide synthesis by Brabham and Fascione in Chapter 5. Finally, Bohé and Crich describe how glycosyl sulfonates permit the construction of the so‐called difficult linkages through SN2‐like glycosylations in Chapter 6. The development of methods for catalytic activation of glycosyl donors is the focus of Part III. This part begins with a description of methods for the construction C‐glycans, often through transition‐metal‐mediated processes, by Liu et al. in Chapter 7. This is followed by a comprehensive overview of recent approaches for catalytic activation of donors for O‐glycosylation by Benito‐Alfonso and Galan in Chapter 8. In Chapter 9, Nguyen and coworkers provide a case study in catalytic activation, focusing on their Ni‐catalyzed 1,2‐cis‐glycoside synthesis. This part concludes with an introduction to the increasingly popular field of photochemical glycosylation by Ragains in Chapter 10. In addition to the challenges in controlling the stereochemical outcome of glycosyla- tion reactions, regioselectivity is a problem the synthetic chemist must attend to when dealing with glycosides. Current state‐of‐the‐art approaches to addressing this issue are outlined in Part IV of the volume. In Chapter 11, Taylor provides us with a discussion of methods for regioselectively glycosylating unprotected glycosyl acceptors. This is fol- lowed by a discussion of methods for one‐pot protection and functionalization of unprotected glycans by Kulkarni in Chapter 12. The final part of the volume provides the reader with examples of classes of glycans where standard approaches to glycosylation do not always apply. This begins with an overview of recent advances in 2‐deoxy‐sugar synthesis by Bennett in Chapter 13. Gallo‐Rodriguez and Kashiwagi follow this up with an introduction to the challenging issue of controlling selectivity in glycosylations with furanoside donors (Chapter 14). Next, Shi and O’Doherty provide us with a description of how the de novo synthesis can permit the construction of a number of carbohydrate natural products in Chapter 15. Finally, Lih and Wu provide an overview of the state of the art in the synthesis of sialic acids in Chapter 16. The goal of this volume is to try to provide a holistic view of chemical glycosylation. Our target audience is not limited to individuals who are currently engaged in carbohy- drate chemistry but extends to the larger synthetic community, many of whom may be new to the field. Our hope is that this volume will inspire investigators to make new, and ideally unforeseen, contributions to the field. We do this because we believe that it will be necessary to engage as many investigators as possible if we are to achieve the long‐ term goal of developing technologies that will permit the routine and rapid construction of oligosaccharide libraries that are desperately needed for the study of glycobiology.

Medford, MA, USA Clay S. Bennett Tufts University 1

Part I

Introduction

3

1

Stereoselective Glycosylations – Additions to Oxocarbenium Ions Bas Hagen, Stefan van der Vorm, Thomas Hansen, Gijs A. van der Marel, and Jeroen D.C. Codée

1.1 Introduction

Tremendous progress has been made in the construction of oligosaccharides, and many impressive examples of large and complex oligosaccharide total syntheses have appeared over the years [1]. At the same time, the exact mechanism underlying the union of two carbohydrate building blocks often remains obscure, and optimization of a glycosyla- tion reaction can be a time- and labor-intensive process [2, 3]. This can be explained by the many variables that affect the outcome of a glycosylation reaction: the nature of both the donor and acceptor building blocks, solvent, activator and activation protocol, temperature, concentration, and even the presence and the type of molecular sieves. The large structural variety of carbohydrates leads to building blocks that differ signifi- cantly in reactivity, with respect to both the nucleophilicity of the acceptor molecule and the reactivity of the donor species. The reactivity of a donor is generally related to the capacity of the donor to accommodate developing positive charge at the anomeric center, upon expulsion of the anomeric . This also determines the amount of character in the transition state leading to the products. Most glycosyla- tion reactions will feature characteristics of both SN1- and SN2-type pathways in the transition states leading to the products. It is now commonly accepted that the exact mechanism through which a glycosidic linkage is formed can be found somewhere in the continuum of reaction mechanisms that spans from a completely dissociative SN1 mechanism on one side to an associative SN2 pathway on the other side (Figure 1.1) [4–6]. On the SN1-side of the spectrum, glycosyl oxocarbenium ions are found as prod- uct-forming intermediates. On this outer limit of the reaction pathway continuum, the oxocarbenium ions will be separated from their counterions by solvent molecules (sol- vent-separated ion pairs, SSIPs), and there will be no influence of the counterion on the selectivity of the reaction. Moving toward the SN2 side of the spectrum contact (or close) ion pairs (CIPs) are encountered, and in reactions of these species, the counterion will have a role to play. Because glycosylation reactions generally occur in apolar sol- vents (dichloromethane is by far the most used one), ionic intermediates have very limited lifetimes, and activated donor species will primarily be present as a pool of

Selective Glycosylations: Synthetic Methods and Catalysts, First Edition. Edited by Clay S. Bennett. © 2017 Wiley-VCH Verlag GmbH & Co. KGaA. Published 2017 by Wiley-VCH Verlag GmbH & Co. KGaA. 4 1 Stereoselective Glycosylations – Additions to Oxocarbenium Ions

O PO LG Activation E-X

− X − + + + + O O O O O X O PO PO PO PO PO PO X − X − X X Covalent Contact Solvent-separated Solvent-separated Contact Covalent species ion pair ion pair ion pair ion pair species (α) (α-CIP) (α-SSIP) (β-SSIP) (β-CIP) (β)

Increasing SN1-character Increasing SN1-character SN1

Increasing SN2-character SN2 SN2

Figure 1.1 Continuum of mechanisms to explain the stereochemical course of glycosylation reactions.

covalent intermediates. The stability, lifetime, and reactivity of an oxocarbenium ion depend – besides the nature of the counterion – on the nature and orientation of the functional groups present on the carbohydrate ring. This chapter explores the role of oxocarbenium ions (and CIPs, featuring a glycosyl cation) in chemical glycosylation reactions. While it was previously often assumed that glycosylations, proceeding via an oxocarbenium ion intermediate, show poor stereoselectivity, it is now clear that oxocar- benium ions can be at the basis of stereoselective glycosylation events. The first part of this chapter deals with the ­stability, reactivity, and conformational behavior of glycosyl oxocarbenium ions, whereas the second part describes their intermediacy in the assem- bly of (complex) oligosaccharides.

1.2 Stability, Reactivity, and Conformational Behavior of Glycosyl Oxocarbenium Ions

Amyes and Jencks have argued that glycosyl oxocarbenium ions have a short but signifi- cant lifetime in aqueous solution [7]. They further argued that in the presence of prop- erly positioned counterions (such as those derived of expulsion of an aglycon), CIPs will rapidly collapse back to provide the covalent species and that the “first stable intermedi- ate for a significant fraction of the reaction” should be the solvent-separated oxocarbe- nium ion. By extrapolation of these observations to apolar organic solvents, Sinnott reached the conclusion that intimate ion pairs have no real existence in an apolar envi- ronment, such as used for glycosylation reactions [8]. Hosoya et al. have studied CIPs by quantum mechanical calculations in dichloromethane as a solvent [9]. In these cal- culations, they have included four solvent molecules to accurately mimic the real-life situation. In many of the studied cases, CIPs turned out to be less stable than the cor- responding solvent-separated ions, as will be described next [10]. Yoshida and cowork- ers have described that activation of thioglucoside 1 with a sulfonium salt activator, featuring the bulky nonnucleophilic tetrakis(pentafluorophenyl) borate counterion, in a continuous-flow microreactor, provides a reactive species (2) that has a lifetime on the 1.2 Stability, Reactivity, and Conformational Behavior of Glycosyl Oxocarbenium Ions 5

OBn + OBn − ArS SAr X O BnO O ArS BnO + BnO SAr BnO BnO Ar = p-FC H 6 4 BnO ArS SAr 1 X = B(C6F5)4 2 B(C6F5)4 −

Scheme 1.1 Generation of glucosyl oxocarbenium ions in a continuous-flow microreactor. order of a second (Scheme 1.1) [11]. They argued that this species was a glucosyl oxo- carbenium ion, “somewhat stabilized” by the disulfide generated from the donor agly- con and the activator. The stability of a glycosyl oxocarbenium ion is largely influenced by the substituents on the carbohydrate ring. The electronegative substituents (primarily oxygen, but also nitrogen-based) have an overall destabilizing effect on the carbocation, and the desta- bilizing effect can be further enhanced by the presence of electron-withdrawing pro- tecting groups, such as acyl functions. The exact position of the substituent on the ring and its orientation influence the stability of the anomeric cation. The combined influ- ence of all substituents on the ring determines the reactivity of a glycosyl donor, and the extensive relative reactivity value (RRV) charts, drawn up by the Ley and Wong groups for a large panel of thioglycosides, clearly illustrate these functional group effects [12–14]. From these RRV tables, it is clear that the donor reactivity spectrum spans at least eight orders of magnitude. To investigate the influence of the carbohydrate ring substituents on the stereochemical outcome of a glycosylation reaction, Woerpel and coworkers have systematically studied C-glycosylation reactions of a set of furanosides and pyranosides, featuring a limited amount of ring substituents [15–20]. Their studies in the furanose series are summarized in Scheme 1.2a [15, 17]. As can be seen, the alkoxy groups at C2 and C3 have a strong influence on the stereochemical outcome of the reaction, where the alkoxy group at C5 appears to have less effect on the reaction. The presence of an alkoxy or alkyl group at C3 leads to the formation of the allylglyco- sides 11 and 12 with opposite stereoselectivity. Woerpel and coworkers have devised a model to account for these stereodirecting substituent effects that takes into account the equilibrium between two possible envelope oxocarbenium ion conformers (13 and 14, Scheme 1.2b) [17]. Attack on these oxocarbenium ion conformers by the occurs from the “inside” of the envelopes, because this trajectory avoids unfavorable eclipsing interactions with the substituent at C2, and it leads, upon rehy- bridization of the anomeric carbon, to a fully staggered product (15 and 16), where attack on the “outside” would provide the furanose ring with an eclipsed C1─C2 con- stellation. The spatial orientation of the alkoxy groups influences the stability of the oxocarbenium ions. An alkoxy group at C3 can provide some stabilization of the carbo- cation when it takes up a pseudo-axial position. Stabilization of the oxocarbenium ion featuring a C2-alkoxy group is best achieved by placing the electronegative substitu- ent in a pseudo-equatorial position to allow for the hyperconjugative stabilization by the properly oriented C2─H2 bond. Alkyl substituents at C3 prefer to adopt a pseudo-­ equatorial position because of steric reasons. With these spatial substituent prefer- ences, the stereochemical outcome of the C-allylation reactions in Scheme 1.2 can be explained. Activation of the C3-benzyloxyfuranosyl acetate with SnBr4 can provide an 6 1 Stereoselective Glycosylations – Additions to Oxocarbenium Ions

O O OAc AllyI-TMS, SnBr , DCM R 4 R 3 R: 3-OBn 8–12 4 R: 2-OBn 5 R: 5-CH2OBn 6 R: 3-OBn, 5-CH2OBn 7 R: 3-Me, 5-CH2OBn BnO BnO O O O O

BnO OBn R 8 94 : 6 9 85 : 15 10 57 : 43 11 R = OBn: 1,3-cis (>99 : 1) cis-Stereoselectivity cis-Stereoselectivity 12 R = Me: 1,3-trans (5 : 95) (a)

O Nu

15

Inside attack Nu Nu R′ R′ + O + + H O O H O + R R 13 14 3 E Nu E 3 Nu

Inside attack

O Nu

(b) 16

Scheme 1.2 (a) Diastereoselective C-allylations of furanosyl acetates. (b) “Inside” attack model.

oxocarbenium ion intermediate that preferentially adopts an E3 conformation, as in 14. Nucleophilic attack on this conformer takes place from the diastereotopic face that leads to the 1,3-cis product. In a similar vein, inside nucleophilic attack on the C2-benzyloxy furanosyl oxocarbenium ion E3 conformer, derived from furanosyl ace- tate 4, accounts for the stereochemical outcome of the C-allylation leading to product 9. To accurately gauge the combined effect of multiple substituents on a furanosyl ring, van Rijssel et al. [21, 22] used a quantum mechanical calculation method, originally developed by Rhoad and coworkers [23], to map the energy of furanosyl oxocarbenium ions related to the complete conformational space they can occupy. Energy maps for all four possible diastereoisomeric, fully decorated furanosyl oxocarbenium ions were gen- erated revealing the lowest energy conformers for the ribo-, arabino-, xylo-, and lyxo- configured furanosyl oxocarbenium ions 17–21 (Scheme 1.3). It became apparent that the orientation of the C5-substituent, having a gg, gt, or tg relation to the substituents at H OMe MeO H H OMe H H OMe MeO H O 0.5 O O 1.0 MeO MeO 1.5 2.0 MeO OMe 2.5 3.0 3.5 4.0 4.5 5.0 5.5 6.0 6.5 7.0 7.5 8.0 8.5 9.0 kcal mol−1 H H MeO OMe H H OMe O O MeO

OMe (a) MeO

TMSOTf O O OAc TES-d D 22–25 26–29 R DCM R

BnO BnO BnO BnO O D O D O D O D

BnO OBn BnO OBn BnO OBn BnO OBn 26 ribo 27 arabino 28 xylo 29 lyxo >98 : 2 >98 : 2 85 : 15 >98 : 2 (b) 1,2-cis 1,2-cis 1,2-cis 1,2-cis

Scheme 1.3 Free energy surface maps of fully decorated furanosyl oxocarbenium ions and diastereoselective reductions of furanosyl acetates. 8 1 Stereoselective Glycosylations – Additions to Oxocarbenium Ions

C4, was of profound influence on the stability of the oxocarbenium ions, and differences up to 4 kcal mol−1 were observed for structures only differing in their C4─C5 rotation. These stereoelectronic effects have also been described in the pyranose series, where a C4─C6 acetal can restrict the C6-oxygen in a tg position, for manno- and gluco-config- ured systems, or in a gg position for galacto-configured constellations [24–26]. The tg orientation represents the most destabilizing orientation because in this situation, the O6 atom is farthest away from the electron-depleted anomeric center, not allowing for any electron density donation for stabilization. With the lowest energy furanosyl oxo- carbenium ion conformers found by the free energy surface (FES) mapping method, the stereochemical outcome of reduction reactions at the anomeric center of the four dias- tereoisomeric furanosyl acetates 22–25 could be explained (Scheme 1.3). Interestingly, all four furanosides reacted in a 1,2-cis selective manner with the incoming nucleophile (tri-ethylsilane-d). Only xylofuranosyl acetate 24 provided some of the 1,2-trans addi- tion products, which could be related to the stability of the 3E gt oxocarbenium ion intermediate 20. The stereoelectronic substituent effects found in the furanose series are paralleled in the pyranose system, where the following substituent effects have been delineated: the stability of pyranosyl oxocarbenium ions benefits from an equatorial orientation of the C2-alkoxy groups (allowing for hyperconjugative stabilization by the σC2─H2 bond) and an axial orientation of the C3 and C4 alkoxy groups [20]. The C5-alkoxymethylene group has a slight preference for an equatorial position because of steric reasons [18]. These substituent preferences have been used to explain the stereochemical outcome of a series of C-allylations, using a two-conformer model. Woerpel and coworkers rea- soned that six-membered oxocarbenium ions preferentially adopt a half-chair structure to accommodate the flat [C1=O5]+ oxocarbenium ion moiety (Scheme 1.4a) [20]. These half-chair intermediates are attacked by incoming following a trajectory 3 that leads to a chair-like transition state. Thus, attack of a H4 half chair 30 preferentially occurs form the β-face (in the case of a d-pyranoside), where attack on the opposite half 4 chair 31 (the H3) leads to the α-product. With the described spatial substituent prefer- ences and mode of nucleophilic attack, the stereoselectivities in the C-allylation reac- tions shown in Scheme 1.4b can be accounted for: the C4─OBn is trans-directing, where the C3 and C2─O─Bn promote the formation of the cis-product. In the lyxo- pyranosyl oxocarbenium ion, these three substituent preferences can be united, and the allylation of 2,3,4-tri-O-benzyl lyxopyranosyl acetate 35 proceeds in a highly stereose- lective manner to provide the 1,2-cis product 40. When a C5 benzyloxymethyl group is added to this system, as in a mannosyl cation, 3 4 it can be reasoned that the H4 oxocarbenium ion is more stable than its H3 counter- part (see Scheme 1.5): the C2, C3, and C4 groups are all positioned properly to provide maximal stabilization of the electron-depleted anomeric center, and only the C5 sub- stituent, in itself not a powerful stereodirecting group, is not positioned favorably [18]. However, the axial orientation of this group does lead to a significant 1,3-diaxial inter- action with the axially positioned C3-alkoxy group. The allylation of mannose proceeds 3 with α-selectivity, indicating that nucleophilic attack on the β-face of the H4 oxocarbe- nium is not a favorable reaction pathway. To account for this stereochemical outcome, Woerpel and coworkers have suggested a Curtin–Hammett kinetic scenario, in which 3 the two half chairs 42 and 43 are in rapid equilibrium. Attack on the H4 conformer suffers from unfavorable steric interactions between the incoming nucleophile and the 1.2 Stability, Reactivity, and Conformational Behavior of Glycosyl Oxocarbenium Ions 9

Nu

+ O O +

Nu 30 31 3 4 (a) H4 H3

AllylTMS, O OAc O BF3⋅OEt2, DCM

BnO BnO 32–36 37–41

OBn O O O O O

BnO OBn BnO OBn BnO OBn BnO OBn OBn 37 38 39 40 41 (b) dr 99 : 1 dr 89 : 11 dr 83 : 7 dr 92 : 8 dr >96 : 4

Scheme 1.4 (a) Two-conformer model to explain the stereoselectivity in pyranosyl C-allylations. (b) Observed diastereoselectivity in reactions of (partially) substituted pyranosyl acetates (major products are shown).

Nu OBn OBn BnO OBn 1,2-cis BnO + 1,2-trans O O BnO + OBn

OBn Nu 42 43 3 4 H4 H3

3 4 Scheme 1.5 The H4 and H3 mannosyl oxocarbenium ions and the trajectories of incoming nucleophiles. substituents at C3 and C5, in addition to the destabilizing C3─C5 interaction, already 4 present in the system. Attack on the α-face of the H3 oxocarbenium ion, on the other hand, is devoid of these unfavorable steric interactions, making this transition state overall more favorable. With strong nucleophiles, the two-conformer oxocarbenium ion model falls short, and SN2-type pathways come into play [27, 28]. In a continuation of their efforts to understand the stereoselectivities of C-glycosylation reactions of (partially) substi- tuted pyranosyl donors, the Woerpel laboratory studied the addition reactions of a series of C-nucleophiles, ranging from weak nucleophiles (such as allyl trimethylsilane) 10 1 Stereoselective Glycosylations – Additions to Oxocarbenium Ions

to relatively strong nucleophiles (such as silyl ketene acetals) [27, 28]. Table 1.1 sum- marizes the stereochemical outcome of the reactions of 2-deoxy glucopyranosyl acetate donor 44 with these nucleophiles under the agency of TMSOTf as a Lewis acid catalyst [28], together with their relative nucleophilicity, as established by Mayr and coworkers [29]. The α-selectivity in the reaction with allyl trimethylsilane can be accounted for by 4 invoking the H3 oxocarbenium ion (55, Scheme 1.6a) as most likely product-forming 3 intermediate. Nucleophilic attack on the alternative H4 half chair 54 again suffers from prohibitively large steric interactions to be a reasonable pathway. With reactive nucleophiles, such as silyl ketene acetals 52 and 53 (Table 1.1, entries 4 and 5), the most likely product-forming pathway proceeds with significant SN2-character taking place on the α-triflate intermediate 56 (Scheme 1.6b) [28]. Of note, no attempts were under- taken to characterize this triflate[30].

1.3 Computational Studies

To better understand the conformational behavior, reactivity, and stability of glycosyl oxocarbenium ions, several quantum mechanical studies have been undertaken (see Table 1.2) [9, 10, 31–35]. Whitfield and coworkers have reported many computational

Table 1.1 Changing diastereoselectivity in the addition of C-nucleophiles of increasing reactivity.

Nuc, MeO TMSOTf, MeO O MeO DCM, 0 °C MeO O MeO MeO 44 OAc 45–49 Nu α/β Entry Nucleophile N product (yield)

89:11 1 TMS 1.8 45 (57%)

50:50 2 4.4 46 TMS (73%) 50

OTMS 68:32 3 6.2 47 (94%) Ph 51

OTMS 27:73 4 8.2 48 (78%) OPh 52 OTMS 19:81 5 9.0 49 (68%) OMe

53 1.3 Computational Studies 11

Nu Nu OMe OMe MeO MeO OMe O O + MeO MeO Nu O O MeO MeO + OMe MeO OTf OMe 54 Nu 55 56 57 3 4 H4 H3

(a)(βαb)

Scheme 1.6 Reactive intermediates in SN1-type (a) and SN2-type (b) pathways. studies in which they investigated the conformational behavior of, among others, tetra- O-methyl gluco- and mannopyranosyl triflates as well as their 4,6-O-benzylidene conge- ners upon ionization (i.e., expulsion of the triflate leaving group) and the conformational behavior of the resulting oxocarbenium ions [32]. To prevent collapse of the initially formed ion pair, they used lithium cations to stabilize the departing anionic leaving

Table 1.2 A selection of oxocarbenium ions and their calculated energies (determined by DFT calculations).

Hosoya Hosoya Whitfield CIPs SSIP

OMe tetra-O-Me Glc MeO OMe + H MeO O O OMe MeO OMe − + OMe MeO − MeO OMe TfO MeO + MeO O OMe OTf MeO O MeO + 2 64 (E ) H 63 ( H3) 3 58 (4H ) 71 (2S ) 3 −1 O +9.5 kcal mol −1 OMe − +8.5 kcal mol MeO OTf + OMe OMe MeO O OMe MeO O MeO O OMe MeO + MeO + MeO 4 65 (4H ) 72 ( H3) 4 3 59 ( E) +10.6 kcal mol−1 +10.6 kcal mol−1

tetra-O-Me Man MeO OMe + MeO O + OMe H MeO O MeO + OMe H O OMe OMe MeO O O OMe MeO − OMe MeO + MeO − OMe OMe OTf OMe MeO + TfO H OMe H H 4 3 O 4 O 60 ( H3) 66 ( H2) 67 ( S2) 73 ( H3) 74 ( S2) +7.3 kcal mol−1 +10.9 kcal mol−1 +11.1 kcal mol−1

Benzylidene-Glc Formylidene-Glc − O O OTf O + Ph O O + MeOO O + MeO O MeO O MeO MeO MeO 68 (4E) 75 (4E) −1 61 (4E) + 13.2 kcal mol + 10.9 kcal mol−1

Benzylidene-Man Formylidene-Man − OMe OMe MeO OTf MeO MeO Ph O O O O O O O O O O MeO O MeO O MeO O MeO + MeO O + + + O + H H − H H OTf H 62 (B ) 69 (B ) 70 (B ) 1 77 (B ) 2,5 2,5 2,5 76 ( S5) 2,5 +11.7 kcal mol−1 +14.3 kcal mol−1 +9.8 kcal mol−1 12 1 Stereoselective Glycosylations – Additions to Oxocarbenium Ions

group. These calculations revealed that ionization of the tetra-O-methyl gluco- and 4 mannopyranosyl α-triflates initially provides H3 (58 and 60, respectively) or closely 4 related E-like oxocarbenium ions 59 (see Stoddart’s hemisphere representation [36] for pseudo-rotational itineraries shown in Figure 1.2a). Expulsion of the anomeric triflate from the β-isomers requires a conformational change, where the glucose and mannose 1 pyranosyl rings distort to an S3-like structure [32]. In this constellation, the anomeric leaving group can be expelled by assistance of one of the ring oxygen lone pairs leading 4 4 to an E (for the glucose) or H3 half-chair (for the mannose) oxocarbenium ion. The stability of these ions is primarily governed by sterics, since they lack the electronic sta- bilization described earlier. Interestingly, similar itineraries have been established to be operational in glycosyl hydrolases. Rovira and coworkers have determined that the hydrolysis of β-glucosides by retaining glucosyl hydrolases, belonging to the GH5, GH7, and GH16 families, proceeds via a trajectory, in which the substrate is first placed in a conformation that allows expulsion of the aglycon (Figure 1.2b) [37, 38]. Then passing 4 1 through H3 transition state 80, which is close in conformational space to the starting S3 4 geometry 79, the C1 product 81 (the covalent enzyme–glucose adduct) is obtained. This catalytic itinerary was visualized using a combination of X-ray crystallography, free-energy landscape mapping (to determine the intrinsically favorable ground-state conformations), and quantum mechanics/molecular mechanics reaction simulations. Further calculations of the Whitfield group showed that the 4,6-O-benzylidene glucose

3 S1 3, oB B1,4

5 oS o S 2 o H1 1 E E1 o 2 H5 H1 B E 4 2 2,5 2,5 5 C1 E B 4H 2 5 H3

4 E 2 1S E 4 3 S 5 H3 o

1,4 B B 3,o (a) 1 S3

‡ − OOOOO O OO H − H + H HO RO HO OR HOR O HO OR OH HO H OH HO HO HO O HO O HO HO + HO HO O OH HO 78 − 79 − 80 81 4 O O 1 O O 4 − O O 4 O O C1 S3 H3 C1

(b)

Figure 1.2 (a) Stoddart’s hemisphere representation for conformational interconversions (only the Northern hemisphere is shown). (b) Conformational itinerary of the substrate as used by various β-glucosidases. The trajectory has been highlighted in the Stoddart diagram in (a) and was also calculated to be the lowest energy pathway of the ionization of a β-glycosyl triflate.