design, synthesis and biological utility of polysaccharide

276
Design, Synthesis and Biological Utility of Polysaccharide Chain-Terminating Glycosides By Christopher D. Brown A dissertation submitted in partial fulfillment of the requirements for the degree of Doctor of Philosophy (Chemistry) At the University of Wisconsin—Madison 2013 Final Oral Defense: July 19, 2012 This dissertation is approved by the following members of the Final Oral Committee Laura L. Kiessling, Professor, Chemistry Samuel H. Gellman, Professor, Chemistry Deane F. Mosher, Professor, Biomolecular Chemistry Eric Strieter, Assistant Professor, Chemistry Jennifer M. Schomaker, Assistant Professor, Chemistry

Upload: others

Post on 19-Feb-2022

0 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Design, Synthesis and Biological Utility of Polysaccharide

Design, Synthesis and Biological Utility of Polysaccharide

Chain-Terminating Glycosides

By

Christopher D. Brown

A dissertation submitted in partial fulfillment of

the requirements for the degree of

Doctor of Philosophy

(Chemistry)

At the

University of Wisconsin—Madison

2013

Final Oral Defense: July 19, 2012

This dissertation is approved by the following members of the Final Oral Committee

Laura L. Kiessling, Professor, Chemistry

Samuel H. Gellman, Professor, Chemistry

Deane F. Mosher, Professor, Biomolecular Chemistry

Eric Strieter, Assistant Professor, Chemistry

Jennifer M. Schomaker, Assistant Professor, Chemistry

Page 2: Design, Synthesis and Biological Utility of Polysaccharide

i

Design, Synthesis and Biological Utility of Polysaccharide Chain-Terminating Glycosides

Christopher D. Brown

Under the supervision of Laura L. Kiessling At the University of Wisconsin—Madison

Polysaccharides are ubiquitous in nature, yet their assembly is poorly understood at the

molecular level. Unlike their nucleic or amino acid counterparts, carbohydrate polymers are

assembled via a template independent process. Glycosyltransferases, the enzymes required for

polymer assembly, are able to control both the length and patterning of a polysaccharide chain.

Structural contributions from the enzyme itself, as well as the substrates it recognizes, play a

central role in the regulation of assembly. From this, it follows that perturbation to the

structure of the enzyme or substrates would result in concomitant perturbation of the activity

of the system. This approach was used to develop a new method for assessing patterning

fidelity of one such carbohydrate polymerase.

The enzyme GlfT2 from Mycobacterium tuberculosis constructs a linear polysaccharide

of galactofuranose (Galf) residues termed the galactan from the donor sugar UDP-Galf. The

galactan is an integral part of the Mycobacterial cell wall and aids in the organism’s innate

ability to resist and evade its human host immune response. A better understanding of this

biosynthetic process may aid in the development of new treatments for this global disease. A

closer examination of galactan structure reveals that the monosaccharides are linked via an

alternating pattern of β-(1,5) and β-(1,6) glycosidic linkages. For this bifunctional enzyme, we

Page 3: Design, Synthesis and Biological Utility of Polysaccharide

ii

wondered if it was faithful to the alternating pattern and devised a method to challenge this

fidelity.

The pivotal hydroxyl groups at the C5 or C6 positions of Galf are the requisite

nucleophiles for chain extension by GlfT2. We reasoned that a chemical substitution of either

hydroxyl group for a non-nucleophilic functionality such as fluorine or hydrogen would block

extension at that position. We hypothesized that this block would force the enzyme to continue

polymerization via the remaining (unsubstituted) hydroxyl group or induce chain termination,

depending on its fidelity to alternating linkage deposition. A divergent synthesis of fluorinated

UDP-Galf analogs was developed, and the donors induced chain termination in a sequence-

specific manner. This suggested GlfT2 operates with high fidelity for constructing the

alternating pattern. Further studies with chain-terminating probes promise to yield insight into

the mechanism of galactan assembly.

Page 4: Design, Synthesis and Biological Utility of Polysaccharide

iii

Acknowledgements

A PhD is a community effort. No one produces a new body of work in a vacuum. Many

people supported me professionally and privately and I reserve this space to thank them. While

it feels almost like bringing the cast out before the performance to take a bow, this is as it

should be: Without these people, there would be no show, no spectacle, and no mystery

unraveled.

Laura Kiessling is an advisor who values unwavering rigor and the excitement born only

of scientific inquiry. I stand today on a firm scientific foundation because I chose to emulate her

drive for discovery. She has been a mentor to me in how to approach scientific thinking and

scientific writing. I have enjoyed learning from her a great deal.

The Kiessling Group operates with a level of achievement and excellence that has me in

awe on a regular basis. All of my coworkers are intelligent, hard-working and clever. The results

of these traits are apparent in the work culture: It is collaborative and fun, engaging and an

easy place to learn. I want to thank Kenzo Yamatsugu, Rebecca Splain, Matt Levengood, Raja

Annamali, Darryl Wesener and John May in particular, all of whom have worked on GlfT2 and

improved our understanding of its function in fundamental ways. I have learned laboratory

skills from each of them and benefited from materials they have generated for my own studies.

I also want to thank my office mates Joe Grim, Mario Martinez and Jonathan Hudon for sharing

the workspace. Laughter is my favorite de-stressor and we had more than our fair share in

5128.

Page 5: Design, Synthesis and Biological Utility of Polysaccharide

iv

For two years I had the privilege of working with an undergraduate chemistry major,

Max Rusek. I feel lucky to have had the chance to mentor him. He worked furiously to supply

me with important chemical intermediates for many of the studies performed in this thesis. He

accelerated the work we accomplished and did it with a smile every day. I wish him well in his

post-bachelor degree endeavors.

The staff members who keep the chemistry building running are top rate. In particular

the facilities directors Charlie Fry, Monika Ivancic and Martha Vestling have been excellent

resources for NMR and mass spectrometry issues. We have top notch facilities because of the

top notch people that run them. I want to thank Kat Myhre, my go-to person for logistical help

regarding event planning as well as a fellow gardening enthusiast. Theresa, Lauren and

Stephanie were always willing to stop and ask about how my day was going. I will miss our

talks, friends.

Lastly, but most importantly, I want to thank my friends and family. My best friend Kelly

Kraft kept me sane by insisting on annual backpacking trips which were always great

adventures. Deb Heilert was great at finding fun things to do in Madison which kept me

balanced. My partner for life, my boyfriend, Peter Gruett, cannot be thanked sufficiently in an

acknowledgement. You are the best part of every single day, and have been since our first date.

I love you so much. I thank my uncles Dick and John and my aunt Bobbi for reminding me where

I come from and what my values are. Finally, my mom and my brother have been my core, my

family, my driving force. Mom, I wrote the words one thing at a time on my lab bench because

Page 6: Design, Synthesis and Biological Utility of Polysaccharide

v

when I was most stressed by graduate school your long-ago words of counsel were my source

of strength, as they have always been. We did it.

Page 7: Design, Synthesis and Biological Utility of Polysaccharide

vi

Table of Contents

Abstract i

Acknowledgements iii

Table of Contents vi

List of Figures ix

List of Abbreviations xii

Chapter 1: The Regulation of Bacterial Polysaccharide Assembly 1

1.1 Introduction 2

1.2 Biosynthesis 5

1.2.1 Genomic organization 5

1.2.2 The location of assembly 6

1.3 Substrate Recognition 7

1.3.1 The Glycosyltransferases 7

1.3.2 Lipid Binding 9

1.3.3 Acceptor binding and the role of primases 13

1.3.4 Donor Binding 14

1.3.5 Determining the order of binding 16

1.4 Polymerization 17

1.4.1 Directionality 17

1.4.2 Processivity 18

1.4.3 Subsites 23

Page 8: Design, Synthesis and Biological Utility of Polysaccharide

vii

1.4.4 Length Control 24

1.4.4.1 Intrinsic factors 24

1.4.4.2 Extrinsic factors 26

1.4.4.3 Chemical chain termination 29

1.4.5 Pattern control 30

1.5 Conclusions 32

Chapter 2: Design and Synthesis of Fluorinated UDP-Galf Analogs 34

2.1 Introduction 35

2.2 GlfT2: A bifunctional polymerase from Mycobacterium tuberculosis 37

2.3 Synthetic probes 39

2.3.1 Synthetic acceptors 39

2.3.2 Synthetic donors 39

2.4 Chemical probes of pattern fidelity 40

2.5 Synthesis of fluorinated Galf analogs 43

2.5.1 Accessing vinyl Galf 43

2.5.2 Setting the stereochemistry at C5 45

2.5.3 Installation of fluorine 46

2.5.4 Installation of phosphate 47

2.5.5 Phosphate coupling 48

2.6 Experimental procedures 51

Page 9: Design, Synthesis and Biological Utility of Polysaccharide

viii

Chapter 3: Evaluation of Fluorosugars as Probes of Carbohydrate Polymerase Activity 69

3.1 Introduction 70

3.2 GlfT2 operates by a sequence specific mechanism 72

3.3 The sequence specificity of GlfT2 is general 74

3.4 A mistake on initiation? 74

3.5 GlfT2 struggles to recognize the terminal linkage of short acceptors 77

3.6 Fluorinated acceptors are dead-end substrates 82

3.7 Experimentals 85

Chapter 4: Expanding the Utility of Polysaccharide Chain-Terminating Glycosides 89

4.1 Introduction 90

4.2 A comparison of fluorinated and deoxygenated Galf donors 93

4.2.1 As sequence specificity probes 93

4.2.2 As Inhibitors 94

4.3 Fluorinated acceptors 97

4.3.1 Synthesis 97

4.3.2 Evaluation as inhibitors 99

4.4 GlfT2 makes two linkages with one active site 100

4.5 Experimentals 102

Page 10: Design, Synthesis and Biological Utility of Polysaccharide

ix

List of Figures

Figure 1.01 Bacterial polysaccharides 3

Figure 1.02: Location of polymerization 7

Figure 1.03: Glycosyltransferase mechanism 8

Figure 1.04: Glycosyltranserfase crystal structures 10

Figure 1.05: Lipid specificity 12

Figure 1.06: Donor binding 15

Figure 1.07: Common Bi Bi mechanisms 16

Figure 1.08: Determining reducing versus non-reducing chain elongation 19

Figure 1.09: Distraction assay for polymerization 22

Figure 1.10: TagF is distributive 22

Figure 1.11: Length control 28

Figure 1.12: Pattern control 32

Table 1.01 4

Figure 2.01: Homopolymers of glucose 35

Figure 2.02: Polysaccharides with alternating linkages 36

Figure 2.03: Galactan biosynthesis 38

Figure 2.04: Synthetic acceptors 40

Figure 2.05: Synthetic donors 41

Figure 2.06: Fidelity assessment assay 41

Figure 2.07: Retrosynthesis of fluorinated UDP-Galf analogs 43

Figure 2.08: Determining the stereochemistry of the 5F furanose 48

Scheme 2.01: Accessing vinyl Galf 44

Page 11: Design, Synthesis and Biological Utility of Polysaccharide

x

Scheme 2.02: A kinetic separation of diastereomers 45

Scheme 2.03: Installation of fluorine 46

Scheme 2.04: Installation of phosphate for 6F-Galf 49

Scheme 2.05: Installation of phosphate for 5F-Galf 49

Scheme 2.06: Phosphate coupling 50

Table 2.01: A comparison of physical parameters for C-F and C-OH bonds 42

Figure 3.01: A mass spectrometry method to observe polymerization by GlfT2 in vitro 71

Figure 3.02: Chain termination is sequence specific with a hexasaccharide acceptor 73

Figure 3.03: Chain termination is seuqnece specific with a tetrasaccharide acceptor 75

Figure 3.04: Synthetic tetrasaccharide behaves the same as chemoenzymatic 76

tetrasaccharide

Figure 3.05: Model for carbohydrate-directed binding 78

Figure 3.06: Disaccharides are poor substrates for extension with F-Galfs 81

Figure 3.07: Trisaccharides reveal importance of lipid in acceptor orientation 83

Figure 3.08: Fluorinated acceptors are dead-end substrates 84

Scheme 3.01: Allyl acceptor synthesis 79

Scheme 3.02: Synthesis of disaccharide acceptors 80

Figure 4.01: Panel of UDP-Galf analogs 90

Figure 4.02: Competition experiment to evaluate fluorinated donors as inhibitors of 94

polymerization

Figure 4.03: Coupled assay reveals drop in UDP production in the presence of 96

fluorinated donors

Figure 4.04: Assessing the polymerization competency of fluorinated acceptors 100

Figure 4.05: Mutant GlfT2 polymerization assays 101

Page 12: Design, Synthesis and Biological Utility of Polysaccharide

xi

Figure 4.06: GlfT2 D371E retains high sequence fidelity 102

Scheme 4.01: Synthesis of deoxy and azido Galf 92

Scheme 4.02: Fluorinated acceptor synthesis 97

Table 4.01: Summary of chain termination results 93

Page 13: Design, Synthesis and Biological Utility of Polysaccharide

xii

List of Abbreviations

ABC ATP binding cassette Ac Acetyl Aq aqueous ATP adenosine triphosphate B. subtilis Bacillus subtilis BSIT benzenesulfonylimidazolium triflate CAN Acetonitrile AcOH acetic acid Bn benzyl Bz benzoyl C. jejuni Campylobacter jejuni CMP cytosine monophosphate CSA camphor sulfonic acid d doublet DAST diethylaminosulfur trifluoride DCM dichloromethane dd doublet of doublets d.r. diastereomeric ratio DIPEA diisopropylethylamine DMF dimethylformamide DNA deoxyribonucleic acid E. coli Escherichia coli EMM exact mass measurement ESI Electrospray ionization Et ethyl Galf galactofuranose GalNAc N-acetylgalactosamine GDP guanine diphosphate Glc glucose GlcNAc N-acetylglucosamine GlfT2 Galactofuranosyltransferase 2 G. stearothermophilus Geobacilus stearothermophilus GT glycosyltransferase HEPES (4-(2-hydroxyethyl)-1-piperazineethanesulfonic acid HPLC high performance liquid chromatography IPTG isopropylthiogalactopyranoside K4CP chondroitin polymerase from E. coli K4 LDH lactate dehydrogenase

Page 14: Design, Synthesis and Biological Utility of Polysaccharide

xiii

m multiplet MALDI-TOF matrix-assisted laser desorption ionization time of flight Man mannose ManNAc N-acetylmannosamine mCPBA meta-chloroperoxybenzoic acid Me methyl MeOH methanol MRSA methicillin resistant staphylococcus aureus NADH nicotinamide adenine dinucleotide NMR nuclear magnetic resonance P. aeruginosa Pseudomonas aeruginosa PBP penicillin binding protein PDB protein data bank PET positron emission tomography PK pyruvate kinase q quartet RT room temperature s singlet S. aureus Staphylococcus aureus SDS-PAGE sodium dodecyl sulfate polyacrylamide gel electrophoresis S. enterica Salmonella enterica S. equisimilis Streptococcus equisimilis S. pneumonia Steptococcus pneumoniae STD NMR saturation transfer difference nuclear magnetic resonance STP standard temperature and pressure TEA triethylamine TFA trifluoroacetic acid TFAA trifluoroacetic anhydride THF tetrahydrofuran TLC thin layer chromatography UDP uridine diphosphate UMP uridine monophosphate UMP-Im. uridine monophosphate imidazolium UV ultraviolet

Page 15: Design, Synthesis and Biological Utility of Polysaccharide

1

Chapter 1

The Regulation of Bacterial Polysaccharide Assembly

This chapter is intended for submission to the Annual Review of Biochemistry

Brown, Christopher D. and Kiessling, Laura L., The Regulation of Bacterial Polysaccharide Assembly. 2012

Page 16: Design, Synthesis and Biological Utility of Polysaccharide

2 1.1 Introduction

Polysaccharides are pervasive in nature. They exist in intracellular, pericellular and extracellular

environments and participate in a wealth of biological processes ranging from energy storage to cell

structural integrity to pathogenesis. Their range spans from short oligomeric sequences to long chains

thousands of monomers in length. These carbohydrate chains can likewise be linear or branched,

charged or neutral, viscous or slippery, rigid or flexible. Indeed, polysaccharides are so diverse in size,

shape and physical characteristics that research into their biosynthesis has demanded an equally

multifaceted approach.

While both eukaryotic and prokaryotic organisms produce polysaccharides, the bacteria are the

true masters of the art. Bacterial polysaccharides are can be covalently bound to the cell wall, anchored

to the membrane, attached to proteins, or secreted into the extracellular matrix. Structures include

peptidoglycan,1 capsular polysaccharide,2 exopolysaccharide,3 O-antigen,4,5 teichoic acid,6 N- and O-

linked glycans7,8 and various other unique structures that decorate the exterior of bacterial cells (Figure

1). They project outward into the environment and serve several important functions. They preserve

cellular integrity by regulating osmotic pressure and cell shape by providing a rigid structural coat. They

serve as docking sites for bacteriophage9 and are involved in colonization, pathogenesis and virulence in

a range of hosts. Many excellent reviews have been written on aspects of polysaccharide biosynthesis,

usually in the context of a specific organism or type of polymer. This review describes the regulation of

polysaccharide assembly, with a focus on the chemical structure of the enzyme and substrates it utilizes

as well as the functional characterization of the polymerization event itself.

Many bacterial polysaccharides are of great medical importance. They are causal elements

underlying much of the morbidity and mortality of bacterial infections. Lipopolysaccharide (LPS) is

involved in septic shock and multiple organ failure. Capsular polysaccharides in streptococcal

Page 17: Design, Synthesis and Biological Utility of Polysaccharide

3

Figure 1. Representative examples of bacterial polysaccharides exhibit great structural diversity. They are found anchored to cellular structures, attached to proteins, lipids, other carbohydrates, or are free in the extracellular matrix. Carbohydrate polymers exposed to the external environment are the point of first contact for many important biological interactions.

Page 18: Design, Synthesis and Biological Utility of Polysaccharide

4 bacteria facilitate colonization of the nasopharyngeal or meningeal epithelium.2 Some have been

identified as highly immunogenic and co-opted for vaccine development.10 Teichoic acid polymers from

Gram-positive organisms including methicillin-resistant Staphylococcus aureus (MRSA) are thought to

facilitate septation during cell division by harboring a reserve of divalent cations for glycosyltransferase

activity.11 The polyanionic polymer alginate from Pseudomonas aeruginosa is integral to formation of

biofilms that render its infection in the lungs of cystoic fibrosis patients so recalcitrant to antibiotics.3 A

current push to generate large quantities of chemically-defined heparin has given rise to the possibility

of using bacterial enzymes involved in capsular polysaccharide assembly for chemoenzymatic

production.12

Table 1.01. Representative bacterial polysaccharides vary in sequence and length.

Variation of polysaccharide sequences and lengths in bacteria compounds the complexity

observed in Nature. Some bacteria employ homopolymers with simple repeating patterns wherease

others construct complicated and rare glycans taylor-made for the required actions of the organism.

Length also varies from short oligomeric sequences, such as the N-linked glycan of Campylobacter jejuni,

to the very large polymers that mimic glycosaminoglycans such as the hyaluronic acid capsular

Page 19: Design, Synthesis and Biological Utility of Polysaccharide

5 polysaccharide of Streptococcus pneumoniae. Length and sequence are so fundamental to

polysaccharide function that most research involving carbohydrate polymer biosynthesis has focused on

understanding the regulation of these two structural features (Table 1.01).

An effort to understand what factors influence glycan localization, length and patterning has

illuminated the features of polysaccharide biosynthesis. The polymerization event is at the heart of this

assembly process, in which activated monosaccharides are appended iteratively end to end to generate

the polysaccharide chain. Elongation of a repeating sequence of sugars occurs in a variety of contexts.

Sometimes the process is the result of the actions of a single polymerase while other systems employ a

division of labor strategy, utilizing the specific activities of several enzymes in sequence or in concert.

This review highlights the biochemical strategies used by bacteria to construct these carbohydrate

polymers and the methods researchers have used to define them. It focuses on the biochemical

regulation of polymerization and highlights concepts used to better define the complicated assembly

process. The review will also highlight techniques that have proven effective to study polymerizations

and identify areas that are in need of further method development.

1.2.1 Biosynthesis: Genomic organization

The methods used to reveal aspects of assembly have been cross-disciplinary. Polysaccharides

isolated from nature are sequenced using degradation and chemical modification techniques that date

back a century. Older methods have been supplemented with newer techniques such as X-ray

crystallography and NMR spectroscopy and mass spectrometry to elucidate structure of enzyme and

polymer alike. High resolution techniques like electron microscopy, radiation inactivation and atomic

force microscopy have facilitated our understanding of higher order organization of these cell surface

structures. A continued drive to sequence biologically important genomes has yielded rich information

about bacterial genomic organization. In bacteria, polysaccharide biosynthesis is organized into gene

Page 20: Design, Synthesis and Biological Utility of Polysaccharide

6 clusters. Genes involved in polysaccharide assembly include sugar nucleotide biosynthetic enzymes,

glycosyltransferases, proteins involved in transport of polymers or precursors across the plasma

membrane, and other regulatory elements. Mutagenesis and deletion studies revealed the importance

of polysaccharides on bacterial behavior, ultrastructure and physiology. Assays for polymerase activity

were generated. Newer techniques that combine genetics, molecular biology, biochemistry, chemical

biology and bioengineering have marshaled in a new era of study into carbohydrate structure and

function.

1.2.2 Biosynthesis : The location of assembly

Bacterial polysaccharide chains are formed in many cellular environments. Consider the three

predominant methods for assembly of O-antigen (Figure 1.02).13 Researchers have classified O-antigen

biosynthesis based on the location of assembly and mechanism of export. The three categories are

Wzx/Wzy, the ATP binding cassette (ABC) and the synthase pathways. Polymers assembled via the

Wzx/Wzy pathway are built up from undecaprenyl phosphate (Und-P) carrier lipids bearing the

fundamental repeat sequence. These monomers are assembled on the cytoplasmic face of the

membrane, exported to the periplasm via the Wzx protein, polymerized by Wzy, ligated to the lipid A

core and transported to the outer membrane. Polymerization in the ABC pathway also occurs on Und-P

carriers, but on the cytosolic face, prior to transport via an ATP Binding Cassette transporter protein for

which the pathway is named. The synthase pathway generates the polysaccharide coupling

polymerization to export. (Figure 1.02) Capsular polysaccharide biosynthesis uses a similar set of

polymerization/export strategies.2

The location of assembly determines the resources available to build the polymer. How

glycosyltransferases utilize these resources to build polymers requires an understanding of enzyme

structure, substrate availability and the key recognition events that bring together the various

Page 21: Design, Synthesis and Biological Utility of Polysaccharide

7 precursors. The structural features of enzyme, donor and acceptor dictate the substrate recognition

events necessary for efficient biosynthesis.

Figure 1.02. Polymerization can take place on the cytosolic, periplasmic or intramembrane portions of the cell. E. coli O9 is assembled via the ABC pathway on the cytosolic side of the membrane (A). E. coli O86 uses the Wzx/Wzy pathway and polymerizes modified Und-P lipids on the periplasmic face of the inner membrane (B). Synthases such as WbbF from S. enterica couple polymerization to export (C). 1.3.1 Substrate Recognition: The Glycosyltransferases

All glycosyltransferases must bring together an acceptor substrate and activated donor to

mediate catalysis. Efforts to understand how these enzymes bind their substrates have given us a

molecular picture of polymer assembly. Glycosyltransferases are classified according to sequence

homology into families numbered GT-1, GT-2 up to GT-110 at present.14 A second method of

classification is based on predicted and solved structural domains; glycosyltransferases predominantly

have either GT-A or GT-B domains. A third classification defines glycosyltransferases mechanistically as

either retaining or inverting. (Figure 1.03) Inverting transferases generate a glycosidic linkage that has

the opposite stereochemistry from the stereochemistry of the anomeric position of the donor. This

mechanism usually involves deprotonation of the acceptor via a general base mechanism followed by

direct displacement of the donor leaving group. For retaining glycosyltransferases, an enzyme-donor

covalent intermediate is formed in a two-step double-inversion process resulting in a glycosidic bond

Page 22: Design, Synthesis and Biological Utility of Polysaccharide

8 with the same stereochemistry as the donor. A comprehensive review on these classifications and

implications the structures have a glycosyltransferase function was published recently.14

Figure 1.03. Glycosyltransferases are either inverting or retaining, depending on whether the stereochemistry at the donor sugar switches or is maintained, respectively. For example, a donor with an axial (red) NDP group can be used to generate a glycosidic bond with either equatorial (blue) or axial stereochemistry.

The most direct technique for understanding how glycosyltransferases recognize their donor

and acceptor substrates is X-ray crystallography.(Figure 1.04) Few structures of polymerases have been

solved, precluding generalizations about regulation of polysaccharide assembly, however the structures

that are reported provide a framework for illuminating the molecular mechanisms at work.(Refs) The

number of polymerases with solved crystal structures is still too small to make broad generalizations

about the regulation of polysaccharide assembly, but the structures that are reported go a long way

toward the illumination of the key molecular mechanisms. For instance, K4CP—a bifunctional

polymerase that generates a chondroitin polymer—is a member of the GT-A family in which two

Rossmann-like folds create a continuous twisting beta sheet surrounded by alpha helices.15

Glycosyltransferases of this family bind a nucleotide donor sugar and an acceptor substrate to engage

polymerization. The crystal structure of the staphylococcal polymerase LtaS, which is responsible for

polyglycerol lipoteichoic acid biosynthesis,16 was solved as a cocrystal with glycerol phosphate and a

Page 23: Design, Synthesis and Biological Utility of Polysaccharide

9 Mn2+ atom.17 The enzyme was found to adopt a sulfatase-like fold in which the conserved polar and

charged residues bind the metal cation. The cation in turn interacts with the phosphate moiety of the

glycerol phosphate group.17 Crystal structures of peptidoglycan transglycosylases, which catalyze the

transfer of sequential disaccharide units to build up the carbohydrate backbone of peptidoglycan

strands, reveal large, multifunctional enzymes with transmembrane, transglycosylation and

transpeptidase domains.18,19 A co-crystal of E.coli transglycosylase bound to the inhibitor moenomycin

(A mimic of the endogenous substrate lipid IV) demonstrates how the enzyme may bind two molecules

of lipid II, condense them to generate lipid IV, and then extend the polymerization via elongation at the

reducing end of lipid IV, maintaining contact with the growing chain as it slips along a groove on the

enzyme’s surface.19

The structures of a variety of carbohydrate polymerases gives an image of how the enzyme

preorganizes the ternary complex to generate a new bond and what structural features of the enzyme

and substrates are essential for binding and catalysis. What imparts an enzyme with specific or broad

reactivity and can that be exploited? Two thematic areas of study have emerged. The first is concerned

with substrate recognition, the second with measuring the resulting activity. A discussion regarding

these two areas is framed around specific examples from the literature and focused on the techniques

used to investigate the structural and functional underpinnings of polysaccharide biosynthesis.

1.3.2 Substrate Recognition: Lipid binding

Most bacterial glycans are assembled on a lipid phosphate carrier, undecaprenyl phosphate

(Und-P) in virtually all cases. Und-P consists of 7 cis, 3 trans and one terminal isoprene units. O-antigens,

capsular polysaccharides and bacterial N- and O-glycans are assembled on this carrier. Mycobacteria use

the related lipid decaprenyl phosphate to assemble the arabinogalactan, a cell wall component.

Page 24: Design, Synthesis and Biological Utility of Polysaccharide

10

Figure 1.04. Active site of chondroitin polymerase K4CP with UDP-GalNAc and Mn2+ (A). LtaS from Bacillus subtilis with glycerol phosphate and Mg2+ (B). Ribbon (C) and surface (D) projections of E. coli PBP1b transglycosylase bound to a lipid IV analog. PDB codes: 2Z87 (K4CP),2W5S (LtaS) 3FWL (PBP1b)

Membrane anchored glycans such as the lipoteichoic acids and some streptococcal capsular

polysaccharides are assembled directly onto phosphatidylglycerol lipids.

Bacterial polysaccharides are appended to a wide range of structures including lipids in the

cellular membrane, the peptidoglycan and proteins. The ultimate ligation site is dictated not by the

polysaccharide chain, but by the lipid carrier on which it was built. To ensure that polysaccharide

assembly results in ligation to the appropriate cellular structure, the process is regulated at some point

by a glycosyltransferase with high specificity for the carrier lipid. In this sense, the specificity of an

enzyme for a lipid carrier acts essentially as a targeting sequence, allowing bacteria to synthesize

complicated glycan structures and deliver them en bloc to their appropriate location. Strict adherence to

Page 25: Design, Synthesis and Biological Utility of Polysaccharide

11 lipid identity is not observed at every step of polysaccharide assembly and different organisms have

adopted specificity at unique points in the biosynthetic sequence (Figure 1.05). Thus,

glycosyltransferases that recognize and act on a range of structural variants have been exploited by

chemists who have designed simplified glycolipid acceptors for enzymatic studies.20,21 This has prompted

scientists interested in glycan engineering to explore the potential for tailored glycan displays

constructed with promiscuous biosynthetic machinery (Figure 1.05).22

The Wang group evaluated the lipid dependence of the polymerase Wzy in the construction of

the E. coli O86 antigen and found that unsaturation or trans geometry in the isoprene unit closest to the

glycan all but abolished the enzymatic activity of Wzy.23 Interestingly, the lipid specificity is relaxed for

the transferases responsible for assembling the repeating unit. The ligase WaaL which catalyzes the

transfer of the fully polymerized O-antigen to Lipid A core, exhibits the highest degree of promiscuity in

the system.24

The Imperiali Group undertook a thorough investigation of the lipid dependence of the pgl

glycosyltransferases responsible for N-glycan assembly in Campylobacter jejuni.25 Biosynthesis of the N-

glycan produced in Campylobacter jejuni is general and organized into a gene cluster containing

nucleotide donor sugar synthases, glycosyltransferases and gene products involved in export and

ligation. An undecaprenyl phosphate lipid is modified sequentially by the enzymes PglC and PglA which

add a 2,4-diacetamido-2,4,6-trideoxyglucose and an α-(1,3) linked GalNAc, respectively.26 The primase

PglJ adds an α-(1,4) linked GalNAc to generate the trisaccharide primer substrate recognized by the

polymerase PglH.27 The three priming enzymes exhibit a strong preference for lipids bearing alpha

unsaturation with cis configuration.25 Saturation of the proximal bond reduced the enzyme reaction rate

of PglC, PglJ and PglB to approximately 25%, 37% and 39% of normal, respectively. Switching from cis to

trans geometry reduced the primase activities to 2%, 30% and 13% respectively. If the proximal

Page 26: Design, Synthesis and Biological Utility of Polysaccharide

12 geometry was preserved, but the length reduced, the enzymes only exhibited mild reductions in

reaction rate. The lipid specificity of the polymerase PglH has not been evaluated. Likewise, the ligase

PglB has not been investigated for lipid specificity, but it has been noted that it exhibits relaxed

specificity for both the glycan and acceptor substrates.22,28,29 In fact, if PglB is expressed in E. coli ΔWaaL,

the enzyme will transfer the accumulated O-antigen to proteins.30,31 The possibility for engineering

glycoproteins using the pgl gene cluster is of broad interest.22,30,31

Figure 1.05. The polymerase Wzy from E. coli has high specificity for the undecaprenyl phosphate carrier lipid whereas the preceding primases exhibit higher promiscuity of the lipid. The reverse is true for N-glycan assembly in C. jejuni. The promiscuity of the ligase PglB has been exploited to generate non-natural proteoglycans.

Exopolysaccharides are a class of polysaccharides which do not require priming of a lipid carrier.

Polymerization of these structures occurs directly from the nucleotide sugar precursors and oligomeric

acceptors, and is often coupled to export. P. aeruginosa produces the exopolysaccharide alginate which

is present in the extracellular matrix of biofilm-forming infections like those in the lungs of cystic fibrosis

patients.3,32 The membrane protein Alg8 polymerizes a homopolymer of mannuronic acid and couples

Page 27: Design, Synthesis and Biological Utility of Polysaccharide

13 elongation to transport of the nascent chain to the periplasm.33,34 The epimerase AlgG present in the

periplasm randomly converts some mannuronic acids to the C5 epimer, guluronic acid.35 The guluronic

acids protect nascent alginate chains from degradation by a lyase AlgL.36 Additional post-polymerization

acylations are added to the 2-O and 3-O positions of ManA. Export across the outer membrane is

mediated in part by additional chaperones like Alg44,37,38 AlgK39 and AlgX.40 Without a lipid anchor to

target the polymer for ligation to the membrane, alginate is released into the extracellular matrix where

it exerts its effects through regulation of the biofilm. In this sense, it is its lack of targeting sequence

which directs it to the appropriate extracellular location.

1.3.3 Substrate recognition: Acceptor binding and the role of primases

In addition to lipid requirements, the acceptor is also recognized by the polymerase at the

carbohydrate terminus. Almost all carbohydrate polymerases recognize a short primer of sugar residues

constructed by one or more glycosyltransferases acting upstream.41,42 These primases have a conserved

role in glycan biosynthesis. They act as gatekeepers for their promiscuous polymerizing counterparts by

synthesizing the carbohydrate primer required for substrate recognition. One intriguing example of the

importance of priming glycosyltransferases involves the biosynthesis of wall teichoic acids (WTAs) in

Staphylococcus aureus and Bacillus subtilis.43 The glycosyltransferases which assemble teichoic acid exist

in large multienzyme complexes and produce polyphosphate chains on undecaprenyl carriers on the

cytosolic face of the membrane.44 The WTAs are polyribitol or polyglycerol phosphate polymers

anchored to the peptidoglycan of gram positive bacteria. Disruption of WTA biosynthesis results in

abnormal growth and septation defects.45 Because teichoic acids act as divalent cation reservoirs, their

absence may cause the altered phenotype via decreased glycosyltransferase activity. Since most

glycosyltransferases require a divalent cation for catalytic activity, having teichoic acids in close

proximity to sites of cell wall biosynthesis may be beneficial.46,47 Deficient cells also exhibit increased

Page 28: Design, Synthesis and Biological Utility of Polysaccharide

14 sensitivity to β-lactam antibiotics, implicating WTA polymers in the regulation of peptidoglycan

crosslinking.48

The teichoic acids of S. aureus and B. subtilis come in two varieties: the wall teichoic acids

attached to the peptidoglycan and the lipoteichoic acids, which are membrane anchored via a

phospholipid. Biosynthesis of WTAs in S. aureus occurs on an Und-P carrier on the cytoplasmic face of

the membrane.6 After a short disaccharide sequence is assembled on the lipid, the enzyme TagB

transfers a glycerol phosphate residue from CDP-glycerol to the C6-OH of the disaccharide. The enzyme

TarF then acts by transferring a second glycerol phosphate, and polymerization follows by the actions of

TarL.49

In B. subtilis the nomenclature for the WTA biosynthetic machinery is similar to S. aureus, but

the roles are somewhat shifted. After TagB transfers a glycerol phosphate to the same disaccharide

acceptor, the transferase TarK transfers a ribitol phosphate to the chain. Polymerization then

commences through the actions of B. subtilis’ TarL analog. Interestingly, the Walker Group identified an

analog of TarF in B. subtilis that was capable of generating oligosaccharides of glycerol phosphate in

vitro, however attempts to isolate WTAs with polyglycerol phosphate backbones were unsuccessful.49

1.3.4 Substrate recognition: Donor binding

By localizing polysaccharide synthesis to specific intra- or extracellular environments the

availability of component parts for polymer assembly is altered. The activated monosaccharide building

blocks found on the cytosolic side of the membrane are soluble nucleotide diphosphate (NDP) sugars

such as UDP-Glc or GDP-Man or CDP-glycerol (CDP-Gro). These NDP donors are recognized by

glycosyltransferases inside the cell. For catalysis of glycosylation, cytosolic glycosyltransferases have

evolved two strategies for binding and orienting the donor sugars. The first strategy uses acidic or polar

residues to coordinate a divalent metal cation like magnesium or manganese in the donor binding site.

Page 29: Design, Synthesis and Biological Utility of Polysaccharide

15 (Figure 1.06) The pyrophosphate moiety also coordinates to the metal, positioning the donor for

nucleophilic attack. The mycobacterial galactan, gram negative O-antigens in the ABC pathway and wall

teichoic acids are examples of polymers synthesized from NDP precursors. The second strategy is to use

a basic residue to position the donor pyrophosphate. On the cell exterior, NDP sugars are not present in

controlled amounts. Glycosyltransferases operating extracellularly favor a strategy of coupling two lipid-

linked glycans, using one as an acceptor and the other as a donor. Peptidoglycan synthesis, O-antigens

assembled via the Wzx/Wzy pathway and lipoteichoic acids favor this extracellular assembly strategy.

Figure 1.06. Glycosyltransferases bind NDP donor sugars by coordinating the pyrophosphate group, positioning it for nucleophilic displacement. In order to traffic polymers with the same composition to different destinations, organisms will

exploit the intracellular and extracellular assembly pathways. Consider the biosynthesis of lipoteichoic

acid and wall teichoic acid from B. subtilis. Both are linear repeating sequences of glycerol phosphate.

Lipoteichoic acids are anchored to the cell membrane via a phosphatidyl glycerol lipid. Wall teichoic

acids are ligated to the C6-OH of MurNAc residues on peptidoglycan. Lipoteichoic acids are assembled

on the extracellular side of the membrane using phosphatidyl glycerol lipids as donors for

polymerization. Wall teichoic acids are assembled in the cytosol on Und-P carrier lipids. This physical

Page 30: Design, Synthesis and Biological Utility of Polysaccharide

16 separation of biosynthesis is necessary for targeting to the appropriate structure and effected via the

utilization of chemically unique donors.

1.3.5 Substrate recognition: Determining the order of binding

In generic terms, a glycosyltransferase binds an acceptor and a donor and converts those two substrates

into two products: the glycosylated acceptor and the byproduct of donor hydrolysis. This reaction type is

termed a Bi Bi reaction and it can have several distinct kinetic mechanisms, the most common of which

are ping pong, sequential ordered or sequential random (Figure 1.07). In ping pong enzyme kinetics, the

enzyme binds one substrate followed by generation of the first product, then the second substrate binds

and the second product is made. In sequential ordered reactions, the binding of one substrate is

contingent on the binding of the second substrate and the release of the products follows a similarly

ordered pattern. Sequential random enzyme mechanisms are such that each substrate binds

independently of the other and the products are released in a similarly random order.

Figure 1.07. Common Bi Bi mechanisms proceed via a ping pong (top left), sequential ordered (bottom left) or random ordered (top right) binding mechanism. A fourth mechanism, Theorell-Chance, is a variant of ordered binding in which the kinetic parameters indicate turnover is rapid upon assembly of the ternary complex.

Page 31: Design, Synthesis and Biological Utility of Polysaccharide

17

Understanding the binding order of substrates for glycosylation is important for several reasons.

Crystals for x-ray crystallography, high throughput assays for inhibitors, or information regarding binding

surfaces can all be aided by this mechanistic inquiry. Enzyme kinetics can reveal which mechanism is

operational for a specific glycosyltransferase. This strategy has been successfully employed to determine

the mechanism of many glycosyltransferases.50-52 Many polymerizing glycosyltransferases employ a

sequential ordered mechanism.50-52 Attempts at isolating co-crystals and developing high throughput

inhibitor screens have also been aided by knowing substrate binding order.

1.4.1 Polymerization: Directionality

In polysaccharide synthesis, the polymerization event itself is mediated by one or more enzymes

that add a repeating sequence of residues to an acceptor. The sequence can be a simple homopolymer

of identical monosaccharides or a heteropolymer of repeating monosaccharides of varying complexity.

Elongation can occur on the cytoplasmic face of the membrane, on the extracytoplasmic face, or as a

nascent chain passages through the bilayer in the pore of a transmembrane-bound polymerase. Many

features of enzyme-specific polymerization events have been defined, including the direction of

elongation, processivity, length and patterning control. By defining the mechanism of polymer assembly

within these categories, researchers are able to focus on how to enhance, augment, re-engineer or

inhibit production of a polysaccharide. Complete understanding of the biosynthetic process promises to

provide tremendous leverage in our efforts to elucidate polysaccharide function.

A polysaccharide chain can be extended from its monomeric building blocks at either the

reducing or non-reducing terminus. The two mechanisms differ in which component acts as nucleophile

and which acts as electrophile, or in other words, which acts as donor and which as acceptor. (Figure

1.08) Experiments which determine elongation direction either rely on chemical block at one direction

or differentially label the reducing and non-reducing termini. The Walker Group determined the

Page 32: Design, Synthesis and Biological Utility of Polysaccharide

18 direction of elongation in peptidoglycan transglycosylation reactions by developing a compound which

blocked elongation at the non-reducing end.53 By swapping the native glucose residue for a galactose, a

subtle switch which preserved substrate binding, they generated a substrate that could only be

elongated from one direction. Observation of elongated products proved the directionality of growth to

be from the reducing end.53

The Yother group developed a strategy to distinguish between elongation directions of capsular

polysaccharide production in S. pneumoniae. In a variation on a pulse-chase experiment, the

bifunctional polymerase Cps3S was incubated with tritiated UDP-Galp and then with 14C-labeled UDP-

GalA.54 By treating the elongated polymers with a depolymerase known to act at the non-reducing

terminus of cellubiuronan polymers, they were able to isolate short oligomers from the reducing end of

depolymerized chains. They found that the reducing end was enriched in tritium indicating elongation

occurrs from the non-reducing terminus.54 Both examples highlight how careful selection of the

appropriate chemical label can yield insight into the mechanism of polymerization.

1.4.2 Polymerization : Processivity

A glycosyltransferase capable of polymerization can do so in a processive or distributive fashion. In

processive polymerization, the enzyme maintains contact with the nascent chain between individual

elongation events. In distributive polymerization, each intermediate-length chain dissociates from the

enzyme and must re-bind prior to subsequent elongation. The mode of elongation has consequences

for the regulatory features of polymer synthesis: length and pattern control. Distinguishing between

these mechanisms involves biasing the experimental conditions to statistically favor or disfavor binding

of an acceptor to the enzyme. Several strategies have been devised to elucidate processivity in

carbohydrate polymerases. Many of the techniques used to assess processivity employ isotopically-

Page 33: Design, Synthesis and Biological Utility of Polysaccharide

19

Figure 1.08. Polymer elongation can occur at the reducing or non-reducing end of the growing chain. In general, when both donor and substrate are lipid-linked glycans, elongation occurs at the reducing end. Enzymes which utilize NDP-sugar donors favor elongation at the non-reducing terminus. One method for determining elongation direction of cellubiuronan biosynthesis involved a double labeling experiment followed by treatment with a depolymerase known to act at the non-reducing terminus. The radioisotope enriched to a further degree on isolated oligomers suggested elongation occurred via the non-reducing terminus.

Page 34: Design, Synthesis and Biological Utility of Polysaccharide

20 labeled substrates. Classically, these experiments are similar to those performed by Messelson and Stahl

in which “heavy” and “light” DNA could be tracked to elucidate the mechanism of DNA replication.55,56

The Imperali Group investigated the processivity of PglH, a polymerase involved in N-glycan

assembly.51 PglH transfers three GalNAc residues to a primed Und-P carrier. PglH was incubated with

high concentrations of tritiated acceptor and then the +1, +2 and +3 adducts were separated and

quantified. If the enzyme-substrate complex dissociated after each glycosylation, then the chance of re-

binding the same acceptor molecule would be low. Thus, in this “single hit” assay, the observation of

long polymers at short time points indicated the enzyme was processive.51 Under conditions where the

starting acceptor concentration was high relative to enzyme concentration, the predominant product

was the +3 adduct at early time points, and the +1 adduct at late time points. Early accumulation of +3

product suggested that the enzyme is processive, but the number of sequential glycotransfers is

controlled through a product inhibition mechanism: As the concentration of +3 adduct increased, the

accumulation of +1 and +2 intermediates was afforded via competitive inhibition.51

An alternative strategy for assessing processivity was employed by the Yother Group to examine

the type 3 capsular polysaccharide from Streptococcus pneumoniae.2 The polysaccharide is a

cellubiuronan polymer anchored to the membrane by a phosphatidyl glycerol lipid.57 Isolated

membranes containing the membrane-bound polymerase Cps3S were treated in a pulse chase

experiment with 14C-labeled UDP-GlcA donor.54 The reaction products were separated by size exclusion

chromatography and each fraction was measured for radioactivity. Rapid increases in large molecular

weight polymer appeared at short time points (5 and 10 minutes). Identification of long polymers at

early time points suggested Cps3S was highly processive.

The Walker Group used a similar strategy to assess the processivity of peptidoglycan

glycosyltransferase PBP1A from Aquifex aeolicus.53 The enzyme condenses two molecules of lipid II to

Page 35: Design, Synthesis and Biological Utility of Polysaccharide

21 generate lipid IV and then continues to append disaccharides to the reducing end of the growing chain.

By incubating PBP1A with 14C-labeled lipid IV and a 100-fold excess of unlabeled lipid II, reaction

products could be observed using SDS-PAGE, which separates the polysaccharides by molecular weight.

The observation that large polymers were appearing at short time points indicated that PBP1A operates

via a processive mechanism. The crystal structure has been solved, and the authors interpreted a mobile

loop near the active site to be responsible for maintaining contact with the elongating substrate

between rounds of glycotransfer.18

The mycobacterial galactan is assembled on a decaprenyl carrier lipid and polymerized by the

glycosyltransferase GlfT2.58-61 The Kiessling Group has investigated the processivity of the enzyme using

several techniques. First, by incubating the enzyme with synthetic disaccharide acceptors at 1,000-fold

excess, polymer length was measured at very short time points.62 To quantify the degree of processivity,

an isotope labeling strategy akin to the classic Messelson-Stahl experiment was performed. In this assay,

GlfT2 is preincubated with normal acceptor for a time t1 before treatment with a heavy pentadeuterated

acceptor.56 If the enzyme is processive, the heavy acceptor will not have access to enzyme occupied by

the light acceptor. After a time t2, the product distribution for heavy and light acceptors can be

compared directly. (Figure 1.09) For each polymer elongated by n residues, a larger intensity peak for

the light isotope indicated that GlfT2 was highly processive. The processivity could be quantified by

measuring the ratio of heavy to light peak intensities.

Not every carbohydrate polymerase is processive. The Brown Group analyzed the polymerase

TagF, which generates the polyglycerol phosphate wall teichoic acid in B. subtilis.63,64 Under initial rate

conditions a 14C-labeled acceptor was found to be elongated by TagF in the presence of the donor, CDP-

Page 36: Design, Synthesis and Biological Utility of Polysaccharide

22

Figure 1.09. A distraction assay using isotopes analogous to the Messelson-Stahl experiments allowed quatitative assessment of GlfT2’s processivity.

Figure 1.10. TagF was shown to be perfectly distributive by demonstrating that saturating amounts of labeled acceptor generated the +1 product in an amount identical to the amount of hydrolyzed donor. glycerol. (Figure 1.10) Surprisingly, the amount of CMP produced was found to be equal to the amount

of +1 product. Because the starting acceptor was in 100-fold excess of the enzyme, the system was

biased against observing a rebinding event. Thus TagF operates by a distributive mechanism.52 The

question is raised concerning how a distributive polymerase can regulate length control without

Page 37: Design, Synthesis and Biological Utility of Polysaccharide

23 maintaining contact with the nascent strand and what evolutionary advantage is born out of a

distributive mechanism of action. Such advantages and discussed in this system below (vide infra).

1.4.3 Polymerization: Subsites and the role of primases

Processive enzymes must make transient contacts with the saccharide residue acted upon

chemically, as well as with several saccharide residues upstream of the action as a way to maintain

contact as the growing polymer slips over the surface. This is true for all processive enzymes, including

for example DNA and RNA polymerases and ribosomes. Efficient processive polymerization appears to

depend on subsite binding.62 This suggests an important role for primases in the assembly of a

polysaccharide. Perhaps these priming glycosyltransferases produce short oligosaccharide primer

sequences that fill the glycosyltransferase subsites. Acceptor substrates with appropriate priming

eliminate any kinetic lag phase and give rise to rapid polymerization.

The cellubiuronan capsular polysaccharide from Streptococcus pneumoniae is important for

colonization and survival of the pathogen. The polymerase Cps3S exhibits a transition from

oligosaccharide to polysaccharide synthesis triggered by generation of and binding to an oligosaccharide

approximately 6 – 8 residues in length.65 By incubating the enzyme and unlabeled acceptor with 14C

UDP-Glc, separating the products by paper chromatography and then quantifying by liquid scintillation

they observed oligomers rapidly reach a steady state levels while polysaccharide production lags at first

and then exhibits an exponential increase. The polymerase becomes highly processive once it binds an

acceptor approximately 8 monosaccharides in length.65,66 This length dependence suggests that

acceptors capable of filling subsites on the enzyme surface exhibit the highest level of processivity

because they are able to interact with – and thus remain bound to – the enzyme most efficiently.

Page 38: Design, Synthesis and Biological Utility of Polysaccharide

24 1.4.4 Polymerization: Length control

Regulation of polymer length is a central feature of polymer synthesis. Unlike their nucleic acid

and amino acid counterparts, carbohydrate polymerases act without a template to direct them. This

template-independent feature of polysaccharide synthesis raises the questions of just how the

polymerization regulates length. Polymer length imparts biological relevance in several systems. Chain

length is proportional to cation buffering capacity in teichoic acids.67 Differences in signal transduction

pathway activation are observed depending on the length of glycosaminoglycan present. The degree of

polymerization and crosslinking of the peptidoglycan renders cells resilient or sensitive to osmotic

pressures. How then, is length controlled? The structure of the polymerase itself always exerts some

level of control. There are features intrinsic to the enzyme, the substrates and the ternary complex they

form that dictates the resultant polymer length. Likewise there are extrinsic factors that impart a

regulatory function over polymer length. Finally, regulation at the gene expression level has an influence

as well.

1.4.4.1 Intrinsic factors

The Kiessling group recently put forth a model for length control based on a lipid tethering

mechanism in galactan assembly.62 Curiously, longer polymers were observed following incubation with

a lower concentration of acceptor. Moreover, increasing lipid length also resulted in longer

polymers.21,62 Inhibition with the geranylgeraniol lipid further supported a model wherein the enzyme

binds not only the carbohydrate portion of the acceptor, but the lipid itself in a bivalent manner.

Bivalent binding explains how template independent polymerases are able to regulate length by

increasing the likelihood of rebinding after passive dissociation of the glycan, particularly when the Galf

oligosaccharide is short. At longer galactan lengths, the efficacy of the tether would be reduced and

resemble monovalent binding, at which point a passive dissociation would terminate polymerization.

Page 39: Design, Synthesis and Biological Utility of Polysaccharide

25 Recently, a crystal structure of GlfT2 was solved.68 The structure has the classic Rossmann-type fold that

glycosyltransferases of the GT-A family share, but has no distinct lipid binding site. To reconcile the

structure with the lipid-dependence of the enzyme, it may be necessary to evoke some interaction with

the lipid bilayer. Indeed the membrane itself could serve as the lipid tethering site.

The Walker group discovered that the lipid tethering model is also valid for PBP1A, the

peptidoglycan glycosyltransferase which catalyzes the construction of the carbohydrate backbone of

peptidoglycan.53 To monitor the effect that lipid length would have on polymerization, the group

synthesized a galactosylated lipid II analog that was incompetent as an acceptor, but still bound the

enzyme as a donor. This allowed them to vary the chemical features of the donor and assess the effect

the changes had on polymerization. The lipid length of the donor, but not the acceptor, determines the

degree of polymerization. They postulate that the lipid tether enhances the processivity of the system.

Because PBP1A elongates from the reducing end, the nascent chain must translocate after each round

of glycosylation to accommodate a new lipid II acceptor. When the lipid length becomes too short,

polymer length decreases to oligomers only 4 – 8 carbohydrates in length.

The capsular polysaccharides of gram-positive bacteria are bound to the peptidoglycan or

anchored to the plasma membrane. Many of these polysaccharides resemble or are identical to the

glycosaminoglycan structures produced in mammalian cells.69 The hyaluronan capsular polysaccharide

from Streptococcus equisimilis is produced by a single enzyme, hyaluronan synthase (HAS). The

hyaluronan polymer has a repeating alternating sequence GlcNAc-β-(1,3)-GlcA-β-(1,4) which is

assembled from UDP donor sugars onto a short hyaluronan oligomeric primer. HAS is an integral

membrane protein that has a strong dependence on phospholipids for its activity.70 Radiation

inactivation studies were performed to determine the monomer is in complex with 16 cardiolipin

phospholipids in its active state.71,72 Polymer size, which can range from 0.2 to greater than 2 MDa, can

Page 40: Design, Synthesis and Biological Utility of Polysaccharide

26 be measured by a combination of size exclusion chromatography and laser light scattering.73 A western

blot based enzyme capture assay using a biotinylated hyaluronan binding protein in conjunction with

streptavidin-agarose allowed measurement of the percentage of active synthases present in the

membrane at any one time.74

The Weigel Group investigated the role of length control using a combination of sequence

analysis and mutagenesis.75 The predicted topology of HAS included two conserved polar residues, K48

and E327, within two transmembrane domains. Mutation at these conserved residues resulted in

severely reduced enzymatic activity and polymer length. Surprisingly, the double mutant K48E/E327K

restored activity, yet still produced shorter polymers. This led the authors to suggest that perhaps these

residues form a salt bridge that creates a ratcheting mechanism to regulate passage of the growing

chain through the channel during export.

1.4.4.2 Extrinsic factors

A second approach to length control can be found in the O-antigen assembly process, which is

part of the wzy/wzz system. The Wang Group has established biochemically that the wzz gene product

imparts a modality to length. Similarly, lipoteichoic acid length in B. subtilis is regulated by differential

gene expression of four synthases .76 Shigella flexneri uses variable expression of two Wzz proteins,

WzzB and Wzzp-HS2, to generate O-antigens with 17 or 90 repeating units, respectively.77 By using a dual

inducible promoter system, researchers were able to observe that production of the very long O-antigen

required concomitant high levels of Wzy polymerase expression. Low expression of Wzy favored the

shorter repeat. Cross-linking experiments were unable to identify a Wzy-WzzB binding interaction,

suggesting that length may be regulated by an indirect mechanism.77

Two hypotheses are discussed for how Wzz proteins modulate chain length: the molecular

ruler78 and the molecular stopwatch.79 The molecular ruler hypothesis is supported by the observation

Page 41: Design, Synthesis and Biological Utility of Polysaccharide

27 that Wzz oligomerizes into higher order structures on the membrane surface. These structures mark out

a predetermined polymer length by virtue of the quaternary structure, though supportive biochemical

evidence is lacking. The Whitfield group used full-length Wzz in lipid membranes and observed regular

hexamers by electron microscopy.80 This regularity may support the molecular ruler hypothesis by

negating earlier studies that demonstrated high variability in quaternary structure.

The molecular stopwatch hypothesis put forward by the Reeves Group suggests that Wzz

regulates length by binding either directly to the polymerase or to the polymer for a discrete amount of

time.79 Thus, the kinetics of the complex would directly impart a variation of length control, either

through stabilization or destabilization of the polymerization reaction. The authors expressed skepticism

that a single class of proteins could somehow measure the chain lengths to such variable degrees and

favor the stopwatch hypothesis, but acknowledge that experimental evidence that identifies the binding

interaction and correlates binding affinity to chain length would be necessary to confirm the stopwatch

model.

For the enzyme TagF from B. subtilis, an association with the membrane is necessary for length

control.63,81 Solubilized TagF was found to exhibit a constant Km and reaction velocity regardless of the

degree of polymerization of the acceptor substrate, generating very long polymers of glycerol

phosphate. Membranes from B. subtilis pre-treated with a general proteinase were found to mitigate

length control in TagF polymerization. This suggests that TagF interacts with membrane lipids, although

the specific interaction is not known at this time.

Finally, LTA biosynthesis in B. subtilis is less well understood.11,82 Genetic studies appear to

indicate that there are as many as four synthases expressed.76 One synthase, YvgJ, appears to act as a

primase, exhibiting low hydrolytic activity with a PG substrate and no detectable LTA production in vitro.

The three remaining synthases – YqgS, YfnI and LtaSBS— generate short, medium and long LTAs,

Page 42: Design, Synthesis and Biological Utility of Polysaccharide

28 respectively, but do so independently of the primase. In systems with all synthases expressed, medium

length LTAs predominate, suggesting that LtaSBS is the major active polymerase. Thus, polymer length

may be determined in part by differential expression. The Grundling Group hypothesizes that different

length LTAs may be beneficial to the cell depending on the different environmental stresses acting on it.

Figure 1.11. Intrinsic and extrinsic factors contribute to polysaccharide length control. Extrinsic factors include interactions with membrane lipids or protein cofactors. Intrinsic factors include conserved salt bridges or lipid binding sites.

Page 43: Design, Synthesis and Biological Utility of Polysaccharide

29 1.4.4.3 Length control by chemical chain termination

Termination of polymerization is often passive, requiring only a failure to elongate. This kinetic

model is defined by conditions in which the ternary complex becomes unfavorable. Often, termination

is a function of donor sugar concentration. When the concentration falls, elongation is aborted due to

unfavorable kinetics of bringing the acceptor and donor together in a productive turnover. For capsular

polysaccharide type 3, a nascent chain is ejected from the enzyme when the concentration of UDP-GlcA

or UDP-Glc is low. The Yother Group has demonstrated in detail how this dual donor specificity is

leveraged to finely tune the percentage of acceptor cellubiuronan that is elongated.83 Cps3S is

exquisitely sensitive to UDP-GlcA concentration and can be used to predict average chain length.

Chemical modifications to the polysaccharide are a second mechanism used to terminate

polymerization. For the E. coli O9a antigen, a competition between WbdB and WbdD determines

whether the chain is elongated (WbdB) or terminated with a dual phosphorylation-methylation event

(WbdD).84 The Whitfield Group used a fluorescent acceptor in an in vitro assay with purified WbdD to

demonstrate that chemical modification of a non-reducing mannose at the 3-position terminated

polymeriztaion. NMR studies definitively revealed a methylphosphate modification.85 The mechanism by

which this competition between elongation and termination is regulated is not understood, but it has

been shown that WbdA binds to WbdD.86 The product of WbdA is the substrate for WbdD. Thus,

sequestration of the chain terminating enzyme in the proximity of the elongating chain likely plays a part

in biasing the system toward chain termination.

The prokaryotic surface layer (S-layer) is a semi-crystalline array of protein that self assembles to

coat the surface of some bacterial cells.87 These proteins are subsequently modified with an O-glycan. In

Geobacillus stearothermophilus, a gram-positive bacterium, the O-glycan is polymer of rhamnose with a

Rha-α-(1,2)-Rha-α-(1,3)-Rha- β-(1,2) repeating sequence. Assembly of the polysaccharide takes place on

Page 44: Design, Synthesis and Biological Utility of Polysaccharide

30 the primed carrier lipid Rha-α-(1,3)-Rha-α-(1,3)-Gal-α-1-PP-Und on the cytoplasmic face of the lipid

bilayer.20 The rhamnan is terminated by a 2-OMe group on nonreducing Rha-α-(1,3)-Rha linkage. The

modularity of glycan biosynthesis coupled with the regularity of an S-layer array promises to be an

excellent platform for developing high throughput glycan interrogation technologies.88

The Messner and Schäffer Groups examined the putative glycosyltransferases present in the slg

gene cluster and assigned their functional roles.89 The undecaprenyl lipid is modified by WsaP, WsaD

and WsaC in turn to generate the Rha-Rha-Gal primer. A pair of enzymes, WsaE and WsaF, then act to

generate the repeating trisaccharide sequence. Incubation of the expressed proteins with synthetic octyl

glycoside acceptors revealed that WsaE is a multifunctional enzyme capable of generating the

regiochemical α-(1,2) and α-(1,3) linkages as well as the 2-O-methylation. Thus, WsaE must choose

between elongating a non-reducing Rha-α-(1,3)-Rha via α-(1,2) glycotransfer or chain termination by

methylation. The conditions under which this decision is made have yet to be explored.

1.4.5 Polymerization: Pattern control

One question to consider during polysaccharide assembly is what factors regulate the repeating

patterning during elongation. In template-dependent polymerizations, the sequence fidelity is dictated

by the template; when DNA polymerase encounters cytosine on the template, then guanine is added to

the nascent chain, and when the ribsosome encounters the codon GCC, an alanyl-tRNA binds and

alanine is added. For template-independent polymerases like polysaccharide synthases, the fidelity of

pattern deposition has not been evaluated.

The Kido Group has been studying a polymannose polymer present in the E. coli O9 antigen,

consisting of a repeating sequence of mannose composed of 3 α-(1,2) linkages followed by 2 α-(1,3)

bonds.90 The enzyme WbdA catalyzes the addition of the α-(1,2) linkages while WbdB installs the α-(1,3)

linked sugars. However, a point mutation in WbdA, R55C, confers a specificity slip in which the mutant

Page 45: Design, Synthesis and Biological Utility of Polysaccharide

31 can only transfer 2 residues with α-(1,3) regiochemistry.90 This suggests that the tertiary structure of

glycosyltransferases is carefully tuned to deliver a certain number of residues.

The polymerase GlfT2 is a bifunctional enzyme with a single active site.91 Bifunctional

glycosyltransferases are capable of generating more than one chemical linkage. This bifunctionality can

arise from transfer of two different monosaccharides, or transfer of the same monosaccharide but with

different regio- or stereoselectivity. The Kiessling group developed a method to monitor sequence

fidelity for bifunctional enzymes that alternate between regiochemical linkages using a chain

termination agent strategy.92 The strategy requires synthetic donors in which one of the hydroxyl groups

that participate in glycosidic bond formation is substituted with a fluorine atom while the other hydroxyl

group is left intact. The fluorine substitution is a close stereoelctronic mimic of the hydroxyl group, so

perturbations to substrate binding are minimized. Incubation of these mimics with a bifunctional

enzyme followed by mass analysis for chain termination can reveal the enzyme’s linkage fidelity: Chain

termination would suggest that the enzyme cannot deviate from the alternating linkage pattern,

whereas polymer detection would suggest that it possesses low fidelity.

Other systems with alternating linkages and bifunctional polymerases could be analyzed using

similar chemistry.93-95 The polysialyltransferase from E. coli capsular polysaccharide assembly alternates

between α-(2,8) and α-(2,9) NeuAc residues. When incubated with 9-azido CMP-NeuAc donors,

polymerization has been shown to terminate.10,96 The Withers Group has recently developed a synthesis

for 8- and 9-fluoro CMP-Sialic acid donors suggesting the technology is present to interrogate the fidelity

of the polysialyltransferase.97 Another capsular polysaccharide, O54 serovar Boreze from Salmonella

enterica, has alternating β-(1,3) and β-(1,4) ManNAc linkages. Polymerization by the enzyme WbbF is

another intriguing target for this approach.93

Page 46: Design, Synthesis and Biological Utility of Polysaccharide

32

Figure 1.12. The fidelity of GLfT2 was assessed using fluorinated chain termination agents. Analysis of terminated products suggested GlfT2 is chaste with respect to alternating linkage formation. A point mutation in WbdA was found to cause a fidelity slip in the construction of a polymannan O-antigen suggesting that small intrinsic structural features on the glycosyltransferases dictate their ability to orient a nascent chain for appropriate patterning.

1.5 Conclusions

The structural and functional requirements of the biosynthetic machinery engaged in

polysaccharide assembly are varied and complex. Inquiry into the mechanistic detail of many such

systems has yielded an enzyme by enzyme picture of the process. Decades of groundwork has allowed

us to line up the events like movie stills and infer the machinations of these polymerases. An effort has

been made here to categorize the mechanistic underpinnings of polymer construction. Binding order,

processivity, length control, pattern control and termination are important features of biosynthesis

Page 47: Design, Synthesis and Biological Utility of Polysaccharide

33 because they help us identify discrete events surrounding polymerization that are not yet understood in

toto. Our ability to piece together the molecular mechanisms that regulate polymerization will depend

on leveraging our increasing level of comfort with these technically and chemically challenging systems.

The research frontier should have many fronts: we still have only a partial understanding of the

biological role of polysaccharides, partially because we have yet to connect biosynthesis to structure and

structure to function. With our increasing ability to manipulate and synthesize non-natural variants, we

should be able to delve deeply into the importance of carbohydrate polymer structure and relate it to

function. Chemical biology methods which combine synthetic compounds with well-defined biochemical

systems have proven to be a powerful technique for understanding the step by step assembly process.

Continuing this process and applying it to multiprotein complexes and systems as well as to glycan

engineering projects will be essential.

The future promises to herald some important developments for bacterial polysaccharide

research and technology. Perhaps the most exciting is that our understanding of the requirements for

polysaccharide biosynthesis paves the way for glycan engineering. Our ability to install the glycan of our

choosing onto the protein or array of our choosing will allow for the generation of high throughput

platforms. These platforms are suited to the “omics” world where large amounts of data generated

rapidly and cheaply can be interrogated for discovery. The medical community will benefit from this on

two fronts. Our ability to screen for and identify glycans of medical importance will give rise to new

vaccination strategies and novel approaches to antibiotic treatment.98 The effects polysaccharides have

on virulence and pathogenesis will be better understood through a better understanding of their

biosynthesis. Lastly, research that sheds light onto how multiprotein complexes create and transport

these biomacromolecules will come as new strategies and technologies for studying complex systems

are developed.

Page 48: Design, Synthesis and Biological Utility of Polysaccharide

34

Chapter 2

Design and Synthesis of Fluorinated UDP-Galf Analogs

Portions of this work are published:

Brown, Christopher D., Rusek, Max S. and Kiessling, Laura L. Fluorosugar Chain Termination Agents as Probes of the Sequence-Specificity of a Carbohydrate Polymerase. J. Am. Chem. Soc. 2012, 6552-6555

Contributions

Max Rusek synthesized provided scale-up quantities of vinyl Galf intermediates

Kenzo Yamatsugu provided all trisaccharide acceptors and the synthetic tetrasaccharide

Page 49: Design, Synthesis and Biological Utility of Polysaccharide

35 2.1 Introduction

The diversity of carbohydrate polymers in Nature is vast. Polysaccharide structure is defined by

the basic repeating series of sugars that comprise its sequence. Polysaccharide sequences ultimately

determine the functional properties of the polymer. Consider, for instance, the simple case of glucose

(Glc) homopolymers. Cellulose, a structural polymer from plants, is a homopolymer of Glc with β-(1,4)

linkages. Glycogen, a polymer utilized by cells for energy reserves, is also a glucan homopolymer, but

with α-(1,4) linkages. Zymosan is a glucan from yeast that is routinely used as a TLR2 agonist, in which

the glucose monomers are linked via β-(1,3) bonds. Finally, chitin, a structural component of insect

exoskeletons, is a homopolymer of N-acetyl glucose (GlcNAc) with β-(1,4) glycosidic bonds. Each subtle

structural change dramatically changes the function of the polymer. (Figure 2.01)

Figure 2.01: Homopolymers of glucose. Minor structural variations in the polymer impart dramatically different functions to a sugar chain.

At the other extreme, polysaccharides such as bacterial O-antigens can have complicated and

exotic structure with upwards of eight monosaccharides linked together to generate the basic repeating

pattern. Polysaccharides with alternating linkage motifs are common. Most glycosaminoglycans, like

hylauronan for instance, are built on an alternating pattern of N-acetyl and uronic acid functionalized

Page 50: Design, Synthesis and Biological Utility of Polysaccharide

36 hexoses. The capsular polysaccharide from E. coli K92, which has been investigated for meningococcal

vaccine development, alternates between α-(2,8) and α-(2,9) sialic acid linkages. Agarose, an industrially

important gelling agent, is a polymer of alternating D- and L-galactose derivatives. The O-antigen from S.

enterica O54 serovar Boreeze is a polymer of ManNAc sugars with alternating α-(1,3) and α-(1,4) bonds.

The glycoprotein α-dystroglycan requires an oligomer of alternating xylose and GlcA residues to function

properly. (Figure 2.02)

Figure 2.02: Heteropolysaccharides with alternating sugars or linkages.

For a long time, the field of carbohydrate research held a paradigm: one enzyme, one linkage.

The thinking went that for each glycosylation, a unique glycosyltransferase recognizes the cognate

substrate, binds the appropriate donor sugar and appends that sugar to the acceptor substrate with

regio- and stereochemical control. For each glycosidic bond found in a glycan, a different enzyme would

be responsible for its construction. This rule is broken for glycosyltransferases which are able to act

upon more than one substrate or bind and utilize more than one donor sugar. Some

heteropolysaccharide chains which alternate between two monosaccharides or linkage types are

synthesized by these bifunctional glycosyltransferases. For these enzymes, questions are raised as to

how the two catalytic activities are used in tandem to generate the alternating pattern.

Page 51: Design, Synthesis and Biological Utility of Polysaccharide

37 2.2 GlfT2: A bifunctional polymerase from Mycobacterium tuberculosis

To study the patterning fidelity of polymerizing bifunctional glycosyltransferases we used the

enzyme GlfT2 as a model system. GlfT2 is a galactofuranosyltransferase from Mycobacterium

tuberculosis.61 It polymerizes a portion of the mycobacterial cell wall termed the galactan. The galactan

is a linear polysaccharide comprised of alternating β-(1,5) and β-(1,6) linked galactofuranose (Galf)

residues. It is built on a primed decaprenyl carrier lipid on the cytosolic side of the membrane, exported

via a putative ABC transport mechanism, modified with additional carbohydrate and lipid structures and

ultimately ligated to the peptidoglycan.58 (Figure 2.03) Besides tethering the mycolic acids to the

peptidoglycan, the functional significance of the galactan is unknown. Deletion studies identify it as an

essential gene: deletion mutants are not viable.99,100 Small molecule inhibitors of the mutase UGM,

which generates UDP-Galf, were shown to be bactericidal.101 Thus, understanding more about the

molecular mechanisms of galactan polymerization is the first step toward designing high throughput

screens and ultimately inhibitors which could serve as new antibiotics.

The galactan was identified as part of a serologically active compound with non-reducing

arabinose residues in the 1930’s. Not until the early 90’s though did GC-MS and FAB-MS experiments

help identify the structure of the polymer as exclusively furanose, the thermodynamically less stable

isomer of galactose. Early biochemical studies using overexpressed enzymes in membrane preparations

indicated that the gene product of Rv3808c (GlfT2) was a glycosyltransferase that utilized UDP-Galf to

modify a membrane phospholipid acceptor.102 Later studies would reveal that synthetic acceptors with

terminal β-(1,5) or β-(1,6) linkages were elongated by up to 4 Galf sugars. Chemical analysis of the +1

products for a β-(1,6) linked disaccharide suggested that the new linkage was β-(1,5). Analysis of the +1

Page 52: Design, Synthesis and Biological Utility of Polysaccharide

38

Figure 2.03: The galactan is synthesized on a decaprenyl pyrophosphate carrier lipid on the cytosolic face on the membrane. Following polymerization, the polymer is transported to the extracellular face and modified with arabinan and mycolic acids.

product for a β-(1,5) linked disaccharide revealed it to be a β-(1,6) linkage suggesting that there was an

intrinsic ability of the enzyme to set the alternating linkage pattern.59 Expression and purification of

GlfT2 in the Lowary Group confirmed these activities were specific to GlfT2.41 The primase GlfT1

was discovered through careful experiments with isolated natural acceptors which showed that that

native decaprenyl pyrophosphate linked disaccharides and trisaccharides were not substrates for GlfT2,

but were elongated by Rv3782 (GlfT1).61,103 The product tetrasaccharide was found to be elongated to a

pentasaccharide by GlfT2.61 (Figure 2.03) Overall, the biosynthetic sequence of events has been

accepted, however the mechanisms that underlie the regulation of galactan biosynthesis remain

somewhat more mysterious.

Page 53: Design, Synthesis and Biological Utility of Polysaccharide

39 2.3.1 Synthetic acceptors

The putative natural acceptor of the polymerase GlfT2 is a tetrasaccharide with the structure

Galf-β-(1,5)-Galf-β-(1,4)-Rhap-α-(1,3)-GlcNAc-α-decaprenyl pyrophosphate. The efforts to isolate this

natural product from biological sources have not been high yielding. Additionally, the compound is likely

somewhat unstable and susceptible to hydrolysis at the pyrophosphate bond. For reasons of synthetic

ease and acceptor chemical robustness, efforts in many labs have yielded simplified synthetic GlfT2

acceptors for activity-based study. Generally, GlfT2 activity has been assessed using disaccharides,21,59

trisaccharides61,104,105 and tetrasaccharides106 bearing octyl, decenyl or phenoxy-capped dodecenyl lipids.

This variety of viable structures that GlfT2 acts upon is a convenient way to leverage the promiscuity of

the enzyme for rapid structure-activity relationships. (Figure 2.04)

2.3.2 Synthetic donors

The first synthesis of UDP-Galf was reported by Tsvetkov in 2000 and employs a stereospecific

phosphorylation-phosphate coupling sequence to install the UDP group.107 The strategy for generating

the donor has been improved synthetically108,109 and chemoenzymatically110,111 since then, but the

overall approach has remained similar to the original report. The de Lederkremer lab reported a

synthesis of a C6-3H-labeled UDP-Galf synthesis which has seen moderate use as a measureable probe in

activity assays.112 Similarly, a 14C-labeled UDP-Galf is routinely generated chemoenzymatically by

incubating UDP-14C-Galp with UGM. We noted the paucity of examples with substitution at the C5

position. Fewer reports have been published on the permissiveness of GlfT2 toward donor modification,

but it seemed clear that the synthetic accessibility of acceptor and donor substrates makes possible the

study of galactan biosynthesis using a chemical approach. (Figure 2.05)

Page 54: Design, Synthesis and Biological Utility of Polysaccharide

40

Figure 2.04. Representative disaccharide and trisaccharide synthetic acceptors recognized by GlfT2 for elongation. Synthetic acceptors with simplified chemical structure have allowed interrogation of GlfT2 activity.

2.4 Chemical probes of pattern fidelity in a bifunctional polymerase

Considering the synthetic feasibility of generating donors and acceptors with specific functional groups,

we returned to the question of pattern fidelity in a bifunctional carbohydrate polymerase. For an

enzyme that can generate two distinct glycosidic linkages in an alternating and repeating pattern, what

regulates the deposition of the sequence, and is polymerization dependent on appropriate patterning?

If an activated glycoside with a bond-blocking functional group was recognized by the enzyme and

Page 55: Design, Synthesis and Biological Utility of Polysaccharide

41

Figure 2.05. UDP-Galf analogs reported in the literature. Isotopically labeled sugars and substitutions at the C2, C3, C5 and C6 positions have been reported sporadically. Reports of substitution at C5 are lacking.

incorporated into a nascent chain, would the enzyme polymerize a galactan chain with an “incorrect”

pattern? We envisioned such an assay would serve as a measurement of the enzyme’s patterning

fidelity by observing the product profile of a reaction with a non-natural donor. Short polymers would

indicate the enzyme was chaste with respect to pattern deposition whereas long polymers would reveal

its linkage promiscuity. (Figure 2.07)

Figure 2.06. Fidelity assessment assay.

Page 56: Design, Synthesis and Biological Utility of Polysaccharide

42

Table 2.01. Fluorine is an excellent mimic of the hydroxyl group. Size, shape and electronic contributions to molecular structure and conformation are well watched.113

In determining what the X-group should be, we considered several features of the mimic that

would improve our chances to successfully assay polymerization. Our X group had to adequately

substitute for a hydroxyl to maximize the chances that GlfT2 would recognize the synthetic donor as a

substrate. The substitution of fluorine in place of a hydroxyl group has several advantages. (Table 2.01)

First, fluorine is small relative to a hydroxyl group, so it is a minimal perturbation from a sterics

perspective. In addition, the bond lengths for C-F and C-O are identical. Perhaps the most promising

feature of a fluorine substitution is the stereoelectronic properties. Both C-O and C-F bonds are highly

polarized, as evidenced from Pauling electronegativity values of 3.5 and 4.0, respectively. Strongly

electron withdrawing groups vicinal to hydroxyl groups favor gauche interactions, so it is likely that the

conformational preference of a fluorosugar would be similar to that of a hydroxyl group. Further

evidence in support of this comes from modeling and spectroscopy of 2-fluoroethanol and ethylene

glycol which demonstrates that the conformation and even the hydroxyl proton orientation are

identical.114-115The inductive effects on the vicinal hydroxyl group are not only conformational: they also

Page 57: Design, Synthesis and Biological Utility of Polysaccharide

43 influence hydroxyl pKA. Since hydroxyl group pKa is thought to be important to the enzyme’s

mechanism of action91 it seemed prudent to select an electron withdrawing functional group that would

mimic the properties of a vicinal diol.

Fluorosugars have been used as mechanistic probes to study glycosyltransferases. The location

of the fluorine on the sugar can have a dramatic impact on the glycosyltransferase activity. The Withers

Group investigated the effect a C2 or C5-Fluorinated glucopyranose donor would have on the

oxocarbenium transition states of glycosylation reactions and discovered that inductive effects were

destabilizing.116 This physical property rendered fluorosugars useful as competitive inhibitors. Metabolic

precursors of fluorosugars can block glycan formation by disrupting nucleotide sugar biosynthesis.

Fluorinated carbohydrates have been used in PET and 19F NMR studies to understand protein-

carbohydrate interactions.117 The evidence of biological compatibility suggested that fluorinated Galf

donors would serve as ideal substrates for GlfT2 interrogation.

2.5 Synthesis of fluorinated Galf analogs

Figure 2.08. Retrosynthesis of UDP-6F-Galf and UDP-5F-Galf to a common intermediate, a vinyl Galf.

2.5.1 Accessing vinyl Galf

We required access to both C6 and C5 fluorinated Galf donor structures 2.01 and 2.02. A 6F-Galf

substrate was accessed by generating a cyclic sulfate from a Galf diol using a published method.

(Ferrieres) The route is useful for substitution at the C6 position, but is difficult to access C5-modified

Page 58: Design, Synthesis and Biological Utility of Polysaccharide

44 Galf structures. To our knowledge, there are no reports of a 5-fluorinated UDP-Galf. Thus, we devised a

divergent synthetic route that could generate both 5- and 6- fluorosubstituted regioisomers 2.03 and

2.04 from a common intermediate, vinyl Galf 2.05. (Figure 2.08) Accessing the common intermediate

involved a short synthesis from commercially available α-methylgalactopyranose 2.06. Benzylidene

protection of the 4- and 6-OH groups followed by benzyl protection afforded fully protected pyranose

intermediate 2.08. Selective removal of the benzylidene acetal in the presence of the methyl acetal

using iodine in refluxing methanol afforded diol 2.09. Regioselective iodination of the primary alcohol

generated the key intermediate 2.10 for accessing the vinyl furanose scaffold. By treating the

Scheme 2.01. Synthesis of the vinyl Galf intermediate.

iodopyranose with n-butyl lithium and carefully controlling the temperature of the reaction, the alkyl

lithium underwent a self-elimination reaction in excellent yield. Separation of the anomers 2.12 and

2.13 was accomplished by treating the cyclic hemiacetal mixture 2.11 with methyl iodide under phase

transfer conditions (Scheme 2.01). Their isolation allowed access to the vinyl Galf intermediate that we

anticipated could be acted upon by a variety of chemical transformation and allow us to generate non-

natural furanosides by functionalizing the double bond.

Page 59: Design, Synthesis and Biological Utility of Polysaccharide

45 2.5.2 Setting the stereochemistry at C5

Selective chemical transformations at a specific position are the central challenge for

carbohydrate chemists. While it is possible to differentiate the C6 position based on sterics and the C1

position on oxidation state, it is challenging to selectively operate on the secondary alcohols at C2, C3

and C5. Strategies to circumvent this reactivity problem include either a de novo approach118,119 or tying

the reactivity of the either the C6 or C1 position to one of the undifferentiated positions. The

benzylidene acetal protecting group, which ties together the C6 and C4 positions, is one example of this

approach. The Danishefsky group has pioneered a similar strategy for glycals, which allows selective

functionalization at C2. Other strategies are pervasive in the literature.109,120,121 The vinyl Galf is a

powerful addition to this strategy by allowing selective chemical transformation at the C5 and C6

positions of hexofuranosides.

We recognized a key challenge of this divergent synthesis would be the need to develop

diastereoselective reactions to both C5 epimers. Because nucleophilic fluorination operates via the SN2

mechanism, a synthetic intermediate with the opposite C5 stereochemistry is required for C5

fluorination. The asymmetric fluorination is still a transformation at the synthetic frontier.122 To

Scheme 2.02. Kinetic separation of C5 epimers.

Page 60: Design, Synthesis and Biological Utility of Polysaccharide

46 circumvent this limitation, we adapted the classic hydrolytic kinetic resolution to effect a kinetically

controlled separation of diastereomers.123 The resolution would have two purposes: It would allow us to

avoid stereoselective additions, and it also provided a chromatographic way to separate the two.

Epoxidation of common intermediate 2.12 using mCPBA generated 2.14 as an inseparable 1:1

mixture of diastereomers. Treatment of this mixture with Jacobsen’s cobalt salen catalyst and 0.5

equivalents of water resulted in clean conversion of the S-diastereomer to the diol (2.16), leaving the R-

diastereomer (2.15) unreacted. Diastereoselectivity varied from 10:1 to > 20:1. The diol and epoxide

were easily separated by silica chromatography. This kinetically controlled separation of diastereomers

allowed for isolation of both stereodefined intermediates necessary to complete the synthesis without

necessitating a large screen for two diastereoselective transformations. (Scheme 2.02)

Scheme 2.03. Installation of fluorine at C5 and C6.

2.5.3 Installation of fluorine

To install fluorine at the C6 position, R-epoxide 2.15 was treated with HF-TEA complex in a

completely regioselective transformation to access fluorinated Galf derivative 2.17. (Figure 2.03)

Previous reports on sugar epoxides had indicated that this acid-catalyzed epoxide opening proceeds via

nucleophilic attack at the less hindered carbon. Hydrogenolysis afforded 6F-methyl-Galf 2.03. The

Page 61: Design, Synthesis and Biological Utility of Polysaccharide

47 installation of fluorine at the C5 position of Galf proceeded from the diol 2.16. Treatment of the diol

with dibutyl tin oxide in refluxing toluene generated the nucleophilic cyclic tin acetal which was trapped

with benzyl bromide to generate a key intermediate with only the C5-hydroxyl free (2.18). Treatment of

2.18 with diethylamino sulfur trifluoride (DAST) generated a 5-fluoro furanose (2.19). To confirm the

stereochemistry at the C5 position, a benzoylated derivative of 2.04 was prepared and compared to the

hydroxylated analog our group had made previously.21 (Figure 2.09) The 3JHH coupling constant between

H4 and H5 was identical. Because the magnitude of a coupling constant is proportional to the dihedral

angle between nuclei, this suggested that the fluorofuranose had Galf stereochemistry at the C5

position.

2.5.4 Installation of phosphate

Fluorosugar intermediates in hand, we turned to the synthetic endgame: installation of the UDP group.

The two main challenges for this part of the synthesis are controlling the anomeric stereochemistry and

working with the hydrolytic lability of sugar phosphates and pyrophosphates. 6F-methyl-Galf 2.03 was

perbenzoylated using benzoyl chloride in pyridine to generate protected intermediate 2.20. Acetolysis of

2.20 removed the stable methyl acetal in favor of the more activated acetyl anomer 2.21. Treatment

with HBr generated a Galf bromide 2.22. The bromide was displaced with dibenzyl phosphate to form a

protected Galf monophosphate with 4:1 selectivity for the α-anomer (2.23) Hydrogenolysis and mild

saponification yielded 6F-Galf monophosphate 2.25 in 78% yield over two steps. (Scheme 2.06) An

analogous sequence was performed on 5-fluoro Galf intermediate 2.04 to generate 5F-Galf

monophosphate 2.31 (Scheme 2.07)

Page 62: Design, Synthesis and Biological Utility of Polysaccharide

48

Figure 2.09. A comparison of 1H NMR spectra between bezoylated Galf (left) and 5-fluorofuranose 2.xx (right) revealed both H4 signals had identical coupling constants to H5. Since the magnitude of the coupling constant is proportional to the dihedral angle, the equivalent values suggest the stereochemistry at C5 is also equivalent.

2.5.5 Phosphate coupling

Coupling of Galf phosphates to UMP completed the synthesis of UDP-Galf derivatives. The Kiessling

group had recently published an improved coupling strategy over the original UDP-Galf synthesis that

used an N-methylimidazolium-activated UMP (2.35).108 To generate the N- methylimidazolium, UMP

(2.34) was treated with freshly distilled trifluoroacetic anhydride followed by N-methylimidazole. This

activated UMP solution was added to a solution of Galf phosphate and stirred at room temperature for

4 hours. For the reaction to proceed, it was necessary to add 4+ equivalents of UMP. Unfortunately,

purification by HPLC was hindered by a UMP byproduct that coeluted with the UDP-6F-Galf product. To

circumvent this problem, a different approach to generate activated UMP was persued using

Page 63: Design, Synthesis and Biological Utility of Polysaccharide

49

Scheme 2.04. Stereoselective phosphorylation of 6F-Galf.

Scheme 2.05. Stereoselective phosphorylation of 5F-Galf

Page 64: Design, Synthesis and Biological Utility of Polysaccharide

50 benzenesulfonylimidazolium triflate (BSIT).124 BSIT is advantageous over the traditional activation

because it does not use trifluoroacetic anhydride and allowed the equivalents of UMP to be dropped

from 4 to 2. This allowed separation of the UDP-Galf products (2.01 and 2.02) in ~ 10% yield, a 4-fold

improvement. (Scheme 2.08)

Scheme 2.06. Phosphate coupling to generate UDP-Galf analogs.

The yield for generating UDP-Galf using this coupling strategy is typically around 35%. The lower

isolated yields for the fluorinated analogs may be a result of inductive effects on the nucleophilicity of

the Galf phosphates. In future, it may be prudent to abandon the synthetic methodology in favor of a

chemoenzymatic approach. The Field Group reported that an uridyltransferase could generate UDP-Galf

from Galf monophosphate in 70% yield110 and others have used the system to generate UDP-6F-

Galf.111,125 Another potential chemoenzymatic synthesis involves generating a Galf aryl glycoside that

Page 65: Design, Synthesis and Biological Utility of Polysaccharide

51 may be accepted as a substrate for the uridyltransferase.126 For this purpose however, the material

isolated was used in subsequent biological studies.

In summary, we have developed a divergent synthesis to install fluorine atoms as hydroxyl

group mimics. This strategy allowed access to both the C6 and C5-fluorinated donor sugar analogs of

UDP-Galf. Central to the success of the synthetic route was our use of a kinetic separation of epoxide

diastereomers using the Jacobsen cobalt salen catalyst in a reaction analogous to a hydrolytic kinetic

resolution. This allowed isolation of stereodefined intermediates that could be elaborated in parallel to

the fluorosugar targets. Such a strategy promises to be general for the synthesis of non-natural

furanoside sugars.

2.6 Experimental Procedures

I. General Procedures and Materials

All compounds were purchased from Sigma Aldrich (Milwaukee, WI) or Fischer Scientific (Pittsburgh,

PA). Tetrahydrofuran (THF) and toluene were distilled from sodium/benzophenone ketyl.

Diisopropylethylamine (DIEA), triethylamine (TEA) and dichloromethane (CH2Cl2) were distilled from

calcium hydride. All reactions were run under nitrogen atmosphere unless otherwise stated. Analytical

thin layer chromatography (TLC) was carried out on E. Merck (Darmstadt) TLC plates pre-coated with

silica gel 60 F254 (250 μm layer thickness). Analyte visualization was accomplished using a UV lamp and

charring with p-anisaldehyde solution. Flash chromatography was performed on Scientific Adsorbents

Incorporated silica gel (32-63 μm, 60 Å pore size) using distilled reagent grade hexanes and ACS grade

ethyl acetate (EtOAc) or methanol, CH2Cl2 and acetic acid. UDP sugars were purified by semi-preparative

HPLC using a Dionex carbopak column eluting in a 300mM triethylammonium acetate buffer at neutral

pH.1H. 19F, 31P and 13C nuclear magnetic resonance (NMR) spectra were recorded on Bruker AC-300 or

Page 66: Design, Synthesis and Biological Utility of Polysaccharide

52 Varian Inova-500 spectrometers, and chemical shifts are reported relative to tetramethylsilane or

residual solvent peaks in parts per million (CHCl3: 1H: 7.26, 13C: 77.0; MeOH: 1H: 3.31, 13C: 49.15; D2O: 1H:

4.79, 13C: referenced to 1H. 19F and 31P referenced to 1H) Peak multiplicity is reported as singlet (s),

doublet (d), doublet of doublets (dd), triplet (t), doublet of triplets (dt), etc. High resolution electrospray

ionization mass spectra (HRESI-MS) were obtained on a Micromass LCT.

II. Experimental Procedures

4,6-O-benzylidene-α-methylgalactopyranose (2.07): α-methylgalactopyranose (10g, 50.5 mmol) was

dissolved in DMF (100mL) in a 1L flame dried RBF and treated with camphor sulfonic acid (1.2g, 5.05

mmol) under N2. Benzaldehyde dimethyl acetal (9.5mL, 60.6 mmol) was added dropwise and the

reaction was stirred at room temperature overnight. The first reaction can be quenched with methanol

and the product precipitated upon addition of IPA to characterize the product. Alternatively, the DMF

solution can be treated with the reagents for the subsequent reaction directly in a two-step one pot

procedure.

2,3-di-O-benzyl-4,6-O-benzylidene-α-methylgalactopyranose (2.08): The solution of benzyldiene acetal

in DMF from the previous reaction was chilled to 0 °C and treated with an NaH mineral oil suspension

(12.6g, 328 mmol). Addition was paced so that the bubbly emulsion did not overflow the container and

stirring remained possible (the reaction gets thick). Benzyl bromide (46mL, 328 mmol) was added

dropwise, the reaction was allowed to warm to room temperature and stirred overnight. The reaction

was quenched with methanol and place in a freezer overnight. White crystals formed. Filtration and

washing with cold methanol yielded 13.8g of pure product. (59% yield over 2 steps) 1H NMR (300 MHz,

CDCl3, δ (ppm), 13C NMR (75 MHz, CDCl3, δ (ppm))

Page 67: Design, Synthesis and Biological Utility of Polysaccharide

53 2,3-di-O-benzyl-α-D-methylgalactopyranose (2.09): 2,3-di-O-benzyl-4,6-O-benzylidene-α-D-

methylgalactopyranose (15.6g, 33.73 mmol) was dissolved in freshly distilled methanol (250mL) under

N2. Iodine (3.5g, 13.5 mmol) was added and the contents were raised to reflux and stirred overnight. The

contents were cooled to RT and quenched with saturated Na2S2O3 solution until the brown color

dissipated. Methanol was removed by rotovap and the crude mixture was extracted 3x with DCM, dried

over MgSO4, filtered and concentrated in vacuo. Purification by silica chromatography (15% 60%

100% EtOAc in hexane) gave 10.2g of white solid. (81% yield) 1H NMR (300 MHz, CDCl3, δ (ppm), 13C

NMR (75 MHz, CDCl3, δ (ppm))

2,3-di-O-benzyl-6-deoxy-6-iodo-β-D-methylgalactopyranose (2.10): 2,3-di-O-benzyl-β-D-

methylgalactopyranose (4.4 g, 11.8 mmol) was dissolved in dry THF (100 mL) under N2 atmosphere.

Triphenylphosphine (4.6 g, 17.6 mmol) and imidazole (2.5 g, 36.6 mmol). The flask was fitted with a

flame dried condenser and the contents warmed to 50 ˚C. A solution of iodine (4.5 g, 17.6 mmol) in dry

THF (20 mL) was added dropwise over 2h. Once the color of the reaction was a persistent yellow the

contents were cooled to RT, quenched with saturated Na2S2O3 solution until the yellow color dissipated.

THF was removed in vacuo then the contents were diluted with water, extracted 2x into DCM, dried

over MgSO4, filtered and concentrated in vacuo. Purification by silica chromatography (gradient 10%

40% EtOAc in Hex) gave 4.61 g of a colorless oil (81%). 1H NMR (300 MHz, CDCl3, δ (ppm)) 7.42 – 7.22 (m,

10H, Ar-H), 4.89 (d, J= 11.1, 1H, OCH2Ph), 4.73 (m, 3H, OCH2Ph, OCH2Ph), 4.26 (d, J= 7.7, 1H, H-1), 4.11

(dd, J= 3.3, 0.9, 1H, H-4), 3.59 (s, 3H, OCH3), 3.61 – 3.46 (m, 2H, H-2, H-5), 3.50 (dd, J= 9.4, 3.5, 1H, H-3),

3.41 (m, 2H, H-6a, H-6b); 13C NMR (75 MHz, CDCl3, δ (ppm)) 138.75, 137.88, 128.73, 128.53, 128.24,

128.07, 127.85, 124.71, 80.59, 78.70, 75.24, 74.87, 72.96, 67.41, 57.28.

Page 68: Design, Synthesis and Biological Utility of Polysaccharide

54

2,3-di-O-benzyl-5-eno-β-D-methylgalactofuranose (2.12): Iodogalactopyanose 2.10 (4.61 g, 9.52 mmol)

was dissolved in anhydrous THF (25 mL) under N2 atmosphere and chilled to -78 ˚C. A 2.5M solution of n-

butyllithium in hexanes (8.4 mL, 20.9 mmol) was added and the reaction was warmed to -40 ˚C over 1hr

and then stirred for another 2h. The remaining n-butyllithium was quenched with a 10% w/v solution of

NH4Cl, the product was extracted 3x into ether, the organics dried over MgSO4, filtered and

concentrated in vacuo. The resulting oil was dissolved in DCM (16mL) and treated with 12M NaOH (2

mL). A large stir bar was used to vigorously mix the biphasic solution. Methyl iodide (1.5 mL, 24.1 mmol)

and tetrabutylammonium iodide (358 mg, .97 mmol) were added and the reaction stirred at RT

overnight. The reaction was diluted with DCM, washed 1x with 5% citric acid, 1x with H2O, and 1x with

brine, dried over MgSO4, filtered and concentrated in vacuo. Purification by silica gel chromatography

(gradient 5% 20% EtOAc) allowed isolation of the β anomer (1.54 g, 47%). Characterization of β

anomer: 1H NMR (300 MHz, CDCl3, δ (ppm)) 7.41 – 7.21 (m, 10H, Ar-H), 5.92 (ddd, J= 17.2, 10.3, 7.1, 1H,

H-5), 5.41 (dt, J= 17.2, 1.0, 1H, H-6 trans), 5.24 (dt, J= 10.3, 1.0, 1H, H-6 cis), 4.92 (s, 1H, H-1), 4.56 (s, 2H,

OCH2Ph), 4.53 (ABq, J=11.7, 2H, OCH2Ph), 4.43 (t, J= 6.9, 1H. H-4), 4.00 (dd, J=3.4, 1.1, 1H, H-2), 3.77 (dd,

J= 7.0, 3.4, 1H, H-3), 3.39 (s, 3H, OCH3); 13C NMR (75 MHz, CDCl3, δ (ppm)) 138.05, 137.81, 136.23,

128.69, 128.62, 128.13, 128.03, 118.32, 107.27, 88.88, 87.67, 82.36, 72.56, 72.32, 55.13

5,6-anhydro-2,3-di-O-benzyl-β-D-methylgalactofuranose and 5,6-anhydro-2,3-di-O-benzyl-β-L-

methylaltrofuranose (2.14): 5-eno-β-D-methylgalactofuranoside 2.12(1.1 g, 3.23 mmol) was dissolved in

reagent grade DCM (32 mL) and chilled to 0 ˚C. mCPBA (2.17 g, 9.69 mmol) was added and the contents

stirred at RT for 2 days. The reaction was quenched with saturated Na2S2O3 and NaHCO3 solution and

Page 69: Design, Synthesis and Biological Utility of Polysaccharide

55 extracted 1x into DCM. The organics were dried over MgSO4, filtered and concentrated in vacuo.

Purification by silica gel chromatography (gradient 5% 20% EtOAc in hexane) gave 947 mg of colorless

oil as a 1:1 diastereomeric mixture of epoxides (97%). 1H NMR (300 MHz, CDCl3, δ (ppm)) 7.41 – 7.21 (m,

20H, Ar-H), 4.95 (d, J= 3.7, 2H, H-1g, H-1a), 4.63 – 4.44 (m, 8H, 2x OCH2Phg, 2x OCH2Pha), 4.03 – 3.97 (m,

2H, H-2g, H-2a), 3.96 – 3.88 (m, 2H,H-3g, H-3a), 3.80 (t, J= 6.6, 2H, H-4g, H-4a), 3.39 (s, 3H, OCH3), 3.38

(s, 3H, OCH3), 3.11 (m, 2H, H-5g, H-5a), 2.75 (m, 4H, H-6g, H-6a); 13C NMR (75 MHz, CDCl3, δ (ppm))

137.79, 137.64, 128.70, 128.69, 128.50, 128.26, 128.15, 128.10, 128.06, 107.69, 107.57, 101.31, 88.09,

87.91, 84.62, 83.82, 82.14, 82.02, 72.65, 72.34, 72.22, 72.15, 55.28, 55.21, 52.17, 51.65, 45.65, 44.03

5,6-anhydro-2,3-di-O-benzyl-beta-D-methylgalactofuranose (2.15) and 2,3-di-O-benzyl-beta-L-

methylaltrofuranose (2.16): (S,S)-N,N’-Bis(3,5-di-tert-butylsalicylidene)-1,2-cyclohexanediaminocobalt

(II) (334mg, .5528 mmol) was dissolved in 200 μL of reagent grade DCM and treated with glacial acetic

acid (25 μL, .442 mmol). The solution stirred open to air for 30 minutes. DCM was removed in vacuo. A

solution of epoxide 2.14 (1.97g, 5.528 mmol) in THF (8mL, 0.7M) and water (50 μL, 2.763 mmol) was

added to the brown residue and the reaction was capped and stirred overnight. The reaction was

stopped when TLC indicated 50% conversion to diol. Quenched with 5 mL of 2M NH3 in MeOH and

stirred for 30 minutes. Concentrated in vacuo. Purified by silica gel chromatography (DCM 60%

EtOAc in hexane, 10% stepwise gradient). Isolated 1.083 g (55%) of epoxide and 958mg (46%) of diol.

Average isolated yields for three separate experiments were 48% for epoxide, 45% for diol.

Characterization of epoxide 8: 1H NMR (300 MHz, CDCl3, δ (ppm)) 7.4-7.2 (m, 10H, Ar-H), 4.93 (s, 1H, H-

1), 4.54 (ABq, 2H, J = 12.1, -OCH2Ph), 4.53 (ABq, 2H, J = 11.6 Hz, -OCH2Ph), 3.99 (dd, J = 3.1, 1.1, 1H, H-2),

3.93 (dd, J = 6.8, 2.9, 1H, H-3), 3.79 (dd, J = 6.7, 6.0, 1H, H-4), 3.37 (s, 3H, OCH3), 3.09 (ddd, J = 5.9, 4.1,

Page 70: Design, Synthesis and Biological Utility of Polysaccharide

56 2.7, 1H, H-5), 2.74 (AB part of an ABX, JAX = 4.1, JBX = 2.7, 2H, H-6a, H-6b); 13C NMR (75 MHz, CDCl3, δ

(ppm)) 137.80, 137.64, 128.70, 128.69, 128.15, 128.05, 107.57, 88.11, 84.63, 82.02, 72.64, 72.33, 55.21,

52.16, 44.02; MS (EMM)+: m/z: 379.1503 (M+Na)+ (M+Na+ calcd 379.1516) Characterization of diol (#):

1H NMR (300 MHz, CDCl3, δ (ppm)) 7.4 – 7.2 (m, 10H, Ar-H), 4.93 (s, 1H, H-1), 4.53 (ABq, J = 11.9, 2H,

OCH2Ph) 4.50 (ABq, J = 11.8, 2H, OCH2Ph), 4.07 (m, 2H, H-2, H-3), 3.99 (broad s, 1H, H-4), 3.92 (m, 1H, H-

5), 3.64 (m, 2H, H-6a, H-6b), 3.37 (2, 3H, OCH3), 2.56 (d, J = 3.9, 1H, 5-OH), 2.20 (t, J = 6.0, 1H, 6-OH); 13C

NMR (75 MHz, CDCl3, δ (ppm)) 137.56, 137.38, 128.75, 128.73, 128.36, 128.28, 128.26, 128.22, 107.39,

87.56, 82.76, 82.51, 72.49, 72.18, 71.51, 63.57, 55.15; MS (EMM)+: m/z: 397.1634 (M+Na)+ (M+Na+

calcd. 397.1622)

2,3-di-O-benzyl-6-deoxy-6-fluoro-β-D-methylgalactofuranose (2.17): Epoxide 2.15 (45 mg, .126 mmol)

was placed in a 3mL Eppendorf tube and dissolved in HF-triethylamine (180 μL, .7M). The tube was

capped, warmed to 70 °C using a sand bath and incubated for 4 days. The contents were cooled to room

temperature, diluted with acetone and carefully filtered through a silica plug. Concentrated in vacuo.

Purified by silica gel chromatography (5% 20% EtOAc in hexane gradient). Isolated 44 mg of

fluoroalcohol as a pale yellow oil (94%). 1H NMR (300 MHz, CDCl3, δ (ppm)) 7.4 – 7.2 (m, 10H, Ar-H), 4.54

(m, 5H, 2x OCH2Ph, H-6a), 4.34 (1/2 of the AB part of an ABMX, 1H, H-6b), 4.14 (dd, J = 5.8, 2.9, 1H, H-4),

4.05 ( dd, J = 5.7, 2.0, 1H, H-3), 3.98 (d, J = 1.5, 1H, H-2) 3.90 (dtd, J = 14.3, 5.9, 3.1, 1H, H-5), 3.37 (s, 3H,

OCH3), 2.23 (broad s, 1H, 5-OH); 19F NMR (280 MHz, CDCl3, δ (ppm)) -233.7 (td, 48.7, 14.1); 13C NMR (75

MHz, CDCl3, δ (ppm)) 137.71, 137.33, 128.74, 128.29, 128.23, 128.19, 128.16, 107.72, 87.11, 85.25,

83.17, 82.99, 81.31, 81.24, 72.62, 72.16, 70.01, 69.74, 55.18; MS (EMM)+: m/z: 399.1579 (M+Na)+

(M+Na+ calcd. 399.1579)

Page 71: Design, Synthesis and Biological Utility of Polysaccharide

57

6-deoxy-6-fluoro-beta-D-methylgalactofuranose (2.03): 2,3-di-O-benzyl-6-deoxy-6-fluoro-β-D-

methylgalactofuranose (44 mg, .117 mmol) was dissolved in reagent grade methanol (4.5 mL, .025M)

and treated with 20 wt% palladium hydroxide on carbon (44mg). The flask was evacuated and backfilled

with H2 from a balloon three times. Stirred at STP overnight. Filtered through celite, washed with

methanol and concentrated in vacuo. Isolated 22 mg of a colorless oil (96%). 1H NMR (300 MHz, CDCl3, δ

(ppm)) 4.91 (s, 1H, H-1), 4.55 (AB part of an ABMX, J = , 2H, H-6a, H-6b), 4.5 (m, 3H, H-2, H-3, H-5), 4.1 (s,

1H, H-4), 3.40 (s, 3H, OCH3); 19F NMR (280 MHz, CDCl3, δ (ppm)) -233.5 (td, 48.1, 14.0) 13C NMR (75 MHz,

CD3OD, δ (ppm)) 110.72, 86.51, 84.27, 83.84 (d, J = 27), 83.30, 78.68, 70.48 (d, J = 89), 55.45; MS

(EMM)+: m/z: 219.0630 (M+Na)+ (M+Na+ calcd. 219.0640)

2,3,5-tri-O-benzoyl-6-deoxy-6-fluoro-β-D-methylgalactofuranose (2.20): Fluoro Galf 2.03 (250mg, 1.274

mmol) was dissolved in dry pyridine (12 mL) under N2 atmosphere and chilled to 0 °C. Benzoyl chloride

(565 μL) was added dropwise over 5 minutes. The contents were warmed to RT and stirred for 3 hours.

The reaction was diluted with toluene and azeotroped 3x . Placed on high vacuum line to dry. Purified by

silica chromatography (5% 20% EtOAc in hexane). Isolated 560 mg of white crystalline powder (86%).

1H NMR (300 MHz, CDCl3, δ (ppm)) 8.1 (dd, J = 9.1, 7.2, 4H, Ar-H), 7.90 (d, J = 7.2, 2H, Ar-H), 7.58 (m, 3H,

Ar-H), 7.42 (t, J = 7.8, 2H, Ar-H), 7.30 (t, J = 7.5, 4H, Ar-H), 5.85 (dq, J = 16.0, 5.2, 1H, H-5), 5.59 (d, J = 5.4,

1H, H-3), 5.45 (d, J = 0.9, 1H, H-2), 5.20 (s, 1H, H-1), 4.80 (ddq, J = 46.7, 9.8, 5.9, 2H, H-6a, H-6b), 4.63

(dd, J = 5.2, 3.9, 1H, H-4), 3.47 (s, 3H, OCH3); 19F NMR (280 MHz, CDCl3, δ (ppm)) -235.5 (td, 46.6, 16.0)

13C NMR (75 MHz, CDCl3, δ (ppm)) 165.92, 165.87, 165.71, 133.72, 133.61, 133.56, 130.22, 130.18,

Page 72: Design, Synthesis and Biological Utility of Polysaccharide

58 130.05, 129.54, 129.20, 128.64, 107.01, 82.50, 82.38, 80.51, 80.44, 80.20, 77.75, 70.99, 70.71, 55.18; MS

(EMM)+: m/z: 531.1450 (M+Na+ calcd. 531.1426)

1-O-acetyl-2,3,5-tri-O-benzoyl-6-deoxy-6-fluoro-D-galactofuranose (2.21): 2,3,5-tri-O-benzoyl-6-deoxy-

6-fluoro-β-D-methylgalactofuranose (100 mg, .197 mmol) was placed in a flame-dried round bottom

flask and treated with a 1.4% v/v solution of concentrated H2SO4 in acetic anhydride (2 mL, .1M). The

reaction stirred at RT for 2 hours. Quenched with saturated NaHCO3 solution, diluted with brine and

extracted into DCM 3x. Dried organic layer over MgSO4, filtered and concentrated in vacuo. Azeotroped

remaining acetic acid with toluene 2x. Purified by silica gel chromatography (5% 20% EtOAc in

hexane) to give 97 mg of white crystalline solid (92%, 5:1 mix of β:α). 1H NMR (300 MHz, CDCl3, δ (ppm))

8.10 (m, 6.5H, Ar-H), 7.90 (d, J = 7.5, 2H, Ar-H), 7.6 – 7.2 (m, 13H, Ar-H), 6.61 (d, J = 4.8, .2H, H-1α), 6.50

(s, 1H, H-1β), 6.17 (t, J = 6.6, .2H, H-2α), 5.89 (dq, J = 16.2, 5.0, 1H, H-5β), 5.75 (m, .4H, H-3α, H-5α), 5.63

(d, J = 4.7, 1H, H-3β), 5.59 (s, 1H, H-2β), 5.0 – 4.8 (m, ~4H, H-4α, H-4-β, H-6aα, H-6aβ, H-6bα, H-6bβ),

2.20 (s, 3H, C(O)CH3 β), 2.13 (s, .6H, C(O)CH3 α); 19F NMR (280 MHz, CDCl3, δ (ppm)) -235.5 (td, J = 46.3,

15.7, β anomer), -237.5 (td, J = 46.1, 19.3, α anomer); 13C NMR (75 MHz, CDCl3, δ (ppm)) 169.35, 165.84,

165.72, 165.50, 133.92, 133.80, 133.65, 130.24, 130.10, 129.34, 129.07, 128.89, 128.77, 128.68, 128.65,

99.55, 93.65, 82.85, 82.79, 82.27, 81.64, 79.97, 77.68, 76.49, 70.74, 70.46, 21.29, 21.23; MS (EMM)+:

m/z: 559.1380 (M+Na+ calcd. 559.1375)

Dibenzyl 2,3,5-tri-O-benzoyl-6-deoxy-6-fluoro-α-D-galactofuranose monophosphate (2.23): 1-O-

acetyl-2,3,5-tri-O-benzoyl-6-deoxy-6-fluoro-D-galactofuranose (100 mg, .186 mmol) was placed in a

flame dried round bottom flask and dissolved in dry DCM (2.0 mL) under N2 atmosphere. The solution

Page 73: Design, Synthesis and Biological Utility of Polysaccharide

59 was chilled to 0 °C and treated with a 33 wt% HBr in AcOH solution (120 μL). After stirring for 20 minutes

the reaction was warmed to RT and stirred for 4 hours. Azeotroped using freshly distilled toluene 3x to

isolate the crude glycosyl bromide as an orange oil. The oil was dissolved in dry toluene (1.0 mL) under

N2. The solution was added dropwise over 30 minutes to a stirring solution of dibenzyl phosphate (208

mg) in dry toluene (1.0 mL). The contents stirred at RT for 2.5 hours. Concentration in vacuo followed by

silica gel chromatography (12% EtOAc in toluene) allowed isolation of 97 mg of the α-anomer as a

slightly cloudy viscous oil (69%). 1H NMR (300 MHz, CDCl3, δ (ppm)) 8.14 (d, J = 7.0, 2H, Ar-H), 8.0 (t, J =

7.1, 4H, Ar-H), 7.60 – 7.00 (m, 19, Ar-H), 6.32 (dd, J = 5.6, 4.7, 1H, H-1), 6.13 (t, J = 7.2, 1H, H-3), 5.70 (m,

2H, H-2, H-5), 5.05 – 4.55 (m, 7H, 2x OCH2Ph, H-6a, H-6b, H-4); 19F NMR (280 MHz, CDCl3, δ (ppm)) -

237.2 (td, J = 46.8, 17.3); 31P NMR (122 MHz, CDCl3, δ (ppm)) -1.88 (m); 13C NMR (75 MHz, CDCl3, δ

(ppm)) 165.80, 165.76, 165.72, 135.65, 135.54, 133.92, 133.84, 133.58, 130.33, 130.24, 130.15, 130.04,

129.34, 128.87, 128.73, 128.65, 128.00, 127.89, 97.85, 97.79, 81.69, 79.37, 79.29, 76.68, 73.43, 71.53,

71.25, 69.62, 69.55, 69.45; MS (EMM)+: m/z: 777.1890 (M+Na+ calcd. 777.1872)

Triethylammonium 2,3,5-tri-O-benzoyl-6-deoxy-6-fluoro-α-D-galactofuranose monophosphate (2.24):

Dibenzyl 2,3,5-tri-O-benzoyl-6-deoxy-6-fluoro-α-D-galactofuranose monophosphate (40 mg, .053 mmol)

was dissolved in ethyl acetate (500 μL, .1M) and triethylamine (50 μL). 10% Pd on carbon (20 mg) was

added. The stoppered flask was evacuated and back filled with H2 from a balloon. The contents stirred at

STP for 90 minutes. The reaction was filtered through celite, washed with ethyl acetate and

concentrated in vacuo. Isolated 33 mg of cloudy, viscous oil (90%). 1H NMR (300 MHz, CDCl3, δ (ppm))

8.10 (t, J = 7.2, 4H, Ar-H), 7.90 (d, J = 7.2, 2H, Ar-H), 7.5 – 7.2 (m, 9H, Ar-H), 6,21 (t, J = 7.2, 1H, H-1), 6.13

(dd, J = 7.4, 4.5, 1H, H-3), 5.69 (dq, J = 18.7, 4.8, 1H, H-5), 5.60 (ddd, J = 7.7, 4.5, 1.4), 5.00 (m, 2H, H-6a,

Page 74: Design, Synthesis and Biological Utility of Polysaccharide

60 H-6b), 4.54 (t, J = 5.4, 1H, H-4), 2.85 (q, J = 7.2, NCH2CH3), 1.2 (t, J = 7.3, NCH2CH3);

19F NMR (280 MHz,

CDCl3, δ (ppm)) -238.1 (td, J = 48.7, 18.9); 31P NMR (122 MHz, CDCl3, δ (ppm)) -1.88 (d, J = 7.9); 13C NMR

(75 MHz, CDCl3, δ (ppm)) 174.93, 66.03, 165.70, 133.43, 133.23, 133.16, 130.41, 130.03, 129.83, 129.54,

129.40, 128.50, 128.44, 128.26, 95.98, 82.61, 80.34, 74.16, 72.44, 72.19, 45.46, 8.57; MS (EMM)-: m/z:

573.0941 ([M-H]-, calcd. 573.0967)

Tributylammonium 6-deoxy-6-fluoro-α-D-galactofuranose monophosphate (2.25): Benzoylated galf

monophosphate (#) (42 mg, .06 mmol) was dissolved in a 5:2:1 mixture of methanol:water:triethylamine

(2.3 mL) and warmed to 30 °C. The reaction stirred for 4 days. Concentrated in vacuo then azeotroped

3x with toluene. Purified by anion exchange chromatography (Biorad AG1-X8 resin, gradient elution 250

mM 500 mM NH4HCO3 solution) using a peristaltic pump. Column fractions were combined, frozen

on dry ice and lyophilized. The residue was treated with ~200 μL Dowex-50WX8-200 resin (pH 7,

pretreated with 2.0M HNBu3Cl) for 4 hours. The resin was removed by filtration through a cotton plug

and the filtrate was frozen on dry ice and lyophilized. Isolated 18 mg of fluffy white solid (56%, 1.1

equivalents of tributylammonium). 1H NMR (300 MHz, D2O, δ (ppm)) 5.45 (t, J = 4.8, 1H, H-1), 4.65 – 4.40

(AB part of an ABMX, J = 47.0, 5.7, 3.7, 2H, H-6a, H-6b), 4.21 (dd, J = 8.6, 7.2, 1H, H-3), 4.06 (ddd, J = 8.4,

4.3, 2.3, 1H, H-2), 3.88 (dtd, 20.8, 5.4, 3.9, 1H, H-5), 3.83 (dd, J = 7.2, 5.6, 1H, H-4), 3.10 (m, 7H), 1.65 (m,

7H), 1.38 (sextet, J = 7.3, 7H), 0.9 (t, J = 7.3, 10H); 19F NMR (280 MHz, D2O, δ (ppm)) -235.6 (td, J = 46.9,

21.1); 31P NMR (122 MHz, D2O, δ (ppm)) -0.3 (d, J = 3.5); 13C NMR (75 MHz, D2O, δ (ppm)) 97.14 (d, J =

4.9), 85.20, 82.98, 81.07 (d, J = 7.0), 76.63 (d, J = 7.5), 73.59, 70.76 (d, J = 19), 52.85, 25.36, 19.43, 12.92;

MS (EMM)-: m/z: 261.0176 ([M-H]-, calcd. 261.0181)

Page 75: Design, Synthesis and Biological Utility of Polysaccharide

61 Uridine diphosphate 6-deoxy-6-fluoro-α-D-galactofuranose (2.01): A solution of tributylammonium

UMP (16mg, .03 mmol) in anhydrous DMF (470 μL) was stirred with activated 4 Å molecular sieves under

N2 for 1 hour. The molecular sieves were removed and the solution was treated with

diisopropylethylamine (17 μL, .1 mmol) and 1-methyl-3-benzenesulfonylimidazolium triflate (12.4 mg,

0.4 mmol) and stirred for 1 minute. The solution was added dropwise over 30 seconds to a solution of

tributylammonium 6-deoxy-6-fluoro-α-D-galactofuranose monophosphate (10mg, .02 mmol) and MgCl2

(3mg, .03 mmol) in anhydrous DMF (220 μL) at 0 °C. The contents were raised to room temperature and

stirred for 2 hours. The reaction was quenched with 1.0 mL of 100 mM ammonium acetate and washed

1x with 1.0 mL of DCM. The DCM was extracted with 1.0 mL of 100 mM ammonium acetate, the

aqueous layers were combined, frozen on dry ice and stored at -20 °C until purification by HPLC. Purified

by HPLC using a semiprep carbopak dionex column, 300mM tirthylammonium acetate, pH = 6.0-7.0,

flow rate = 1mL/min. Collected peak eluting at 40 – 45 minutes, froze on dry ice and lyophilized to give a

light brown solid. 1H NMR (500 MHz, D2O, δ (ppm)) 7.98 (d, J = 8.3, 1H, H-5u), 5.99 (m, 2H, H-1r, H-6u)

5.67 (t, J = 4.2, 1H, H-1g) 4.62 (m, 2H, H-6ag, H-6bg) 4.15 – 4.42 (m, 7H, H-2g, H-2r, H-3g, H-3r, H-4r, H-

5ar, H-5br) 4.01 (dq, J = 20.5, 5.0, 1H, H-5g) 3.91 (t, J = 7.5, 1H, H-4g); 19F NMR (280 MHz, D2O, δ (ppm)) -

234.8; 13C NMR (125 MHz, D2O, δ (ppm)) 183.52, 168.86, 154.46, 144.27, 105.28, 100.30, 90.95, 87.08,

85.93, 85.85, 85.75, 83.82, 83.77, 79.33, 79.27, 76.38, 76.19, 73.52, 73.38, 72.28, 67.50, 49.28, 10.81;

MS (EMM)+: m/z: 591.0406 ([M+Na]+, calcd. 591.0400)

2,3,6-tri-O-benzyl-β-L-methylaltrofuranose (2.18): Dibutyltin oxide (244 mg, .979 mmol) was placed in a

flame-dried round bottom flask flushed with N2. A solution of dibenzyl Altf 2.16 (282 mg, .753 mmol) in

hot toluene (2.5 mL) was added and the flask was equipped with a claisen adapter, thermometer and

Page 76: Design, Synthesis and Biological Utility of Polysaccharide

62 reflux condenser. The ensemble was tilted 45° such that the bend in the claisen adapter could collect

refluxing solvent. A spatula tip of MgSO4 was placed in the adapter and the reaction was heated to

reflux. After 5 hours, the claisen adapter was removed and the condenser was fitted directly to the

reaction flask. Benzyl bromide (116 uL, .979 mmol) and tetrabutylammonium bromide (24 mg, .075

mmol) were added the reflux continued overnight. The contents were cooled to RT and quenched with

aqueous KF. Toluene was isolated, washed with brine, dried over MgSO4, filtered and concentrated in

vacuo. Purification by silica gel chromatography (5% 20% EtOAc in hexane) gave 237 mg of colorless

oil (70%). 1H NMR (300 MHz, CDCl3, δ (ppm)) 7.4 – 7.2 (m, 15H, Ar-H), 4.94 (s, 1H, H-1), 4.50 (m, 6H, 3x

OCH2Ph), 4.22 (t, J = 4.9, 1H, H-4), 4.15 (dd, J = 4.8, 1.1, 1H, H-3), 3.99 (m, 1H, H-5), 3.95 (d, J = 1.2, 1H,

H-2), 3.55 (AB part of an ABX, JAX = 3.9, JBX = 6.7, JAB = 9.9, 2H, H-6a, H-6b), 3.37 (s, 3H, OCH3), 2.50 (d, J =

3.5, 1H, 5-OH); ); 13C NMR (75 MHz, CDCl3, δ (ppm)) 138.02, 137.55, 128.69, 128.64, 128.61, 128.30,

128.16, 128.10, 128.02, 127.96, 107.46, 87.32, 83.31, 82.65, 73.68, 72.22, 71.90, 71.13, 70.82, 55.07; MS

(EMM)+: m/z: 487.2116 ([M+Na]+, calcd. 487.2092)

2,3,6-tri-O-benzyl-5-deoxy-5-fluoro-β-D-methylgalactofuranose (2.19): 2,3,6-tri-O-benzyl-β-L-

methylaltrofuranose (26 mg, .056 mmol) was dissolved in dry DCM (500 μL) under N2 and chilled to -10

°C. DAST (11 μL, .084 mmol) was added dropwise. After stirring for 30 minutes the reaction was warmed

to RT and stirred for an additional hour. The contents were chilled to 0 °C and stirred with aqueous

NaHCO3 (2.0 mL) for 30 minutes. Extracted into chloroform, washed with water, dried over MgSO4,

filtered and concentrated in vacuo. Purification by silica gel chromatography (5% 10% EtOAc in

hexane) gave 12 mg of a colorless oil (46%). 1H NMR (300 MHz, CDCl3, δ (ppm)) 7.4 – 7.2 (m, 15H, Ar-H)

4.93 (s, 1H, H-1), 4.75 (ddt, J = 48.4, 6.7, 3.7, 1H, H-5), 4.51 (m, 6H, 3x OCH2Ph), 4.12 (ddd, J = 22.6, 6.9,

Page 77: Design, Synthesis and Biological Utility of Polysaccharide

63 3.5, 1H, H-4), 4.05 (m, 1H, H-3), 3.98 (m, 1H, H-2), 3.7 (m, 2H, H-6a, H-6b), 3.37 (s, 3H, OCH3);

13C NMR

(75 MHz, CDCl3, δ (ppm)) 138.03, 137.74, 137.63, 128.67, 128.14, 127.96, 127.91, 107.51, 92.03, 89.65,

87.91, 82.78 (d, J = 6.5), 80.21 (d, J = 17.9), 73.74, 72.62, 72.25, 69.69, 69.37, 55.20, 29.92; 19F NMR (280

MHz, CD3OD, δ (ppm)) -204.9 (dq, J = 44.9, 23.7); MS (EMM)+: m/z: 489.2036 ([M+Na]+, calcd. 489.2048)

5-deoxy-5-fluoro-beta-D-methylgalactofuranose (2.04): 2,3,6-tri-O-benzyl-5-deoxy-5-fluoro-β-D-

methylgalactofuranose (51 mg, .109 mmol) was dissolved in reagent grade methanol (4.5 mL, .025M)

and treated with 10 wt% palladium hydroxide on carbon (25mg). The flask was evacuated and backfilled

with H2 from a balloon three times. Stirred at STP overnight. Filtered through celite, washed with

methanol and concentrated in vacuo. Isolated 22 mg of a colorless oil (100%). 1H NMR (300 MHz,

CD3OD, δ (ppm)) 4.71 (d, J = 1.6, 1H, H-1), 4.58 (ddt, J = 48.4, 6.9, 3.8, 1H, H-5), 4.0 – 3.6 (m, 5H, H-2, H-

3, H-4, H-6a, H-6b), 3.33 (s, 3H, OCH3); 19F NMR (280 MHz, CD3OD, δ (ppm)) -209.3 (dq, J = 48.6, 20.1);

13C NMR (75 MHz, CD3OD, δ (ppm)) 109.25, 93.37, 91.01, 82.35, 81.30, 81.05, 76.92, 76.85, 61.49, 61.17,

54.22; MS (EMM)+: m/z: 219.0638 ([M+Na]+, calcd. 219.0640)

2,3,6-tri-O-benzoyl-5-deoxy-5-fluoro-β-D-methylgalactofuranose (2.26): Fluoro Galf 2.04 (20mg, .099

mmol) was dissolved in dry pyridine (1.0 mL) under N2 atmosphere and chilled to 0 °C. Two crystals of

dimethylamino pyridine were added. Benzoyl chloride (60 μL) was added dropwise over 5 minutes. The

contents were warmed to RT and stirred for 1 hour. The reaction was diluted with toluene and

azeotroped 3x . Placed on high vacuum line to dry. Purified by silica chromatography (5% 10% EtOAc

in hexane). Isolated 35 mg of colorless oil (70%). 1H NMR (300 MHz, CDCl3, δ (ppm)) 8.08 (m, 6H, Ar-H),

7.58 (m, 3H, Ar-H), 7.45 (m, 6H, Ar-H), 5.63 (d, J = 4.5, 1H, H-3), 5.54 (s, 1H, H-2), 5.38 (m, 1H, H-5), 5.20

Page 78: Design, Synthesis and Biological Utility of Polysaccharide

64 (s, 1H, H-1), 4.70 (m, 2H, H-6a, H-6b), 4.43 (ddd, J = 27.0, 4.7, 2.1), 3.48 (s, 3H, OCH3);

19F NMR (280 MHz,

CDCl3, δ (ppm)) -208.0 (dq, J = 48.6, 27.0); 13C NMR (75 MHz, CDCl3, δ (ppm))166.42, 166.22, 165.70,

139.90, 133.83, 133.77, 133.48, 130.40, 130.19, 130.01, 129.79, 129.31, 129.17, 128.73, 128.64, 107.40,

90.59, 88.17, 82.65, 82.41, 81.34, 77.89, 77.81, 64.30, 63.98, 55.35; MS (EMM)+: m/z: 531.1399

([M+Na]+, calcd. 531.1426)

1-O-acetyl-2,3,6-tri-O-benzoyl-5-deoxy-5-fluoro-D-galactofuranose (2.27): 2,3,6-tri-O-benzoyl-5-deoxy-

5-fluoro-β-D-methylgalactofuranose (84 mg, .165 mmol) was placed in a flame-dried round bottom flask

and treated with a 1.4% v/v solution of concentrated H2SO4 in acetic anhydride (2.0 mL). The reaction

stirred at RT for 1.5 hours. Quenched with 2.0 mL of saturated NaHCO3 solution, diluted with brine and

extracted into DCM 3x. Dried organic layer over MgSO4, filtered and concentrated in vacuo. Azeotroped

remaining acetic acid with toluene 2x. Purified by silica gel chromatography (5% 20% EtOAc in

hexane) to give 81 mg of white crystalline solid (91%). 1H NMR (300 MHz, CDCl3, δ (ppm)) 8.08 (m, 6H,

Ar-H), 7.58 (m, 3H, Ar-H), 7.45 (m, 6H, Ar-H), 6.60 (d, J = 4.5, .2H, H-1α), 6.52 (s, 1H, H-1β), 6.14 (t, J =

6.9, .2H, H-3α), 5.77 (dd, J = 7.4, 4.6, .2H, H-2α), 5.69 (m, 2H, H-2β, H-3β), 5.36 (m, 1H, H-5β), 5.20 (m,

.2H, H-5α), 4.65 (m, 3.4H, H-6aα, H-6-aβ, H-6bα, H-6bβ, H-4β), 4.43 (ddd, J = 26.2, 6.4, 2.8, 1H, H-4α),

2.20 (s, 3H, C(O)CH3 β), 2.09 (s, .6H, C(O)CH3 α); 19F NMR (280 MHz, CDCl3, δ (ppm)) -206.1 (dq, J = 51.6,

21.2, α) -207.8 (dq, J = 48.7, 24.5, β); 13C NMR (75 MHz, CDCl3, δ (ppm)) 169.69, 169.20, 166.34, 166.28,

166.15, 165.99, 165.44, 134.05, 133.96, 133.49, 130.26, 130.16, 130.09, 130.01, 129.81, 129.74, 129.04,

128.96, 128.86, 128.80, 128.64, 99.80, 93.14, 90.49, 90.40, 88.07, 87.96, 85.02, 84.77, 80.73, 80.46,

77.71, 77.63, 75.70, 74.29, 74.22, 64.16, 63.83, 63.58, 21.27, 21.17; MS (EMM)+: m/z: 559.1364

([M+Na]+, calcd. 559.1375)

Page 79: Design, Synthesis and Biological Utility of Polysaccharide

65

Dibenzyl 2,3,6-tri-O-benzoyl-5-deoxy-5-fluoro-α-D-galactofuranose monophosphate (2.29): 1-O-

acetyl-2,3,6-tri-O-benzoyl-5-deoxy-5-fluoro-D-galactofuranose (22 mg, .041 mmol) was placed in a flame

dried round bottom flask and dissolved in dry DCM (450 μL) under N2 atmosphere. The solution was

chilled to 0 °C and treated with a 33 wt% HBr in AcOH solution (28 μL). After stirring for 20 minutes the

reaction was warmed to RT and stirred for 4 hours. Azeotroped using freshly distilled toluene 3x to

isolate the crude glycosyl bromide as an orange oil. The oil was dissolved in dry toluene (450 μL) under

N2. The solution was added dropwise over 30 minutes to a stirring solution of dibenzyl phosphate (49

mg, .176 mmol) in dry toluene (500 μL). The contents stirred at RT for 2.5 hours. Concentration in vacuo

followed by silica gel chromatography (12% EtOAc in toluene) allowed isolation of 28 mg of the α-

anomer as a slightly cloudy viscous oil (90%). 1H NMR (300 MHz, CDCl3, δ (ppm)) 8.00 (m, 6H, Ar-H), 7.55

(m, 3H, Ar-H), 7.4 – 7.2 (m, 16H, Ar-H), 6.31 (t, J = 4.7, 1H, H-1), 6.14 (t, J = 6.9, 1H, H-3), 5.70 (ddd, J =

7.1, 4.3, 2.2, 1H, H-2), 5.17 (dtd, J = 47.1, 6.4, 3.2, 1H, H-5), 5.07 (AB part of an ABX, J = 6.5, 3.5, 2H,

POCH2Ph), 4.90 (AB part of an ABX, J = 11.6, 6.9, 2H, POCH2Ph), 4.61 (m, 2H, H-6a, H-6b), 4.50 (ddd, J =

24.4, 6.4, 3.5, 1H , H-4); 19F NMR (280 MHz, CDCl3, δ (ppm)) -203.8 (dq, J = 56.9, 28.9); 31P NMR (122

MHz, CDCl3, δ (ppm)) -2.4 (m); 13C NMR (75 MHz, CDCl3, δ (ppm)) 166.13, 165.96, 165.76, 134.00,

133.89, 133.46, 130.26, 130.15, 129.96, 129.60, 128.76, 128.70, 128.63, 128.58, 128.05, 127.85, 101.02,

97.61, 97.54, 90.59, 88.15, 80.86, 80.60, 76.54, 76.44, 73.56, 73.47, 69.56, 69.47, 63.32; MS (EMM)+:

m/z: 777.1898 ([M+Na]+, calcd. 777.1872)

Triethylammonium 2,3,6-tri-O-benzoyl-5-deoxy-5-fluoro-α-D-galactofuranose monophosphate (2.30):

Dibenzyl 2,3,6-tri-O-benzoyl-5-deoxy-5-fluoro-α-D-galactofuranose monophosphate (29 mg, .038 mmol)

Page 80: Design, Synthesis and Biological Utility of Polysaccharide

66 was dissolved in ethyl acetate (1.0 mL, .05M) and triethylamine (38 μL). 10% Pd on carbon (15 mg) was

added. The stoppered flask was evacuated and back filled with H2 from a balloon. The contents stirred at

STP overnight. The reaction was filtered through celite, washed with ethyl acetate and concentrated in

vacuo. Isolated 14 mg of cloudy, viscous oil (54%).1H NMR (300 MHz, CDCl3, δ (ppm)) 8.1 (d, J = 7.5, 2H,

Ar-H), 7.9 (m, 3H, Ar-H), 7.5 – 7.2 (m, 9H, Ar-H), 6.2 (m, 2H, H-1, H-3), 5.62 (m, 1H, H-2), 5.30 (m, 1H, H-

5), 4.7 (m, 2H, H-6a, H-6b), 4.5 (dt, J = 19.3, 6.2, 1H, H-4), 3.00 (q, J = 7.3, 3.8H, NCH2CH3), 1.2 (t, J = 7.5,

10.7H, NCH2CH3); 19F NMR (280 MHz, CDCl3, δ (ppm)) -199.7 (dq, J = 47.1, 25.1); 31P NMR (122 MHz,

CDCl3, δ (ppm)) 0.0 (d, J = 3.5); 13C NMR (75 MHz, CDCl3, δ (ppm)) 164.83, 164.74, 164.58, 132.26,

132.06, 131.90, 129.14, 128.93, 128.81, 128.58, 128.47, 128.44, 127.92, 127.25, 127.1795.17, 95.11,

91.02, 78.54, 78.24, 75.80, 75.70, 73.09, 72.98, 62.71, 62.43, 44.38, 7.37; MS (EMM)-: m/z: 573.0966

([M]-, calcd. 573.0967)

Tributylammonium 5-deoxy-5-fluoro-α-D-galactofuranose monophosphate (2.31): Triethylammonium

2,3,6-tri-O-benzoyl-5-deoxy-5-fluoro-α-D-galactofuranose monophosphate (14 mg, .021 mmol) was

dissolved in a 5:2:1 mixture of methanol:water:triethylamine (1.0 mL) and warmed to 30 °C. The

reaction stirred for 4 days. Concentrated in vacuo then azeotroped 3x with toluene. Purified by anion

exchange chromatography (Biorad AG1-X8 resin, gradient elution 250 mM 500 mM NH4HCO3

solution) using a peristaltic pump. Column fractions were combined, frozen on dry ice and lyophilized.

The residue was treated with ~200 μL Dowex-50WX8-200 resin (pH 7, pretreated with 2.0M HNBu3Cl)

for 4 hours. The resin was removed by filtration through a cotton plug and the filtrate was frozen on dry

ice and lyophilized. Isolated 10 mg of fluffy white solid (82%, 1.5eq. of HNBu3); 1H NMR (300 MHz, D2O,

δ (ppm)) 5.51 (dd, J = 6.2, 4.2, 1H, H-1), 4.69 (dtd, J = ~50, 5.8, 3.9, 1H, H-5), 4.19 (t, J = 7.8, 1H , H-3),

Page 81: Design, Synthesis and Biological Utility of Polysaccharide

67 4.15 (ddd, J = 8.0, 4.4, 2.0, 1H, H-2), 3.95 (dt, J = 19.5, 6.3, 1H, H-4), 3.84 (m, 2H, H-6a, H-6b), 3.10 (m,

9H, NBu), 1.62 (m, 9H, NBu), 1.33 (sextet, J = 7.3, 10H, NBu), 0.9 (t, J = 7.2, 13H); 19F NMR (280 MHz,

D2O, δ (ppm)) -199.7 (dq, J = 46.4, 23.4); 31P NMR (122 MHz, D2O, δ (ppm)) -0.4 (broad s); 13C NMR (75

MHz, D2O, δ (ppm)) 96.91, 96.85, 95.41, 93.10, 79.82, 79.57, 76.74, 76.69, 73.59, 73.50, 60.71, 60.43,

52.85, 25.36, 19.44, 12.93; MS (EMM)-: m/z: 261.0182 ([M]-, calcd. 261.0181)

5-deoxy-5-fluoro-α-D-galactofuranose uridine diphosphate (2.02 ): A solution of tributylammonium

UMP (5mg, .01 mmol) in anhydrous DMF (150 μL) was stirred with activated 4 Å molecular sieves under

N2 for 1 hour. The molecular sieves were removed and the solution was treated with

diisopropylethylamine (5 μL, .03 mmol) and 1-methyl-3-benzenesulfonylimidazolium triflate (4 mg, 0.01

mmol) and stirred for 1 minute. The solution was added dropwise over 30 seconds to a solution of

tributylammonium 5-deoxy-5-fluoro-α-D-galactofuranose monophosphate (3mg, .007 mmol) and MgCl2

(1 3mg, .01 mmol) in anhydrous DMF (75 μL) at 0 °C. The contents were raised to room temperature and

stirred for 2 hours. The reaction was quenched with 1.0 mL of 100 mM ammonium acetate and washed

1x with 1.0 mL of DCM. The DCM was extracted with 1.0 mL of 100 mM ammonium acetate, the

aqueous layers were combined, frozen on dry ice and stored at -20 °C until purification by HPLC. Purified

by HPLC using a semiprep carbopak dionex column, 300mM tirthylammonium acetate, pH = 6.0-7.0,

flow rate = 1mL/min. Collected peak eluting at 40 – 45 minutes, froze on dry ice and lyophilized to give a

light brown solid. 1H NMR (500 MHz, D2O, δ (ppm)) 7.83 (d, J = 8.0, 1H, H-5u), 5.86 (m, 2H, H-1r, H-6u)

5.55 (t, J = 4.8, 1H, H-1g) 4.56 (m, 1H, H-5g) 4.00 – 4.25 (m, 7H, H-2g, H-2r, H-3g, H-3r, H-4r, H-5ar, H-

5br) 3.89 (dt, J = 18.9, 6.5, 1H, H-4g) 3.71- 3.80 (m, 2H, H-6ag, H-6bg); 19F NMR (280 MHz, D2O, δ (ppm)) -

Page 82: Design, Synthesis and Biological Utility of Polysaccharide

68 232.0; 13C NMR (125 MHz, D2O, δ (ppm)) 144.25, 105.25, 102.84, 100.32, 90.97, 85.90, 85.83, 79.02,

76.41, 75.63, 72.20, 67.42, 49.28, 10.81; MS (EMM)+: m/z: 591.0406 ([M+Na]+, calcd. 591.0400).

Page 83: Design, Synthesis and Biological Utility of Polysaccharide

69

Chapter 3

Evaluation of Fluorosugars as Probes of Carbohydrate Polymerase Activity

Portions of this work are published:

Brown, Christopher D., Rusek, Max S. and Kiessling, Laura L. Fluorosugar Chain Termination Agents as Probes of the Sequence-Specificity of a Carbohydrate Polymerase. J. Am. Chem. Soc. 2012, 6552-6555

Contributions

Max Rusek Synthesized disaccharide acceptors

Kenzo Yamatsugu provided all trisaccharide acceptors and the synthetic tetrasaccharide

Page 84: Design, Synthesis and Biological Utility of Polysaccharide

70 3.1 Introduction

GlfT2, the polymerase that constructs the mycobacterial galactan polymer, has seen an increase

in attention since its discovery over ten years ago. Its intriguing ability to generate a polymer of

appropriate length and alternating Galf-β-(1,5)-Galf-β-(1,6) linkages has caught the interest of many

laboratories. How does a single glycosyltransferase regulate the length, regiochemistry and

stereochemistry of carbohydrate polymer synthesis? Other polymerases such as DNA polymerase or the

ribosome construct their polymers off of a template strand. Glycosyltransferases which polymerize

carbohydrates into polysaccharide chains have no such template. While some studies have shed light

onto length control and other elements of polysaccharide assembly (chapter 1), essentially no work has

been done on understanding pattern control. We intended to use our fluorinated UDP-Galf analogs in an

in vitro assay of GlfT2 previously developed in this laboratory to shed light on this aspect of polymer

biosynthesis.

We have previously published a report that details an in vitro assay of GlfT2 activity using

recombinant His6-GlfT2, synthetic acceptors and UDP-Galf.62 A magnesium-dependent polymerization

takes place when enzyme is incubated with acceptor and donor. The polymeric products are observed

by matrix assisted laser desorption ionization time of flight (MALDI-TOF) mass spectrometry. (Figure

3.01) This previous report provided the first evidence that GlfT2 was an active polymerase intrinsically

capable of generating galactan polymers of physiological length. This activity assay of polymerization

proved tremendously useful for understanding the molecular underpinnings of polysaccharide chain

synthesis in this system. By systematically varying the chemical structure of the acceptor, our lab was

able to deduce some of the requirements for efficient polymerization.

Page 85: Design, Synthesis and Biological Utility of Polysaccharide

71

Figure 3.01. Incubation of GlfT2 with disaccharide acceptor 3.xx and UDP-Galf generates full length galactan polymers. The crude mixture can be analyzed by MALDI-TOF mass spectrometry.

For template independent polymerases like GlfT2, there is an outstanding question for how the

enzyme regulates the length of polymer chain it produces. By varying the length of lipid aglycone

appended to a Galf disaccharide motif and observing longer polymer chains in the mass spectrum, the

group was able to put forth a new mechanism for length control.62 Specifically, the observation that

longer lipid chains promoted synthesis of longer galactan chains suggested that the acceptors were

binding in a bivalent manner and that this “tethering” of the acceptor controlled the length of the

Page 86: Design, Synthesis and Biological Utility of Polysaccharide

72 polymer by increasing the chances of rebinding after passive dissociation of the non-reducing

terminus.62 The assay was adapted again using isotopically labeled acceptor to quantify the degree of

processivity using a molecular distraction approach.56 For these reasons, the assay was considered to be

a robust readout of the enzyme’s polymerizing activity and was used as the basis for assessing pattern

control and fidelity.

3.2 GlfT2 operates by a sequence specific mechanism

To determine if fluorinated Galf donors are substrates for GlfT2, the enzyme was incubated with

hexasaccharide acceptor 3.04 and the natural donor UDP-Galf (3.02) and a typical polymer distribution

was observed. (Figure 3.02A) Next, the enzyme was incubated with acceptor 3.04 and 6-fluorinated

donor UDP-6F-Galf 2.01 for 18 hours. The hexasaccharide acceptor contains a non-reducing Galf-β-(1,6)-

Galf motif. Perpetuation of the alternating linkage pattern would begin with the generation of a new β-

(1,5) linkage in which the C5 hydroxyl group of the acceptor attacks the anomeric carbon of the UDP-

sugar. If the electrophile was the 6-fluoro donor, the resulting +1 Galf acceptor heptasaccharide 3.06

would form. This new acceptor is unable to generate a β-(1,6) linkage, however the C5 hydroxyl group

remains available. If the enzyme is able to use the C5 hydroxyl, it would generate a galactan chain with

consecutive β-(1,5) linkages: a mistake. If, however, the enzyme is faithful to the alternating linkage

pattern, chain termination should occur. Analysis of the crude reaction products by MALDI-TOF mass

spectrometry following the 18 hour incubation showed a strong signal for the +1 product 3.06 and no

other significant product peaks. (Figure 3.02B, top)

To determine if chain termination by 6F-Galf was a consequence of GlfT2’s sequence specificity

or just a result of passive dissociation of a fluorinated acceptor, the analogous experiment with GlfT2,

hexasaccharide 3.04 and UDP-5F-Galf (2.02) was performed. Again, the first linkage formed is a-β-(1,5)

Page 87: Design, Synthesis and Biological Utility of Polysaccharide

73 bond generated by nucleophilic attack of the hexasaccharide C5-OH on the anomeric carbon of UDP-5F-

Galf. This heptasaccharide has a non-reducing β-(1,5) linkage and a C6-OH. Incorporation of a second

fluoro-Galf would generate a β-(1,6) linkage. Only after this +2 product forms would the enzyme

encounter a fluorine atom blocking the position where the next linkage is formed. Analysis of the

Figure 3.02. When GlfT2 is incubated with a hexasaccharide acceptor and UDP-Galf, a normal distribution of polymers is generated (A). When GlfT2, hexasaccharide and UDP-6F-Galf are incubated together, a single fluorinated Galf residue is added and then polymerization terminates abruptly. (B, top) When the same hexasaccharide is acted on in the presence of UDP-5F-Galf, two fluorinated Galf residues are added. (B, bottom) The chain termination occurs prior to generation of a mistaken linkage in both cases.

Page 88: Design, Synthesis and Biological Utility of Polysaccharide

74 reaction products from the incubation of UDP-5F-Galf showed a signal corresponding to the +2 F-Galf

octasaccharide 3.07. The presence of this peak indicates that GlfT2 is able to recognize and elongate a

fluorinated acceptor, but only to the extent that the alternating linkage pattern is maintained. (Figure

3.02B, bottom)

3.3 Sequence-specific chain termination is general

That GlfT2 so carefully maintains the alternating linkage pattern is remarkable considering it

recognizes and acts on a variety of acceptors including monosaccharide Galf lipids, disaccahrides,

trisaccharides and tetrasaccharides . (Beccas paper) We wondered if this result was general for galactan

acceptor fragments with a different number of starting Galf residues. In other words, was this sequence

specific termination a general observation, or just happenstance for this specific hexasaccharide

acceptor? We performed a similar experiment using fluorinated donors and tetrasaccharide acceptor

3.08. This acceptor, like the hexasaccharide, is elongated with Galf residues by GlfT2. (Figure 3.03A)

Sequence specific chain termination would be expected to follow the same pattern on this shorter

acceptor because it also has a non-reducing β-(1,6) Galf disaccharide motif. Results of MALDI-TOF

analysis for incubation with the 6F and 5F donors again showed +1 and +2 product peaks, respectively.

This suggested that the termination was a mechanism-specific result. (Figure 3.03B)

3.4 A mistake on initiation?

The polymerization assays depicted in Figures 3.02 and 3.03 indicate the enzyme is forced to

chain terminate polymerization when incorporation of a fluorinated Galf blocks the alternating linkage

pattern. In every case, however, a residual peak with mass corresponding to a first mistake appears to

grown in as well (+2 for 6F-Galf, +3 for 5F-Galf). Two explanations are possible to explain these peaks.

Page 89: Design, Synthesis and Biological Utility of Polysaccharide

75

Figure 3.03. Incubation of GlfT2 with tetrasaccharide and UD-Galf gave the expected polymer distribution (A). Repeating the experiment with UDP-6F-Galf resulted in glycotransfer of one 6F-Galf residue. (B, top) The same tetrasaccharide was elongated by two 5F-Galf residues. (B, bottom) Chain termination appears to be sequence-specific and general.

First, it may represent a small amount of polymer in which a fluorinated Galf was able to add in and

generate a mistake. Alternatively, the tetrasaccharide and hexasaccharide acceptors- both of which are

generated chemoenzymatically- may have a subpopulation with terminal β-(1,5) linkages. This is

possible considering the high promiscuity of the enzyme may pre-dispose it to misread the terminal

linkage and make a mistake upon initiation, and the chemoenzymatic method used to generate the

starting material. A subpopulation with the wrong linkage could be contributing to the residual mistaken

Page 90: Design, Synthesis and Biological Utility of Polysaccharide

76 peak. To distinguish between these two possibilities, a tetrasaccharide acceptor was accessed

synthetically (courtesy of Kenzo Yamatsugu). This synthetic tetrasaccharide has the same linkage pattern

as the one used in figure 3.03, but because it was not generated chemoenzymatically, there is no

subpopulation of acceptors with the wrong terminal linkage. If the residual peaks were a result of such a

subpopulation, then incubation with the synthetic acceptor would not give a mass spectrum with those

Figure 3.04. Synthetic tetrasaccharide behaves in an identical manner to tetrasaccharide isolated from chemoenzymatic sources. The “doublet” appearance in the 6F-Galf MALDI spectrum is due to a significant amount of M+K product peaks, an occasional occurrence.

Page 91: Design, Synthesis and Biological Utility of Polysaccharide

77 peaks. If the residual peaks are a result of some intrinsic low-level ability of the enzyme to generate a

galactan fragment with consecutive β-(1,5) or β-(1,6) linkages, then the peaks should still be present in

an experiment with the synthetic tetrasaccharide. Incubation of GlfT2 with this synthetic tetrasaccharide

acceptor and UDP-6F-Galf gave the same product profile: A dominant +1 F-Galf peak and a residual +2

peak. Likewise, the incubation with UDP-5F-Galf produced a dominant +2 peak and a residual +3 peak

(very faint). This finding suggests the enzyme may be able to generate a small but detectable amount of

galactan with consecutive β-(1,6) or β-(1,5) bonds, but cannot sustain polymerization in a meaningful

way. These data support a model in which the alternating linkage pattern is necessary for

polymerization, in which ideal conditions for polymerization are unable to generate even short polymers

with mistaken linkages. (Figure 3.04)

3.5 GlfT2 struggles to identify the terminal linkage of short acceptors

How, from a molecular standpoint, does GlfT2 ensure alternating linkage deposition. Recently, a

crystal structure of GlfT2 was published and the authors put forward a model for bifunctional pattern

control based on acceptor binding.127 The enzyme was crystallized as a tetramer with a molecule of UDP

bound. No evidence was provided for the functional significance of the tetramer, but the monomers do

exhibit a classic GT-A Rossmann-like fold. In the model, an acceptor binds to the enzyme active site

shallowly or deeply, depending on the non-reducing glycosidic bond geometry. This, the authors claim,

aligns the acceptor hydroxyl groups for deprotonation at either C5 or C6, respectively. The model does

not take into account the effect of the lipid on acceptor orientation. This is unusual because reports

have shown that the nature of the lipid has a profound effect on polymer length control.62 If the lipid

had no effect on acceptor orientation, then all acceptors with identical non-reducing glycosidic linkages

should behave the same in our chain termination assay.

Page 92: Design, Synthesis and Biological Utility of Polysaccharide

78

Figure 3.05. Proposed carbohydrate-directed binding model for bifunctional linkage control. A fluorinated acceptor with fluorine positioned to disrupt the alternating pattern would interact unfavorably with the carboxylate base. The red line defines the latitude where the catalytic carboxylate is located.

The synthesis of disaccharide acceptors bearing either β-(1,6) or β-(1,5) glycosidic bonds was

completed following a published procedure. (Schemes 3.01, 3.02) Briefly, a diisopropylidene

galactopyranose 3.16 was protected as benzyl ether at C6. Reflux of 3.17 in acidic methanol allowed

isolation of a C6-protected furanose scaffold 3.18 which could be ester-protected with benzoyl chloride.

Treatment of 3.19 with tin tetrachloride yielded bicyclic intermediate 3.20 which was opened by

Page 93: Design, Synthesis and Biological Utility of Polysaccharide

79 acetolysis to generate 1,6-bis acetylated Galf 3.21. This sequence provides a way to distinguish the C5

and C6 positions of a galactofuranoside that is capable at first of participating in a glycosylation reaction

as a donor, then selective unmasking to access an acceptor scaffold. Glycosylation with allyl alcohol gave

differentially protected Galf 3.22 in moderate yield. Deprotection of the C6-acyl group with workup

conditions that allowed benzoyl group migration afforded both the C6-OH (3.23) and C5-OH (3.24)

products. Both migrated and unmigrated products could be used as acceptors in a glycosylation with

thioglycoside 3.25. (Scheme 3.02) The allyl disaccharides were subsequently cross-metathesized with

11-phenoxy-undecene and deprotected to afford GlfT2 acceptors 3.28 and 3.31.

Scheme 3.01. Disaccharide acceptors were assembled using a published procedure. Both regiochemical linkages are accessible using this strategy. Synthesis completed by Max Rusek.

Incubation of GlfT2 with β-(1,6) disaccharide acceptor 3.26 and UDP-Galf produced a typical

distribution of polymer peaks. A similar distribution was seen with β-(1,5) disaccharide acceptor 3.31. To

Page 94: Design, Synthesis and Biological Utility of Polysaccharide

80 test whether acceptor binding was directed by the non-reducing glycosidic linkage the enzyme was

incubated with 3.28 or 3.31 and UDP-6F-Galf. After 18 hours, a small +1 product peak formed for the β-

(1,6) acceptor, but no product formed for the β-(1,5) acceptor. (Figure 3.06) When the experiment was

repeated with UDP-5F-Galf, no product for either acceptor was observed. (Figure 3.06) Apparently, the

combination of short acceptors and fluorinated donors is too high a kinetic barrier to efficient

glycosylation. (Figure 3.06) These results suggest the carbohydrate-directing model for pattern control

Scheme 3.02. Silver triflate-mediated glycosylations to generate β-(1,5) and β-(1,6) linked Galf disaccharides. Synthesis completed by Max Rusek.

Page 95: Design, Synthesis and Biological Utility of Polysaccharide

81 should be amended: An acceptor binds in a lipid-directed manner such that the active site aspartate is

poised to deprotonate the acceptor hydroxyl group to perpetuate the alternating linkage pattern

provided the carbohydrate primer is long enough.

Figure 3.06. Disaccharide acceptors are poor substrates for glycotransfer with fluorinated donors.

Page 96: Design, Synthesis and Biological Utility of Polysaccharide

82

We then looked to trisaccahrides 3.32 and 3.33 (synthesized by Kenzo Yamatsugu). These longer

acceptors, like their cognate disaccharides, are elongated by GlfT2 (unpublished results). This is

unsurprising since 3.32 is the +1 Galf product of 3.31 and 3.33 the +1 Galf product of 3.28. If the enzyme

binds acceptors by recognizing the non-reducing glycosidic linkage and positioning it appropriately for

subsequent nucleophilic attack, then the fluorinated donors should terminate sequence-specifically.

Incubation of GlfT2 and 3.32 with UDP-6F-Galf gave a +1 product. A +1 product also was generated upon

incubation with UDP-5F-Galf. (Figure 3.07) These results are not indicative of sequence-specific chain

termination. Surprisingly, neither UDP-6F-Galf nor UDP-5F-Galf were capable of adding into the β-(1,5)

trisaccharide at all. (Figure 3.07) This suggests the enzyme does not bind the acceptor through

recognition of the non-reducing glycosidic bond and positioning it for attack. A more favorable model is

one in which the lipid directs binding of the acceptor. Since chain terminators act in a sequence specific

manner for longer acceptors, their deviation from this behavior with short acceptors implies that

initiation is the time when a mistake is most likely, when promiscuity is at its highest. Fidelity in linkage

construction is likely imparted to the system via the processivity of the enzyme. In other words, the

enzyme does not generate alternating linkages simply by binding the acceptor with appropriate pre-

organization for attack, but by remaining bound to the acceptor after the initial glycosylation. This may

partially explain why the enzyme has evolved a processive mechanism for polymerization.

3.6 Fluorinated products are dead-end substrates

Fluorinated carbohydrates have been known to bind with lower affinity than their fully

hydroxylated counterparts, and it is possible that the reaction between a fluorinated acceptor and a

fluorinated donor is kinetically disfavored because of this potential loss of affinity. Thus, we wondered if

the fluorinated products could recover from the chain-terminating pattern block if they were exposed to

Page 97: Design, Synthesis and Biological Utility of Polysaccharide

83

Figure 3.07. Trisaccharide acceptors with β-(1,6) linkages at the reducing end are substrates for fluorinated donors, (top two lines) but chain termination is not sequence-specific. Acceptors with β-(1,5) reducing end stereochemistry are not substrates for fluorinated donors. (bottom two lines) The enzyme is not binding the acceptors at their non-reducing end.

fresh enzyme and UDP-Galf. Solutions of the fluorinated acceptors generated in the polymerization with

hexasaccharide acceptor were treated with fresh GlfT2 and UDP-Galf. After incubation for 18 hours, the

Page 98: Design, Synthesis and Biological Utility of Polysaccharide

84 contents were analyzed by MALDI. Although the typical lipid-less galactan series did arise, no additional

Galf residues were added to the fluorinated acceptors. This result may indicate that chain terminators

can be used as inhibitors. (Figure 3.08)

Figure 3.08. Isolated Fluoro-Galf capped acceptors do not elongate when incubated with GlfT2 and very high UDP-Galf concentrations. (red peaks) A byproduct polymer series of lipid-less non-fluorinated galactan forms. (black) Fluorinated acceptors appear to be dead end substrates.

Inhibition of galactan polymerization is a long term goal of researchers in the field. A chain

terminating UDP-Galf analog would potentially have mixed modes of substrate inhibition (both acceptor

and donor inhibition). These studies using fluorinated donors have created new questions concerning

the properties of an ideal chain termination agent and whether a chain termination approach would be

effective at inhibiting polymerization. These areas of study are an exciting new direction for the field and

are addressed in the next chapter.

We have devised a strategy for assessing sequence specificity for a carbohydrate polymerase

using fluorosugars as chain termination agents, expanding the chemical glycobiologist’s toolbox. This

Page 99: Design, Synthesis and Biological Utility of Polysaccharide

85 approach invokes the work of Frederick Sanger, in which chain-terminating nucleosides were used to

rapidly and accurately sequence nucleic acid polymers, and reimagines the concept to gain insight into

another class of biopolymer: the polysaccharide.128 In polymerizations such as DNA replication,

sequence specificity is determined by a template. In contrast, polysaccharide polymerization has no

obvious mechanism for ensuring sequence fidelity. Synthetic probes which challenge a bifunctional

carbohydrate polymerase’s patterning specificity were designed (chapter 2). Both probes exhibited

chain termination capability of GlfT2-catalyzed galactan polymerization. Moreover, the chain

termination occurred only when GlfT2 was challenged to deviate from the normal alternating pattern,

suggesting that polymerization occurs with sequence specific dependence. We anticipate that this

method can be generalized to study repeating motif specificity in other carbohydrate polymerases.

3.7 Experimental

General procedure for GlfT2 activity assay

Polymerization reactions consisted of 10 μL total volume containing final concentrations of 0.2 μM His6-

GlfT2, approximately 200 μM acceptor, 1.0 mM UDP donor sugar in 50 mM Hepes, pH 7.0, 25 mM

MgCl2, and 100 mM NaCl. Reactions were incubated at room temperature for 18 h, then quenched with

30 μL of a 1:1 mixture of CHCl3/MeOH. Quenched reaction mixtures were evaporated to dryness under

vacuum in a SpeedVac SC100 (Varian) then resuspended in 10 μL of 50% MeCN in MQ water for MALDI

MS analysis. Samples for MALDI MS analysis were spotted as a 1:3 mixture with α-cyano-4-

hydroxycinnamic acid matrix and spectra were recorded in positive linear mode using a Bruker Ultraflex

III mass spectrometer.

Page 100: Design, Synthesis and Biological Utility of Polysaccharide

86 Isolation of tetrasaccharide and hexasaccharide acceptors

A polymerization reaction was run in the presence of GlfT2, β-(1,6) disaccharide and UDP-Galf. The

reaction was quenched after 45 minutes and suspended in 50% ACN/H2O. The solution was filtered

through a milipore syringe filter and purified by RP-HPLC (10% - 80% ACN in 0.1% AcOH in H2O).

Monitored absorbance at 195nm and collected peaks at retention times 18 – 25. Analysis by MALDI-TOF

identified pure tetrasaccharide and hexasaccharide fractions. Used in polymerizations assays without

further purification.

Synthesis of (1,6) and (1,5)-disaccharide acceptors

2,3,5,6-tetra-O-acetyl-β-D-Galfuranosyl-(1,5)-2,3,6-tri-O-benzoyl-β-O-allylgalactofuranose (3.26):

2,3,5,6-tetra-O-acetyl -1-(thioethyl)-β-D-galactofuranose (20 mg, .037 mmol) and allyl 2,3,6-tri-O-

benzoyl-β-D-galactofuranose (15.2 mg, .028 mmol) were dissolved in 1 mL dry DCM, treated with

activated molecular sieves and allowed to stir under N2 for 15 minutes at RT. The solution was chilled to

0 °C and silver triflate (1.4 mg, .006 mmol) and N-iodosuccinimide (8.8 mg, .039 mmol) were added. A

red-brown color developed. After 30 minutes of stirring at 0 °C the reaction was filtered through celite

and washed with DCM until all the pink color was washed into the filtrate. The filtrate was quenched

with 5 mL of Na2S2O3 solution (saturated), 5 mL of saturated NaHCO3 solution, 5 mL of brine. The organic

layer was dried over MgSO4, filtered and concentrated in vacuo. Purification by silica gel

chromatography gave 21 mg of a colorless oil. (70% yield) ); 1H and 13C NMR spectra matched those

previously reported.

2,3,5,6-tetra-O-acetyl-β-D-galactofuranosyl-(1,5)-2,3,6-tri-O-benzoyl-β-D-(12-phenoxy-2-

undecenyl)galactofuranose (3.27): 2,3,5,6-tetra-O-acetyl-β-D-galfuranosyl-(1,5)-2,3,6-tri-O-benzoyl-β-O-

allylgalactofuranose (18 mg, .017 mmol) was dissolved in a solution of 11-phenoxy-undecene (19 mg,

Page 101: Design, Synthesis and Biological Utility of Polysaccharide

87 .077 mmol) in DCM (200 μL). The solution was placed under N2 atmosphere, treated with Grubbs’ 1st

generation catalyst (1 mg, .0019 mmol) and stirred at RT overnight. Added 1 more mg of catalyst in the

morning then concentrated on rotovap. Purification by silica gel chromatography (5% 10% 20%

EtOAc in hexane) gave 7 mg of cross metathesis product (33% yield). 1H and 13C NMR spectra matched

those previously reported.

β-D-galactofuranosyl-(1,5) -β-D-(12-phenoxy-2-undecenyl)galactofuranose (3.28): 2,3,4,5-tetra-O-

acetyl-β-D-galactofuranosyl-(1,5)-2,3,6-tri-O-benzoyl-β-D-(12-phenoxy-2-undecenyl)galactofuranose (5

mg, .004 mmol) was treated with 500 μL of 0.5M NaOMe in methanol and stirred for 3h at RT under N2.

The solution was neutralized with acidic resin, filtered off through a cotton plug, washed with methanol

and concentrated in vacuo. Purification by silica gel chromatography (10% MeOH in DCm) gave 2.4mg of

colorless film (100% yield). 1H and 13C NMR spectra matched those previously reported.

2,3,5,6-tetra-O-acetyl -1-(thioethyl)-β-D-galactofuranose (3.25): 1-O-acetyl-2,3,5,6-tri-O-benzoyl -D-

galactofuranose (20 mg, ..037 mmol) was dissolved in dry DCM (100 μL) under N2 and chilled to 0 °C.

trimethylsilyl ethyl thioether (9 μL, .056 mmol) was added via syringe. Zinc iodide (6 mg, .018 mmol) was

quickly weighed and added to the reaction. The reaction was warmed to RT and stirred 2h. TEA (2 drops)

was added to quench the reaction. A precipitate was removed by filtering through a cotton plug,

washing with DCM and the filtrate was concentrated in vacuo. Purification by silica gel chromatography

(5% 10% EtOAc in hexane) gave an oily white residue, 15 mg (75% yield); 1H and 13C NMR spectra

matched those previously reported.

2,3,5,6-tetra-O-acetyl-β-D-Galfuranosyl-(1,6)-2,3,5-tri-O-benzoyl-β-O-allylgalactofuranose (3.29):

2,3,5,6-tetra-O-acetyl -1-(thioethyl)-β-D-galactofuranose (15 mg, .027 mmol) and allyl 2,3,5-tri-O-

benzoyl-β-D-galactofuranose (11 mg, .021 mmol) were dissolved in 1 mL dry DCM (1.0 mL), treated with

Page 102: Design, Synthesis and Biological Utility of Polysaccharide

88 activated molecular sieves and allowed to stir under N2 for 15 minutes at RT. The solution was chilled to

0 °C and silver triflate (1.3 mg, .005 mmol) and N-iodosuccinimide (7 mg, .029 mmol) were added. A red-

brown color developed. After 30 minutes of stirring at 0 °C the reaction was filtered through celite and

washed with DCM until all the pink color was washed into the filtrate. The filtrate was quenched with 5

mL of Na2S2O3 solution (saturated), 5 mL of saturated NaHCO3 solution, 5 mL of brine. The organic layer

was dried over MgSO4, filtered and concentrated in vacuo. Purification by silica gel chromatography

gave 10 mg of a colorless oil. (48% yield) ); 1H and 13C NMR spectra matched those previously reported.

2,3,5,6-tetra-O-acetyl-β-D-galactofuranosyl-(1,6)-2,3,5-tri-O-benzoyl-β-D-(12-phenoxy-2-

undecenyl)galactofuranose (3.30): 2,3,5,6-tetra-O-acetyl-β-D-Galfuranosyl-(1,6)-2,3,5-tri-O-benzoyl-β-

O-allylgalactofuranose (10 mg, .010 mmol) was dissolved in a solution of 11-phenoxy-undecene (11 mg,

.045 mmol) in DCM (100 μL). The solution was placed under N2 atmosphere, treated with Grubbs’ 1st

generation catalyst (1 mg, .001 mmol) and stirred at RT overnight. Added 1 more mg of catalyst in the

morning then concentrated on rotovap. Purification by silica gel chromatography (5% 10% 20%

EtOAc in hexane) gave 5.8 mg of cross metathesis product (40% yield). 1H and 13C NMR spectra matched

those previously reported.

β-D-galactofuranosyl-(1,6) -β-D-(12-phenoxy-2-undecenyl)galactofuranose (3.31): 2,3,5,6-tetra-O-

acetyl-β-D-galactofuranosyl-(1,6)-2,3,5-tri-O-benzoyl-β-D-(12-phenoxy-2-undecenyl)galactofuranose (5.8

mg, .004 mmol) was treated with 600 μL of 0.5M NaOMe in methanol and stirred for 3h at RT under N2.

The solution was neutralized with acidic resin, filtered off through a cotton plug, washed with methanol

and concentrated in vacuo. Purification by silica gel chromatography (10% MeOH in DCm) gave 2.2 mg of

colorless film (92% yield). 1H and 13C NMR spectra matched those previously reported.

Page 103: Design, Synthesis and Biological Utility of Polysaccharide

89

Chapter 4

Expanding the Utility of Chain-Terminating Glycosides

Portions of this work were published:

May, J.F., Levengood, M. R., Splain, R.A., Brown, C.D., Kiessling, L.L. A processive carbohydrate polymerase that mediates bifunctional catalysis using a single active site, Biochemisry, 2012, 1148-1159

Contributions

Max Rusek and Kenzo Yamatsugu aided in the preparation of 5-deoxy UDP-Galf

John May and Rebecca Splain constructed the mutant GlfT2 plasmids

Page 104: Design, Synthesis and Biological Utility of Polysaccharide

90 4.1 Introduction

Reports of chain terminating glycosyl donors are appearing with rising frequency in the

literature. For example, the polysialyltransferase from E.coli K12 is a carbohydrate polymerase that

generates a capsular polysaccharide polysialic acid with alternating α-(2,8) and α-(2,9) alternating

linkages.10,96 A 9-azido analog of the natural donor, CMP-sialic acid, was synthesized and found to

incorporate into acceptor strands in a chain-terminating manner.96 Deoxy Galf donors have been

reported to chain terminate the galactan. In some cases, a fluorinated donor sugar was thought to act as

a chain termination agent in cell-based assays, but later discovered to act as a competitive inhibitor of

components of the Leloir pathway instead.129,130

Figure 4.01. Chain terminator donor sugar panel.

We realized there was a need to comprehensively address the nature of chain terminating

glycosyl donors. For GlfT2 in particular, understanding the influence that a bond-blocking functional

group has on the nascent chain is important. The physical properties that led us to design and synthesize

fluorinated probes would be different in many respects from analogous deoxy or azido probes. For

instance, the inductive effects on vicinal hydroxyl pKa would vary considerably from a polarized C-F

bond to a non-polar C-H or C-N3 bond. This pKa is relevant for Galf transfer as GlfT2 operates via a

general base mechanism that deprotonates either the C5 or C6 hydroxyl group. This in turn may affect

Page 105: Design, Synthesis and Biological Utility of Polysaccharide

91 the behavior of the chain terminator, and would be an important consideration depending on the

potential use of the compound as either an inhibitor or as a sequence fidelity probe. (Figure 4.01)

Sequence fidelity is discussed at length in chapter 3. An inhibitor could potentially be useful in several

ways. Chain terminators such as AZT, which disrupts the activity of the HIV enzyme reverse transcriptase

are used directly as antibiotics.131 Alternatively, inhibitors can be used as probes in enzyme kinetics

investigations to yield information such as substrate binding order. More generally, it would be

important for research on the mycobacterial cell wall to know what sort of chemical probes can be

incorporated into various structures, as this would be useful for cell-based studies as a potential imaging

tool. Thus, a broad comparison of fluorinated, deoxygenated and azido Galf donor would be beneficial

to a wide range of future study.

In order to compare the fluorinated, deoxygenated and azido donor behavior as patterning

fidelity probes and inhibitors we sought to synthesize the panel of inhibitors from Figure 4.01. We

returned to the vinyl Galf intermediate and found that beta vinyl methyl-Galf 2.12 could be modified

under standard hydroboration conditions to yield the 5-deoxy intermediate 4.04. It was possible to

access the 6-deoxy scaffold be treatment of resolved epoxide 2.15 with sodium borohydride, but it was

easier to access it directly from fucose (4.11). 6-azido Galf can be synthesized from epoxide 2.15

(deprotected). Longer reaction times and lower temperatures improved the regioselectivity of the ring

opening. Cognizant that the azide functionality reduces under standard hydrogenation/hydrogenolysis

conditions, we modified the stereoselective phosphorylation from dibenzyl to diallyl phosphate.

Treatment of protected 6-N3-Galf phosphate with palladium tetrakis triphenylphosphine should remove

the allyl protecting groups easily. These approaches highlight the versatility of the vinyl group to

carbohydrate mimic synthesis. (Scheme 4.01)

Page 106: Design, Synthesis and Biological Utility of Polysaccharide

92

Scheme 4.01. Synthesis of deoxygenated and azido UDP-Galf analogs.

Page 107: Design, Synthesis and Biological Utility of Polysaccharide

93 4.2 A comparison of fluorinated and deoxygenated Galf donors.

4.2.1 As sequence specificity probes

We first wanted to investigate whether or not fluorinated and deoxygenated donors were

equally well equipped to report on sequence fidelity for GlfT2. As discussed in chapter 3, the

tetrasaccharide acceptor 3.12 was incubated with GlfT2 and the fluorinated donors (2.01 or 2.02) to give

the characteristic +1/+2 sequence-specific chain termination result. When the 6-deoxy donor was

incubated with the same tetrasaccharide acceptor and GlfT2, a strong signal corresponding to the +1

deoxy-Galf intermediate was observed. This result is in agreement with sequence-specific chain

termination observed with UDP-6F-Galf. To confirm the chain termination is a result of pattern blocking

rather than passive diffusion of the resulting acceptor, a similar experiment with UDP-5H-Galf is

planned. Reports in the literature suggest that 5-deoxy compounds add a single 5H-Galf residue to

trisaccharides. This result is similar for our trisaccharide study, but trisaccharides are difficult acceptor

substrates to interpret: they behave idiosyncratically, potentially because they are too short to fill the

enzyme subsites. (See discussion chapter 3, Table 4.01)

Table 4.01. Summary of chain termination results for various acceptors with fluorinated and deoxygenated donors. *Results reported in Reference68

Page 108: Design, Synthesis and Biological Utility of Polysaccharide

94 4.2.2 As inhibitors

We assessed the ability of our chain terminators to inhibit galactan formation in two ways. First,

we ran polymerization assays in which our fluorinated or deoxy donors were forced to compete for

incorporation with the natural donor, UDP-Galf. These polymerizations were run under the same

conditions as the single-donor experiments from chapter 3. A chain terminator with strong inhibition

would be expected to incorporate into a polymerizing chain rapidly, impairing polymerization

completely, or nearly so. Evaluation of such chain terminated products by MALDI-TOF would be

expected to reveal abruptly terminated galactan oligomers. Alternatively, a chain terminator that was

unable to effectively compete with UDP-Galf to inhibit polymerization would reveal a normal polymer

Figure 4.02. When GlfT2 is incubated with tetrasaccharide, UDP-Galf and UDP-6F-Galf, the polymerization is not inhibited, even at 10:1 fluorinated donor concentrations. Close inspection shows fluorinated Galf residues are able to incorporate into the growing chains. (top) The same experiment with UDP-5F-Galf again fails to inhibit polymerization. No fluorinated Galf incorporation was observed. (bottom)

Page 109: Design, Synthesis and Biological Utility of Polysaccharide

95 distribution in our assay. The results of the competition experiments between UDP-Galf and the

fluorinated donors suggested they were weak inhibitors of polymerization. Interestingly, the 6F donor

was able to incorporate into the galactan chains, but not inhibit polymerization completely even at 10

fold excess relative to UDP-Galf. The 5F-Galf donor was unable to compete effectively even at 10-fold

excess and no incorporation was observed. (Figure 4.02)

The second method we used to demonstrate inhibition involved monitoring the rate of UDP

production in a continuous coupled assay our lab has used earlier.(May) The fluorinated donor must

bind the same active site as UDP-Galf and thus, it is likely that there is some competition for biding

between the two. Likewise, a dead-end fluorinated acceptor would compete with non-fluorinated

acceptors as well. The combined effect should be a reduction in the initial rate of turnover by GlfT2. The

coupled assay allows us to monitor that turnover and observe whether or not a reduction in

polymerization is occurring in a more quantitative sense. In this assay UDP is consumed by the enzyme

pyruvate kinase to generate the phosphopyruvate substrate for lactate dehydrogenase. In the presence

of NADH, this relay of enzymatic transformations can be observed by measuring the absorbance of

NADH over time. Absorbance is directly related to UDP consumption and so measurement of UDP can

serve as readout of enzyme activity.

GlfT2 was incubated with tetrasaccharide acceptor 3.12, UDP-Galf and the cocktail of coupled

assay enzymes and substrates. To observe inhibition, NADH absorbance was followed over time at UDP-

6F-Galf donor concentrations of 0μM, 500μM and 5 mM, which corresponded to a 0, 1, 10-fold ratio

relative to the natural donor. The rate of UDP production was extrapolated from the absorbance

change. This measurement serves as a crude readout of enzyme activity. Of course, one problem with

monitoring inhibition in this way is that UDP-6F-Galf is also a substrate for the enzyme, so the UDP

Page 110: Design, Synthesis and Biological Utility of Polysaccharide

96

Figure 4.03. GlfT2-catalyzed polymerization of tetrasaccharide in the presence of the inhibitor UDP-6F-Galf. The rate of reaction decreases as the ratio of UDP-6F-Galf to UDP-Galf increases.

produced from 6F-Galf incorporation is also measured. (Figure 4.04) We observed a reduction in UDP-

production that was dose-dependent. There are potentially multiple modes of inhibition operating in

Page 111: Design, Synthesis and Biological Utility of Polysaccharide

97 this system as the fluorodonor probably acts as a competitive inhibitor, a chain terminator and the

products of chain termination themselves act as competitive inhibitors of the acceptor. This mixed

activity makes it difficult to measure specific kinetic parameters such as Ki. However the reduced

turnover does suggest a degree of inhibition is taking place. A comparison to the remaining panel of

donors is planned.

4.3 Fluorinated Galf can be used to generate fluorinated acceptors

All glycosyltransferases must bring together a donor and acceptor in a ternary complex. The

mechanism of substrate binding typically involves one of three pathways: ping pong, sequential ordered

and random ordered. (Figure 1.07) Binding order is an important consideration for researchers

interested in growing x-ray crystallography quality protein crystals because it determines the order of

substrate soaking that should be trialed. High throughput screens also must consider binding order

when they are designed or the screen risks identifying hits that are inactive in a biologically relevant

context. Classic enzymology experiments can be used to distinguish between these mechanisms,

provided a competitive inhibitor is available.

Fluorinated acceptors have been used as competitive inhibitors to determine binding order for a

glycosyltransferase before. For example, to elucidate the binding order for the glycosyltransferase

MurG,50 the Walker Group synthesized a fluorinated lipid II analog that was found to competitively

inhibit the natural substrate. Another substrate analog-based approach was used to study the binding

order for the polymerase PglH.51 In that system product inhibition was found to regulate chain length.

Because the product was acting as a competitive inhibitor it was used to determine the binding order of

the polymerase. We recognized that the fluoro-capped dead-end acceptors generated in our chain

termination assays might be competitive inhibitors so we devised a synthesis of a fluorinated

Page 112: Design, Synthesis and Biological Utility of Polysaccharide

98 disaccharide based on our acceptor synthesis. The 1-O-acetyl-6F-Galf intermediate 2.21 could readily be

converted to the thioethyl glycoside donor 4.22. Similarly, the 1-O-acetyl-5F-Galf intermediate 2.27 was

converted to thioglycoside 4.26. (Scheme 4.02) These thioglycosides could be used in glycosylation

reactions analogous to the ones used to generate disaccharide acceptors 3.28 and 3.31 Cross metathesis

and deprotection afforded 5F-Galf-β-(1,6)-Galf and 6F-Galf-β-(1,5)-Galf disaccharides 4.25 and 4.29,

respectively.

Scheme 4.02. Synthesis of fluorinated Galf disaccharides.

Page 113: Design, Synthesis and Biological Utility of Polysaccharide

99

These analogs were designed so that the fluorine atoms were positioned to interrupt the

alternating linkages. STD NMR studies on discaccharides and trisaccharides suggest that the non-

reducing terminus has the fewest contacts with the protein, making substitution at this end the least

perturbing.105,132 The cognate non-fluorinated substrates 3.28 and 3.31 were previously found to be

elongated. In order to use the fluorinated derivative as inhibitors, we needed to confirm that these

fluorinated acceptors were incompetent for elongation by the polymerase. Incubation of GlfT2 with

fluorinated acceptor 4.25 and UDP-Galf in our standard polymerization assay surprisingly yielded

polymeric products. (Figure 4.04) Because the fluorine atom was positioned to block elongation at this

position, the only way polymeric products could form was if the enzyme initiated polymerization with a

mistaken consecutive β-(1,6) linkage. This result confirms that the enzyme is capable of making a

mistake on initiation and recovering from such an event. Additionally, the data also support the

assertion that the enzyme controls the alternating linkages during processive elongation and not via

differential binding on initiation (vida supra). When the polymerization experiment was repeated with

5F acceptor 4.29, no polymerization was observed. Studies to further assess 4.29 as a potential

competitive inhibitor are planned. To determine if the fluorinated acceptor was acting as a competitive

inhibitor, our coupled assay will be used under initial rate conditions at various concentrations of

fluorinated acceptor. If the compound is acting as a competitive inhibitor then plotting the double

reciprocal graph (1/[S] vs 1/v) for each inhibitor concentration should yield a series of lines that

converge at the Y-axis. Once a competitive inhibitor is identified, the kinetic analysis to determine

substrate binding mechanism is straightforward.

Page 114: Design, Synthesis and Biological Utility of Polysaccharide

100

Figure 4.04. Assessing polymerization competency with fluorinated acceptors.

4.4 One active site and 2 linkages

One longstanding question concerning GlfT2’s bifuncionality is whether the enzyme utilizes one

active site for both linkages, or two distinct sites, one for each linkage. Inverting glycosyltransferases

typically use a so-called “DXD” motif in their active sites. This motif acts as a general base to

deprotonate the acceptor hydroxyl proton rendering it nucleophilic for attack on the donor. In

particular, GlfT2 has two such DXD motifs. Homology modeling to a related glycosyltransferase, SpsA,

indicated that 256DXD258 was positioned to coordinate the magnesium ion and bind UDP-Galf whereas

371DDA373 was positioned to act as the base.

We reasoned that mutation of this catalytic base would help us discern the number of active

sites. If Glft2 had a single active site, mutagenesis should abolish all activity. If GlfT2 possesses two

active sites, one for each linkage, then mutagenesis should abolish one, but not the other bond-forming

activity. By incubating mutant enzymes with acceptor substrates and observing the product profile, we

Page 115: Design, Synthesis and Biological Utility of Polysaccharide

101 would be able to distinguish between these possibilities. Thus, alanine and glutamate mutants were

constructed and expressed in the same E. coli expression system used for the wild type enzyme.

Incubation of D371A, D371E, D372A or D372E with tetrasaccharide acceptors either abolished

polymerization entirely (D371A, D372A, D372E), or was permissive (D371E) to polymerization. These

results strongly indicate that the enzyme has a single active site. (Figure 4.05)

Figure 4.05. Mutant GlfT2 polymerization assays. Only D371E was active. These results suggest GlfT2 uses a single active site for bifunctional catalysis.

Recently, a crystal structure of GlfT2 was published.127 Satisfyingly, the enzyme does indeed

possess a single active site and D371 appears to be positioned to act as catalytic base. Interestingly, the

enzyme crystalized as a tetramer with C4 symmetry, forming a central pore. The authors were unable to

Page 116: Design, Synthesis and Biological Utility of Polysaccharide

102 demonstrate that the tetramer was the functionally active state of the enzyme. The report also suggests

that bifunctionality arises from how far the acceptor extends into the binding pocket. Since the D371E

mutant is active, but contains an additional CH2 group, the “ruler” would be altered. We wondered if the

change in length would have an effect on the enzyme’s linkage fidelity so we incubated D371E with

tetrasaccharide and the fluorinated Galf donors. The same chain termination pattern resulted: 6F-Galf

added once and 5F-Galf, twice. This may not be consistent with the bifunctional ruler model. (Figure

4.06) Modeling of this change using the crystal structure coordinates is planned once the structure is

deposited in the PDB.

Figure 4.06. GlfT2 D371E maintains high sequence fidelity. The altered length arising from an aspartate to glutamate mutation appears to have little effect on patterning regulation.

4.5 Experimentals

Synthesis of UDP-5-deoxy-Galf (UDP-5H-Galf)

2,3-di-O-benzyl-5-deoxy-β-D-methylgalactofuranose (4.04): 2,3-di-O-benzyl-5-vinyl-β-D-

methylgalactofuranose (122mg) was dissolved in anhydrous THF (2.44mL) under N2. A solution of

Page 117: Design, Synthesis and Biological Utility of Polysaccharide

103 borane in THF (1.0M, 1.1 mL) was added dropwise and the contents stirred overnight. The contents

were heated to 50 °C then 367μL of 3.0M NaOH solution and a solution of peroxide (30 wt % in H2O)

were aded. Waited 5 minutes, then stirred at RT for 45 minutes. The product was extracted 2x into

DCM, dried over MgSO4, filtered and concentrated in vacuo. Purification by silica gel chromatography

(10% 20% 50% EtOAc in hexane) yielded 35mg of product (27% yield). 1H NMR (300 MHz, CDCl3, δ

(ppm)); ); 13C NMR (75 MHz, CDCl3, δ (ppm))

5-deoxy-β-D-methylgalactopyranose (4.05): 2,3-di-O-benzyl-5-deoxy-β-D-methylgalactofuranose

(49mg, .137 mmol) was dissolved in reagent grade methanol (1,4 mL) and treated with Pd/C catalyst

(25mg, 50% by wt). The atmosphere was evacuated and back filled with H2 from a balloon 3x and stirred

for 2 hours. The catalyst was removed by filtering through celite and washing with methanol. The filtrate

was concentrated in vacuo to give a colorless oil, 18mg (74% yield). 1H NMR (300 MHz, CDCl3, δ (ppm)) ;

13C NMR (75 MHz, CDCl3, δ (ppm))

2,3,6-tri-O-benzoyl-5-deoxy-β-D-methylgalactofuranose (4.06): 5-deoxy-β-D-methylgalactopyranose

(18mg, .101 mmol) was dissolved in anhydrous pyridine (1.0 mL) under N2 atmosphere and treated

dropwise with benzoyl chloride (41 μL, .354 mmol) and stirred at RT for 2h. The reaction stalled so

another 3.5 equivalents of benzoyl chloride were added and the reaction was complete instantaneously.

Azeotroped 2x with toluene then purified by silica chromatography (5% 10% 20% EtOAc in

hexane). Isolated 50mg of colorless oil (100% yield). 1H NMR (300 MHz, CDCl3, δ (ppm)); 13C NMR (75

MHz, CDCl3, δ (ppm))

1-O-acetyl-2,3,6-tri-O-benzoyl-5-deoxy-β-D-methylgalactofuranose (4.07): 2,3,6-tri-O-benzoyl-5-deoxy-

β-D-methylgalactofuranose was dissolved in a 1.4% v/v solution of H2SO4 in Ac2O and stirred for 2h at

RT. The reaction was quenched with a 1:1 mixture of saturated NaHCO3 and brine, extracted 3x into

Page 118: Design, Synthesis and Biological Utility of Polysaccharide

104 DCM, dried over MgSO4, filtered and concentrated in vacuo. Residual acetic acid was removed by

azeotrope with toluene. Purification by silica chromatography (15% 10% 20% EtOAc in hexane)

gave a colorless oil, 54mg (57% yield). 1H NMR (300 MHz, CDCl3, δ (ppm)); 13C NMR (75 MHz, CDCl3, δ

(ppm))

Dibenzyl 2,3,6-tri-O-benzoyl-5-deoxy -α-D-galactofuranose phosphate (4.09): 1-O-acetyl-2,3,6-tri-O-

benzoyl-5-deoxy-β-D-methylgalactofuranose (26mg, .05 mmol) was dissolved in dry DCM (500μL) under

N2 and chilled to 0 °C. A solution of HBr in AcOH (33 wt.%, 33 μL, .125 mmol) was added dropwise. After

30 minutes at 0 °C, the reaction was warmed to RT and stirred for 2 hours. The contents were

azeotroped 3x using anhydrous toluene. Separately, a solution of dibenzyl phosphate (56mg, .2mmol) in

dry toluene (500 μL) and TEA (28μL, .2 mmol) was prepared under N2. The crude oil of Galf bromide was

dissolved in dry toluene and added dropwise to the stirring solution of dibenzyl phosphate. The contents

were stirred at RT and monitored by TLC. Complete at 1h, concentrated in vacuo. Purification by silica

chromatography (10% EtOAc in toluene) gave 24mg of pure product as a cloudy oil (66% yield, 2 steps).

1H NMR (300 MHz, CDCl3, δ (ppm)); 13C NMR (75 MHz, CDCl3, δ (ppm)); 31P NMR (122 MHz, CDCl3, δ

(ppm))

Triethylammonium 2,3,6-tri-O-benzoyl-5-deoxy-α-D-galactofuranose phosphate(4. 9 ¾): Dibenzyl

2,3,6-tri-O-benzoyl-5-deoxy-α-D-galactofuranose phosphate (20mg, .027 mmol) was dissolved in reagent

grade ethyl acetate (300 μL) and TEA (30 μL) and treated with Pd/C catalyst (10% by weight, 10mg). The

flask was evacuated under vacuum and backfilled with H2 from a balloon. The reaction stirred at RT

overnight. The contents were filtered through celite, washed with ethyl acetate and concentrated in

vacuo to give a white residue, 11mg (65% yield). 1H NMR (300 MHz, CDCl3, δ (ppm)); 13C NMR (75 MHz,

CDCl3, δ (ppm)); 31P NMR (122 MHz, CDCl3, δ (ppm))

Page 119: Design, Synthesis and Biological Utility of Polysaccharide

105 Tributylammonium 5-deoxy-α-D-galactofuranose phosphate (4.10): Triethylammonium 2,3,6-tri-O-

benzoyl-5-deoxy-α-D-galactofuranose phosphate (11mg, .014 mmol) was dissolved in a 5:2:1 smixture of

MeOH:H2O:TEA, raised to 30 °C and stirred for 7d. Concentrated on rotovap to remove solvents and

purified by anion exchange chromatography (Biorad AG1-X8 resin, gradient elution 250 mM 500 mM

NH4HCO3 solution) using a peristaltic pump. Column fractions were combined, frozen on dry ice and

lyophilized. The residue was treated with ~200 μL Dowex-50WX8-200 resin (pH 7, pretreated with 2.0M

HNBu3Cl) for 4 hours. The resin was removed by filtration through a cotton plug and the filtrate was

frozen on dry ice and lyophilized. Isolated 5mg of a white solid (72%, 2 steps); 1H NMR (300 MHz, CDCl3,

δ (ppm)); 13C NMR (75 MHz, CDCl3, δ (ppm)); 31P NMR (122 MHz, CDCl3, δ (ppm))

UDP-5-deoxy-Galf (4.02): Trifluoroacetic anhydride was distilled from P2O5 (BP = 36 °C). UMP

tributylammonium was placed in an oven-dried pear flask with stir bar and placed under N2 and chilled

to 0°C. ACN (1.2mL), N,N-dimethylanaline (183 μL) and TEA (50 μL) were added. A solution of

trifluoroacetic anhydride (254 μL) in CAN (450 μL) was prepared at 0 °C under N2 then cannulated to the

flask with UMP, which turned yellow. After 15 minutes, the excess TFAA was evacuated using a needle

equipped to a vacuum line with N2 backfilling from a balloon. A solution of 1-methylimidazole (86 μL) in

ACN (900 μL) and TEA (250 μL) was prepared under N2 and at 0 °C. This solution was cannulated to the

activated UMP solution and stirred at 0 °C for 30 minutes. The tributylammonium 5-deoxy-Galf

monophosphate (5mg, .008 mmol) was suspended in ACN (100 μL). Activated molecular sieves were

added. The UMP-imidazolium solution (200 μL) was added dropwise over 30 minutes. The reaction

stirred cold for 1.5h then at RT for 3.5h. The coupling was quenched with 500 μL of 100mM NH4OAc

solution, washed with 500 μL DCM, back extracted into another 500μL 100mM NH4OAc, the aqueous

layers were combined, frozen on dry ice and stored at -20 °C until purification by HPLC. HPLC purification

using a Dionex carbopak column, 300mM NH4OAc, 100μL injections gave 2.7mg of product (63% yield).

Page 120: Design, Synthesis and Biological Utility of Polysaccharide

106 1H NMR (300 MHz, CDCl3, δ (ppm)); 13C NMR (75 MHz, CDCl3, δ (ppm)); 31P NMR (122 MHz, CDCl3, δ

(ppm))

D-methylfucofuranose: Fucose (1g, 6.09 mmol) was placed in a 100 mL flame-dried RBF and suspended

in dry methanol (12.5 mL) and chilled to 0 °C under N2 atmosphere. A solution of HCl in methanol was

prepared by adding acetyl chloride (348μL, 4.87 mmol) to a second flask of cold methanol (12.5mL). A

slow cannulation under positive N2 pressure was performed to transfer the HCl to the solution of sugar.

By the end of the transfer the reaction had become homogenous. Stirred at RT overnight. Added 500mg

of NaHCO3(s) and let stir 5 minutes. Some bubbles formed. Concentrated in vacuo and used immediately

in the next reaction without purification.

2,3,5-tri-O-benzoyl-D-methylfucofuranose: Methylfucofuranose (~1.1g, 6.09 mmol) was dissolved in dry

pyridine (30mL) and chilled to 0 °C under N2. Benzoyl chloride (2.7mL, 23.5 mmol) was added dropwise

and the reaction stirred at RT overnight. Warmed to 30 °C, added 20 μL of benzoyl chloride and stirred

overnight again. Filtered through fluted filter paper, azeotroped filtrate with toluene until all pyridine

was removed. Purified by silica gel chromatography (100% DCM) gave 239mg of product as a mix of

anomers (10% yield). 1H NMR (300 MHz, CDCl3, δ (ppm)); 13C NMR (75 MHz, CDCl3, δ (ppm));

1-O-acetyl-2,3,5-tri-O-benzoyl-D-fucofuranose (4.12): Tri-O-benzoyl methylfucofuranose (239mg, .487

mmol) was dissolved in a 1.4% v/v H2SO4 in Ac2O (4.9mL) and stirred at RT for 1h. The reaction was

chilled to 0 °C, quenched with a 1:1 saturated NaHCO3:brine solution, extracted 3x into DCM, dried over

MgSO4, filtered and concentrated in vacuo. Remaining AcOH was removed by azeotrope with toluene.

Purified by silica gel chromatography (5% 10% 20% EtOAc in hexane) to give 151mg of white solid

(60% yield); 1H NMR (300 MHz, CDCl3, δ (ppm)); 13C NMR (75 MHz, CDCl3, δ (ppm));

Page 121: Design, Synthesis and Biological Utility of Polysaccharide

107 2,3,5-tri-O-benzoyl-1-bromo -β-D-fucofuranose (4.13): 1-O-acetyl-2,3,5-tri-O-benzoyl-D-fucofuranose

(161 mg, .311 mmol) was dissolved in dry DCM (3.1 mL) and chilled to 0 °C under N2. A solution of HBr in

AcOH (30 wt.%, 200 μL, .778 mmol) was added dropwise. After stirring cold for 30 minutes, the contents

were warmed to RT and stirred for 4h. The reaction was azeotroped with dry toluene and used in the

next reaction without further purification or characterization.

Dibenzyl 2,3,5-tri-O-benzoyl-α-D-fucofuranose phosphate (4.14): 2,3,5-tri-O-benzoyl-1-bromo -D-

fucofuranose (~168 mg, .311 mmol) was dissolved in dry toluene (1.0 mL). Separately, a solution of

dibenzyl phosphate (346mg, 1.244 mmol) in dry toluene (2.1 mL) and TEA (172 μL, 1.244 mmol) was

prepared. The bromoglycoside solution was added dropwise over 10 minutes and the reaction stirred at

RT for 2h then was concentrated on rotovap. Purification by silica gel chromatography (10% EtOAc in

toluene) allowed isolation of 77mg of pure α-anomer (34% yield, 2 steps). 1H NMR (300 MHz, CDCl3, δ

(ppm)); 13C NMR (75 MHz, CDCl3, δ (ppm)); 31P NMR (122 MHz, CDCl3, δ (ppm))

Triethylammonium 2,3,5-tri-O-benzoyl-α-D-fucofuranose phosphate: Dibenzyl 2,3,5-tri-O-benzoyl-α-D-

fucofuranose phosphate (77 mg, .104 mmol) was dissolved in ethyl acetate (1.0 mL) and TEA (100 μL)

and treated with Pd/C catalyst (10 wt.%, 35mg). The flask was evacuated and backfilled with H2 from a

balloon 3x and the reaction stirred overnight. The solution was filtered through celite, washed with ethyl

acetate and concentrated on rotovap to give 66mg of cloudy oil (96% yield). 1H NMR (300 MHz, CDCl3, δ

(ppm)); 13C NMR (75 MHz, CDCl3, δ (ppm)); 31P NMR (122 MHz, CDCl3, δ (ppm))

Tributylammonium α-D-fucofuranose phosphate (4.15): Triethylammonium 2,3,5-tri-O-benzoyl-α-D-

fucofuranose phosphate (66 mg, .1 mmol) was dissolved in a 5:2:1 MeOH:H2O:TEA mixture and stirred at

30 °C for 6d. Removed solvent on rotovap and purified by anion exchange chromatography (Biorad AG1-

X8 resin, gradient elution 250 mM 500 mM NH4HCO3 solution) using a peristaltic pump. Column

Page 122: Design, Synthesis and Biological Utility of Polysaccharide

108 fractions were combined, frozen on dry ice and lyophilized. The residue was treated with ~200 μL

Dowex-50WX8-200 resin (pH 7, pretreated with 2.0M HNBu3Cl) for 4 hours. The resin was removed by

filtration through a cotton plug and the filtrate was frozen on dry ice and lyophilized. Isolated 20 mg of a

white solid (58%, 2 steps); 1H NMR (300 MHz, CDCl3, δ (ppm)); 13C NMR (75 MHz, CDCl3, δ (ppm)); 31P

NMR (122 MHz, CDCl3, δ (ppm))

UDP-6-deoxy-Galf (4.01): Trifluoroacetic anhydride was distilled from P2O5 (BP = 36 °C). UMP

tributylammonium was placed in an oven-dried pear flask with stir bar and placed under N2 and chilled

to 0°C. ACN (1.2mL), N,N-dimethylanaline (183 μL) and TEA (50 μL) were added. A solution of

trifluoroacetic anhydride (254 μL) in CAN (450 μL) was prepared at 0 °C under N2 then cannulated to the

flask with UMP, which turned yellow. After 15 minutes, the excess TFAA was evacuated using a needle

equipped to a vacuum line with N2 backfilling from a balloon. A solution of 1-methylimidazole (86 μL) in

ACN (900 μL) and TEA (250 μL) was prepared under N2 and at 0 °C. This solution was cannulated to the

activated UMP solution and stirred at 0 °C for 30 minutes. The tributylammonium 6-deoxy-Galf

monophosphate (20 mg, .027 mmol) was suspended in ACN (250 μL). Activated molecular sieves were

added. The UMP-imidazolium solution (900 μL) was added dropwise over 30 minutes. The reaction

stirred cold for 1.5h then at RT for 3.5h. The coupling was quenched with 1 mL of 100mM NH4OAc

solution, washed with 1 mL DCM, back extracted into another 1 mL 100mM NH4OAc, the aqueous layers

were combined, frozen on dry ice and stored at -20 °C until purification by HPLC. HPLC purification using

a Dionex carbopak column, 300mM NH4OAc, 100μL injections gave 21 mg of product (46% yield). 1H

NMR (300 MHz, CDCl3, δ (ppm)); 13C NMR (75 MHz, CDCl3, δ (ppm)); 31P NMR (122 MHz, CDCl3, δ (ppm))

5,6-anhydro-β-D-methylgalactofuranose: 2,3-di-O-benzyl-5,6-anhydro-β-D-methylgalactofuranose (50

mg, .140 mmol) was dissolved in reagent grade methanol and treated with Pd(OH)2/C (20wt.%, 25 mg)

Page 123: Design, Synthesis and Biological Utility of Polysaccharide

109 catalyst. The flask was evacuated with H2 backfilling from a balloon 3x and stirred overnight. The

contents were filtered through celite, washed with methanol and concentrated to give 31 mg of a faint

gray oil (> 100% - catalyst contaminating); 1H NMR (300 MHz, CDCl3, δ (ppm)); 13C NMR (75 MHz, CDCl3,

δ (ppm));

6-azido-β-D-methylgalactofuranose (4.16): 5,6-anhydro-β-D-methylgalactofuranose (31 mg, ~

.140mmol) was dissolved in dry DMF under N2. Sodium azide (23 mg, .35 mmol) was added, the contents

were capped and parafilmed and heated to 100 °C to stir overnight. The reaction was cooled to RT and

concentrated on rotovap. Purification by silica gel chromatography (5% 10% DCM in methanol) 12mg

of pure product (39% yield). ); 1H NMR (300 MHz, CDCl3, δ (ppm)); 13C NMR (75 MHz, CDCl3, δ (ppm));

6-azido-2,3,5-tri-O-benzoyl-β-D-methylgalactofuranose (4.17): 6-azido-β-D-methylgalactofuranose (12

mg, .055 mmol) was dissolved in anhydrous pyridine (550 μL) and treated with benzoyl chloride (22 μL,

.191 mmol). The contents stirred overnight. Pyridine was removed by toluene azeotrope and the

product was purified by silica gel chromatography (5% 10% 20% EtOAc in hexane) to give a

colorless oil, 3mg (5% yield); 1H NMR (300 MHz, CDCl3, δ (ppm)); 13C NMR (75 MHz, CDCl3, δ (ppm));

1-acetyl-6-azido-2,3,5-tri-O-benzoyl-D-galactofuranose (4.18): 6-azido-2,3,5-tri-O-benzoyl-β-D-

methylgalactofuranose (3 mg, .007 mmol) was dissolved in a 1.4% v/v solution of H2SO4 in Ac2O (100 μL)

and stirred at RT for 2h. Quenched with a 1:1 NaHCO3:NaCl solution, extracted 3x into DCM, dried over

MgSO4, filtered and concentrated in vacuo. Purification by silica gel chromatography (5% 10% 20%

EtOAc in hexane) gave 4mg of a colorless oil (100% yield). 1H NMR (300 MHz, CDCl3, δ (ppm)); 13C NMR

(75 MHz, CDCl3, δ (ppm));

6-azido-2,3,5-tri-O-benzoyl-1-bromo-β-D-galactofuranose (4.18): 1-acetyl-6-azido-2,3,5-tri-O-benzoyl-

D-galactofuranose (4 mg, .007 mmol) was dissolved in dry DCM, chilled to 0 °C and treated dropwise

Page 124: Design, Synthesis and Biological Utility of Polysaccharide

110 with a solution of HBr in AcOH (33 wt.%, 5 μL, .018 mmol). After 30 minutes of stirring cold, the reaction

was warmed to RT and stirred for 2 hours. Azeotrope with dry toluene gave a crude orange oil used in

the next reaction without further purification or characterization.

Diallyl 6-azido-2,3,5-tri-O-benzoyl-α-D-galactofuranose phosphate (4.19): 6-azido-2,3,5-tri-O-benzoyl-

1-bromo-β-D-galactofuranose (~ 4 mg, .007 mmol) was dissolved in dry toluene (100 μL). Separately, a

solution of diallyl phosphate (5 mg, .028 mmol) in dry toluene (50 μL) and TEA (4 μL, .028 mmol) was

prepared. The Galf bromide solution was added dropwise and the reaction stirred at RT for 2 hours. The

contents were concentrated on rotovap and purified by silica gel chromatography (30% 40% EtOAc in

toluene) to give pure α-anomer as a cloudy oil, 1mg (~20% yield, 2 steps). ). 1H NMR (300 MHz, CDCl3, δ

(ppm)); 13C NMR (75 MHz, CDCl3, δ (ppm)); 31P NMR (122 MHz, CDCl3, δ (ppm))

2,3,5-tri-O-benzoyl-6-fluoro-1-(thioethyl)-β-D-galactofuranose (4.22): 1-O-acetyl-2,3,5-tri-O-benzoyl-6-

fluoro-D-galactofuranose (85 mg, .158 mmol) was dissolved in dry DCM under N2 and chilled to 0 °C.

trimethylsilyl ethyl thioether (50 μL, .29 mmol) was added via syringe. Zinc iodide (30 mg, .095 mmol)

was quickly weighed and added to the reaction. The reaction was warmed to RT and stirred 2h. TEA (50

μL) was added to quench the reaction. A precipitate was removed by filtering through a cotton plug,

washing with DCM and the filtrate was concentrated in vacuo. Purification by silica gel chromatography

(5% 10% EtOAc in hexane) gave an oily white residue, 73 mg (86% yield); 1H NMR (300 MHz, CD3OD, δ

(ppm)); 19F NMR (280 MHz, CD3OD, δ (ppm)) ; 13C NMR (75 MHz, CD3OD, δ (ppm))

6-fluoro -2,3,5-tri-O-benzoyl-β-D-Galfuranosyl-(1,5)-2,3,6-tri-O-benzoyl-β-O-allylgalactofuranose

(4.23): 2,3,5-tri-O-benzoyl-6-fluoro-1-(thioethyl)-β-D-galactofuranose (20 mg, .037 mmol) and allyl 2,3,6-

tri-O-benzoyl-β-D-galactofuranose (15.2 mg, .028 mmol) were dissolved in 1 mL dry DCM, treated with

activated molecular sieves and allowed to stir under N2 for 15 minutes at RT. The solution was chilled to

Page 125: Design, Synthesis and Biological Utility of Polysaccharide

111 0 °C and silver triflate (1.4 mg, .006 mmol) and N-iodosuccinimide (8.8 mg, .039 mmol) were added. A

red-brown color developed. After 30 minutes of stirring at 0 °C the reaction was filtered through celite

and washed with DCM until all the pink color was washed into the filtrate. The filtrate was quenched

with 5 mL of Na2S2O3 solution (saturated), 5 mL of saturated NaHCO3 solution, 5 mL of brine. The organic

layer was dried over MgSO4, filtered and concentrated in vacuo. Purification by silica gel

chromatography gave 21 mg of a colorless oil. (70% yield) ); 1H NMR (300 MHz, CD3OD, δ (ppm)); 19F

NMR (280 MHz, CD3OD, δ (ppm)) ; 13C NMR (75 MHz, CD3OD, δ (ppm))

6-fluoro-2,3,5-tri-O-benzoyl-β-D-galactofuranosyl-(1,5)-2,3,6-tri-O-benzoyl-β-D-(12-phenoxy-2-

undecenyl)galactofuranose (4.24): 6-fluoro -2,3,5-tri-O-benzoyl-β-D-Galfuranosyl-(1,5)-2,3,6-tri-O-

benzoyl-β-O-allylgalactofuranose (18 mg, .017 mmol) was dissolved in a solution of 11-phenoxy-

undecene (19 mg, .077 mmol) in DCM (200 μL). The solution was placed under N2 atmosphere, treated

with Grubbs’ 1st generation catalyst (1 mg, .0019 mmol) and stirred at RT overnight. Added 1 more mg of

catalyst in the morning then concentrated on rotovap. Purification by silica gel chromatography (5%

10% 20% EtOAc in hexane) gave 7 mg of cross metathesis product (33% yield). 1H NMR (300 MHz,

CD3OD, δ (ppm)); 19F NMR (280 MHz, CD3OD, δ (ppm)) ; 13C NMR (75 MHz, CD3OD, δ (ppm))

6-fluoro -β-D-galactofuranosyl-(1,5) -β-D-(12-phenoxy-2-undecenyl)galactofuranose (4.25): 6-fluoro-

2,3,5-tri-O-benzoyl-β-D-galactofuranosyl-(1,5)-2,3,6-tri-O-benzoyl-β-D-(12-phenoxy-2-

undecenyl)galactofuranose (5 mg, .004 mmol) was treated with 500 μL of 0.5M NaOMe in methanol and

stirred for 3h at RT under N2. The solution was neutralized with acidic resin, filtered off through a cotton

plug, washed with methanol and concentrated in vacuo. Purification by silica gel chromatography (10%

MeOH in DCM) gave 2.4mg of colorless film (100% yield). 1H NMR (300 MHz, CD3OD, δ (ppm)); 19F NMR

(280 MHz, CD3OD, δ (ppm)) ; 13C NMR (75 MHz, CD3OD, δ (ppm))

Page 126: Design, Synthesis and Biological Utility of Polysaccharide

112 2,3,6-tri-O-benzoyl-5-fluoro-1-(thioethyl)-β-D-galactofuranose (4.26): 1-O-acetyl-2,3,6-tri-O-benzoyl-5-

fluoro-D-galactofuranose (20 mg, ..037 mmol) was dissolved in dry DCM (100 μL) under N2 and chilled

to 0 °C. trimethylsilyl ethyl thioether (9 μL, .056 mmol) was added via syringe. Zinc iodide (6 mg, .018

mmol) was quickly weighed and added to the reaction. The reaction was warmed to RT and stirred 2h.

TEA (2 drops) was added to quench the reaction. A precipitate was removed by filtering through a

cotton plug, washing with DCM and the filtrate was concentrated in vacuo. Purification by silica gel

chromatography (5% 10% EtOAc in hexane) gave an oily white residue, 15 mg (75% yield); 1H NMR

(300 MHz, CD3OD, δ (ppm)); 19F NMR (280 MHz, CD3OD, δ (ppm)) ; 13C NMR (75 MHz, CD3OD, δ (ppm))

5-fluoro -2,3,6-tri-O-benzoyl-β-D-Galfuranosyl-(1,6)-2,3,5-tri-O-benzoyl-β-O-allylgalactofuranose

(4.27): 2,3,6-tri-O-benzoyl-5-fluoro-1-(thioethyl)-β-D-galactofuranose (15 mg, .027 mmol) and allyl 2,3,5-

tri-O-benzoyl-β-D-galactofuranose (11 mg, .021 mmol) were dissolved in 1 mL dry DCM (1.0 mL), treated

with activated molecular sieves and allowed to stir under N2 for 15 minutes at RT. The solution was

chilled to 0 °C and silver triflate (1.3 mg, .005 mmol) and N-iodosuccinimide (7 mg, .029 mmol) were

added. A red-brown color developed. After 30 minutes of stirring at 0 °C the reaction was filtered

through celite and washed with DCM until all the pink color was washed into the filtrate. The filtrate was

quenched with 5 mL of Na2S2O3 solution (saturated), 5 mL of saturated NaHCO3 solution, 5 mL of brine.

The organic layer was dried over MgSO4, filtered and concentrated in vacuo. Purification by silica gel

chromatography gave 10 mg of a colorless oil. (48% yield) ); 1H NMR (300 MHz, CD3OD, δ (ppm)); 19F

NMR (280 MHz, CD3OD, δ (ppm)) ; 13C NMR (75 MHz, CD3OD, δ (ppm))

5-fluoro-2,3,6-tri-O-benzoyl-β-D-galactofuranosyl-(1,6)-2,3,5-tri-O-benzoyl-β-D-(12-phenoxy-2-

undecenyl)galactofuranose (4.28): 5-fluoro -2,3,6-tri-O-benzoyl-β-D-Galfuranosyl-(1,6)-2,3,5-tri-O-

benzoyl-β-O-allylgalactofuranose (10 mg, .010 mmol) was dissolved in a solution of 11-phenoxy-

Page 127: Design, Synthesis and Biological Utility of Polysaccharide

113 undecene (11 mg, .045 mmol) in DCM (100 μL). The solution was placed under N2 atmosphere, treated

with Grubbs’ 1st generation catalyst (1 mg, .001 mmol) and stirred at RT overnight. Added 1 more mg of

catalyst in the morning then concentrated on rotovap. Purification by silica gel chromatography (5%

10% 20% EtOAc in hexane) gave 5.8 mg of cross metathesis product (40% yield). 1H NMR (300 MHz,

CD3OD, δ (ppm)); 19F NMR (280 MHz, CD3OD, δ (ppm)) ; 13C NMR (75 MHz, CD3OD, δ (ppm))

5-fluoro -β-D-galactofuranosyl-(1,6) -β-D-(12-phenoxy-2-undecenyl)galactofuranose (4.29): 5-fluoro-

2,3,6-tri-O-benzoyl-β-D-galactofuranosyl-(1,6)-2,3,5-tri-O-benzoyl-β-D-(12-phenoxy-2-

undecenyl)galactofuranose (5.8 mg, .004 mmol) was treated with 600 μL of 0.5M NaOMe in methanol

and stirred for 3h at RT under N2. The solution was neutralized with acidic resin, filtered off through a

cotton plug, washed with methanol and concentrated in vacuo. Purification by silica gel chromatography

(10% MeOH in DCM) gave 2.2 mg of colorless film (92% yield). 1H NMR (300 MHz, CD3OD, δ (ppm)); 19F

NMR (280 MHz, CD3OD, δ (ppm)) ; 13C NMR (75 MHz, CD3OD, δ (ppm))

Page 128: Design, Synthesis and Biological Utility of Polysaccharide

114

References

(1) Sauvage, E.; Kerff, F.; Terrak, M.; Ayala, J. A.; Charlier, P. FEMS Microbiol. Rev. 2008, 32, 234. (2) Yother, J. Annu. Rev. Microbiol. 2011, 65, 563. (3) Ramsey, D. M.; Wozniak, D. J. Molecular Microbiology 2005, 56, 309. (4) Liu, B.; Knirel, Y. A.; Feng, L.; Perepelov, A. V.; Senchenkova, S. y. N.; Wang, Q.; Reeves, P. R.; Wang, L. FEMS Microbiol. Rev. 2008, 32, 627. (5) Samuel, G.; Reeves, P. Carbohydrate Research 2003, 338, 2503. (6) Swoboda, J. G.; Campbell, J.; Meredith, T. C.; Walker, S. ChemBioChem 2010, 11, 35. (7) Nothaft, H.; Szymanski, C. M. Nat Rev Micro 2010, 8, 765. (8) Guerry, P.; Szymanski, C. M. Trends in Microbiology 2008, 16, 428. (9) Eugster, M. R.; Haug, M. C.; Huwiler, S. G.; Loessner, M. J. Molecular Microbiology 2011, 81, 1419. (10) Shen, G.-J.; Datta, A. K.; Izumi, M.; Koeller, K. M.; Wong, C.-H. J. Biol. Chem. 1999, 274, 35139. (11) Schmidt, R. R.; Pedersen, C. M.; Qiao, Y.; Zahringer, U. Org. Biomol. Chem. 2011, 9, 2040. (12) Zhang, Z.; McCallum, S. A.; Xie, J.; Nieto, L.; Corzana, F.; Jim nez-Barbero, J. s.; Chen, M.; Liu, J.; Linhardt, R. J. J. Am. Chem. Soc. 2008, 130, 12998. (13) Raetz, C. R. H.; Whitfield, C. Annu. Rev. Biochem. 2002, 71, 635. (14) Lairson, L. L.; Henrissat, B.; Davies, G. J.; Withers, S. G. Annu. Rev. Biochem. 2008, 77, 521. (15) Osawa, T.; Sugiura, N.; Shimada, H.; Hirooka, R.; Tsuji, A.; Shirakawa, T.; Fukuyama, K.; Kimura, M.; Kimata, K.; Kakuta, Y. Biochemical and Biophysical Research Communications 2009, 378, 10. (16) Gründling, A.; Schneewind, O. Proc. Natl. Acad. Sci. U. S. A. 2007, 104, 8478. (17) Lu, D.; Wörmann, M. E.; Zhang, X.; Schneewind, O.; Gründling, A.; Freemont, P. S. Proc. Natl. Acad. Sci. U. S. A. 2009, 106, 1584. (18) Yuan, Y.; Barrett, D.; Zhang, Y.; Kahne, D.; Sliz, P.; Walker, S. Proc. Natl. Acad. Sci. U. S. A. 2007, 104, 5348. (19) Huang, C.-Y.; Shih, H.-W.; Lin, L.-Y.; Tien, Y.-W.; Cheng, T.-J. R.; Cheng, W.-C.; Wong, C.-H.; Ma, C. Proc. Natl. Acad. Sci. U. S. A. 2012, 109, 6496. (20) Steiner, K.; Novotny, R.; Werz, D. B.; Zarschler, K.; Seeberger, P. H.; Hofinger, A.; Kosma, P.; Schäffer, C.; Messner, P. J. Biol. Chem. 2008, 283, 21120. (21) Splain, R. A.; Kiessling, L. L. Bioorg. Med. Chem. 2010, 18, 3753. (22) Schwarz, F.; Lizak, C.; Fan, Y.-Y.; Fleurkens, S.; Kowarik, M.; Aebi, M. Glycobiology 2011, 21, 45. (23) Woodward, R.; Yi, W.; Li, L.; Zhao, G.; Eguchi, H.; Sridhar, P. R.; Guo, H.; Song, J. K.; Motari, E.; Cai, L.; Kelleher, P.; Liu, X.; Han, W.; Zhang, W.; Ding, Y.; Li, M.; Wang, P. G. Nat. Chem. Biol. 2010, 6, 418. (24) Han, W.; Wu, B.; Li, L.; Zhao, G.; Woodward, R.; Pettit, N.; Cai, L.; Thon, V.; Wang, P. G. J. Biol. Chem. 2012, 287, 5357. (25) Chen, M. M.; Weerapana, E.; Ciepichal, E.; Stupak, J.; Reid, C. W.; Swiezewska, E.; Imperiali, B. Biochemistry 2007, 46, 14342.

Page 129: Design, Synthesis and Biological Utility of Polysaccharide

115 (26) Morrison, J. P.; Troutman, J. M.; Imperiali, B. Bioorganic & Medicinal Chemistry 2010, 18, 8167. (27) Glover, K. J.; Weerapana, E.; Imperiali, B. Proceedings of the National Academy of Sciences of the United States of America 2005, 102, 14255. (28) Li, L.; Woodward, R.; Ding, Y.; Liu, X.-w.; Yi, W.; Bhatt, V. S.; Chen, M.; Zhang, L.-w.; Wang, P. G. Biochemical and Biophysical Research Communications 2010, 394, 1069. (29) Kelly, J.; Jarrell, H.; Millar, L.; Tessier, L.; Fiori, L. M.; Lau, P. C.; Allan, B.; Szymanski, C. M. Journal of Bacteriology 2006, 188, 2427. (30) Feldman, M. F.; Wacker, M.; Hernandez, M.; Hitchen, P. G.; Marolda, C. L.; Kowarik, M.; Morris, H. R.; Dell, A.; Valvano, M. A.; Aebi, M. Proceedings of the National Academy of Sciences of the United States of America 2005, 102, 3016. (31) Wacker, M.; Linton, D.; Hitchen, P. G.; Nita-Lazar, M.; Haslam, S. M.; North, S. J.; Panico, M.; Morris, H. R.; Dell, A.; Wren, B. W.; Aebi, M. Science 2002, 298, 1790. (32) Ryder, C.; Byrd, M.; Wozniak, D. J. Current Opinion in Microbiology 2007, 10, 644. (33) Remminghorst, U.; Rehm, B. H. A. Applied and Environmental Microbiology 2006, 72, 298. (34) Remminghorst, U.; Hay, I. D.; Rehm, B. H. A. Journal of Biotechnology 2009, 140, 176. (35) Franklin, M. J.; Chitnis, C. E.; Gacesa, P.; Sonesson, A.; White, D. C.; Ohman, D. E. Journal of Bacteriology 1994, 176, 1821. (36) Schiller, N. L.; Monday, S. R.; Boyd, C. M.; Keen, N. T.; Ohman, D. E. Journal of Bacteriology 1993, 175, 4780. (37) Remminghorst, U.; Rehm, B. H. A. FEBS Letters 2006, 580, 3883. (38) Oglesby, L. L.; Jain, S.; Ohman, D. E. Microbiology 2008, 154, 1605. (39) Aarons, S. J.; Sutherland, I. W.; Chakrabarty, A. M.; Gallagher, M. P. Microbiology 1997, 143, 641. (40) Robles-Price, A.; Wong, T. Y.; Sletta, H.; Valla, S.; Schiller, N. L. Journal of Bacteriology 2004, 186, 7369. (41) Rose, N. L.; Completo, G. C.; Lin, S. J.; McNeil, M.; Palcic, M. M.; Lowary, T. L. J. Am. Chem. Soc. 2006, 128, 6721. (42) Webb, A. J.; Karatsa-Dodgson, M.; Gründling, A. Molecular Microbiology 2009, 74, 299. (43) Brown, S.; Zhang, Y.-H.; Walker, S. Chemistry & Biology 2008, 15, 12. (44) Formstone, A.; Carballido-López, R.; Noirot, P.; Errington, J.; Scheffers, D.-J. Journal of Bacteriology 2008, 190, 1812. (45) Holland, L. M.; Conlon, B.; O'Gara, J. P. Microbiology 2011, 157, 408. (46) Allison, S. E.; D'Elia, M. A.; Arar, S.; Monteiro, M. A.; Brown, E. D. J. Biol. Chem. 2011, 286, 23708. (47) Atilano, M. L.; Pereira, P. M.; Yates, J.; Reed, P.; Veiga, H.; Pinho, M. G.; Filipe, S. R. Proc. Natl. Acad. Sci. U. S. A. 2010, 107, 18991. (48) Campbell, J.; Singh, A. K.; Santa Maria, J. P.; Kim, Y.; Brown, S.; Swoboda, J. G.; Mylonakis, E.; Wilkinson, B. J.; Walker, S. ACS Chem. Biol. 2010, 6, 106. (49) Brown, S.; Meredith, T.; Swoboda, J.; Walker, S. Chemistry & Biology 2010, 17, 1101. (50) Chen, L.; Men, H.; Ha, S.; Ye, X.-Y.; Brunner, L.; Hu, Y.; Walker, S. Biochemistry 2002, 41, 6824. (51) Troutman, J. M.; Imperiali, B. Biochemistry 2009, 48, 2807. (52) Sewell, E. W. C.; Pereira, M. P.; Brown, E. D. J. Biol. Chem. 2009, 284, 21132.

Page 130: Design, Synthesis and Biological Utility of Polysaccharide

116 (53) Perlstein, D. L.; Andrew Wang, T.-S.; Doud, E. H.; Kahne, D.; Walker, S. J. Am. Chem. Soc. 2009, 132, 48. (54) Cartee, R. T.; Forsee, W. T.; Schutzbach, J. S.; Yother, J. J. Biol. Chem. 2000, 275, 3907. (55) Meselson, M.; Stahl, F. W. Proc. Natl. Acad. Sci. U. S. A. 1958, 44, 671. (56) Levengood, M. R.; Splain, R. A.; Kiessling, L. L. J. Am. Chem. Soc. 2011, 133, 12758. (57) Cartee, R. T.; Forsee, W. T.; Yother, J. Journal of Bacteriology 2005, 187, 4470. (58) Besra, G. S.; Khoo, K. H.; McNeil, M. R.; Dell, A.; Morris, H. R.; Brennan, P. J. Biochemistry 1995, 34, 4257. (59) Kremer, L.; Dover, L. G.; Morehouse, C.; Hitchin, P.; Everett, M.; Morris, H. R.; Dell, A.; Brennan, P. J.; McNeil, M. R.; Flaherty, C.; Duncan, K.; Besra, G. S. J. Biol. Chem. 2001, 276, 26430. (60) Berg, S.; Kaur, D.; Jackson, M.; Brennan, P. J. Glycobiology 2007, 17, 35R. (61) Beláňová, M.; Dianišková, P.; Brennan, P. J.; Completo, G. C.; Rose, N. L.; Lowary, T. L.; Mikušová, K. Journal of Bacteriology 2008, 190, 1141. (62) May, J. F.; Splain, R. A.; Brotschi, C.; Kiessling, L. L. Proc. Natl. Acad. Sci. U. S. A. 2009, 106, 11851. (63) Schertzer, J. W.; Brown, E. D. J. Biol. Chem. 2003, 278, 18002. (64) Pereira, M. P.; Schertzer, J. W.; D'Elia, M. A.; Koteva, K. P.; Hughes, D. W.; Wright, G. D.; Brown, E. D. ChemBioChem 2008, 9, 1385. (65) Forsee, W. T.; Cartee, R. T.; Yother, J. J. Biol. Chem. 2006, 281, 6283. (66) Forsee, W. T.; Cartee, R. T.; Yother, J. J. Biol. Chem. 2009, 284, 11826. (67) Schirner, K.; Marles-Wright, J.; Lewis, R. J.; Errington, J. EMBO J 2009, 28, 830. (68) Poulin, M. B.; Zhou, R.; Lowary, T. L. Org. Biomol. Chem. 2012, 10, 4074. (69) Weigel, P. H.; DeAngelis, P. L. J. Biol. Chem. 2007, 282, 36777. (70) Weigel, P. H.; Kyossev, Z.; Torres, L. C. J. Biol. Chem. 2006, 281, 36542. (71) Tlapak-Simmons, V. L.; Kempner, E. S.; Baggenstoss, B. A.; Weigel, P. H. J. Biol. Chem. 1998, 273, 26100. (72) Tlapak-Simmons, V. L.; Baggenstoss, B. A.; Kumari, K.; Heldermon, C.; Weigel, P. H. J. Biol. Chem. 1999, 274, 4246. (73) Baggenstoss, B. A.; Weigel, P. H. Analytical Biochemistry 2006, 352, 243. (74) Kyossev, Z.; Weigel, P. H. Analytical Biochemistry 2007, 371, 62. (75) Kumari, K.; Baggenstoss, B. A.; Parker, A. L.; Weigel, P. H. J. Biol. Chem. 2006, 281, 11755. (76) Wörmann, M. E.; Corrigan, R. M.; Simpson, P. J.; Matthews, S. J.; Gründling, A. Molecular Microbiology 2011, 79, 566. (77) Carter, J. A.; Jiménez, J. C.; Zaldívar, M.; Álvarez, S. A.; Marolda, C. L.; Valvano, M. A.; Contreras, I. Microbiology 2009, 155, 3260. (78) Morona, R.; van den Bosch, L.; Manning, P. A. Journal of Bacteriology 1995, 177, 1059. (79) Bastin, D. A.; Stevenson, G.; Brown, P. K.; Haase, A.; Reeves, P. R. Molecular Microbiology 1993, 7, 725. (80) Larue, K.; Kimber, M. S.; Ford, R.; Whitfield, C. J. Biol. Chem. 2009, 284, 7395. (81) Schertzer, J. W.; Brown, E. D. Journal of Bacteriology 2008, 190, 6940. (82) Sutcliffe, I. C. Molecular Microbiology 2011, 79, 553. (83) Forsee, W. T.; Cartee, R. T.; Yother, J. J. Biol. Chem. 2009, 284, 11836. (84) Clarke, B. R.; Cuthbertson, L.; Whitfield, C. J. Biol. Chem. 2004, 279, 35709. (85) Clarke, B. R.; Richards, M. R.; Greenfield, L. K.; Hou, D.; Lowary, T. L.; Whitfield, C. J. Biol. Chem. 2011, 286, 41391.

Page 131: Design, Synthesis and Biological Utility of Polysaccharide

117 (86) Clarke, B. R.; Greenfield, L. K.; Bouwman, C.; Whitfield, C. J. Biol. Chem. 2009, 284, 30662. (87) Messner, P.; Steiner, K.; Zarschler, K.; Schäffer, C. Carbohydrate Research 2008, 343, 1934. (88) Kiessling, L. L.; Splain, R. A. Annu. Rev. Biochem. 2010, 79, 619. (89) Steiner, K.; Hagelueken, G.; Messner, P.; Schäffer, C.; Naismith, J. H. Journal of Molecular Biology 2010, 397, 436. (90) Kido, N.; Kobayashi, H. Journal of Bacteriology 2000, 182, 2567. (91) May, J. F.; Levengood, M. R.; Splain, R. A.; Brown, C. D.; Kiessling, L. L. Biochemistry 2012. (92) Brown, C. D.; Rusek, M. S.; Kiessling, L. L. J. Am. Chem. Soc. 2012, 134, 6552. (93) Keenleyside, W. J.; Perry, M.; Maclean, L.; Poppe, C.; Whitfield, C. Molecular Microbiology 1994, 11, 437. (94) Keenleyside, W. J.; Whitfield, C. J. Biol. Chem. 1996, 271, 28581. (95) Keenleyside, W. J.; Clarke, A. J.; Whitfield, C. Journal of Bacteriology 2001, 183, 77. (96) Vionnet, J.; Vann, W. F. Glycobiology 2007, 17, 735. (97) Morley, T. J.; Withers, S. G. J. Am. Chem. Soc. 2010, 132, 9430. (98) Campbell, J.; Singh, A. K.; Swoboda, J. G.; Gilmore, M. S.; Wilkinson, B. J.; Walker, S. Antimicrobial Agents and Chemotherapy 2012, 56, 1810. (99) Pan, F.; Jackson, M.; Ma, Y.; McNeil, M. J. Bacteriol. 2001, 183, 3991. (100) Sassetti, C. M.; Rubin, E. J. Proc. Natl. Acad. Sci. U. S. A. 2003, 100, 12989. (101) Dykhuizen, E. C.; May, J. F.; Tongpenyai, A.; Kiessling, L. L. J. Am. Chem. Soc. 2008, 130, 6706. (102) Mikusova, K.; Yagi, T.; Stern, R.; McNeil, M. R.; Besra, G. S.; Crick, D. C.; Brennan, P. J. J. Biol. Chem. 2000, 275, 33890. (103) Mikušová, K.; Beláňová, M.; Korduláková, J.; Honda, K.; McNeil, M. R.; Mahapatra, S.; Crick, D. C.; Brennan, P. J. Journal of Bacteriology 2006, 188, 6592. (104) Gandolfi-Donadío, L.; Gallo-Rodriguez, C.; de Lederkremer, R. M. J. Org. Chem. 2003, 68, 6928. (105) Szczepina, M. G.; Zheng, R. B.; Completo, G. C.; Lowary, T. L.; Pinto, B. M. ChemBioChem 2009, 10, 2052. (106) Completo, G. C.; Lowary, T. L. J. Org. Chem. 2008, 73, 4513. (107) Tsvetkov, Y. E.; Nikolaev, A. V. J. Chem. Soc. 2000, 889. (108) Marlow, A. L.; Kiessling, L. L. Org. Lett. 2001, 3, 2517. (109) Peltier, P.; Daniellou, R.; Nugier-Chauvin, C.; Ferrières, V. Org. Lett. 2007, 9, 5227. (110) Errey, J. C.; Mukhopadhyay, B.; Kartha, K. P. R.; Field, R. A. Chem. Commun. 2004, 2706. (111) Peltier, P.; BeláHová, M.; Dianiaková, P.; Zhou, R.; Zheng, R. B.; Pearcey, J. A.; Joe, M.; Brennan, P. J.; Nugier-Chauvin, C.; Ferrières, V.; Lowary, T. L.; Daniellou, R.; Mikuaová, K. Chem. Biol. 2010, 17, 1356. (112) Bordoni, A.; Lima, C.; Mariño, K.; de Lederkremer, R. M.; Marino, C. Carbohydrate Research 2008, 343, 1863. (113) Anslyn, E. a. D., D. Modern Physical Organic Chemistry; University Science Books: Sausalito, 2006. (114) Hagen, K.; Hedberg, K. J. Am. Chem. Soc. 1973, 95, 8263. (115) Guvench, O.; MacKerell, A. D. J. Phys. Chem. A 2006, 110, 9934. (116) Withers, S. G.; Street, I. P.; Bird, P.; Dolphin, D. H. J. Am. Chem. Soc. 1987, 109, 7530.

Page 132: Design, Synthesis and Biological Utility of Polysaccharide

118 (117) Allman, S. A.; Jensen, H. H.; Vijayakrishnan, B.; Garnett, J. A.; Leon, E.; Liu, Y.; Anthony, D. C.; Sibson, N. R.; Feizi, T.; Matthews, S.; Davis, B. G. ChemBioChem 2009, 10, 2522. (118) Northrup, A. B.; MacMillan, D. W. C. Science 2004, 305, 1752. (119) Covell, D. J.; Vermeulen, N. A.; Labenz, N. A.; White, M. C. Angewandte Chemie International Edition 2006, 45, 8217. (120) Halcomb, R. L.; Danishefsky, S. J. J. Am. Chem. Soc. 1989, 111, 6661. (121) Sanders, W. J.; Kiessling, L. L. Tet. Lett. 1994, 35, 7335. (122) Furuya, T.; Kamlet, A. S.; Ritter, T. Nature 2011, 473, 470. (123) Schaus, S. E.; Brandes, B. D.; Larrow, J. F.; Tokunaga, M.; Hansen, K. B.; Gould, A. E.; Furrow, M. E.; Jacobsen, E. N. J. Am. Chem. Soc. 2002, 124, 1307. (124) Mohamady, S.; Desoky, A.; Taylor, S. D. Org. Lett. 2011, 14, 402. (125) Peltier, P.; Guégan, J.-P.; Daniellou, R.; Nugier-Chauvin, C.; Ferrières, V. Eur. J. Org. Chem. 2008, 2008, 5988. (126) Moretti, R.; Thorson, J. S. J. Biol. Chem. 2007, 282, 16942. (127) Wheatley, R. W.; Zheng, R. B.; Richards, M. R.; Lowary, T. L.; Ng, K. K. S. J. Biol. Chem. 2012. (128) Sanger, F.; Nicklen, S.; Coulson, A. R. Proc. Natl. Acad. Sci. U. S. A. 1977, 74, 5463. (129) Pouilly, S.; Bourgeaux, V.; Piller, F.; Piller, V. ACS Chem. Biol. 2012. (130) Barthel, S. R.; Antonopoulos, A.; Cedeno-Laurent, F.; Schaffer, L.; Hernandez, G.; Patil, S. A.; North, S. J.; Dell, A.; Matta, K. L.; Neelamegham, S.; Haslam, S. M.; Dimitroff, C. J. J. Biol. Chem. 2011, 286, 21717. (131) Ruprecht, R. M.; O'Brien, L. G.; Rossoni, L. D.; Nusinoff-Lehrman, S. Nature 1986, 323, 467. (132) Szczepina, M. G.; Zheng, R. B.; Completo, G. C.; Lowary, T. L.; Pinto, B. M. Bioorganic & Medicinal Chemistry 2010, 18, 5123.

Page 133: Design, Synthesis and Biological Utility of Polysaccharide

119

Appendix I

NMR Spectra

Page 134: Design, Synthesis and Biological Utility of Polysaccharide

120

Page 135: Design, Synthesis and Biological Utility of Polysaccharide

121

Page 136: Design, Synthesis and Biological Utility of Polysaccharide

122

Page 137: Design, Synthesis and Biological Utility of Polysaccharide

123

Page 138: Design, Synthesis and Biological Utility of Polysaccharide

124

Page 139: Design, Synthesis and Biological Utility of Polysaccharide

125

Page 140: Design, Synthesis and Biological Utility of Polysaccharide

126

Page 141: Design, Synthesis and Biological Utility of Polysaccharide

127

Page 142: Design, Synthesis and Biological Utility of Polysaccharide

128

Page 143: Design, Synthesis and Biological Utility of Polysaccharide

129

Page 144: Design, Synthesis and Biological Utility of Polysaccharide

130

Page 145: Design, Synthesis and Biological Utility of Polysaccharide

131

Page 146: Design, Synthesis and Biological Utility of Polysaccharide

132

Page 147: Design, Synthesis and Biological Utility of Polysaccharide

133

Page 148: Design, Synthesis and Biological Utility of Polysaccharide

134

Page 149: Design, Synthesis and Biological Utility of Polysaccharide

135

Page 150: Design, Synthesis and Biological Utility of Polysaccharide

136

Page 151: Design, Synthesis and Biological Utility of Polysaccharide

137

Page 152: Design, Synthesis and Biological Utility of Polysaccharide

138

Page 153: Design, Synthesis and Biological Utility of Polysaccharide

139

Page 154: Design, Synthesis and Biological Utility of Polysaccharide

140

Page 155: Design, Synthesis and Biological Utility of Polysaccharide

141

Page 156: Design, Synthesis and Biological Utility of Polysaccharide

142

Page 157: Design, Synthesis and Biological Utility of Polysaccharide

143

Page 158: Design, Synthesis and Biological Utility of Polysaccharide

144

Page 159: Design, Synthesis and Biological Utility of Polysaccharide

145

Page 160: Design, Synthesis and Biological Utility of Polysaccharide

146

Page 161: Design, Synthesis and Biological Utility of Polysaccharide

147

Page 162: Design, Synthesis and Biological Utility of Polysaccharide

148

Page 163: Design, Synthesis and Biological Utility of Polysaccharide

149

Page 164: Design, Synthesis and Biological Utility of Polysaccharide

150

Page 165: Design, Synthesis and Biological Utility of Polysaccharide

151

Page 166: Design, Synthesis and Biological Utility of Polysaccharide

152

Page 167: Design, Synthesis and Biological Utility of Polysaccharide

153

Page 168: Design, Synthesis and Biological Utility of Polysaccharide

154

Page 169: Design, Synthesis and Biological Utility of Polysaccharide

155

Page 170: Design, Synthesis and Biological Utility of Polysaccharide

156

Page 171: Design, Synthesis and Biological Utility of Polysaccharide

157

Page 172: Design, Synthesis and Biological Utility of Polysaccharide

158

Page 173: Design, Synthesis and Biological Utility of Polysaccharide

159

Page 174: Design, Synthesis and Biological Utility of Polysaccharide

160

Page 175: Design, Synthesis and Biological Utility of Polysaccharide

161

Page 176: Design, Synthesis and Biological Utility of Polysaccharide

162

Page 177: Design, Synthesis and Biological Utility of Polysaccharide

163

Page 178: Design, Synthesis and Biological Utility of Polysaccharide

164

Page 179: Design, Synthesis and Biological Utility of Polysaccharide

165

Page 180: Design, Synthesis and Biological Utility of Polysaccharide

166

Page 181: Design, Synthesis and Biological Utility of Polysaccharide

167

Page 182: Design, Synthesis and Biological Utility of Polysaccharide

168

Page 183: Design, Synthesis and Biological Utility of Polysaccharide

169

Page 184: Design, Synthesis and Biological Utility of Polysaccharide

170

Page 185: Design, Synthesis and Biological Utility of Polysaccharide

171

Page 186: Design, Synthesis and Biological Utility of Polysaccharide

172

Page 187: Design, Synthesis and Biological Utility of Polysaccharide

173

Page 188: Design, Synthesis and Biological Utility of Polysaccharide

174

Page 189: Design, Synthesis and Biological Utility of Polysaccharide

175

Page 190: Design, Synthesis and Biological Utility of Polysaccharide

176

Page 191: Design, Synthesis and Biological Utility of Polysaccharide

177

Page 192: Design, Synthesis and Biological Utility of Polysaccharide

178

Page 193: Design, Synthesis and Biological Utility of Polysaccharide

179

Page 194: Design, Synthesis and Biological Utility of Polysaccharide

180

Page 195: Design, Synthesis and Biological Utility of Polysaccharide

181

Page 196: Design, Synthesis and Biological Utility of Polysaccharide

182

Page 197: Design, Synthesis and Biological Utility of Polysaccharide

183

Page 198: Design, Synthesis and Biological Utility of Polysaccharide

184

Page 199: Design, Synthesis and Biological Utility of Polysaccharide

185

Page 200: Design, Synthesis and Biological Utility of Polysaccharide

186

Page 201: Design, Synthesis and Biological Utility of Polysaccharide

187

Page 202: Design, Synthesis and Biological Utility of Polysaccharide

188

Page 203: Design, Synthesis and Biological Utility of Polysaccharide

189

Page 204: Design, Synthesis and Biological Utility of Polysaccharide

190

Page 205: Design, Synthesis and Biological Utility of Polysaccharide

191

Page 206: Design, Synthesis and Biological Utility of Polysaccharide

192

Page 207: Design, Synthesis and Biological Utility of Polysaccharide

193

Page 208: Design, Synthesis and Biological Utility of Polysaccharide

194

Page 209: Design, Synthesis and Biological Utility of Polysaccharide

195

Page 210: Design, Synthesis and Biological Utility of Polysaccharide

196

Page 211: Design, Synthesis and Biological Utility of Polysaccharide

197

Page 212: Design, Synthesis and Biological Utility of Polysaccharide

198

Page 213: Design, Synthesis and Biological Utility of Polysaccharide

199

Page 214: Design, Synthesis and Biological Utility of Polysaccharide

200

Page 215: Design, Synthesis and Biological Utility of Polysaccharide

201

Page 216: Design, Synthesis and Biological Utility of Polysaccharide

202

Page 217: Design, Synthesis and Biological Utility of Polysaccharide

203

Page 218: Design, Synthesis and Biological Utility of Polysaccharide

204

Page 219: Design, Synthesis and Biological Utility of Polysaccharide

205

Page 220: Design, Synthesis and Biological Utility of Polysaccharide

206

Page 221: Design, Synthesis and Biological Utility of Polysaccharide

207

Page 222: Design, Synthesis and Biological Utility of Polysaccharide

208

Page 223: Design, Synthesis and Biological Utility of Polysaccharide

209

Page 224: Design, Synthesis and Biological Utility of Polysaccharide

210

Page 225: Design, Synthesis and Biological Utility of Polysaccharide

211

Page 226: Design, Synthesis and Biological Utility of Polysaccharide

212

Page 227: Design, Synthesis and Biological Utility of Polysaccharide

213

Page 228: Design, Synthesis and Biological Utility of Polysaccharide

214

Page 229: Design, Synthesis and Biological Utility of Polysaccharide

215

Page 230: Design, Synthesis and Biological Utility of Polysaccharide

216

Page 231: Design, Synthesis and Biological Utility of Polysaccharide

217

Page 232: Design, Synthesis and Biological Utility of Polysaccharide

218

Page 233: Design, Synthesis and Biological Utility of Polysaccharide

219

Page 234: Design, Synthesis and Biological Utility of Polysaccharide

220

Page 235: Design, Synthesis and Biological Utility of Polysaccharide

221

Page 236: Design, Synthesis and Biological Utility of Polysaccharide

222

Page 237: Design, Synthesis and Biological Utility of Polysaccharide

223

Page 238: Design, Synthesis and Biological Utility of Polysaccharide

224

Page 239: Design, Synthesis and Biological Utility of Polysaccharide

225

Page 240: Design, Synthesis and Biological Utility of Polysaccharide

226

Page 241: Design, Synthesis and Biological Utility of Polysaccharide

227

Page 242: Design, Synthesis and Biological Utility of Polysaccharide

228

Page 243: Design, Synthesis and Biological Utility of Polysaccharide

229

Page 244: Design, Synthesis and Biological Utility of Polysaccharide

230

Page 245: Design, Synthesis and Biological Utility of Polysaccharide

231

Page 246: Design, Synthesis and Biological Utility of Polysaccharide

232

Page 247: Design, Synthesis and Biological Utility of Polysaccharide

233

Page 248: Design, Synthesis and Biological Utility of Polysaccharide

234

Page 249: Design, Synthesis and Biological Utility of Polysaccharide

235

Page 250: Design, Synthesis and Biological Utility of Polysaccharide

236

Page 251: Design, Synthesis and Biological Utility of Polysaccharide

237

Page 252: Design, Synthesis and Biological Utility of Polysaccharide

238

Page 253: Design, Synthesis and Biological Utility of Polysaccharide

239

Page 254: Design, Synthesis and Biological Utility of Polysaccharide

240

Page 255: Design, Synthesis and Biological Utility of Polysaccharide

241

Page 256: Design, Synthesis and Biological Utility of Polysaccharide

242

Page 257: Design, Synthesis and Biological Utility of Polysaccharide

243

Page 258: Design, Synthesis and Biological Utility of Polysaccharide

244

Page 259: Design, Synthesis and Biological Utility of Polysaccharide

245

Page 260: Design, Synthesis and Biological Utility of Polysaccharide

246

Page 261: Design, Synthesis and Biological Utility of Polysaccharide

247

Page 262: Design, Synthesis and Biological Utility of Polysaccharide

248

Page 263: Design, Synthesis and Biological Utility of Polysaccharide

249

Page 264: Design, Synthesis and Biological Utility of Polysaccharide

250

Page 265: Design, Synthesis and Biological Utility of Polysaccharide

251

Page 266: Design, Synthesis and Biological Utility of Polysaccharide

252

Page 267: Design, Synthesis and Biological Utility of Polysaccharide

253

Page 268: Design, Synthesis and Biological Utility of Polysaccharide

254

Page 269: Design, Synthesis and Biological Utility of Polysaccharide

255

Page 270: Design, Synthesis and Biological Utility of Polysaccharide

256

Page 271: Design, Synthesis and Biological Utility of Polysaccharide

257

Page 272: Design, Synthesis and Biological Utility of Polysaccharide

258

Page 273: Design, Synthesis and Biological Utility of Polysaccharide

259

Page 274: Design, Synthesis and Biological Utility of Polysaccharide

260

Page 275: Design, Synthesis and Biological Utility of Polysaccharide

261

Page 276: Design, Synthesis and Biological Utility of Polysaccharide

262