Modern Synthetic Methods in Carbohydrate Chemistry -  - E-Book

Modern Synthetic Methods in Carbohydrate Chemistry E-Book

0,0
160,99 €

-100%
Sammeln Sie Punkte in unserem Gutscheinprogramm und kaufen Sie E-Books und Hörbücher mit bis zu 100% Rabatt.
Mehr erfahren.
Beschreibung

The fields of glycochemistry and glycoscience are rich and varied and where much can be learned from Nature. As Nature is not always able to produce carbohydrates in quantities useful for not only in research but also as therapeutic agents, new ways need to be found to optimize the yield. This book presents an overview of the latest developments in the field of carbohydrates, ranging from de-novo approaches via cyclodextrin chemistry to the synthesis of such highly complex glycoconjugates as glycosphingolipids and GPI anchors.

The main emphasis remains on the synthetic aspects making the book an excellent source of information for those already involved in carbohydrate chemistry, as well as for those organic chemists who are beginners in this field. Equally of interest to synthetic chemists, as well as medicinal chemists and biochemists.

Sie lesen das E-Book in den Legimi-Apps auf:

Android
iOS
von Legimi
zertifizierten E-Readern

Seitenzahl: 550

Veröffentlichungsjahr: 2013

Bewertungen
0,0
0
0
0
0
0
Mehr Informationen
Mehr Informationen
Legimi prüft nicht, ob Rezensionen von Nutzern stammen, die den betreffenden Titel tatsächlich gekauft oder gelesen/gehört haben. Wir entfernen aber gefälschte Rezensionen.



Table of Contents

Related Titles

Title Page

Copyright

Foreword

Preface

List of Contributors

Chapter 1: De Novo Approaches to Monosaccharides and Complex Glycans

1.1 Introduction

1.2 De Novo Synthesis of Monosaccharides

1.3 Iterative Pd-Catalyzed Glycosylation and Bidirectional Postglycosylation

1.4 Synthesis of Monosaccharide Azasugar

1.5 Oligosaccharide Synthesis for Medicinal Chemistry

1.6 Conclusion and Outlook

1.7 Experimental Section

List of Abbreviations

Acknowledgments

References

Chapter 2: Synthetic Methodologies toward Aldoheptoses and Their Applications to the Synthesis of Biochemical Probes and LPS Fragments

2.1 Introduction

2.2 Methods to Construct the Heptose Skeleton

2.3 Synthesis of Heptosylated Oligosaccharides

2.4 Synthesis of Heptosides as Biochemical Probes

2.5 Conclusions

2.6 Experimental Part

List of Abbreviations

Acknowledgments

References

Chapter 3: Protecting-Group-Free Glycoconjugate Synthesis: Hydrazide and Oxyamine Derivatives in N-Glycoside Formation

3.1 Introduction

3.2 Glycosyl Hydrazides (1-(Glycosyl)-2-acylhydrazines)

3.3 O-Alkyl-N-Glycosyl Oxyamines

3.4 N,O-Alkyl-N-Glycosyl Oxyamines

3.5 Concluding Remarks and Unanswered Questions

3.6 Procedures

List of Abbreviations

Acknowledgment

References

Chapter 4: Recent Developments in the Construction of cis-Glycosidic Linkages

4.1 Introduction

4.2 Cis-Glycosylation

4.3 Conclusion

Acknowledgments

List of Abbreviations

References

Chapter 5: Stereocontrol of 1,2-cis-Glycosylation by Remote O-Acyl Protecting Groups

5.1 Introduction

5.2 Stereodirecting Influence of Acyl Groups at Axial and Equatorial O-3: Opposite Stereoselectivity Proves Anchimeric Assistance

5.3 Acyl Groups at O-4 in the galacto Series: Practical Synthesis of α-Glycosides: Complete Stereoselectivity

5.4 Lack of Stereocontrolling Effect of Acyl Groups at Equatorial O-4 in 4C1 Conformation

5.5 Effect of Substituents at O-6

5.6 Interplay of Stabilized Bicyclic Carbocation and Two H Conformations of Oxocarbenium Ions

5.7 Conclusion

5.8 Key Experimental Procedures

List of Abbreviations

References

Chapter 6: Synthesis of Aminoglycosides

6.1 Introduction

6.2 Amine-Protecting Group Strategies

6.3 Controlled Degradation of Aminoglycosides

6.4 Chemoselective Alcohol-Protecting Group Manipulations

6.5 Strategies for Glycosylation of Aminoglycoside Scaffolds

6.6 Synthesis of Amphiphilic Aminoglycosides

6.7 Chemoenzymatic Strategies for the Preparation of Aminoglycoside Analogs

6.8 Novel Synthetic Strategies to Overcome Resistance to Aminoglycosides

6.9 Conclusions and Future Perspectives

6.10 Selected Synthetic Procedures

Acknowledgments

List of Abbreviations

References

Chapter 7: Synthesis of Natural and Nonnatural Heparin Fragments: Optimizations and Applications toward Modulation of FGF2-Mediated FGFR Signaling

7.1 Introduction

7.2 Total Synthesis of Standard HPN Fragments

7.3 Total Synthesis of Modified HPN Fragments: Some Synthetic Clues

7.4 Alternative Synthetic Methods: Means to Build Libraries

7.5 Biological Evaluation

7.6 Conclusion and Outlook

7.7 Experimental Section (General Procedures)

Acknowledgments

List of Abbreviations

References

Chapter 8: Light Fluorous-Tag-Assisted Synthesis of Oligosaccharides

8.1 Introduction

8.2 Fluorous-Protecting Groups and Tags Amenable to Fluorous Solid-Phase Extraction in Carbohydrate Synthesis

8.3 Light Fluorous-Protecting Groups with Potential Use in Oligosaccharide Synthesis

8.4 “Cap-Tag” Strategies or Temporary Fluorous-Protecting Group Additions

8.5 Double-Tagging Carbohydrates with Fluorous-Protecting Groups

8.6 Other Advantages to Fluorous-Assisted Oligosaccharide Synthesis

8.7 Conclusions and Outlook

8.8 Experimental Section

Acknowledgments

List of Abbreviations

References

Chapter 9: Advances in Cyclodextrin Chemistry

9.1 Introduction

9.2 General Reactivity, Per- and Monofunctionalization

9.3 Capping Reagents for Direct Modification

9.4 Bulky Reagents for Direct Modifications

9.5 Selective Deprotections

9.6 Conclusion and Perspectives

9.7 Experimental Procedures

List of Abbreviations

References

Chapter 10: Design and Synthesis of GM1 Glycomimetics as Cholera Toxin Ligands

10.1 Introduction

10.2 Cholera Toxin and Its Specific Membrane Receptor, the GM1 Ganglioside

10.3 Rational Design of GM1-os Mimics as Cholera Toxin Inhibitors and Synthesis of First-Generation Ligands

10.4 Third Generation of GM1 Ganglioside Mimics: Toward Nonhydrolyzable Cholera Toxin Antagonists

10.5 Conclusions

10.6 Experimental Section

Acknowledgments

List of Abbreviations

References

Chapter 11: Novel Approaches to Complex Glycosphingolipids

11.1 Introduction

11.2 Syntheses of Complex Glycans of Gangliosides

11.3 Total Syntheses of Complex Gangliosides

11.4 Conclusion and Outlook

11.5 Experimental Section

List of Abbreviations

References

Chapter 12: Chemical Synthesis of GPI Anchors and GPI-Anchored Molecules

12.1 Introduction

12.2 Challenges in the Synthesis of GPIs

12.3 Tools for Synthesis of GPIs

12.4 Synthesis of GPIs with Linear Glycan Core

12.5 Synthesis of GPIs with Branched Glycan Core

12.6 GPI Derivatives for Biological Research

12.7 Synthesis of GPI-Anchored Peptides and Proteins

12.8 Conclusions and Outlook

Acknowledgments

List of Abbreviations

References

Index

Related Titles

Boysen, M.M.K. (ed.)

Carbohydrates - Tools for Stereoselective Synthesis

2013

ISBN: 978-3-527-32379-1

Wrolstad, R.E.

Food Carbohydrate Chemistry

Series: Institute of Food Technologists Series

2012

ISBN: 978-0-8138-2665-3

Wang, B., Boons, G.-J. (eds.)

Carbohydrate Recognition

Biological Problems, Methods, and Applications

2011

ISBN: 978-1-118-01756-2

Hanessian, S., Giroux, S., Merner, B.L

Design and Strategy in Organic Synthesis

From the Chiron Approach to Catalysis

2013

ISBN: 978-3-527-33391-2

The Editors

Prof. Dr. Daniel B. Werz

Technische Universität Braunschweig

Institut für Organische Chemie

Hagenring 30

38106 Braunschweig

Germany

Dr. Sébastien Vidal

Université Claude Bernard Lyon 1

ICMBS-UMR-CNRS 5246

43 Boulevard du 11 Nov. 1918

69622 Villeurbanne

France

All books published by Wiley-VCH are carefully produced. Nevertheless, authors, editors, and publisher do not warrant the information contained in these books, including this book, to be free of errors. Readers are advised to keep in mind that statements, data, illustrations, procedural details or other items may inadvertently be inaccurate.

Library of Congress Card No.: applied for

British Library Cataloguing-in-Publication Data

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>.

© 2014 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: 978-3-527-33284-7

ePDF ISBN: 978-3-527-65897-8

ePub ISBN: 978-3-527-65896-1

Mobi ISBN: 978-3-527-65895-4

oBook ISBN: 978-3-527-65894-7

Foreword

The ever-expanding field of glycoscience is driven by the vast diversity of biological processes mediated by carbohydrates, their oligomers, and conjugates, thereby providing enormous opportunities for the use of carbohydrates and their derivatives as research tools and as therapeutic agents. The difficulties in isolating homogeneous glycoforms from nature in any significant quantity and the limitations of such methods to naturally occurring glycoforms underline the need for efficient and effective chemical synthesis of such molecules. The superficially simple mechanism of the glycosidation reaction masks what is, in reality, a very difficult problem and one that drives the need for innovative chemical solutions, which in turn is attracting an increasing number of bright young (and not so young) researchers to the area. Against this background, Sébastien Vidal and Daniel B. Werz have assembled in this book an outstanding collection of work of talented authors to give their individual perspectives on “Modern Synthetic Methods in Carbohydrate Chemistry,” be they at the level of “simple” monosaccharides or complex glycosides. As the title suggests, oligosaccharide synthesis and carbohydrate chemistry, in general, are but one facet of modern organic chemistry. This is richly brought out in the content of the book from which it is clear that many recent advances in glycochemistry and indeed glycoscience depend very heavily on the power and ingenuity of contemporary synthetic organic methodology.

The first chapter, an authoritative contribution by George O'Doherty from Northeastern University in Boston, sets the tone by drawing the reader's attention to the fact that monosaccharides and complex glycans are not necessarily best prepared from actual sugars and that modern synthetic chemistry in the form of de novo synthetic approaches has an important role to play. The second chapter continues the theme as Stéphane Vincent from the University of Namur reminds us of the importance of the aldoheptoses and the need for their synthesis in terms of both biological probes and lipopolysaccharide (LPS) fragments. Mark Nitz from the University of Toronto then expounds on the protecting-group-free synthesis of glycoconjugates and in particular on the use of hydrazides and oxyamine derivatives for the synthesis of N-glycosides, before Alphert Christina, Gijs van der Marel, and Jeroen Codée from the Leiden University laboratory describe the recent evolution of the stereocontrolled synthesis of the 1,2-cis-glycosidic linkages, thereby nicely underlining the power of modern synthetic chemistry. Nikolay Nifantiev and his coauthors from the Zelinsky Institute of Organic Chemistry in Moscow continue with the theme of the stereocontrolled synthesis of the 1,2-cis-glycosides with particular emphasis on the role of participation by remote esters, a process for which the jury has yet to return a clear verdict. A chapter by Micha Fridman from Tel Aviv University then takes us into the realm of the aminoglycoside antibiotics and presents both useful chemistry and a perspective on novel applications of these compounds beyond their current use as antibacterials. Pierre-Alexandre Driguez at Sanofi in France takes us into the realm of complex oligosaccharide synthesis and tackles the vexing problems of the synthesis of heparin fragments, both natural and nonnatural, before Rajarshi Roychoudhury and Nicola Pohl from Indiana University set out the many advantages of light fluorous tag-assisted synthesis of oligosaccharides and the various strategies devised to date in order to take advantage of them. Samuel Guieu and Matthieu Sollogoub at the University of Aveiro and the Université Pierre and Marie Curie in Paris recount recent advances in cyclodextrin chemistry with an emphasis on selective functionalization methods, after which Anna Bernardi from the University of Milan takes up the theme of the design and synthesis of GM1 glycomimetics as cholera toxin ligands. This book continues with a chapter by Hiromune Ando and coworkers from Gifu University on recent advances in the synthesis of glycosphingolipids, from which much is to be learnt about the stereocontrolled synthesis of the once-difficult sialic acid glycosides. Finally, the volume closes with an important chapter from Daniel Varón Silva and his coauthors at the Max Planck Institute of Colloids and Interfaces in Berlin describing the need for, the many challenges in, and their numerous elegant solutions to, the synthesis of glycosylphosphatidylinositol (GPI) anchors and of GPI-anchored molecules.

The discerning reader will gain much from this volume not the least of which, hopefully, will be the recognition of the dynamic nature of the field of glycochemistry and of the many challenges still remaining. When viewed in the broader context, these many challenges are nothing more than problems in modern synthetic organic chemistry whose solution only awaits the arrival and ingenuity of fresh minds to the area.

Wayne State University

David Crich

November 2013

Preface

More than 100 years ago, pioneering achievements in carbohydrate chemistry were awarded the second Nobel Prize in Chemistry to Emil Fischer. Since that time this branch of organic chemistry has lost none of its fascination. The success story began with Fischer's brilliant logical deduction of the glucose structure (based on simple organic reactions performed with sugars) and was followed by basic glycosylation methods (such as the Fischer-Helferich or the Koenigs-Knorr procedures). Later, in 1937, another Nobel Prize was awarded for the investigation of disaccharides and vitamin C to Walter N. Haworth. After a long period of hibernation, carbohydrate chemistry has developed into an arsenal of highly sophisticated chemical methods, which have enabled the carbohydrate chemist to selectively build almost all of the possible glycosidic linkages. Since Nature has created an unbelievable variety of oligosaccharides and we can at least partially understand the biological significance of many of these structures today, it is very necessary to have a specialized chemical toolkit either to create mono- as well as oligosaccharides or to further modify any of these structures.

The present monograph comprises a fine selection of hot topics in carbohydrate chemistry that have found applications in biological studies and the preparation of complex natural glycans. The synthetic methodologies used for the preparation of these substrates have gathered modern skills from general organic chemistry to the more specific field of glycochemistry. Nevertheless, both organic chemistry and glycochemistry benefited from each other's experiences to push further the limits of the molecular architectures attainable.

In contrast to several other books dealing with carbohydrate chemistry, a collection of very recent synthetic developments in the field of carbohydrate chemistry are presented. Recent synthetic achievements including de novo synthesis of carbohydrates, highlights of cyclodextrin chemistry, synthesis of highly complex glycoconjugates such as glycosphingolipids and GPI anchors are treated, always with a strong focus on the synthetic aspects.

The idea for this book project came up at the European Young Investigators Workshop co-organized by us in April 2011 in Lyon (France). After this conference gathering young researchers in glycochemistry, we felt that putting together a book would help the scientific community to identify some key aspects of glycosciences in order to address the next scientific challenges in the post-genomics and post-proteomics era. Glycomics are now the next leap for scientists and this book is intended to bring together techniques from synthetic organic and carbohydrate chemistry so that each domain would benefit from each other.

This book was the collective work of a number of glycochemists. Most importantly, we would like to thank all contributors whose time, efforts, and expertise have made this book a useful scientific resource for beginners and advanced researchers both in organic chemistry and glycochemistry. We are grateful to Drs. Elke Maase and Lesley Belfit at Wiley-VCH for their help and useful advices in preparing this book.

Braunschweig, Germany

Daniel B. Werz

Lyon, France

Sébastien Vidal

November 2013

List of Contributors

Hiromune Ando
Gifu University
Department of Applied Bioorganic Chemistry
Faculty of Applied Biological Sciences
1-1 Yanagido
Gifu 501-1193
Japan
and
Kyoto University
Institute for Integrated Cell-Material Sciences (iCeMS)
Yoshida Ushinomiya-cho
Sakyo-ku
Kyoto 606-8501
Japan
Yifat Berkov-Zrihen
Tel Aviv University
School of Chemistry
Raymond and Beverly Sackler Faculty of Exact Sciences
Ramat Aviv
Tel Aviv 69978
Israel
Anna Bernardi
Università degli Studi di Milano
Dipartimento di Chimica
via Golgi 19
20133 Milano
Italy
Alphert E. Christina
Leiden University
Leiden Institute of Chemistry
Bio-organic Synthesis Group
Einsteinweg 55, P.O. Box 9502
2300 RA Leiden
The Netherlands
Melissa Cochran
University of Toronto
Department of Chemistry
Toronto
ON M5S 3H6
Canada
Jeroen D. C. Codée
Leiden University
Leiden Institute of Chemistry
Bio-organic Synthesis Group
Einsteinweg 55, P.O. Box 9502
2300 RA Leiden
The Netherlands
Michael F. Cuccarese
Northeastern University
Department of Chemistry and Chemical Biology
360 Huntington Ave.
Boston, MA 02115
USA
Pierre-Alexandre Driguez
Sanofi R&D
Early to Candidate Unit
195 Route d'Espagne
BP 13669
31036 Toulouse Cedex
France
and
1 Avenue Pierre Brossolette
91385 Chilly-Mazarin Cedex
France
Micha Fridman
Tel Aviv University
School of Chemistry
Raymond and Beverly Sackler Faculty of Exact Sciences
Ramat Aviv
Tel Aviv 69978
Israel
Sebastian Götze
Max Planck Institute of Colloids and Interfaces
Department of Biomolecular Systems
Am Mühlenberg 1
14476 Potsdam
Germany
and
Freie Universität Berlin
Institut für Chemie und Biochemie
Arnimallee 22
14195 Berlin
Germany
Samuel Guieu
University of Aveiro
CICECO and QOPNA
Department of Chemistry
Aveiro 3810-193
Portugal
Hideharu Ishida
Gifu University
Department of Applied Bioorganic Chemistry
Faculty of Applied Biological Sciences
1-1 Yanagido
Gifu 501-1193
Japan
Makoto Kiso
Gifu University
Department of Applied Bioorganic Chemistry
Faculty of Applied Biological Sciences
1-1 Yanagido
Gifu 501-1193
Japan
and
Kyoto University
Institute for Integrated Cell-Material Sciences (iCeMS)
Yoshida Ushinomiya-cho
Sakyo-ku
Kyoto 606-8501
Japan
Bozhena S. Komarova
N.D. Zelinsky Institute of Organic Chemistry
Russian Academy of Sciences
Laboratory of Glycoconjugate Chemistry
Leninsky prospect 47
Moscow 119991
Russia
Yoshiyuki A. Kwase
University of Toronto
Department of Chemistry
Toronto, ON M5S 3H6
Canada
Jiazhen J. Li
Northeastern University
Department of Chemistry and Chemical Biology
360 Huntington Ave.
Boston, MA 02115
USA
Gijsbert A. van der Marel
Leiden University
Leiden Institute of Chemistry
Bio-organic Synthesis Group
Einsteinweg 55, P.O. Box 9502
2300 RA Leiden
The Netherlands
Nikolay E. Nifantiev
N.D. Zelinsky Institute of Organic Chemistry
Russian Academy of Sciences
Laboratory of Glycoconjugate Chemistry
Leninsky prospect 47
Moscow 119991
Russia
Mark Nitz
University of Toronto
Department of Chemistry
Toronto, ON M5S 3H6
Canada
George A. O'Doherty
Northeastern University
Department of Chemistry and Chemical Biology
360 Huntington Ave.
Boston, MA 02115
USA
Rita Pal
Gifu University
Department of Applied Bioorganic Chemistry
Faculty of Applied Biological Sciences
1-1 Yanagido
Gifu 501-1193
Japan
and
Kyoto University
Institute for Integrated Cell-Material Sciences (iCeMS)
Yoshida Ushinomiya-cho
Sakyo-ku
Kyoto 606-8501
Japan
Nicola L. B. Pohl
Indiana University
Department of Chemistry
212 S. Hawthorne Drive
Bloomington
IN 47405
USA
José J. Reina
Università degli Studi di Milano
Dipartimento di Chimica
via Golgi 19
20133 Milano
Italy
Rajarshi Roychoudhury
Indiana University
Department of Chemistry
212 S. Hawthorne Drive
Bloomington
IN 47405
USA
Peter H. Seeberger
Max Planck Institute of Colloids and Interfaces
Department of Biomolecular Systems
Am Mühlenberg 1
14476 Potsdam
Germany
and
Freie Universität Berlin
Institut für Chemie und Biochemie
Arnimallee 22
14195 Berlin
Germany
Daniel Varón Silva
Max Planck Institute of Colloids and Interfaces
Department of Biomolecular Systems
Am Mühlenberg 1
14476 Potsdam
Germany
and
Freie Universität Berlin
Institut für Chemie und Biochemie
Arnimallee 22
14195 Berlin
Germany
Matthieu Sollogoub
UPMC Univ Paris 06
Sorbonne Université
Institut Universitaire de France
Institut Parisien de Chimie Moléculaire
UMR-CNRS 7201
C. 181, 4 place Jussieu
75005 Paris
France
Abdellatif Tikad
University of Namur (UNamur)
Chemistry Department
rue de Bruxelles 61
5000 Namur
Belgium
Yury E. Tsvetkov
N.D. Zelinsky Institute of Organic Chemistry
Russian Academy of Sciences
Laboratory of Glycoconjugate Chemistry
Leninsky prospect 47
Moscow 119991
Russia
Nadezhda E. Ustyuzhanina
N.D. Zelinsky Institute of Organic Chemistry
Russian Academy of Sciences
Laboratory of Glycoconjugate Chemistry
Leninsky prospect 47
Moscow 119991
Russia
Ivan Vilotijevic
Max Planck Institute of Colloids and Interfaces
Department of Biomolecular Systems
Am Mühlenberg 1
14476 Potsdam
Germany
and
Freie Universität Berlin
Institut für Chemie und Biochemie
Arnimallee 22
14195 Berlin
Germany
Stéphane P. Vincent
University of Namur (UNamur)
Chemistry Department
rue de Bruxelles 61
5000 Namur
Belgium

Chapter 1

De Novo Approaches to Monosaccharides and Complex Glycans

Michael F. Cuccarese, Jiazhen J. Li, and George A. O'Doherty

1.1 Introduction

Over the years, considerable effort has been made toward the development of new synthetic routes to monosaccharides [1]. This interest came primarily from the medicinal chemistry community, as these new routes often provided access to unnatural sugars, which could be of use in structure–activity relationship (SAR) studies. In addition, the synthesis of monosaccharides, and in particular hexoses, has served as a challenge and a measuring stick to the synthetic organic community. Of particular interest are the routes to hexoses that start from achiral starting materials, where asymmetric catalysis is used to install the stereochemistry. In the synthetic organic community, these routes are described as “de novo” or “de novo asymmetric” routes to carbohydrates, whereas in the carbohydrate community, the term de novo takes up other meanings. For the purposes of this review, the term de novo asymmetric synthesis refers to the use of catalysis for the asymmetric synthesis of carbohydrates from achiral compounds [2]. This then precludes the inclusion of de novo process that produced sugars from molecules with preexisting chiral centers (e.g., Seeberger and Reißig) [3, 4].

The challenge of a de novo synthetic approach to carbohydrates has been met by many groups (Scheme 1.1). These approaches begin most notably with the seminal work by Masamune and Sharpless [5] (2 to 5), which utilized iterative asymmetric epoxidation of allylic alcohols to prepare all eight possible hexoses. More recently, Danishefsky [6] demonstrated the power of asymmetric hetero-Diels–Alder reaction for the synthesis of several glycals (3 and 4 to 5), which inspired further studies toward oligosaccharide synthesis. Johnson and Hudlicky [7] turned to the use of enzyme catalysis for the oxidation/desymmetrization of substituted benzene rings to achieve hexopyranoses (1 to 5). Alternatively, Wong and Sharpless [8] used a combination of transition metal catalysis (asymmetric dihydroxylation) and an enzyme-catalyzed aldol reaction for the synthesis of several 2-keto-hexoses. More recently, this challenge has been engaged by MacMillan who utilized an iterative aldol reaction approach (a proline-catalyzed aldol followed by a subsequent diastereoselective aldol reaction) to produce various hexoses (6 to 5) [3]. Of these approaches, only the iterative epoxidation strategy of Masamune and Sharpless provides access to all eight hexoses, but it is also noteworthy that their route required the most steps and protecting groups. It is this latter point, the reduction of steps and avoidance of protecting groups, which guided the development of these synthetic endeavors [9].

Scheme 1.1De novo approaches to hexoses.

We have also developed two practical methods for the de novo synthesis of hexoses. These efforts have resulted in the discovery of two orthogonal approaches to hexopyranoses with variable C-6 substitutions. These approaches entail an iterative dihydroxylation strategy to hexose sugar lactones (9 to 5) [10] and an Achmatowicz strategy that is amenable to all eight hexose diastereomers (7 or 8 to 5) [11] Of the two approaches, the iterative asymmetric dihydroxylation of dienoates (Scheme 1.1) is the most efficient in terms of steps (one step for racemic to three steps for asymmetric) and the minimal use of protecting groups. On the contrary, the Achmatowicz approach is superior in terms of synthetic scope. The potential of this approach can be seen in the highly efficient de novo route to various mono-, di-, tri- tetra-, and heptasaccharide motifs, allowing their use for biological and medicinal structure–activity studies.

Of the many ways to compare these de novo routes (e.g., number of steps, availability of starting materials, and atom economy), clearly the best metric is the scope of its use in synthetic and biological applications. The Achmatowicz approach is distinguished from the other approaches when it comes to practical application to rare sugars, medicinal chemistry, and more specially oligosaccharides. These features result from its compatibility with the Pd-π-allyl-catalyzed glycosylation for the stereospecific formation of the glycosidic bond [12, 13]. As outlined in Scheme 1.2, the Pd(0)-catalyzed glycosylation reaction is both general and stereospecific [14, 15]. The reaction occurs rapidly and in high yields for both the α-10 to α-11 and β-10 to β-11 systems and works best when Pd2(dba)3·CHCl3 is used as the Pd(0) source with triphenylphospine as the ligand in a 1 : 2 Pd/PPh3 ratio. While carboxylate-leaving groups also work, the t-butoxycarbonate group (BocO−) is critical for the successful implementation of this reaction with alcohol nucleophiles. For example, when the t-butoxycarboxy group was replaced with a benzoyl or pivaloyl group, the palladium-catalyzed glycosylation reaction was significantly slower.

Scheme 1.2 Stereospecific Pd-catalyzed glycosylation.

The de novo Achmatowicz approach to hexoses has great potential for preparing various D- and L-sugars because the starting 6-t-butoxycarboxy-2H-pyran-3(6H)-ones (10 and 13) can easily be prepared from optically pure furfuryl alcohols 12 (either (R) or (S) enantiomer) [16] by a one or two step procedure (Scheme 1.3). Depending on the temperature of the second step, the t-butylcarbonate acylation reaction can selectively give the α-Boc pyranones 10α and 13α at −78 °C, whereas at room temperature, a 1 : 1 ratio of the α- and β-Boc protected enones were produced. Thus, this procedure can be used to prepare multigram quantities of both α- and β-pyranones in either enantiomeric form. When the Pd-glycosylation reaction was coupled with the Achmatowicz oxidation of a furan alcohol and diastereoselective t-butylcarbonate acylation reaction, a net three-step stereo-divergent pyranone-forming reaction resulted. Key to the success of the de novo asymmetric Achmatowicz approach is the practical access to all four possible pyranone diastereomers from either furan alcohol enantiomers 12(R) or 12(S).

Scheme 1.3 Achmatowicz approach to the simplest hexoses.

An important aspect of this approach is the ease with which furan alcohols can be prepared in enantiomerically pure form from achiral furans (e.g., 7 and 8). There are many asymmetric approaches to prepare furan alcohols. The two most prevalent approaches are (i) the Noyori reduction of acylfurans (8 to 12) and (ii) the Sharpless dihydroxylation of vinylfurans (7 to 12) (Scheme 1.4) [17]. Both routes are readily adapted to 100 g scale synthesis and use readily available reagents. While the Sharpless route is most amenable to the synthesis of hexoses with a C-6 hydroxy group, the Noyori route distinguishes itself in its flexibility to virtually any substitution at the C-6 position. Herein, we review the development of the Achmatowicz approach to the de novo synthesis of carbohydrates, with application to oligosaccharide assembly and medicinal chemistry studies.

Scheme 1.4De novo asymmetric approaches to chiral furan alcohols.

1.2 De Novo Synthesis of Monosaccharides

Putting this together, a very practical de novo approach to manno-hexoses can be carried out in six steps from achiral acylfuran 15. The route began with a three-step synthesis of Boc-pyranone 18 in an ∼50% overall yield. Thus, in only three highly diastereoselective steps, pyranone 18 was converted into manno-pyranose 21 in 45% overall yield. The three-step sequence consists of a Pd-catalyzed glycosylation (18 to 19) and two postglycosylation reactions, a Luche reduction (NaBH4/CeCl3, 19 to 20) and Upjohn dihydroxylation (OsO4(cat)/NMO, 20 to 21) (NMO, N-methylmorpholine-N-oxide) (Scheme 1.5). This six-step sequence of achiral 15 to partially protected mannose 21 demonstrates the power of asymmetric synthesis. From the point of view of carbohydrate synthesis, this route consists of a three-step asymmetric synthesis of glycosyl donor 18, which in three additional steps is converted into manno-sugar 21. Viewing this approach from a synthetic perspective becomes more relevant in the subsequent schemes. Other hexose congeners were also prepared using other postglycosylation transformations [18]. Critical to the success of this approach is the chemoselective use of functionality of C–C and C–O π-bonds as atom-less protecting groups (i.e., enone as a protected triol) as well as an anomeric-directing group (via a Pd-π-allyl).

Scheme 1.5De novo asymmetric Achmatowicz approach to L-manno-hexoses.

1.3 Iterative Pd-Catalyzed Glycosylation and Bidirectional Postglycosylation

The synthetic efficiency of the approach reveals itself when the Pd(0)-catalyzed glycosylation was applied in an iterative manner for oligosaccharide synthesis (Scheme 1.6 and Scheme 1.7) [19]. The step savings occurred because of the bidirectional use of postglycosylation Luche and Upjohn reactions. While not always shorter, these routes compare favorably with more traditional carbohydrate approaches and offer exclusive access to enantiomers as well as D-/L-sugar diastereomers. In addition to reducing steps, this highly atom-economical approach avoids the extensive use of protection/deprotection steps. For example, the 1,6-manno-trisaccharide 23 was prepared from enone 18 in six steps (nine from achiral furan 15). The synthesis was accomplished by an iterative use of a t-butyldimethylsilyl (TBS)-deprotection/glycosylation strategy to prepare trisaccharide 22, followed by a tris-reduction and tris-dihydroxylation to install the manno-stereochemistry. By simply switching the order of the reduction and glycosylation steps, this route can also be used to prepare 1,4-manno-trisaccharide 23. Key to the success of this sequence was the highly stereoselective reduction and dihydroxylation reaction, which installed six stereocenters in one transformation (25 to 26). This approach was successfully used in the medicinal chemistry SAR study of digitoxin, an anticancer agent [20].

Scheme 1.6De novo asymmetric approach to 1,6- and 1,4-oligo-hexoses.

Scheme 1.7De novo asymmetric approach to branched 1,4- and 1,6-oligo-L-hexoses.

1.3.1 Bidirectional Iterative Pd-Catalyzed Glycosylation and Postglycosylation

The synthetic efficiency was magnified when the glycosylation reaction was also applied in a bidirectional manner [21]. For example, when the TBS group of pyran 20 was removed, the resulting diol 27 can be bis-glycosylated to form tris-pyran 28. This bidirectional application of ketone reduction and TBS deprotection gave tetraol 29. Once again, the tetraol of 29 can be per-glycosylated with excess pyranone 18 to give heptasaccharide 30. Finally, a ketone per-reduction and double bond per-dihydroxylation gave heptasaccharide 31. It is worth noting, while there is similar local symmetry around each alkene, they exist in different stereochemical environments.

1.3.2 Synthesis of Monosaccharide Aminosugar Library

While these bidirectional approaches do have some significant synthetic efficiency over traditional approaches, they do suffer from the fact that the routes tend to be linear in nature and do not readily adapt to convergent synthesis. On the contrary, these routes most readily adapt to divergent synthesis. In particular to divergent synthesis, as it is applied to the synthesis of unnatural sugars. This becomes particularly advantageous when it is being used to address problems associated with medicinal chemistry. An example of this application can be seen in our application of this de novo Achmatowicz approach for the synthesis of a library of glycosylated methymycin analogs for eventual medicinal chemistry SAR studies [22].

Methymycin is one of the several 12-membered ring-glycosylated macrolide antibiotics isolated from Streptomyces venezuelae ATCC 15439 (ATCC, American Type Culture Collection; Scheme 1.8). Similar to other macrolide antibiotics, the rare deoxy-aminosugar portion of methymycin (desosamine) is important for their bioactivity. Thus, its modifications hold promise as a valuable approach toward preparing new macrolide antibiotics with improved and/or altered biological properties. Our de novo approach to a library of methymycin analogs is retrosynthetically outlined in Scheme 1.8, where the macrolide aglycon 10-deoxymethynolide was glycosylated in a stereo-divergent manner (with D- or L-Boc pyranones 44 or ent-44, respectively) to give either α-D-glycoside 45 or its diastereomer α-L-glycoside 46. Subsequent postglycosylation transformations were used to provide various sugar congeners and stereoisomers, in particular, unnatural deoxy-aminosugar isomers.

Scheme 1.8De novo synthesis of a methymycin monosaccharide library.

The installation of amino-functional groups, in practice, was most easily accomplished at the C-4 position (Scheme 1.9). For instance, the α-D-pyranone ring on methymycin analog 45 could be converted into a 4-aminosugar 41 with α-D-rhodino-stereochemistry in four steps, via a reduction, activation of the resulting alcohol, azide inversion, and reduction strategy. Alternatively, this approach is also compatible with the installation of equatorial amino groups at the C-4 position. This is accomplished by means of a net retention of stereochemistry in the substitution reaction at the C-4 position. The reaction occurred via a Pd-catalyzed π-allyl reaction with trimethylsilyl azide (TMSN3) as the nucleophile to give allylic azide 47 from α-D methymycin analog 46. In turn, azide 47 could be converted into azido-/azasugar methymycin analogs with α-L-rhamno- (36 and 37) and α-L-amiceto-stereochemistries (34 and 35).

Scheme 1.9De novo synthesis of a methymycin monosaccharide aminosugar library.

1.4 Synthesis of Monosaccharide Azasugar

Among the polyhydroxylated indolizidine alkaloids, the most well-known member is D-swainsonine 60 (Scheme 1.10) [23–25]. Swainsonine is known as a potent inhibitor of both lysosomal α-mannosidase [26] and mannosidase II [27] and has shown promise as an anticancer agent [28]. Fleet et al. [29] have shown that the enantiomer (L-swainsonine) selectively inhibited narginase (L-rhamnosidase, = 0.45 μM), whereas the D-swainsonine showed no inhibitory activity toward this enzyme. Owing to the biological importance of both D- and L-swainsonines, several epimers and analogs have become attractive targets for syntheses [30, 31].

Scheme 1.10De novo synthesis of D-swainsonine.

Thus, we became interested in the de novo asymmetric synthesis of both enantiomers of swainsonine and various epimers for further biological studies [32]. We envisioned a similar postglycosylation transformation, in which the installation of the 4-amino-manno-stereochemistry in methymycin analog 47 could be instrumental in the development of a de novo Achmatowicz approach to the indolizidine natural product swainsonine 60 (Scheme 1.10). In practice, this required access to acylfuran 49, which was prepared in two steps from 2-lithiofuran and butyrolactone 48. After Noyori reduction, Achmatowicz reaction, and diastereoselective acylation, 49 was converted into pyranone 52. Glycosylation of benzyl alcohol with 52 installed the required C-1-protecting group. As before, Luche reduction, carbonate formation, and Pd-π-allyl allylic azide displacement installed the C-4 azido group in 56. Before the C-2/C-3 double bond was dihydroxylated, the TBS ether is converted into a mesylate-leaving group as in 58. Finally, a diastereoselective dihydroxylation (58 to 59) and exhaustive hydrogenolysis (59 to 60) cleanly provided D-swainsonine in an optically pure form. Thus, in only 13 steps, either D- or L-swainsonine can be prepared from achiral starting material. This is of particular note because both enantiomers have been valued as known inhibitors of glycosidase enzymes (D-swainsonine for α-D-mannosidases and α-L-swainsonine for α-L-rhamnosidases). This route has also been used to prepare various diastereoisomers of swainsonine, which are also known to be effective glycosidase inhibitors [32].

1.5 Oligosaccharide Synthesis for Medicinal Chemistry

As part of the continuing search for new antibiotics against bacterial resistance [33], the cyclic hexapeptide mannopeptimycin-ϵ 61 was isolated from the fermentation broths of Streptomyces hygroscopicus LL-AC98 and related mutant strains [34]. The key structural features of the mannopeptimycins are a cyclic hexapeptide core with alternating D- and L-amino acids, three of which are rare. Two of the amino acids (β-D-hydroxyenuricididine and D-tyrosine) are glycosylated with mannose sugars. The glycosylated amino acids are an N-glycosylated β-hydroxyenuricididine with an α-mannose, and an O-glycosylated tyrosine with an α-(1,4-linked)-bis-manno-pyranose disaccharide portion. The unique structure and unprecedented biological activity have inspired both biological [35] and synthetic studies from us and others.

Our de novo asymmetric Achmatowicz approach was also applied to the synthesis of the glycosylated tyrosine portion of the antibiotic mannopeptimycin-ϵ [36]. Specifically, we targeted a protected tyrosine with bis-manno-1,4-disaccharide with an isovalerate at the C-4′ position. This approach was used to prepare the amino acid portion of the natural product 62 as well as the disaccharide portion in the unnatural L/L-configuration. The linear route involved the application of the iterative bis-glycosylation, acylation, and bis-dihydroxylation of protected tyrosine in only six steps (Scheme 1.11).

Scheme 1.11 Assembly of the mannopeptimycin-ϵ disaccharide.

As part of further SAR studies of the antibiotic mannopeptimycin, access to the C-4 amide analogs 42 was also desired [37]. As before, this was also accomplished using the Pd-π-allyl-catalyzed allylic azide alkylation in conjunction with azide reduction and acylation. Specifically, allylic alcohol 66 was converted into methyl carbonate 67, and the allylic carbonate was converted into allylic azide 68. In turn, the azide was selectively reduced and the corresponding amine 69 was acylated. The remaining double bonds of 70 were dihydroxylated, and the TBS groups were removed to provide the desired target tyrosine disaccharide 71 (Scheme 1.12). Key to the success of this approach was the somewhat surprisingly selective ionization of the equatorial allylic carbonate in 67 without any sign of ionizing the anomeric phenol in the axial configuration.

Scheme 1.12 Elaboration to C-4″ aza-mannopeptidomycin.

1.5.1 Tri- and Tetrasaccharide Library Syntheses of Natural Product

As part of a high-throughput-based search for new natural products with unique structures and interesting biological activity, two partially acetylated trisaccharide (cleistrioside-5/-6, 72 and 73) and six partially acetylated tetrasaccharide (cleistetroside-2/-3/-4/-5/-6/-7, 74–81) natural products were discovered (Figure 1.1) from the leaves and bark of trees with a folk medicinal tradition [38–40]. These dodecanyl tri- and tetrarhamnoside structures with various degrees of acylation were isolated from Cleistopholis patens and Cleistopholis glauca [36, 37]. The structures of the cleistriosides and cleistetrosides were assigned by detailed NMR analysis and later confirmed by total synthesis by us [41] and others [42, 43]. In addition to clarifying the structural issues, our synthetic interest in these oligosaccharides was aimed at supplying sufficient material for SAR-type studies and as a test of our synthetic methodology [44].

Figure 1.1 The cleistriosides, cleistetrosides, and analogs.

The route began with pyranone ent-44, which was easily prepared in three steps from commercially available acetylfuran (Scheme 1.13). In four steps (glycosylation, reduction, dihydroxylation, and acetonide protection), glycosyl donor ent-44 was converted into protected rhamnose 83. A subsequent glycosylation, reduction, acylation, and dihydroxylation gave diol disaccharide 84. Unfortunately, when diol 84 was exposed to our typical Pd-catalyzed glycosylation conditions, the trisaccharide 85 with the wrong regiochemistry was formed. To our delight, this substrate regioselectivity (4 : 1) could be reversed via the formation of tin acetal 86. Thus, a tin-directed regioselective (7 : 1) glycosylation gave trisaccharide 87 with the required carbohydrate at the C-3 position. A subsequent chloro-acylation gave trisaccharide 88, which was ready for further elaboration into both the cleistrioside and cleistetroside natural products (Scheme 1.14, Scheme 1.16, and Scheme 1.17).

Scheme 1.13 Synthesis of trisaccharide 88.

Scheme 1.14 Synthesis of cleistrioside-5 and -6.

The pyranone ring in trisaccharide 88 is perfectly situated for further elaboration into a rhamnose ring with the desired acylation pattern for cleistrioside-5 and -6. This was accomplished by Luche reduction and acylation or chloro-acylation to give 89 and 90. A subsequent dihydroxylation and ortho-ester-mediated C-2 acylation gave trisaccharides 91 and 92, which could be converted into cleistrioside-5 and -6 (72 and 73) by a selective removal of the chloro-acetates over the other acetates (removed with thiourea) and acetonide-protecting groups (hydrolyzed with AcOH/H2O).

In addition, the trisaccharides 91 and 92 can be further converted into eight members of the cleistetroside family of natural products (Scheme 1.16 and Scheme 1.17). This divergent route began with the selective Pd-catalyzed glycosylation of the free C-3 alcohol in 91 and 92 and a subsequent Luche reduction to afford tetrasaccharides 93 and 94 (Scheme 1.15).

Scheme 1.15 Synthesis of tetrasaccharides 93 and 94.

As outlined in Scheme 1.16, tetrasaccharide 93 was divergently converted into five of the six desired cleistetroside-2, -3, -4, -6, and -7. This was accomplished in a range of three to five steps by subtle variations of the postglycosylation reactions (dihydroxylation, acylation, and chloro-acylation), followed by global deprotection (thiourea then AcOH/H2O).

By applying the same postglycosylation/deprotection reaction sequence on tetrasaccharide 94, the remaining cleistetroside-5 was prepared in three steps. In addition, the revised route was also used to prepare two previously unknown analogs, cleistetroside-9 and -10. This was accomplished in a range of one to three steps by modular application of the postglycosylation reactions (dihydroxylation, acylation, and chloro-acylation) (Scheme 1.17).

Scheme 1.16 Synthesis of cleistetroside-2, -3, -4, -6, and -7.

Scheme 1.17 Synthesis of cleistetroside-5, -8, and -10.

It is worth noting that the route to any of the cleistetrosides is quite comparable to the two previously reported routes to cleistetroside-2 in terms of total number of steps. What distinguished it from these more traditional routes is its flexibility to diverge to any of the possible natural product isomers. Thus, we have found that this divergent approach is particularly amenable to medicinal chemistry studies. In this regard, our synthetic access to these eight natural products (two trisaccharides and six tetrasaccharides) and additional two natural product analogs enabled detailed medicinal chemistry SAR studies. The divergent nature of the approach is graphically displayed in Scheme 1.18, where in only 13 steps and in 20% overall yield, the key trisaccharide 88 could be prepared from achiral furan 95. Trisaccharide 88 serves as the linchpin molecule that can, in 6–11 steps, be converted into any of the desired natural products, in sufficient quantities for further studies.

Scheme 1.18 Divergent synthesis retrosynthetic summary.

1.5.2 Anthrax Tetrasaccharide Synthesis

Possibly the most elaborate application of this de novo Achmatowicz approach to oligosaccharides was the synthesis of the anthrax tetrasaccharide 100. The approach to this tetrasaccharide natural product merged well with our efforts to C-4 aminosugars with the synthesis of rhamnose-containing oligosaccharides (Scheme 1.19) [45, 46]. Anthrax is a zoonotic disease caused by the spore-forming bacterium Bacillus anthracis [47]. In an effort to find a unique structural motif associated with the bacterium, the anthrax tetrasaccharide 100 was discovered. The tetrasaccharide 100 consists of three L-rhamnose sugars and a rare sugar, D-anthrose [48]. The uniqueness of the D-anthrose sugar and the resistance of carbohydrate structures to evolutionary change make the anthrax tetrasaccharide an interesting target for synthesis [49]. Several carbohydrate approaches to the anthrax tetrasaccharide and one to a related trisaccharide have been reported [49a, 50], which derive their stereochemistry from the known but less common sugar L-rhamnose and the rare D-fucose. Our de novo approach to the tetrasaccharide 100 was envisioned as occurring through a traditional glycosylation between trisaccharide 96 with

Scheme 1.19 Retrosynthesis of anthrax tetrasaccharide 100.

trichloroacetimidate 97 (Scheme 1.19). In turn, our de novo approach was planned to prepare both of these fragments (96 and 97) from the achiral acetylfuran 95, which it is worth noting are significantly less expensive than either L-rhamnose or D-fucose.

Our synthesis of the anthrose monosaccharide 97 is described in Scheme 1.20 and involved two Pd-π-allylation reactions. The route began with the synthesis of (p-methoxybenzyl) PMB-protected pyranone 101 from 95 via ent-44. Using the Pd-catalyzed C-4 allylic azide chemistry previously described, pyranone ent-44 was converted into allylic azide 104 and dihydroxylated to give rhamno-sugar 105. The 6-deoxy-gluco-stereochemistry is then installed by a protection/C-2 inversion strategy to give anthrose sugar 108. Finally, a Lev-protection, PMB-deprotection strategy, and trichloroacetimidate formation were used to convert 108 into the glycosyl donor sugar 97 (14 steps from achiral acetylfuran 95).

Scheme 1.20De novo synthesis of a D-anthrose sugar trichloroacetimidate.

The de novo Achmatowicz approach to the tris-rhamno portion of the anthrax tetrasaccharide began with the synthesis of disaccharide 115 from pyranone ent-44 and benzyl alcohol (Scheme 1.21). After glycosylation and postglycosylation transformations to install the rhamno-stereochemistry (ent-44 to 111), the 1,2-trans-diol of 111 was then protected with the Ley-spiroketal to provide monosaccharide 112 with a free C-2 hydroxyl group. After a similar three-step glycosylation (112 and 113) and postglycosylation sequence, 113 was converted into disaccharide 114, which in a one-pot ortho-ester protocol was protected to give disaccharide 115 with a free C-3 alcohol.

Scheme 1.21 Synthesis of disaccharide 115.

Simply repeating the same three-step glycosylation and postglycosylation sequence converted disaccharide 115 into trisaccharide 116 (Scheme 1.22). Once again the one-pot ortho-ester formation/acylation/hydrolysis sequence gave trisaccharide 117 with the free C-3 alcohol, ready for glycosylation with an anthrose sugar fragment. Unfortunately, any attempt at glycosylation of trisaccharide 117 failed because of the instability of the Ley-spiroketal to the Lewis acidic nature of

Scheme 1.22 Synthesis of tetrasaccharide 119.

the traditional glycosylation conditions. Undaunted, we turned to an alternative protecting group strategy. Thus, the C-3 hydroxyl group of 117 was protected as a levulinate ester, and the Ley-spiroketal-protecting group was removed to form 118. Then, the anthrose sugar was installed by an acylation, selective levulinate deprotection (using hydrazine), and glycosylation with anthrose monosaccharide 97 delivering the corresponding tetrasaccharide 119.

Finally, we turned to the deprotection of tetrasaccharide 119 into anthrax tetrasaccharide 100. Deprotection of levulinate-protecting groups followed by an etherification (MeI/Ag2O) delivered the methyl ether 120. A one-pot condition was employed to reduce and acylate azide 120 along with global deprotection of the acetate groups to generate the free alcohol (PEt3/LiOH/H2O), which upon selective peptide coupling of primary amine and 3-hydroxy-3-methylbutanoic acid (HBTU/Et3N) (HBTU, O-(benzotriazol-1-yl)-N,N,N′,N′-tetramethyluronium hexafluorophosphate) afforded amide 121. Removal of the benzyl groups in 121 under hydrogenolysis conditions provided the natural product anthrax tetrasaccharide 100 (Scheme 1.23).

Scheme 1.23 Synthesis of anthrax tetrasaccharide 100.

1.6 Conclusion and Outlook

For the last 25 years, various groups have been investigating the use of asymmetric catalysis for the synthesis of hexoses. When beginning with achiral starting materials and when asymmetric catalysis is used for the installation of asymmetry, these syntheses are called “de novo asymmetric” or “de novo” for short. While these de novo approaches have been quite impressive in terms of the scope of products prepared and the brevity of steps, they were lacking in terms of application to oligosaccharide synthesis. As part of these efforts, we have developed two orthogonal de novo asymmetric approaches to hexoses: an iterative dihydroxylation strategy and an Achmatowicz strategy. Owing to its compatibility with a Pd-catalyzed glycosylation reaction, this later Achmatowicz de novo approach to hexoses has seen significant application for the assembly of oligosaccharides.

Of particular note is the flexibility of the de novo Achmatowicz route to a myriad of carbohydrate motifs. The approach couples asymmetric catalysis with the Achmatowicz rearrangement for the synthesis of D- and L-pyranones and utilizes highly diastereoselective glycosylation and postglycosylation reactions for the assembly of oligosaccharides. Heavily featured in the glycosylation and postglycosylation reactions was the Pd-π-allyl-catalyzed allylic substitution reaction for substitution at both the anomeric C-1 and C-4 positions of the hexose. What made these allylic substitution reactions so powerful is the one-step double inversion mechanism of the Pd-catalyzed reaction, which in a highly stereospecific manner reliably provides products with net retention of stereochemistry.

Whether these approaches were used in linear and/or bidirectional manner, highly efficient syntheses of natural and unnatural mono-, di-, and oligosaccharides resulted. The overall efficiency of these approaches was the result of the strategic use of the enone functionality in the pyranone as atom-less protecting groups. While the use of atom-less protecting groups has been under-explored in carbohydrate chemistry, the success of the above-mentioned approaches suggest that this strategy should be given more attention from the synthetic organic chemistry community. The overall practicality of these syntheses can be seen in their ability to provide material for medicinal chemistry studies, often in ways that are not readily available by traditional carbohydrate syntheses [51].

1.7 Experimental Section

(R)-(+)-1-(2-furyl)ethanol: To a 250 ml flask, 2-acetylfuran (22.0 g, 0.20 mol), CH2Cl2 (100 ml), formic acid/triethylamine (1 : 1, 108 ml), and Noyori asymmetric transfer hydrogenation catalyst (R)-Ru(η6-mesitylene)-(R,R)-TsDPEN (585 mg, 0.95 mmol) were added. The resulting solution was stirred at room temperature for 24 h. The reaction mixture was diluted with water (150 ml) and extracted with Et2O (3 × 150 ml). The combined organic layers were washed with saturated aqueous NaHCO3, dried over Na2SO4, and concentrated under reduced pressure. The crude product was purified by silica gel flash chromatography and eluted with 20% Et2O/hexanes to give (R)-(+)-1-(2-furyl)ethanol (21.5 g, 0.19 mmol, 96%) as a clear liquid.
6-Hydroxy-2-methyl-(2R)-2H-pyran-3(6H)-one: (R)-(+)-1-(2-furyl)ethanol (8.9 g, 79.5 mmol): Tetrahydrofuran (THF) (100 ml) and H2O (25 ml) were added to a round bottom flask and cooled to 0 °C. Solid NaHCO3 (13.4 g, 159 mmol), NaOAc·3H2O (10.8 g, 79.5 mmol) and N-bromosuccinimide (NBS) (14.9 g, 83.5 mmol) were added to the solution and the mixture was stirred at 0 °C for 1 h. The reaction was quenched with saturated aqueous NaHCO3 (200 ml), extracted with Et2O (3 × 200 ml), dried over Na2SO4, and concentrated under reduced pressure. The crude product was purified by silica gel flash chromatography and eluted with 35% EtOAc/hexanes to give 6-hydroxy-2-methyl-(2R)-2H-pyran-3(6H)-one (9.67 g, 75.5 mmol, 95%, α:β = 2.6 : 1) as a clear liquid.
Carbonic acid, (2R,6R)-5,6-dihydro-6-methyl-5-oxo-2H-pyran-2-yl-1,1-dimethylethyl ester (ent-44): 6-Hydroxy-2-methyl-(2R)-2H-pyran-3(6H)-one (3.70 g, 28.9 mmol) was dissolved in CH2Cl2 (15 ml) and the solution was cooled to −78 °C. A CH2Cl2 (15 ml) solution of (Boc)2O (9.46 g, 43.4 mmol), and a catalytic amount of dimethylaminopyridine (DMAP) (350 mg, 2.89 mmol) was added to the reaction mixture. The reaction was stirred at −78 °C for 5 h. The reaction was quenched with saturated aqueous NaHCO3 (50 ml), extracted with Et2O (3 × 50 ml), dried over Na2SO4, and concentrated under reduced pressure. The crude product was purified by silica gel flash chromatography and eluted with 7% EtOAc/hexanes to give two diastereomers of tert-butyl((6R)-6-methyl-5-oxo-5,6-dihydro-2H-pyran-2-yl) carbonate (5.72 g, 25.1 mmol, 87%) in 3 : 1 (α:β) ratio ent-44.
(2R,6S)-2H-6-[(4-Methoxyphenyl)methoxy]-2-methyl-pyran-3(6H)-one (101): To a solution of Boc-protected pyranone, ent-44 (13 g, 57.0 mmol), and para-methoxy benzyl alcohol (157.3 g, 114.0 mmol) in dry CH2Cl2 (57 ml), Pd2(dba)3·CHCl3 (294 mg, 1 mol% Pd) and PPh3 (297 mg, 2.0 mol%) at 0 °C were added under argon atmosphere. After stirring for 2 h, the solution was warmed to room temperature, the reaction mixture was quenched with 300 ml of saturated NaHCO3, extracted (3 × 300 ml) with Et2O, dried (Na2SO4), and concentrated under reduced pressure. The crude product was purified using silica gel flash chromatography and eluted with 5% EtOAc/hexane to give PMB ether 101 (13.6 g, 54.7 mmol, 96%) as a colorless oil.
(2R,3S,6S)-2H-3,6-Dihydro-6-[(4-methoxyphenyl)methoxy]-2-methyl-pyran-3-ol (102): A solution of pyranone 101 (13.5 g, 54.4 mmol) in dry CH2Cl2 (54.4 ml) and 0.4 M CeCl3/MeOH (54.4 ml) was cooled to −78 °C. NaBH4 (2.08 g, 55.5 mmol) was added, and the reaction mixture was stirred for 4 h at −78 °C. The resulting solution was diluted with Et2O (400 ml) and was quenched with 200 ml of saturated NaHCO3, extracted (3 × 400 ml) with Et2O, dried (Na2SO4), and concentrated under reduced pressure. The crude product was purified using silica gel chromatography and eluted with 40% EtOAc/hexane to give 12.6 g (50.6 mmol, 93%) of allylic alcohol 102 as a white solid.
Carbonic acid, (2R,3S,6S)-3,6-Dihydro-6-[(4-methoxyphenyl)methoxy]-2-methyl-2H-pyran-3-yl methyl ester (103): To a stirred solution of allylic alcohol 102 (20 g, 80 mmol), pyridine (38.8 ml, 480 mmol), and DMAP (1.96 g) in dry CH2Cl2 (400 ml), methyl chloroformate was added dropwise (33.9 ml, 480 mmol) at 0 °C. After reacting for 1 h at 0 °C, water (300 ml) was added and the reacted mixture was extracted with CH2Cl2 (3 × 400 ml), dried (Na2SO4), and concentrated under reduced pressure. The crude product was purified using silica gel flash chromatography and eluted with 10% EtOAc/hexane to give 24.4 g (79.2 mmol, 99%) of carbonate 103 as colorless oil.
(2R,3S,6S)-2H-3-Azido-3,6-dihydro-6-[(4-methoxyphenyl)methoxy]-2-methyl-pyran (104): To a stirred solution of carbonate 103 (30 g, 97.4 mmol), allylpalladium chloride dimer (378 mg, 1.0 mmol%), and 1,4-bis(diphenylphosphino)butane (1.68 g, 4.0 mmol%) in dry THF (97.2 ml) TMSN3 (15.5 ml, 116.9 mmol) was added under argon atmosphere. The reaction mixture was stirred at room temperature for 0.5 h, the solvent was evaporated under reduced pressure and purified using silica gel flash chromatography, and eluted with 7% EtOAc/hexane to obtain 24.9 g (90.6 mmol, 93%) allylic azide 104 as colorless oil.
(4-Methoxyphenyl)methyl-4-azido-4,6-dideoxy-α-D-mannopyranoside (105): To a mixture of t-butanol, acetone (145.4 ml, 1 : 1 (v/v), 1 M) and solution of allylic azide 104 (20 g, 72.7 mmol) at 0 °C, a solution of N-methyl morpholine N-oxide/water (50% w/v, 50 ml) was added. Crystalline OsO4 (185 mg, 1 mol%) was added and the reaction mixture was allowed to stir for 24 h. The reaction mixture was quenched with 200 ml saturated Na2S2O3 solution, extracted with EtOAc (3 × 500 ml), dried with (Na2SO4), concentrated under reduced pressure, and then purified using silica gel flash chromatography, eluting with 90% EtOAc/hexane to give diol 105 (22.0 g, 71.2 mmol, 98%).

List of Abbreviations

Ac

Acetyl

AD-mix-α

3 equiv K

3

Fe(CN)

6

/K

2

CO

3

, 1 equiv MeSO

2

NH

2

, 5% OsO

4

,

and 6% (DHQ)

2

PHAL

AD-mix-β

3 equiv K

3

Fe(CN)

6

/K

2

CO

3

, 1 equiv MeSO

2

NH

2

, 5% OsO

4

,

and 6% (DHQD)

2

PHAL

ATCC

American Type Culture Collection

Bn

Benzyl

Boc

t

-Butoxycarbonyl

Bu

Butyl

Cbz

Carbobenzoxy

ClAc

Chloroacetyl

CTAB

Cetyltrimethylammonium bromide

DCC

Dicyclohexylcarbodiimide

DDQ

2,3-Dichloro-5,6-dicyano-1,4-benzoquinone

DMAP

Dimethylaminopyridine

DMP

Dess–Martin periodinane

dppb

1,4-Bis(diphenylphosphino)butane

Et

Ethyl

HBTU

O

-(Benzotriazol-1-yl)-

N

,

N

,

N

′,

N

′-tetramethyluronium

hexafluorophosphate

Lev

Levulinoyl

NBS

N

-Bromosuccinimide

NMO

N

-Methylmorpholine-

N

-oxide

Noyori

(

S

,

S

) (

R

)-Ru(η

6

-mesitylene)-(

S

,

S

)-TsDPEN

p

-TsOH

p

-Toluenesulfonic acid

Ph

Phenyl

PMB

p

-Methoxybenzyl

py

Pyridine

SAR

Structure Activity Relationship

TBAF

Tetrabutylammonium fluoride

TBS

t

-Butyldimethylsilyl

Tf

Trifluoromethanesulfonyl

THF

Tetrahydrofuran

TMS

Trimethylsilyl

TsDPEN

PhCH(TsN)CH(NH)Ph

Acknowledgments

We are grateful to NIH (GM090259) and NSF (CHE-0749451) for their support of our research programs. MFC also acknowledges the NSF for his fellowship from the NSF-IGERT Nanomedicine Program (DGE-0965843) at Northeastern University.

References

1. (a) Bertozzi, C.R. and Kiessling, L.L. (2001) Science, 291, 2357–2364.(b) Danishefsky, S.J., McClure, K.F., Randolph, J.T., and Ruggeri, R.B. (1993) Science, 260, 1307–1309.(c) Plante, O.J., Palmacci, E.R., and Seeberger, P.H. (2001) Science, 291, 1523–1527.(d) Sears, P. and Wong, C.-H. (2001) Science, 291, 2344–2350.(e) Yamada, H., Harada, T., Miyazaki, H., and Takahashi, T. (1994) Tetrahedron Lett., 35, 3979–3982.(f) Wang, L.X. (2006) Curr. Opin. Drug Discovery Dev., 9, 194–202.(g) Ni, J., Song, H., Wang, Y., Stamatos, N., and Wang, L.X. (2006) Bioconjugate Chem., 17, 493–500.(h) Zeng, Y., Wang, J., Li, B., Hauser, S., Li, H., and Wang, L.X. (2006) Chem. Eur. J., 12, 3355–3364.(i) Wang, J., Le, N., Heredia, A., Song, H., Redfield, R., and Wang, L.X. (2005) Org. Biomol. Chem., 3, 1781–1786.

2. For reviews of other approaches to hexoses, see: (a) Gijsen, H.J.M., Qiao, L., Fitz, W., and Wong, C.-H. (1996) Chem. Rev., 96, 443–473.(b) Hudlicky, T., Entwistle, D.A., Pitzer, K.K., and Thorpe, A.J. (1996) Chem. Rev., 96, 1195–1220.(c) Yu, X. and O'Doherty, G.A. (2008) ACS Symp. Ser., 990, 3–28.

3. (a) Timmer, M.S.M., Adibekian, A., and Seeberger, P.H. (2005) Angew. Chem. Int. Ed., 44, 7605–7607.(b) Adibekian, A., Bindschädler, P., Timmer, M.S.M., Noti, C., Schützenmeister, N., and Seeberger, P.H. (2007) Chem. Eur. J., 13, 4510–4527.(c) Stallforth, P., Adibekian, A., and Seeberger, P.H. (2008) Org. Lett., 10, 1573–1576.

4. (a) Bouché, L. and Reißig, H.-U. (2012) Pure Appl. Chem., 84, 23–36.(b) Pfrengle, F. and Reißig, H.-U. (2010) Chem. Soc. Rev., 39, 549–557.

5. (a) Ko, S.Y., Lee, A.W.M., Masamune, S., Reed, L.A. III, Sharpless, K.B., and Walker, F.J. (1983) Science, 220, 949–951.(b) Ko, S.Y., Lee, A.W.M., Masamune, S., Reed, L.A. III, Sharpless, K.B., and Walker, F.J. (1990) Tetrahedron, 46, 245–264.

6. (a) Danishefsky, S.J. (1989) Chemtracts, 273–297.For improved catalysis see: (b) Schaus, S.E., Branalt, J., and Jacobsen, E.N. (1998) J. Org. Chem., 63, 403–405.

7. (a) Johnson, C.R., Golebiowski, A., Steensma, D.H., and Scialdone, M.A. (1993) J. Org. Chem., 58, 7185–7194.(b) Hudlicky, T., Pitzer, K.K., Stabile, M.R., Thorpe, A.J., and Whited, G.M. (1996) J. Org. Chem., 61, 4151–4153.

8. Henderson, I., Sharpless, K.B., and Wong, C.H. (1994) J. Am. Chem. Soc., 116, 558–561.

9. For examples of protecting group free synthesis, see: (a) Crabtree, R.H. (2007) Science, 318, 756.(b) Chen, M.S. and White, M.C. (2007) Science, 318, 783.

10. (a) Ahmed, M.M. and O'Doherty, G.A. (2005) J. Org. Chem., 67, 10576–10578.(b) Ahmed, M.M. and O'Doherty, G.A. (2005) Tetrahedron Lett., 46, 4151–4155.(c) Gao, D. and O'Doherty, G.A. (2005) Org. Lett., 7, 1069–1072.(d) Zhang, Y. and O'Doherty, G.A. (2005) Tetrahedron, 61, 6337–6351.(e) Ahmed, M.M. and O'Doherty, G.A. (2005) Tetrahedron Lett., 46, 3015–3019.(f) Ahmed, M.M., Berry, B.P., Hunter, T.J., Tomcik, D.J., and O'Doherty, G.A. (2005) Org. Lett., 7, 745–748.

11. (a) Harris, J.M., Keranen, M.D., Nguyen, H., Young, V.G., and O'Doherty, G.A. (2000) Carbohydr. Res., 328, 17–36.(b) Harris, J.M., Keranen, M.D., and O'Doherty, G.A. (1999) J. Org. Chem., 64, 2982–2983.

12. Babu, R.S. and O'Doherty, G.A. (2003) J. Am. Chem. Soc., 125, 12406–12407.

13. Concurrent with these studies was the similar discovery by Feringa, Lee: (a) Comely, A.C., Eelkema, R., Minnaard, A.J., and Feringa, B.L. (2003) J. Am. Chem. Soc., 125, 8714–8715.(b) Kim, H., Men, H., and Lee, C. (2004) J. Am. Chem. Soc., 126, 1336–1337.

14. The poor reactivity in Pd-catalyzed allylation reaction of alcohols as well as a nice solution to this problem was reported, see: Kim, H. and Lee, C. (2002) Org. Lett., 4, 4369–4372.

15. For a related Rh system, see: Evans, P.A. and Kennedy, L.J. (2000) Org. Lett., 2, 2213–2215.

16. We have had great success at the preparation of optically pure furan alcohols from the Noyori reduction of the corresponding acylfuran, see: Li, M., Scott, J.G., and O'Doherty, G.A. (2004) Tetrahedron Lett., 45, 6407–6411.

17. (a) Li, M., Scott, J.G., and O'Doherty, G.A. (2004) Tetrahedron Lett., 45, 1005–1009.(b) Bushey, M.L., Haukaas, M.H., and O'Doherty, G.A. (1999) J. Org. Chem., 64, 2984–2985.

18. (a) Wang, H.-Y.L. and O'Doherty, G.A. (2011) Chem. Commun., 47, 10251–10253.(b) Shan, M., Xing, Y., and O'Doherty, G.A. (2009) J. Org. Chem., 74, 5961–5966.(c) Zhou, M. and O'Doherty, G.A. (2007) J. Org. Chem., 72, 2485–2493.(d) Haukaas, M.H. and O'Doherty, G.A. (2002) Org. Lett., 4, 1771–1774.(e) Haukaas, M.H. and O'Doherty, G.A. (2001) Org. Lett., 3, 3899–3992.

19. (a) Babu, R.S., Zhou, M., and O'Doherty, G.A. (2004) J. Am. Chem. Soc., 126, 3428–3429.(b) Babu, R.S. and O'Doherty, G.A. (2005) J. Carbohydr. Chem., 24, 169–177.

20. (a) Wang, H.-Y.L., Rojanasakul, Y., and O'Doherty, G.A. (2011) ACS Med. Chem. Lett., 2, 264–269.(b) Wang, H.Y.L., Xin, W., Zhou, M., Stueckle, T.A., Rojanasakul, Y., and O'Doherty, G.A. (2011) ACS Med. Chem. Lett., 2, 73–78.

21. Babu, R.S., Chen, Q., Kang, S.W., Zhou, M., and O'Doherty, G.A. (2012) J. Am. Chem. Soc., 134, 11952–11955.

22. Borisova, S.A., Guppi, S.R., Kim, H.J., Wu, B., Liu, H.W., and O'Doherty, G.A. (2010) Org. Lett., 12, 5150–5153.

23. (a) Hino, M., Nakayama, O., Tsurumi, Y., Adachi, K., Shibata, T., Terano, H., Kohsaka, M., Aoki, H., and Imanaka, H. (1985) J. Antibiot., 38, 926–935.(b) Patrick, M., Adlard, M.W., and Keshavarz, T. (1993) Biotechnol. Lett., 15, 997–1000.

24. (a) Colegate, S.M., Dorling, P.R., and Huxtable, C.R. (1979) Aust. J. Chem., 32, 2257–2264.(b) Colegate, S.M., Dorling, P.R., and Huxtable, C.R. (1985) Plant Toxicol., 249–254.

25. (a) Molyneux, R.J. and James, L.F. (1982) Science, 216, 190–191.(b) Davis, D., Schwarz, P., Hernandez, T., Mitchell, M., Warnock, B., and Elbein, A.D. (1984) Plant Physiol., 76, 972–975.

26. Liao, Y.F., Lal, A., and Moremen, K.W. (1996) J. Biol. Chem., 271, 28348–28358.

27. (a) Elbein, A.D., Solf, R., Dorling, P.R., and Vosbeck, K. (1981) Proc. Natl. Acad. Sci. U.S.A., 78, 7393–7397.(b) Kaushal, G.P., Szumilo, T., Pastuszak, I., and Elbein, A.D. (1990) Biochemistry, 29, 2168–2176.(c) Pastuszak, I., Kaushal, G.P., Wall, K.A., Pan, Y.T., Sturm, A., and Elbein, A.D. (1990) Glycobiology, 1, 71–82.

28. (a) Goss, P.E., Baker, M.A., Carver, J.P., and Dennis, J.W. (1995) Clin. Cancer Res., 1, 935–944.(b) Das, P.C., Robert, J.D., White, S.L., and Olden, K. (1995) Oncol. Res., 7, 425–433.(c) Goss, P.E., Reid, C.L., Bailey, D., and Dennis, J.W. (1997) Clin. Cancer Res., 3, 1077–1086.

29. Davis, B., Bell, A.A., Nash, R.J., Watson, A.A., Griffiths, R.C., Jones, M.G., Smith, C., and Fleet, G.W.J. (1996) Tetrahedron Lett., 37, 8565–8568.

30. For the first syntheses, see: (a) Mezher, H.A., Hough, L., and Richardson, A.C. (1984) J. Chem. Soc., Chem. Commun., 447–448.(b) Fleet, G.W.J., Gough, M.J., and Smith, P.W. (1984) Tetrahedron Lett., 25, 1853–1856.

31. For a review of swainsonine syntheses, see: (a) Nemr, A.E. (2000) Tetrahedron Lett., 56, 8579–8629.For more recent syntheses, see: (b) Martin, R., Murruzzu, C., Pericas, M.A., and Riera, A. (2005) J. Org. Chem., 70, 2325–2328.(c) Heimgaertner, G., Raatz, D., and Reiser, O. (2005) Tetrahedron Lett., 61, 643–655.(d) Song, L., Duesler, E.N., and Mariano, P.S. (2004) J. Org. Chem., 69, 7284–7293.(e) Lindsay, K.B. and Pyne, S.G. (2004) Aust. J. Chem., 57, 669–672.(f) Pearson, W.H., Ren, Y., and Powers, J.D. (2002) Heterocycles, 58, 421–430.(g) Lindsay, K.B. and Pyne, S.G. (2002) J. Org. Chem., 67, 7774–7780.(h) Buschmann, N., Rueckert, A., and Blechert, S. (2002) J. Org. Chem., 67, 4325–4329.(i) Zhao, H., Hans, S., Cheng, X., and Mootoo, D.R. (2001) J. Org. Chem., 66, 1761–1767.(j) Ceccon, J., Greene, A.E., and Poisson, J.F. (2006) Org. Lett., 8, 4739–4712.(k) Au, C.W.G. and Pyne, S.G. (2006) J. Org. Chem., 71, 7097–7099.(l) Dechamps, I., Pardo, D., and Cossy, J. (2007) ARKIVOC, 5, 38–45.

32. (a) Guo, H. and O'Doherty, G.A. (2006) Org. Lett., 8, 1609–1612.(b) Coral, J.A., Guo, H., Shan, M., and O'Doherty, G.A. (2009) Heterocycles, 79, 521–529.(c) Abrams, J.N., Babu, R.S., Guo, H., Le, D., Le, J., Osbourn, J.M., and O'Doherty, G.A. (2008) J. Org. Chem., 73, 1935–1940.

33. Walsh, C.T. (2000) Nature, 406, 775–781.

34. He, H., Williamson, R.T., Shen, B., Grazaini, E.I., Yang, H.Y., Sakya, S.M., Petersen, P.J., and Carter, G.T. (2002) J. Am. Chem. Soc., 124, 9729–9736.

35. (a) Petersen, P.J., Wang, T.Z., Dushin, R.G., and Bradford, P.A. (2004) Antimicrob. Agents Chemother., 48, 739–746.(b) Sum, P.E., How, D., Torres, N., Petersen, P.J., Lenoy, E.B., Weiss, W.J., and Mansour, T.S. (2003) Bioorg. Med. Chem. Lett., 13, 1151–1155.(c) He, H., Shen, B., Petersen, P.J., Weiss, W.J., Yang, H.Y., Wang, T.-Z., Dushin, R.G., Koehn, F.E., and Carter, G.T. (2004) Bioorg. Med. Chem. Lett., 14, 279–282.

36. Babu, R.S., Guppi, S.R., and O'Doherty, G.A. (2006) Org. Lett., 8, 1605–1608.

37. Guppi, S.R. and O'Doherty, G.A. (2007) J. Org. Chem., 72, 4966–4969.

38. Tané, P., Ayafor, J.P., Sondengam, B.L., Lavaud, C., Massiot, G., Connolly, J.D., Rycroft, D.S., and Woods, N. (1988) Tetrahedron Lett., 29, 1837–1840.

39. Seidel, V., Baileul, F., and Waterman, P.G. (1999) Phytochemistry, 52, 465–472.

40. Hu, J.-F., Garo, E., Hough, G.W., Goering, M.G., O'Neil-Johnson, M., and Eldridge, G.R. (2006) J. Nat. Prod., 69, 585–590.

41. Wu, B., Li, M., and O'Doherty, G.A. (2010) Org. Lett., 12, 5466–5469.

42. Zhang, Z., Wang, P., Ding, N., Song, G., and Li, Y. (2007) Carbohydr. Res., 342, 1159–1168.

43. Cheng, L., Chen, Q., and Du, Y. (2007) Carbohydr. Res., 342, 1496–1501.

44. Shi, P., Silva, M., Wu, B., Wang, H.Y.L., Akhmedov, N.G., Li, M., Beuning, P., and O'Doherty, G.A. (2012) ACS Med. Chem. Lett., 3, 1086–1090.

45. Guo, H. and O'Doherty, G.A. (2007) Angew. Chem. Int. Ed., 46, 5206–5208.

46. Guo, H. and O'Doherty, G.A. (2008) J. Org. Chem., 73, 5211–5220.

47. (a) Mock, M. and Fouet, A. (2001) Annu. Rev. Microbiol., 55, 647–671.(b) Sylvestre, P., Couture-Tosi, E., and Mock, M. (2002) Mol. Microbiol., 45, 169–178.

48. Daubenspeck, J.M., Zeng, H., Chen, P., Dong, S., Steichen, C.T., Krishna, N.R., Pritchard, D.G., and Turnbough, C.L. Jr. (2004) J. Biol. Chem., 279, 30945–30953.

49. Several approaches to related tri- and pentasaccharides, see: (a) Saksena, R., Adamo, R., and Kovac, P. (2006) Bioorg. Med. Chem. Lett., 16, 615–617.(b) Saksena, R., Adamo, R., and Kovac, P. (2007) Bioorg. Med. Chem., 15, 4283–4310.(c) Wang, Y., Liang, X., and Wang, P. (2011) Tetrahedron Lett., 52, 3912–3915.(d) Tamborrini, M., Bauer, M., Bolz, M., Maho, A., Oberli, M.A., Werz, D.B., Schelling, E., Jakob, Z., Seeberger, P.H., Frey, J., and Pluschke, G. (2011) J. Bacteriol., 193, 3506–3511.(e) Milhomme, O., Dhenin, S.G.Y., Djedaini-Pilard, F., Moreau, V., and Grandjean, C. (2012) Carbohydr. Res., 356, 115–131.

50. (a) Werz, D.B. and Seeberger, P.H. (2005) Angew. Chem. Int. Ed., 44, 6315–6318.(b) Adamo, R., Saksena, R., and Kovac, P. (2005)