mapping the enzyme specificities of intestinal maltase

258
MAPPING THE ENZYME SPECIFICITIES OF INTESTINAL MALTASE-GLUCOAMYLASE AND SUCRASE-ISOMALTASE by Razieh Eskandari B.Sc. (Chemistry), Yasouj University, 2003 M.Sc. (Chemistry), Shiraz University, 2006 THESIS SUBMITTED IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY in the Department of Chemistry Faculty of Science © Razieh Eskandari 2012 SIMON FRASER UNIVERSITY Spring 2012 All rights reserved. However, in accordance with the Copyright Act of Canada, this work may be reproduced, without authorization, under the conditions for “Fair Dealing.” Therefore, limited reproduction of this work for the purposes of private study, research, criticism, review and news reporting is likely to be in accordance with the law, particularly if cited appropriately.

Upload: others

Post on 03-Feb-2022

7 views

Category:

Documents


0 download

TRANSCRIPT

MAPPING THE ENZYME SPECIFICITIES OF

INTESTINAL MALTASE-GLUCOAMYLASE AND SUCRASE-ISOMALTASE

by

Razieh Eskandari

B.Sc. (Chemistry), Yasouj University, 2003 M.Sc. (Chemistry), Shiraz University, 2006

THESIS SUBMITTED IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF

DOCTOR OF PHILOSOPHY

in the Department of Chemistry

Faculty of Science

© Razieh Eskandari 2012 SIMON FRASER UNIVERSITY

Spring 2012

All rights reserved. However, in accordance with the Copyright Act of Canada, this work may be reproduced, without authorization, under the conditions for “Fair Dealing.” Therefore, limited reproduction of this work for the purposes of private

study, research, criticism, review and news reporting is likely to be in accordance with the law, particularly if cited appropriately.

ii

APPROVAL

Name: Razieh Eskandari

Degree: Doctor of Philosophy

Title of Thesis: Mapping the Enzyme Specificities of Intestinal Maltase-Glucoamylase and Sucrase-Isomaltase Examining Committee: Chair: Dr. Tim Storr Assistant Professor Dr. B. Mario Pinto Professor Senior Supervisor Dr. Steven Holdcroft Professor Supervisor Dr. Peter D. Wilson Associate Professor Supervisor Dr. Andrew J. Bennet Professor Internal Examiner, Department of Chemistry Dr. Jeffrey W. Keillor Professor External Examiner , Department of Chemistry University of Ottawa

Date Defended/Approved: January 20, 2012

Last revision: Spring 09

Declaration of Partial Copyright Licence The author, whose copyright is declared on the title page of this work, has granted to Simon Fraser University the right to lend this thesis, project or extended essay to users of the Simon Fraser University Library, and to make partial or single copies only for such users or in response to a request from the library of any other university, or other educational institution, on its own behalf or for one of its users.

The author has further granted permission to Simon Fraser University to keep or make a digital copy for use in its circulating collection (currently available to the public at the “Institutional Repository” link of the SFU Library website <www.lib.sfu.ca> at: <http://ir.lib.sfu.ca/handle/1892/112>) and, without changing the content, to translate the thesis/project or extended essays, if technically possible, to any medium or format for the purpose of preservation of the digital work.

The author has further agreed that permission for multiple copying of this work for scholarly purposes may be granted by either the author or the Dean of Graduate Studies.

It is understood that copying or publication of this work for financial gain shall not be allowed without the author’s written permission.

Permission for public performance, or limited permission for private scholarly use, of any multimedia materials forming part of this work, may have been granted by the author. This information may be found on the separately catalogued multimedia material and in the signed Partial Copyright Licence.

While licensing SFU to permit the above uses, the author retains copyright in the thesis, project or extended essays, including the right to change the work for subsequent purposes, including editing and publishing the work in whole or in part, and licensing other parties, as the author may desire.

The original Partial Copyright Licence attesting to these terms, and signed by this author, may be found in the original bound copy of this work, retained in the Simon Fraser University Archive.

Simon Fraser University Library Burnaby, BC, Canada

iii

ABSTRACT

In humans, maltase-glucoamylase (MGAM) and sucrase-isomaltase (SI)

are the small intestinal glucosidases responsible for catalyzing the last glucose-

releasing step in starch digestion. MGAM and SI are each composed of

duplicated catalytic domains, N-terminal membrane domains (ntMGAM and ntSI)

and C-terminal luminal domains (ctMGAM and ctSI). They display

complementary substrate specificities for the mixture of short, linear and

branched oligosaccharide substrates that typically make up terminal starch-

digestion products. As they are involved in the breakdown of dietary starch and

sugars into glucose, regulating their activities with α-glucosidase inhibitors is an

attractive approach to control blood glucose levels for the prevention and

treatment of type-2 diabetes.

This thesis work deals with mapping (determination of selectivity and

specificity) of MGAM and SI with synthetic inhibitors. The syntheses and

enzymatic evaluation of sulfonium-ion glucosidase inhibitors, with potent

inhibitory activities against intestinal glucosidases are the main topics of this

thesis. First, an alternative route for the synthesis of kotalanol, a naturally-

occurring sulfonium-ion glucosidase inhibitor isolated from Salacia reticulata, and

its 6'-epimer are described, and the inhibitory activities of these compounds

against ntMGAM are reported. Second, the total syntheses of de-O-sulfonated

iv

ponkoranol, another naturally-occurring sulfonium-ion glucosidase inhibitor

isolated from the same species, its 5'-epimer, and their selenium analogues are

described. The synthetic route is also extended to obtain 3'-O-methylponkoranol.

The inhibitory activities of these latter compounds against the four human

intestinal glucosidase enzymes, ntMGAM, ctMGAM, ntSI, and ctSI are examined.

Finally, from the structural studies of ntMGAM, it was postulated that ctMGAM

might have an extended binding site compared to ntMGAM, which favours

binding of longer inhibitors such as acarbose (an antidiabetic drug that is

currently in use for the treatment of type-2 diabetes). Based on this difference,

the syntheses of candidate inhibitors containing maltose extensions at 3'- and 5'-

of de-O-sulfonated ponkoranol are described.

The inhibition of maltose hydrolysis suggests that selective inhibition of

one enzyme unite over the others is possible despite relatively small structural

changes in the inhibitor. This panel of inhibitors can now be used to turn off

certain enzymes while probing the action of others with respect to starch

digestion.

Keywords: Glucosidase inhibitors; Salacia reticulata; kotalanol; ponkoranol; maltase-glucoamylase; sucrase-isomaltase; type-2 diabetes; sulfonium-ions; selenonium ions.

v

DEDICATION

This work is dedicated to my respected parents and my teachers who

helped me every step

of the way so that I can get to where I am today.

It is also dedicated to my lovely husband, Mehdi

who was extremely patient with me all these years.

vi

ACKNOWLEDGMENTS

All praise is to God Almighty Who continues to bless my life.

I would like to thank my senior supervisor, Dr. B. Mario Pinto, for his

support and guidance and giving me the opportunity to work in his laboratory.

I would like to thank my supervisory committee: Dr. Peter Wilson and Dr.

Steven Holdcroft and my examining committee: Dr. Andrew J. Bennet and Dr.

Jeffery W. Keiller for their valuable advice and feedback. I would also like to

thank our collaborators, Dr. David R. Rose, Ms. Kyra Jones, and Dr. Douglas A.

Kuntz for the enzyme inhibition data. I am grateful to Dr. Andrew Lewis and Mr.

Colin Zhang for the NMR help and Mr. Hongwen Chen for acquiring the high

resolution mass Spectra.

I thank Dr. H. Sharghi, my M.Sc. supervisor, who taught me how to work

in a chemistry laboratory. I would like to thank Dr. Silvia Borrelli for proof reading

my introduction chapter and Dr. Niloufar Choubdar and Dr. Rehana Hossany for

helping me in the lab during the first year of my PhD program and being

wonderful friends. I would like to thank Dr. Jayakanthan Kumarasamy, Dr.

Ravinder Reddy, with whom I collaborated and Dr. Sankar Mohan for his help.

Last but not least, my deepest gratitude to my parents, my sister for their

support and love. My deepest thanks go to my husband, Mehdi, for his support,

patience and encouragement.

vii

TABLE OF CONTENTS

Approval .............................................................................................................................ii Abstract ............................................................................................................................. iii Dedication ......................................................................................................................... v Acknowledgments .............................................................................................................vi Table of Contents ............................................................................................................. vii List of Figures.................................................................................................................... x List of Schemes ............................................................................................................... xiii List of Tables ................................................................................................................... xiv Abbreviations ...................................................................................................................xv

CHAPTER 1: General introduction ................................................................................ 1 1.1 Carbohydrates .......................................................................................................... 1 1.2 Carbohydrates in health and disease ....................................................................... 4 1.3 Starch digesting enzymes ......................................................................................... 5

1.3.1 Glycoside hydrolases .................................................................................... 8 1.3.2 Alpha-amylases ............................................................................................. 8 1.3.3 Brush-border hydrolases ............................................................................... 8 1.3.1 Disaccharidase deficiency in heath and disease ........................................ 12

1.4 Glycosidase mechanism of action .......................................................................... 13 1.4.1 Mechanism of retaining glycosidases ......................................................... 13 1.4.2 Mechanism of inverting glycosidases .......................................................... 15

1.5 Transition-state mimics ........................................................................................... 16 1.6 Glycosidase inhibitors ............................................................................................. 19

1.6.1 Iminosugars................................................................................................. 20 1.6.2 Carbasugars................................................................................................ 24 1.6.3 Marine organosulfates ................................................................................. 25 1.6.4 Sulfonium-sulfate thiosugars ....................................................................... 27

1.7 Thesis overview ...................................................................................................... 34 1.8 References ............................................................................................................. 37

CHAPTER 2: Synthesis of a biologically active isomer of kotalanol, a naturally-occurring glucosidase inhibitor .................................................................. 44 2.1 Keywords ................................................................................................................ 46 2.2 Abstract ................................................................................................................... 46 2.3 Introduction ............................................................................................................. 46 2.4 Results and discussion ........................................................................................... 50 2.5 Experimental ........................................................................................................... 55

2.5.1 General ....................................................................................................... 55 2.5.2 Enzyme inhibition assays ............................................................................ 55

viii

2.5.3 Compound characterization data ................................................................ 56 2.6 Acknowledgments ................................................................................................... 66 2.7 References ............................................................................................................. 67 2.8 Supporting Information ........................................................................................... 69

CHAPTER 3: Potent glucosidase inhibitors: de-O-sulfonated ponkoranol and its stereoisomer ..................................................................................................... 82 3.1 Keywords ................................................................................................................ 84 3.2 Abstract ................................................................................................................... 84 3.3 Introduction ............................................................................................................. 84 3.4 Results and discussion ........................................................................................... 88 3.5 Experimental ........................................................................................................... 93

3.5.1 General methods......................................................................................... 93 3.5.2 Compound characterization data ................................................................ 94

3.6 Acknowledgments ................................................................................................... 98 3.7 References ............................................................................................................. 98 3.8 Supporting Information ......................................................................................... 101

CHAPTER 4: The effect of heteroatom substitution of sulfur for selenium in glucosidase inhibitors on intestinal α-glucosidase activities ................................ 108 4.1 Keywords .............................................................................................................. 110 4.2 Abstract ................................................................................................................. 110 4.3 Introduction ........................................................................................................... 110 4.4 Results and discussion ......................................................................................... 113 4.5 Experimental ......................................................................................................... 118

4.5.1 General methods....................................................................................... 118 4.5.2 Enzyme kinetics ........................................................................................ 118 4.5.3 Compound characterization data .............................................................. 119

4.6 Acknowledgments ................................................................................................. 122 4.7 References ........................................................................................................... 123 4.8 Supporting Information ......................................................................................... 126

CHAPTER 5: Probing the active-site requirements of human intestinal N-terminal maltase-glucoamylase: The effect of replacing the sulfate moiety by a methyl ether in ponkoranol, a naturally-occurring α-glucosidase inhibitor ........................................................................................................................ 132 5.1 Keywords .............................................................................................................. 134 5.2 Abstract ................................................................................................................. 134 5.3 Introduction ........................................................................................................... 135 5.4 Results and discussion ......................................................................................... 139 5.5 Experimental ......................................................................................................... 144

5.5.1 General methods....................................................................................... 144 5.5.2 Enzyme kinetics ........................................................................................ 145 5.5.3 Compound characterization data .............................................................. 145

5.6 Acknowledgments ................................................................................................. 152 5.7 References ........................................................................................................... 152

ix

5.8 Supporting Information ......................................................................................... 155

CHAPTER 6: Selectivity of 3'-O-methylponkoranol for inhibition of N- and C-terminal maltase-glucoamylase and sucrase-isomaltase, potential therapeutics for digestive disorders or their sequelae ........................................... 165 6.1 Keywords .............................................................................................................. 167 6.2 Abstract ................................................................................................................. 167 6.3 Introduction ........................................................................................................... 167 6.4 Results and discussion ......................................................................................... 173 6.5 Acknowledgments ................................................................................................. 176 6.6 References ........................................................................................................... 177

CHAPTER 7: Probing the intestinal α-glucosidase enzyme specificities of starch-digesting maltase-glucoamylase and sucrase-isomaltase: Synthesis and inhibitory properties of 3′- and 5′- maltose-extended de-O-sulfonated ponkoranol ................................................................................................................... 179 7.1 Keywords .............................................................................................................. 181 7.2 Abstract ................................................................................................................. 181 7.3 Introduction ........................................................................................................... 182 7.4 Results and discussion ......................................................................................... 187 7.5 Experimental ......................................................................................................... 199

7.5.1 Compound characterization data .............................................................. 199 7.6 Acknowledgments ................................................................................................. 211 7.7 References ........................................................................................................... 212 7.8 Supporting Information ......................................................................................... 216

CHAPTER 8: Conclusions and future work .............................................................. 231 8.1 Conclusions .......................................................................................................... 231 8.2 Future work ........................................................................................................... 235 8.3 References ........................................................................................................... 239

x

LIST OF FIGURES

Figure 1-1: Glucose (1), lactose (2), sucrose (3). ........................................................ 2 Figure 1-2: Components of starch: amylose and amylopectin. .................................... 3 Figure 1-3: Starch digestion by the action of salivary and pancreatic α–

amylases and small intestinal α–glucosidases. ......................................... 7 Figure 1-4: Pictorial representation of brush-border membrane-bound MGAM

and SI. ....................................................................................................... 9 Figure 1-5: Surface representation of the ntSI and ntMGAM active sites

(Reproduced from J. Biol. Chem. 2010, 285, 17763–17770. © the American Society for Biochemistry and Molecular Biology). ................... 11

Figure 1-6: Transition state proposed for glycosidase catalyzed reactions. .............. 17 Figure 1-7: Proposed half chair and boat conformations of pyranosyl cations

in glycosidase-mediated hydrolysis reactions. ........................................ 18 Figure 1-8: Proposed envelope conformations of furanosyl cations in

glycosidase mediated-hydrolysis reactions. ............................................ 18 Figure 1-9: Positive charge build-up at the exocyclic oxygen, endocyclic

oxygen, and the anomeric carbon during glycosidase-mediated hydrolysis reactions. ................................................................................ 19

Figure 1-10: Skeletal frameworks most commonly found in naturally-occurring iminosugar glycosidase inhibitors and their synthetic derivatives. .......... 20

Figure 1-11: Nojirimycin (12), 1-deoxynojirimycin (13), and isofagomine (14). ............ 21 Figure 1-12: Naturally-occurring polyhydroxylated pyrrolidine-based

glycosidase inhibitors. ............................................................................. 22 Figure 1-13: Structures of selected, naturally-occurring pyrrolizidine alkaloids,

which display α-glucosidase inhibitory activities. ..................................... 23 Figure 1-14: Castanospermine (23), swainsonine (24). ............................................... 24 Figure 1-15: Examples of nortropanes (25, 26). .......................................................... 24 Figure 1-16: Examples of naturally-occurring carbasugar α-glucosidase

inhibitors and the antibiotic validamycin. ................................................. 25 Figure 1-17: Structures of organosulfate α-glucosidase inhibitors isolated from

marine invertebrates. ............................................................................... 26 Figure 1-18: Structures of organosulfate α-glucosidase inhibitors isolated from

marine invertebrates. ............................................................................... 27

xi

Figure 1-19: Structure of an N-oxide analogue of castanospermine and a sulfonium-ion analogue. .......................................................................... 28

Figure 1-20: Structure of compounds 40-43. ............................................................... 29 Figure 1-21: Naturally-occurring zwitterionic sulfonium-sulfate glucosidase

inhibitors. ................................................................................................. 29 Figure 1-22: Initially proposed structure of a naturally-occurring sulfoxide α-

glucosidase inhibitor. ............................................................................... 31 Figure 1-23: Initially proposed structure of the naturally-occurring α-

glucosidase inhibitor neosalacinol. .......................................................... 32 Figure 2-1: Components isolated from Salacia species. ............................................ 48 Figure 2-2: Kotalanol stereoisomer. ........................................................................... 48 Figure 3-1: Components isolated from Salacia species. ............................................ 85 Figure 3-2: Proposed structure of neosalacinol. ........................................................ 86 Figure 3-3: De-O-sulfonated ponkoranol and its 5’-stereoisomer. ............................. 88 Figure 3-4: 1D-NOESY correlations of selected protons in compounds 18 and

21. ............................................................................................................ 92 Figure 4-1: Components isolated from Salacia species. .......................................... 111 Figure 4-2: Structure of the 5′-stereoisomer of de-O-sulfonated ponkoranol. .......... 112 Figure 4-3: Superimposition of the ring carbon atoms of the proposed

intermediate in glucosidase-catalyzed reactions (in green) and the selenonium ion (in blue). ....................................................................... 113

Figure 4-4: Selenium analogues of de-O-sulfonated ponkoranol and its 5’-stereoisomer. ......................................................................................... 113

Figure 4-5: 2D-NOESY correlations of selected protons in compound 13. .............. 116 Figure 5-1: Components isolated from Salacia species. .......................................... 136 Figure 5-2: De-O-sulfonated ponkoranol and its 5’-stereoisomer. ........................... 137 Figure 5-3: Effect of removing the sulfate group. Superposition of kotalanol (4)

(orange) and de-O-sulfonated kotalanol (5) (purple) structures. Double-headed arrows show the proximities of the sulfate group to the surrounding hydrophobic residues Y299, W406 and F575. (Reproduced with permission from Biochemistry, 2010, 49, 443–451. Copyright American Chemical Society). ........................................ 138

Figure 5-4: 3’-O-Methylponkoranol. ......................................................................... 138 Figure 5-5: 1D-NOESY correlations of selected protons in compound 9. ................ 143 Figure 6-1: Components of starch: amylose and amylopectin. ................................ 168 Figure 6-2: Schematic diagram of MGAM and SI indicating their hydrolytic

activities. ................................................................................................ 169 Figure 6-3: Sulfonium-ion α-glucosidase inhibitors 1-7. ........................................... 171

xii

Figure 6-4: Representative Lineweaver-Burk plot of ctMGAM-N2 inhibited by 3 at concentrations of 0 nM, 75 nM, 125 nM, and 200 nM. ................... 174

Figure 7-1: Schematic diagram of MGAM and SI indicating hydrolytic activity. ....... 183 Figure 7-2: Components 1-8 isolated from Salacia species. ................................... 184 Figure 7-3: Structure of acarbose 9, an α-glucosidase inhibitor currently used

in the treatment of type-2 diabetes. ....................................................... 185 Figure 7-4: 3′-O-β-maltosyl-de-O-sulfonated ponkoranol 10 and 5′-O-β-

maltosyl-de-O-sulfonated ponkoranol 11. .............................................. 187 Figure 7-5: Structure of 3'-O-methylponkoranol 26. ................................................. 192 Figure 7-6: NOESY correlations in compound 24. ................................................... 194 Figure 8-1: Structures of compounds 1-8. ............................................................... 232 Figure 8-2: Structures of compounds 9-11. ............................................................. 235 Figure 8-3: Proposed selective inhibitors of ntSI. .................................................... 236

xiii

LIST OF SCHEMES

Scheme 1-1: Proposed mechanism for retaining glycosidases. ............................... 15 Scheme 1-2: Proposed mechanism for inverting glycosidases. ................................ 16 Scheme 2-1: First attempted synthesis of kotalanol 4. ............................................. 49 Scheme 2-2: Synthesis of kotalanol 4. ...................................................................... 50 Scheme 2-3: Retrosynthetic analysis. ....................................................................... 51 Scheme 2-4: Synthesis of the diols 18 and 19. ......................................................... 52 Scheme 2-5: Synthesis of the cyclic sulfates 12 and 22. .......................................... 53 Scheme 2-6: Coupling reactions. .............................................................................. 54 Scheme 3-1: Retrosynthetic analysis. ....................................................................... 89 Scheme 3-2: First attempted synthesis of 8. ............................................................. 89 Scheme 3-3: Second attempted synthesis of 8. ........................................................ 90 Scheme 3-4: Synthesis of compound 8. ................................................................... 91 Scheme 3-5: Synthesis of compound 9. ................................................................... 92 Scheme 4-1: Retrosynthetic analysis. ..................................................................... 114 Scheme 4-2: Synthesis of compounds 7 and 8. ..................................................... 115 Scheme 5-1: Retrosynthetic analysis. ..................................................................... 139 Scheme 5-2: First attempted synthesis of 9. ........................................................... 140 Scheme 5-3: Second attempted synthesis of 9. ..................................................... 141 Scheme 5-4: Synthesis of compound 9. ................................................................. 143 Scheme 7-1: Retrosynthetic analysis. ..................................................................... 188 Scheme 7-2: Synthesis of benzyl 6-O-trifluoromethanesulfonyl-D-

glucopyranoside derivatives 17 and 22, with benzylated maltose units at C-2 or C-4. .............................................................. 190

Scheme 7-3: Coupling reactions to give sulfonium-ions. ........................................ 191 Scheme 7-4: Synthesis of compounds 10 and 11. ................................................. 193 Scheme 8-1: Proposed synthetic route for target compound 12. ............................ 237

xiv

LIST OF TABLES

Table 1-1: Comparison of inhibitory activities (IC50 in µM) of naturally-occurring sulfonium-ion glucosidase inhibitors 44-46, 48 and 49 against rat intestinal α-glucosidases. ............................................................................... 33

Table 3-1: Experimentally determined Ki values (nM) of compounds 2-5, 8 and 9.. ........ 87 Table 4-1: Experimentally determined Ki values (µM) of compounds 5-8 ...................... 117 Table 6-1: IC50 (µM) against disaccharidases. ............................................................... 172 Table 6-2: Experimentally determined Ki values (µM) of 3 and 7. ................................. 175 Table 7-1: Comparison of inhibition profiles against MGAM and SI subunits, Ki

(µM) .............................................................................................................. 198

xv

ABBREVIATIONS

Å Ångström, 10-10 m

AcOH acetic acid

aq aqueous

Ala alanine

Ar aromatic

B boat

Bn benzyl

br broad

C chair

c concentration

Calcd calculated

CAZy carbohydrate-active enzyme database

COSY correlation spectroscopy

ct Carboxyl terminal

ctMGAM C-terminal domain of MGAM

ctMGAM-N2 C-terminal domain of MGAM, isoform N2

ctMGAM-N20 C-terminal domain of MGAM, isoform N20

ctSI C-terminal domain of SI

d doublet

xvi

dd doublet of doublets

ddd doublet of doublets of doublets

dt doublet of triplets

DMF N, N-dimethylformamide

DAB 1,4-dideoxy-1,4-imino-D-arabinitol

DGDP 2,5-dideoxy-2,5-imino-D-glucitol

DMDP 2,5-dideoxy-2,5-imino-D-mannitol

E envelope

F575 phenylalanine575

EtOAc ethyl acetate

GH glycoside hydrolase

Glu glutamic acid

h hour

H half chair

HFIP 1,1,1,3,3,3-hexafluoro-2-propanol

HRMS high resolution mass spectrometry

HSQC heteronuclear single quantum coherence

IC50 concentration required to inhibit 50% of the enzyme activity

J coupling constant in Hertz

KIE kinetic isotope effect

Ki inhibition constant

Km Michaelis constant

xvii

Kmobs Km observed in presence of inhibitor

Leu leucine

m multiplet

Me methyl

MeOH methanol

MGAM maltase-glucoamylase

NMR nuclear magnetic resonance

NOESY nuclear Overhauser effect spectroscopy

nt amino terminal

ntMGAM N-terminal domain of MGAM

ntSI N-terminal domain of SI

Ph phenyl

PMB para-methoxybenzyl

pr propyl

psi lb/inch2

s singlet

SI sucrase-isomaltase

t triplet

TFA trifluoroacetic acid

THF tetrahydrofuran

Thr threonine

TLC thin layer chromatography

xviii

Trp tryptophan

TS transition state

Tyr tyrosine

Tris tris(hydroxymethyl)-aminomethane

V max enzyme saturated velocity

Val valine

W406 tryptophan406

Y 299 tyrosine299

1

CHAPTER 1: GENERAL INTRODUCTION

1.1 Carbohydrates

Carbohydrates are the major source of metabolic energy in the modern

human diet. During the nineteenth century carbohydrates were defined as

molecules made of carbon, oxygen and hydrogen atoms, with the general

formula Cn(H2O)n. However, this definition has been modified to include

derivatives of carbohydrates as well as nitrogen containing carbohydrates that do

not fit into this formula. Today in general, polyhydroxy aldehydes or ketones,

alcohols, acids, their simple analogues and their heterocyclic analogues are also

considered to be carbohydrates.

In the modern human diet, carbohydrates are found in many food sources

such as cereals, fruits and vegetables. Plants, the main source of carbohydrates,

use water and carbon dioxide in the photosynthesis process for the production of

carbohydrates. The most known of the carbohydrate family is glucose (1, Figure

1-1). A single carbohydrate molecules such as glucose (1), is called a

monosaccharide.

When a monosaccharide is covalently linked to another carbohydrate

molecule, it is called a disaccharide. The most common disaccharides are

lactose (2, Figure 1-1), and sucrose (3, Figure 1-1). Lactose (milk sugar) (2), is

the primary source of carbohydrates for nursing infants, and sucrose (table

2

sugar) (3) is found mainly in fruits, vegetables or as an added sweetener for food

and beverages.1

Figure 1-1: Glucose (1), lactose (2), sucrose (3).

Polysaccharides are composed of repeating units of either mono- or di-

saccharides, and are also called glycans.

Carbohydrates can be classified as digestible or non-digestible,

nutritionally.2,3 Digestible carbohydrates include starch and simple mono and

disaccharide sugars. Disaccharides and larger molecules from this class can

break down into their monosaccharide components that can be absorbed in the

upper intestinal tract. Non-digestible carbohydrates such as fibers, inulin and

polyols are resistant to digestion by intestinal enzymes but are instead processed

by bacterial enzymes in the colon.

Starch is the most abundant storage polysaccharide and a major source of

metabolic energy in plants. Starch has been used by humans, even before they

discovered how to write. The Egyptians used it to bake their bread 5000 years

ago and to glue papyrus despite a lack of knowledge of the fermentation

Glucose (1) Lactose (milk sugar, 2)

O

OHOHHO

OHHO

O

OHOHO

OHHO

O

OHHO

OH

HO

O

OHOH

OH

O

OH

O

HO

HO

OH

OH

Sucrose (table sugar, 3)

3

process. Persians and Indians used it to make dishes like wheat halva. Rice

starch as surface treatment of paper has been used in paper production in China,

from 700 AD onwards. Starch is made up of two main structural components:

amylose, a long linear chain joined by α-1,4-linked glucose units and

amylopectin, a larger branched molecule with linear α-1,4-linked glucose chains

with approximately 5% having additional α-1,6-linked branch points (Figure 1-2).4

In contrast to the digestion of simpler carbohydrates such as sucrose and lactose

that require the activities of sucrase and lactase, respectively, starch digestion

requires the action of six enzymes. This is discussed further in section 1.3. The

multiplicity of animal starch-digesting enzymes mirrors the multiplicity of the

starch synthetic enzymes of plants.5

Figure 1-2: Components of starch: amylose and amylopectin.

Amylose

Amylopectin

α-1,4α-1,6

4

Sugar alcohols are another group of carbohydrates which have been used

in recent years as part of “sugar-free” and low-calorie diets. Sugar alcohols such

as sorbitol, xylitol and maltitol are reduced forms of sugars, naturally found in

some fruits and vegetables, and can be used in place of sucrose to sweeten

foods. They are partially hydrolyzed by intestinal enzymes and thus are

incompletely absorbed into the bloodstream.

1.2 Carbohydrates in health and disease

In order to meet the needs of humans for energy produced from glucose

oxidation, carbohydrate rich diets are important.6 The rapid change in diet and

lifestyle of humans, has not occurred without side effects.

In the diet of early humans, when an excess of carbohydrates was

available their storage by conversion to fats was important to provide energy

requirements during times of low food availability.7 However, the present rapid

change in patterns of carbohydrate consumption by humans involves ingestion of

large quantities of carbohydrates in short periods of time.7

In many cases, the energy intake exceeds the requirement, leading to the

storage of carbohydrates in the form of body fat and causing development of

obesity. 8 In addition, high intake of carbohydrates can result the development of

type-2 diabetes. Blood glucose levels are tightly regulated through the secretion

of insulin from β-cells of the pancreas, which, in turn, stimulates uptake of

glucose into muscle cells. The digestion and absorption of large quantities of

carbohydrates, imposes pressure on β-cells for secretion of insulin. In individuals,

5

this chronic stress has been implicated as a damaging factor for pancreatic β-

cells and cellular response mechanisms that contribute to the development of

type-2 diabetes.9 An approach to control high blood sugar levels, is to delay

intestinal glucose absorption through the inhibition of starch-digesting enzymes

with α-glucosidase inhibitors. This is further discussed in Section 1.6. To

understand and manipulate the physiologic responses of the human organism to

carbohydrate feeding, studies to identify the multiple enzyme/starch (as the

highest consumption by the human population) interactions and the glucosidic

activities of the human gastrointestinal tract are necessary.

1.3 Starch digesting enzymes

The starch digesting enzymes belong to the class of enzymes known as

Glycoside Hydrolases (GH). Glycoside hydrolases, also known as glycosidases,

are a prevalent class of enzymes that cleave the glycosidic bond between two

carbohydrate molecules, one of the strongest bonds found in natural polymers.

These enzymes are capable of breaking glycosidic bonds 1017 times faster than

the uncatalyzed reaction.10 The mechanism of action of glycosidases will be

discussed in Section 1.4.

Depending on their point of action in a carbohydrate chain, glycosidases

are classified as endo- or exohydrolases: endohydrolases hydrolyze internal

glycosidic linkages of a carbohydrate chain to give smaller chains, whereas,

exohydrolases cleave the glycosidic linkage at the non-reducing end of a glycan

structure, releasing monomer units.11,12

6

In humans, four different enzymes mediate the digestion of ingested

carbohydrates. First two endo-acting glucosidases, salivary and pancreatic α-

amylases, break down the complex starch molecules into smaller linear and

branched oligomers (limit dextrins) which are then further hydrolyzed into glucose

by two small-intestinal, brush-border exohydrolases, maltase-glucoamylase

(MGAM) and sucrase-isomaltase (SI); glucose is then absorbed into the

bloodstream (Figure 1-3).

7

Figure 1-3: Starch digestion by the action of salivary and pancreatic α–amylases and small intestinal α–glucosidases.

α-Amylase

Limit dextrins

Sucrase-IsomaltaseMaltase-Glucoamylase

Glucose

α-1,4 linkageα-1,6 linkage

α-amylase activities

α-1,4 linkageα-1,6 linkageMGAM & SI activitiesSI activity

8

1.3.1 Glycoside hydrolases

There are currently 125 different GH families, which are defined according

to amino acid sequence similarity, according to the “Carbohydrate Active Enzyme

database (CAZY).11 Salivary and pancreatic α–amylases belong to GH13

whereas, two small-intestinal, brush-border maltase-glucoamylase (MGAM) and

sucrase-isomaltase (SI) belong to GH31.12,13

1.3.2 Alpha-amylases

Amylase was the first enzyme identified by biochemists. The enzymes

require Ca2+ and Cl− ions to display their full activity.14,15 After birth, full production

of amylases occur as a weaning adaptation to carbohydrate feeding, and this

stimulus remains during life as the primary stimulus leading to the synthesis,

processing, and secretion of pancreatic α-amylase and other digestive enzymes.

GH13 salivary and pancreatic α-amylases are believed to catalyze the hydrolysis

of starch via the retaining mechanism (discussed in Section 1.4.1) and cleave the

internal α-1,4 bonds of amylose and amylopectin (Figure 1-3).

1.3.3 Brush-border hydrolases

For effective release of free glucose, glucose oligomers resulting from the

amylase activities have to be further hydrolyzed by four exoglucosidic activities of

the small intestine. These activities consist of the enzymes sucrase-isomaltase

(SI) and maltase-glucoamylase (MGAM), each composed of two subunits

containing catalytic sites that act on the non-reducing ends of linear glucose

oligomers, with substantial release of free glucose monomers (Figure 1-3).16

9

MGAM and SI share the common ancestry therefore, they share many similar

features. The individual MGAM and SI domains were referred to as either N-

terminal (ntSI and ntMGAM) for the domain closest to the membrane anchor or

C-terminal (ctSI and ctMGAM) for the luminal domain to avoid the confusion of

the associated domain activities (Figure 1-4).

Figure 1-4: Pictorial representation of brush-border membrane-bound MGAM and SI.

MGAM is less abundant than SI and comprises about 2% of brush-border

membrane proteins.13 MGAM has very little activity for starch itself but shows

greatly enhanced hydrolytic activities with α-amylase pre-treated starch.17

10

The four mucosal maltase enzymes are referred to as α-glucosidase

activities because they hydrolyze the nonreducing end of all α-1,4-glucose

oligomers to free glucose. In addition, the isomaltase subunit of SI also shows α-

1,6-glucosidic activity that cleaves the linkages present in the branching points of

intermediate branched oligomers and isomaltose (Figure 1-3). To understand the

structural basis for differential substrate specificity observed between ntMGAM

and ntSI, their active sites were compared.

1.3.3.1 Crystal structures of brush-border hydrolases

Thus far, of the four α-glucosidase activities only the crystal structures of

ntMGAM and ntSI have been solved.18,19

From kinetic studies, it was observed that maltose (Glu-α-(1,4)-Glu) could

be hydrolyzed by both ntMGAM and ntSI whereas isomaltose (Glu-α-(1,6)-Glu)

could only be hydrolyzed efficiently by ntSI. The different substrate specificities of

ntMGAM and ntSI seem consistent with their roles in terminal starch digestion.19

The active site of ntMGAM and ntSI were found to be composed of a

substrate-binding pocket comprising –1 and +1 subsites.18,20 Substrates bind to

the pocket via their non-reducing end, with the non-reducing sugar ring

interacting with the –1 subsite and the reducing ring interacting with the +1

subsite. Substrate cleavage occurs between –1 and +1 subsites, following a

catalytic mechanism that results in retention of configuration at the anomeric

center (discussed in Section 1.4.1).

11

Structural superposition of the ntMGAM and ntSI kotalanol-bound

structures (Section 1.6.4) revealed identical conservation of -1 subsite residues.

Therefore, substrate discrimination is most likely mediated by the +1 subsite and

there are accordingly structural differences observed between the ntSI and

ntMGAM in this region. A notable structural difference between ntMGAM and ntSI

subsite architecture includes substitution of smaller residues in ntMGAM (Thr205,

Ala576, Tyr299 and Thr204) by larger residues in ntSI (Leu233, Val605, Trp327

and Gln232) that results in a wide and open ntMGAM +1 subsite compared to the

narrow groove ntSI +1 subsite (Figure 1-5).19

Figure 1-5: Surface representation of the ntSI and ntMGAM active sites (Reproduced from J. Biol. Chem. 2010, 285, 17763–17770. © the American Society for Biochemistry and Molecular Biology).

It seems counterintuitive that the narrow +1 subsite of ntSI would

accommodate both its α-1,4 and α-1,6 substrates, whereas a wider +1 subsite in

ntMGAM would account for specificity for the α-1,4 substrate. However, perhaps

the substrate with α-1,6 branch points requires constraining prior to hydrolysis.

12

Based on sequence comparison of the ntMGAM and SI domains, in ntSI Trp327

may be important in conferring α-1,6 specificity since it is conserved as Tyr in

ntMGAM, ctMGAM and ctSI.20

1.3.1 Disaccharidase deficiency in heath and disease

In the three-month-old human embryo, all of the intestinal disaccharidases

are already active.21 The α-disaccharidase activities reach the normal adult level

in the sixth to seventh month of fetal life. The only exception is maltase, which is

still low in the newborn. But carbohydrate malabsorption is not rare in children.

The symptoms vary and are dependent on diet and age. The common symptoms

are abdominal pain and watery diarrhea especially for lactase deficiency and

congenital sucrase-isomaltase deficiency (CSID) due to the osmotic force of the

undigested sugar that attracts fluid into the small intestinal lumen. Intraluminal

fluid accumulation increases peristalsis and decreases small intestinal transit

time.22, 23 The symptoms of maltase-glucoamylase (MGAM) deficiency are poorly

defined. Glucoamylase deficiency may also cause chronic diarrhea in young

children.22 The osmotic force of starch is less than that of sucrose or lactose

because of its larger molecular weight.24

Congenital sucrase-isomaltase deficiency (CSID) is characterized by

absence or deficiency of the mucosal sucrase-isomaltase enzyme. Specific

diagnosis requires upper gastrointestinal biopsy. Recently, 13C-sucrose breath

tests have been used as a safe, simple and non-invasive method of CSID

diagnosis. In simple terms, 13C-sucrose breath tests involve the oral ingestion of

labelled 13C-glucose and 13C-sucrose as substrates to overnight-fasting patients

13

on two separate days, followed by collecting breath samples. These 13C-labeled

sugars are cleaved by the action of the specific glycosidases. Further oxidation

gives 13C-enriched 13CO2 which is excreted in a patient's breath. 13CO2 breath

enrichments are assayed using an infrared spectrometer. Sucrose digestion and

oxidation are calculated as a mean percent coefficient of glucose oxidation

averaged between 30 and 90 minutes.25

CSID, also known as disaccharide intolerance, in patients is relieved if

sucrose is removed from the diet or an oral supplement of sucrase enzyme is fed

with the sugar.

1.4 Glycosidase mechanism of action

Glycosidase mediated hydrolysis can occur via two main mechanisms:

retaining or inverting. In a retaining glycosidase-catalyzed reaction, the hydrolysis

product will have retention of configuration at the anomeric carbon, whereas the

product from an inverting glycosidase-catalyzed reaction will have inversion of

configuration at the anomeric center. The sugar substrate can be a five-

membered ring (furanose) or a six-membered ring (pyranose), thus leading to the

classification of furanoside or pyranoside hydrolases.26 In 1953, Koshland

proposed the mechanism for these two main groups of glycosidases which are

still widely supported.27

1.4.1 Mechanism of retaining glycosidases

This mechanism is illustrated using α-glycosidase as an example (Scheme

1-1). In this mechanism, glycosidic bond breakage occurs via the double bond-

14

displacement involving the formation and subsequent hydrolysis of a glycosyl-

enzyme intermediate. The two carboxyl groups involved in the mechanism are

5.5 Å apart. 28-30 In the first step, one of the two carboxylic acid residues acts as

a general acid catalyst and protonates the aglycon to make it a better leaving

group, whereas the second carboxylate residue acts as a nucleophile, forming

the glycosyl-enzyme intermediate. In the second step, the carboxylate side chain

acts as a general base to assist deprotonation of an incoming water molecule,

which attacks at the anomeric center of the glycosyl-enzyme intermediate,

liberating the hydrolyzed product with net retention of configuration (α). Both

steps, the formation of the glycosyl-enzyme intermediate and its conversion into

product, occurs via oxacarbenium-ion like transition states as shown below

(Scheme 1-1).

15

Scheme 1-1: Proposed mechanism for retaining glycosidases.

1.4.2 Mechanism of inverting glycosidases

This mechanism is illustrated using a β-glycosidase as an example

(Scheme 1-2). In this mechanism, the hydrolysis reaction takes place via a single

displacement. The two carboxyl groups involved in the mechanism are suitably

placed (10.5 Å apart on average),28-30 allowing for the substrate and water

molecules to bind between them (Scheme 1-2). One carboxyl group serves as a

OHOHO

OR

OH

OH

O O

δ+

δ−

H

OO OHOHO

OR

OH

O O

δ+

H

OO

HO

δ+

δ−

δ−

++

OHOHO

OH

OH

O O

ROH

O

O

glycosyl-enzyme intermediateOHO

HO

OH

OH

O O

δ−

H

OO

HO

δ+

δ−

δ−

++

OHOHO

OH

OH

OH

HO O

OO

oxacarbenium-ion like transition state

oxacarbenium-ion like transition state

nucleophile

general acid

general base

HOH

16

general acid and protonates the aglycon, the other residue acts as a general

base directing a water molecule to attack at the anomeric carbon, forming the

product with inversion of configuration (α) at the anomeric center. This catalytic

mechanism also proceeds through an oxacarbenium-ion like transition state

(Scheme 1-2).

Scheme 1-2: Proposed mechanism for inverting glycosidases.

1.5 Transition-state mimics

Both inverting and retaining glycosidase enzymes mechanisms involve a

transition state (TS) with substantial oxacarbenium ion-like character (Figure 1-

6).31,32 The positive charge on the endocyclic oxygen atom and bond between

this oxygen and the anomeric carbon atom has partial double bond character

(Figure 1-6).

OHOHO

OH

OH

O O

δ−

H

OOOHO

HO

O

OH

O O

δ−

H

OO

HO

δ+

δ−

δ−

++

oxacarbenium-ion like transition state

ORH

O H

δ+

H

Oδ+

R

HOHO

HO

OH

OH

HO O

OO

HO

ROH

general acid

general base

17

Figure 1-6: Transition state proposed for glycosidase catalyzed reactions.

Formation of the oxacarbenium ion has been supported by kinetic isotope

effect (KIE) measurements which indicated various degrees of sp2 character at

the anomeric carbon.33

In enzyme-catalyzed reactions, the transition state is stabilized by

electrostatic and hydrophobic interactions with the enzyme active site. These

interactions are optimized at the transition state so that the activation energy for

the enzyme-catalyzed reactions is less than that of the non-catalyzed reaction.34

Therefore, a highly effective inhibitor for these enzymes can be a stable molecule

which can mimic both charge and shape of the oxacarbenuim-ion transition state.

Design efforts focusing on mimicking the assumed geometry of the

transition state of the oxacarbenium-ion, in the case of a pyranosyl cation (six-

membered ring glycosyl cation) have focused on either half chair (4H3 or 3H4) or

boat conformations (2,5B or B2,5) which enable the C-1, C-2, O-5 and C-5 atoms

to be in a coplanar arrangement, as shown below (Figure 1-7).35

OHOHO

O

OH

δ−

H

HO

δ+

++

H

O

δ+

R

R

δ+

δ−

18

Figure 1-7: Proposed half chair and boat conformations of pyranosyl cations in glycosidase-mediated hydrolysis reactions.

In the case of the furanosyl cation (five-membered ring glycosyl cation),

the ring probably possesses an envelope conformation in which the C-1, C-2, O-

4 and C-4 atoms adopt a coplanar arrangement and hence, the furanose ring is

presumed to adopt an envelope conformation (3E or E3), as shown below for the

ribofuranosyl cation (Figure 1-8).35

Figure 1-8: Proposed envelope conformations of furanosyl cations in glycosidase mediated-hydrolysis reactions.

Designs focusing on charge have mimicked charge build up at three

different positions, the exocyclic oxygen, endocyclic oxygen, and anomeric

OHO

OH

OH

OH1

2

3

4

5

OHO25

OH

OH

HO

3

4

1

3H4

O OHOH

OH

HO

OOH

OHHO

HO

2,5B B2,5

12

34

5 1

2345

4H3

O

OH

HO

HO

O

OH

OHOH3E E3

12

3

4

123

4

19

center. Some inhibitors mimic charge by protonation at the exocyclic oxygen

which is typical of an early transition state and resembles closely the conjugate

acid of the glycoside 4 (Figure 1-9); others have charge build up at the endocyclic

oxygen atom along with its resonance form which is typical of a late transition

state and resemble more closely the resonance structures of an oxacarbenium-

ion intermediate 5 and 6, as shown below (Figure 1-9).36

Figure 1-9: Positive charge build-up at the exocyclic oxygen, endocyclic oxygen, and the anomeric carbon during glycosidase-mediated hydrolysis reactions.

1.6 Glycosidase inhibitors

Molecules that can inhibit the activity of an enzyme are very important in

controlling many biological activities. Glycosidase inhibitors have many potential

therapeutic applications in the treatment of diseases such as diabetes, cancer

and viral infections.37

The number of glycosidase inhibitors is continually growing and it is

outside the scope of this thesis to review them all. Instead, the general features

representative of naturally-occurring glucosidase inhibitors will be described.

OOH

HOHO

HO

4

OH

OOH

HOHO

HO

OOH

HOHO

HO

5 6R

20

1.6.1 Iminosugars

Since the discovery of nojirimycin (12) in 1966 (Figure 1-11),38 the most

popular way of designing glycosidase inhibitors has been a nitrogen atom in the

saccharide ring. Iminosugar-based glycosidase inhibitors have been the subject

of intense study,39 because of their profound effect on glycosidases.

Iminosugars (aza sugars) are moderately basic and are believed to be

protonated in the endocyclic nitrogen atom under physiological conditions;

therefore, they might mimic the positive charge on the ring oxygen at the

transition state, but not the flattened conformation.40,41,36 Alternatively they may

just bind electrostatically to carboxylate residues in the active site as ground-

state analogues.

Iminosugars can generally be categorized into five major structural

classes, namely piperidines (7), pyrrolidines (8), pyrrolizidines (9), indolizidines

(10), and nortropanes (11) (Figure 1-10).

Figure 1-10: Skeletal frameworks most commonly found in naturally-occurring iminosugar glycosidase inhibitors and their synthetic derivatives.

N

N

7

10

N N

N

8 9

11

21

1.6.1.1 Piperidines

Polyhydroxylated piperidines constitute the largest subgroup of α-

glycosidase inhibitors. Nojirimycin (12) was the first natural saccharide isolated

from several strains of Bacillus, Streptomyces, and mulberry tree leaves, and

was identified as an antibiotic. Although capable of inhibiting α- and β-

glucosidases, the compound is relatively unstable.42 Therefore, reduction of

nojirimycin results in a more stable compound, 1- deoxynojirimycin (13), which

has been shown to inhibit glucosidases in a reversible and competitive manner

(Figure 1-11).39

Isofagomine (14) is another example of the same family with about 440

times more inhibitory activity against β-glucosidases compared to 1-

deoxynojirimycin, but is only a moderate inhibitor of α-glucosidases (Figure 1-

11).43

Figure 1-11: Nojirimycin (12), 1-deoxynojirimycin (13), and isofagomine (14).

1.6.1.2 Pyrrolidines

Polyhydroxylated pyrrolidine alkaloids, including 2,5-dideoxy-2,5-imino-D-

mannitol (DMDP, 15), 2,5-dideoxy-2,5-imino-D-glucitol (DGDP, 16), 1,4-dideoxy-

N

OHOHHO

OH

HN

OHOHHO

OH

HHO

Nojirimycin (12) 1-Deoxynojirimycin (13)

NH

OH

HOH2C

HO

Isofagomine (14)

22

1,4-imino-D-arabinitol (DAB-1, 17) are naturally-occurring inhibitors of plant and

mammalian glucosidases as well as being antiviral agents (Figure 1-12).44

Figure 1-12: Naturally-occurring polyhydroxylated pyrrolidine-based glycosidase inhibitors.

1.6.1.3 Pyrrolizidines

Polyhydroxylated natural products namely, (+)-australine (18), (+)-

casuarine (19), hyacinthacine A2 (20), hyacinthacine A1 (21) and (+)-alexine (22)

(Figure 1-13) are members of this family.

(+)-Australine (18), isolated from Castenospermine australe,45 is not only a

competitive inhibitor of α-glucosidase and glycoprotein processing,46 but also

displays anti-HIV activity.47

N

OHHO

H

OHHON

OHHO

H

OHHON

OHHO

H

HO

DMDP (15) DGDP (16) DAB (17)

23

Figure 1-13: Structures of selected, naturally-occurring pyrrolizidine alkaloids, which display α-glucosidase inhibitory activities.

1.6.1.4 Indolizidines

The indolizidine alkaloid (+)-castanospermine (23, Figure 1-14) isolated from the

seeds of the Moreton Bay chestnut tree (Castanospermum australe),48 is one

example of this family that has shown potent inhibition of endoplasmic reticulum

α-glucosidase, and intestinal maltase and sucrase.49 Importantly,

castanospermine (23) exhibits activity against a range of human viral pathogens,

including parainfluenza,50 dengue virus,51 HSV-252 and HIV-1.53

Swainsonine (24, Figure 1-14), another example of this group of compounds,

was isolated from Swansona canescens, Astragalus lentiginosus, and Ipomoea

carnea, and is the first alkaloid with strong inhibition (nM) toward α-mannosidase

II.54

N

H OH

OH

HO

OH(+)-Australine (18)

N

H OH

OH

HO

OH(+)-Casuarine (19)

HON

H OH

OH

OH(+)-Hyacinthacine A2 (20)

N

H OH

OH

HO

OH(+)-Alexine (22)

N

H OH

OH

OH(+)-Hyacinthacine A1 (21)

24

Figure 1-14: Castanospermine (23), swainsonine (24).

1.6.1.5 Nortropanes

This class of compounds is called calystegins and is isolated from plants

such as Calystegia sepium, Ipomoea carnea, and Physalis alkekengi var.

francheti. Calystegin A3 (25) and calystegin B1 (26) are examples of this class

that have shown α,β-glucosidase inhibitory activity (Figure 1-15).55

Figure 1-15: Examples of nortropanes (25, 26).

1.6.2 Carbasugars

Carbasugars (or pseudosugars) are carbocyclic analogues of monosaccharides

in which the ring-oxygen atom has been replaced by a methylene group.56 5a-

Carba-α-D-galactopyranose (27) (Figure 1-16)57 was discovered as the first

carbasugar natural product. Carbasugars and aminocyclitols occur more

N

OHHO

OH

HO

H

Castanospermine (23)

N

OHOHH

Swainsonine (24)

OH

NH

OH

HO

OH

OH

Calystegin B1 (26)

NH

OHOH

OH

Calystegin A3 (25)

HO

25

commonly as components of more complex natural products. Examples include

the aminocyclitols valienamine (28), valiolamine (29), validamine (30) and their

derivatives voglibose (31) and acarbose (32), some with potent α-glucosidase

inhibitory activity. Acarbose (32) and voglibose (31), are approved for the clinical

treatment of type-2 diabetes. Voglibose (31) was shown to possess 20 to 30

times more potent α-glucosidase inhibitory activity than acarbose (32),58 with

fewer side effects (Figure 1-16).59

Figure 1-16: Examples of naturally-occurring carbasugar α-glucosidase inhibitors and the antibiotic validamycin.

1.6.3 Marine organosulfates

During the last decade there has been increasing interest in the discovery

of inhibitors from marine organisms.60 In this regard, a number of novel α-

glucosidase inhibitors have been isolated which resemble iminosugars.61

OH

OHOH

HO

HO

27

NH2

OHOH

HO

HO

Valienamine (28)

NH2

OHOH

HO

HO

Valiolamine (29)

OH

NH

OHOH

HO

HO

Voglibose (31)

OH

OH

OH

NH2

OHOH

HO

HO

Validamine (30)

OHN

HO

OOH

H3C

O

OOH

OH

O

HO OH

HO

HOHO

OH

HO

HO

OHAcarbose (32)

26

Penarolides (33 and 34)62 (Figure 1-17) and the schulzeine family (35–37)

(Figure 1-18)63 are members of this group that structurally comprise an alkaloid

or amino acid residue coupled with an O-sulfated long-chain fatty acid through an

amide group.

1.6.3.1 Penoralide sulfates

Penarolide sulfate A1 (33) and A2 (34) are structurally unique 30- and 31-

membered macrolides which include a trisulfated, lipophilic chain encircled by a

proline residue (Figure 1-17).62 Isolated from a marine sponge Penares sp.,

these compounds inhibit α-glucosidase with IC50 values of 1.2 and 1.5 mg/mL,

making them approximately 30–40 times more active than 1-deoxynojirimycin

(13).

Figure 1-17: Structures of organosulfate α-glucosidase inhibitors isolated from marine invertebrates.

OON

O

NaO3SO

OSO3Na

OSO3Na

Penarolide A1 (33)

OON

O

Penarolide A2 (34)

OSO3Na

OSO3NaNaO3SO

27

1.6.3.2 Schulzeines

The schulzeines (35–37), represent a new class of marine alkaloids,

isolated by Fusetani and co-workers from extracts of the marine sponge Penares

schulzei, they show potent α-glucosidase-inhibitory activity (IC50 = 48–170 nM)

(Figure 1-18).61

Figure 1-18: Structures of organosulfate α-glucosidase inhibitors isolated from marine invertebrates.

1.6.4 Sulfonium-sulfate thiosugars

Castanospermine (23) is thought to derive its glycosidase inhibition properties

from an ability to mimic the charge and the shape of the oxacarbenium ion-like

transition state. A requirment for this transition state mimicry is protonation of the

ring nitrogen atom at physiological pH within the active site of the enzyme.

()5

()9NH

NaO3SO

OSO3Na

ON O

HO

OH

OSO3Na

R1 R2

35 (+) schulzeines A R1= H, R2 = Me36 (+) schulzeines C R1=H, R2 = H

()5

()9NH

NaO3SO

OSO3Na

ON O

HO

OH

OSO3Na

37 (-) schulzeines C

28

Additional efforts to design glycosidase inhibitors involved the synthesis of

candidates which carry a permanent positive charge in order to provide the

required electrostatic interactions between inhibitor and the active site. In this

regard, the N-oxide analogue of castanospermine (38, Figure 1-19) has been

synthesized and was shown to be a weak inhibitor of β-glucosidase in

comparison with the parent compound, castanospermine (23).63

For the purpose of establishing such a permanent positive charge,

introduction of a sulfur atom in the castanospermine scaffold was considered

next. Our group reported the synthesis of a bridgehead sulfonium salt analogue

of the indolizidine alkaloid castanospermine (39) as well as its conformational

study, as a model to test the theory that formation of a sulfonium salt carrying a

permanent positive charge might be advantageous in providing the necessary

electrostatic interactions between the inhibitor and the active site carboxylate

residues (Figure 1-19).64,65

Figure 1-19: Structure of an N-oxide analogue of castanospermine and a sulfonium-ion analogue.

The syntheses of bicyclic sulfonium ion 40 and 41, analogues of a

swainsonine (24) and sulfonium compounds 42 and 43 with structures related to

australine (18) has also been reported by our group.66

S ClO4

OHHO

HON

OHHO

HOO

HO

38 39

29

Figure 1-20: Structure of compounds 40-43.

This approach was strongly validated by Yoshikawa’s discovery, beginning

in 1997, of a series of naturally-occurring glucosidase inhibitors that encompass

a zwitterionic sulfonium-sulfate structure 44-49 (Figure 1-21).67

Figure 1-21: Naturally-occurring zwitterionic sulfonium-sulfate glucosidase inhibitors.

S

OHHO

OSO3 OH

OH OH

HO

OHOH

S

OHHO

OSO3 OH

OH OH

HO

OH

S

OHHO

OSO3

OH OH

HO

S

OHHO

OSO3

OH

HO

Kotalanol (45)Salacinol (44)

Ponkoranol (49)Salaprinol (48)

S

OHHO

OH OH

OH OH

HO

OHOH

CH3OSO3

De-O-sulfonated Kotalanol (46)

S

OHHO

OH

OH OH

HOHCO2

De-O-sulfonated Salacinol (47)

S S

S S

OH OH

OH

OH OH

OHOHOH OH

OHH OHHR R

Cl Cl

OTf OTf

40 41

42 43R = H or OH

30

The compounds were Isolated from the aqueous extracts of the roots and

stems of Salacia reticulata and related plant species and were structurally related

in having a 1,4-anhydro-4-thio-D-arabinitol unit and a polyhydroxylated acyclic

chain of varying length.68-70

1.6.4.1 Salacinol

The first of the sulfonium-sulfate inhibitors to be discovered was salacnol,

isolated by Yoshikawa and co-workers in 1997 from the extracts of the dried

roots and stems of Salacia reticulata, known as kotalahimbutu’ in Singhalese, a

large climbing plant found throughout the forests of Southern India and Sri Lanka

(Figure 1-21).68 Aqueous extracts of this plant, have been traditionally used in

Indian medicine for treating type-2 diabetes. Prepared extracts by soaking the

bark and roots in water overnight, have been employed for the treatment of type-

2 diabetes.71

Salacinol (44) shows strong blood glucose level control in rats72 and

inhibits intestinal α-glucosidases such as maltase, sucrase, and isomaltase

(Table 1-1).

1.6.4.2 Kotalanol

In addition to the isolation of salacinol (44), the aqueous extracts of

Salacia reticulata have yielded a number of other bioactive components,

including kotalanol (45),69 which displays more potent α-glucosidase inhibitory

activity than 44 (Table 1-1). From a structural perspective, kotalanol (45) differs

from salacinol (44) in the length of the polyhydroxylated side chain.

31

Although degradation studies69 of kotalanol (45) led to establishment of the

absolute stereochemistry of the 1-deoxy-4-thiopentofuranosyl moiety, the

absolute stereochemistry of the heptitol side chain remained unknown until our

group reported the structural elucidation and total synthesis of this medicinally

important natural product.73

1.6.4.3 De-O-sulfonated kotalanol

In 2008, Ozaki and co-workers reported the isolation of another α-

glucosidase inhibitor from S. reticulata using the bioassay-guided isolation

technique similar to the one used by Yoshikawa et al. and assigned its structure

as the unusual 13-membered cyclic sulfoxide 50 (Figure 1-22).74 Subsequent

reevaluation of Ozaki’s data by Muraoka et al. has led to the conclusion that this

compound is in fact de-O-sulfonated kotalanol (46),70 which Yoshikawa et al. had

previously prepared via the desulfonation of kotalanol.74 This compound was the

most potent inhibitor of rat intestinal α-glucosidase isolated from S. reticulata.

Figure 1-22: Initially proposed structure of a naturally-occurring sulfoxide α-glucosidase inhibitor.

S

HO

HO

HO

OH

OH

OH

OHOH

O

50

32

1.6.4.4 Neosalacinol

Neosalacinol (51) isolated from the medicinal plant Salacia oblanga by

Asano et al.76 On the basis of NMR spectroscopic analysis, mass spectrometry,

and a degradation study, the identity of this compound was established as a de-

O-sulfonated derivative of salacinol, and its structure was formulated as the

sulfonium–alkoxide zwitterion 51 (Figure 1-23).76 Subsequent synthetic studies by

Muraoka et al. have revealed that this is indeed de-O-sulfonated salacinol (47),

and it is not an inner salt.77 The inhibitory activity was found to be the same as

salacinol (44).

Figure 1-23: Initially proposed structure of the naturally-occurring α-glucosidase inhibitor neosalacinol.

1.6.4.5 Salaprinol

Salaprinol (48) is the least structurally complex of the zwitterionic

thiosugar α-glucosidase inhibitors discovered to date,76 isolated from the

methanolic extract of the roots and stems of the Sri Lankan plant Salacia

prinoides.

S

OHHO

O

OH OH

HO

51

33

1.6.4.6 Ponkoranol

Isolated from Salacia reticulata by Kitamura et al.,78 as a structurally

similar zwitterionic sulfonium sulfate. Ponkoranol is a six-carbon polyhydroxylated

side chain homologue of salacinol (44). The compound was initially named

reticulanol, and was then found to occur in Salacia prinoides by Yoshikawa et al.

In common with other members of the thiosugar sulfonium sulfate family,

ponkoranol (49) inhibits maltase, sucrase and isomaltase with IC50 values in the

low micromolar range (Table 1-1).70 Interestingly this compound was synthesized

in our laboratory prior to its isolation.79

Table 1-1: Comparison of inhibitory activities (IC50 in µM) of naturally-occurring sulfonium-ion glucosidase inhibitors 44-46, 48 and 49 against rat intestinal α-glucosidases.a

aValues in parentheses indicate Ki values (µM).

Inhibitor Maltase Sucrase Isomaltase Ref.

Salacinol (44) 5.2 (0.97) 1.6 (0.2) 1.3 (1.1) 75, 80

Kotalanol (45) 7.2 (0.52) 0.75 (0.42) 5.7 (4.2) 75, 80

De-O-sulfonated kotalanol (46) 0.227 (0.11) 0.186 (0.52) 0.099 (0.42) 77

Salaprinol (48) >100 >100 - 75

Ponkoranol (49) 3.2 0.29 2.6 75

34

In this thesis work an alternative route for the synthesis of kotalanol as

well as the design and synthesis of the 6'-epimer of kotalanol, de-O-sulfonated

ponkoranol, its 5'-epimer, and their selenium analogues will be described. The

thesis will also describe the design and synthesis of 3’-O-methylponkoranol and

C-3′- and C-5′-β-O-maltose-extended de-O-sulfonated ponkoranol. These

candidates will be used to probe the active-site requirements of the intestinal α-

glucosidase enzyme specificities of starch-digesting maltase-glucoamylase and

sucrase-isomaltase. Such understanding is intended in the long term to either

use these inhibitors for controlled release of glucose from starch digestion or use

the enzymes themselves as therapeutic agents in individuals lacking these

enzyme activities.

1.7 Thesis overview

This thesis work is presented primarily in journal article style, with Chapter

1 serving as a general introduction, followed by Chapters 2-7 as journal articles,

and Chapter 8 describing conclusions and future work.

Chapter 1, presents an introduction to carbohydrates, related diseases,

glycosidase enzymes, starch-digesting enzymes, the crystal structure of brush

border enzymes, their deficiency in health and disease, and their mechanism of

action. This is followed by a discussion on the transition states of glycosidase-

mediated hydrolysis reactions. Some examples of glucosidase inhibitors which

belong to one of the categories of iminosugars, carbasugars, marine

organosulfates, thiosugars, and sulfonium salts are then presented.

35

Chapter 2 presents the manuscript (Eskandari, R.; Jayakanthan, K.;

Kuntz, D. A.; Rose, D. R.; Pinto B. M. Bioorg. Med. Chem. 2010, 18, 2829-2835)

that describes the design and synthesis of the naturally-occurring glucosidase

inhibitor, kotalanol and its stereoisomer and their inhibitory activities against

recombinant human N-terminal maltase-glucoamylase (ntMGAM).

Chapter 3 presents the manuscript (Eskandari, R.; Kuntz, D. A.; Rose, D.

R.; Pinto, B. M. Org. Lett. 2010, 12, 1632-1635) that describes the synthesis of

de-O-sulfonated ponkoranol and its 5'-epimer and their evaluation as glucosidase

inhibitors against ntMGAM.

Chapter 4 presents the manuscript (Eskandari, R.; Jones, K.; Rose, D. R.;

Pinto, B. M. J. Chem. Soc., Chem. Commun. 2011, 47, 9134-9136) that

describes the synthesis of the selenium analogues of de-O-sulfonated

ponkoranol and their evaluation as glucosidase inhibitors against four human

intestinal glucosidase enzymes maltase-glucoamylase MGAM (ntMGAM,

ctMGAM) and sucrase-isomaltase (ntSI, ctSI).

Chapter 5 presents the manuscript (Eskandari, R.; Jones, K.; Rose, D.

R.; Pinto, B. M. Bioorg. Med. Chem. Lett. 2010, 20, 5686-5689) that describes

the synthesis of 3'-O-methylponkoranol and its evaluation as an inhibitor of

ntMGAM.

Chapter 6 presents the manuscript (Eskandari, R.; Jones, K.; Rose, D. R.;

Pinto, B. M. Bioorg. Med. Chem. Lett. 2011, 21, 6491-6494) that describes

biological evaluation of 3'-O-methylponkoranol against human intestinal

36

glucosidase enzymes, namely ntMGAM, ctMGAM and sucrase-isomaltase (ntSI,

ctSI).

Chapter 7 presents the manuscript (Eskandari, R.; Jones, K.; Reddy K.

R.; Jayakanthan, K.; Chaudet, M.; Rose, D. R.; Pinto, B. M. Chem. Eur. J. 2011,

17, 14817-14825) that describes the synthesis of two C-3′- and C-5′-β-O-

maltose-extended analogues of the naturally-occurring sulfonium ion inhibitor,

de-O-sulfonated ponkoranol. Evaluation of inhibitory activities against ntMGAM,

ctMGAM, ntSI, ctSI is also reported.

In Chapter 8, the general conclusions and future work are presented.

37

1.8 References

1. Robayo-Torres, C. C.; Quezada-Calvillo, R.; Nichols, B. L. Clin.

Gastroenterol. Hepatol. 2006, 4, 276–287.

2. Asp, N. G. Am. J. Clin. Nutr. 1995, 61, 930S-937S.

3. Englyst, H. N.; Kingman, S. M.; Cummings, J. H. Eur. J. Clin. Nutr. 1992, 46, S33-50.

4. Annison, G.; Topping, D. L. Annu. Rev. Nutr. 1994, 14, 297-320.

5. Quezada-Calvillo, R.; Robayo-Torres, C. C.; Ao, Z.; Hamaker, B. R.;

Quaroni, A.; Brayer, G. D.; Sterchi, E. E.; Baker, S. S.; Nichols, B. L. J.

Pediatr. Gastroenterol. Nutr. 2007, 45, 32–43. b) Quezada-Calvillo, R.;

Robayo-Torres, C. C.; Opekun, A. R.; Sen, P.; Ao, Z.; Hamaker, B. R.;

Quaroni, A.; Brayer, G. D.; Wattler, S.; Nehls, M. C.; Sterchi, E. E.;

Nichols, B. L. J. Nutr. 2007, 137, 1725–1733.

6. Chugani, H. T. Prev. Med., 1998, 27, 184-188.

7. Pond, W. G.; Nichols, B. L.; Brown, D. L. Adeqaute Food for All, Culture,

Science, and Technology of Food in the 21st Century, CRC Press, NY,

2009.

8. Cordain, L.; Eaton, S. B.; Sebastian, A.; Mann, N.; Lindeberg, S.; Watkins,

B. A..; O'Keefe, J. H.; Brand-Miller, J. Am. J. Clin. Nutr. 2005, 81(2), 341-

354.

9. Livesey, G. Proc. Nutr. Soc. 2005, 64, 105–113.

10. Yip, V. L.; Withers, S. G. Org. Biomol. Chem. 2004, 2, 2707-2713.

11. Henrissat, B. Biochemistry J. 1991, 280, 309-316.

12. Marks, V.; Flatt, P.; Dobbing, J.; Ed. Springer-Verlag: Berlin Heidelberg,

1989; pp 135-146.

38

13. Van Beers, E. H.; Buller, H. A.; Grand, R. J.; Einerhand, A. W.; Dekker, J.

Crit. Rev. Biochem. Mol. Biol. 1995, 30, 197-262.

14. Brayer, G. D.; Sidhu, G.; Maurus, R.; Rydberg, E. H.; Braun, C.; Wang,

Y.; Nguyen, N. T.; Overall, C. M.; Withers S. G. Biochemistry 2000, 39,

4778–4791.

15. Maurus, R.; Begum, A.; Kuo, H. H.; Racaza, A.; Numao, S.; Andersen, C.;

Tams, J. W.; Vind, J.; Overall, C. M.; Withers, S. G.; Brayer G. D. Protein

Sci. 2005, 14, 743–755.

16. Gray, G. M. J. Nutr. 1992, 122, 172-177. b) Gray, G. M. N. Engl. J. Med.

1975, 292, 1225-1230. c) Fujita, S.; Fuwa, H. J. Nutr. Sci. Vitaminol.

(Tokyo) 1984, 30,135-142. d) Jones, B. J.; Brown, B. E.; Loran, J. S.;

Edgerton, D.; Kennedy J. F.; Stead, J. A.; Silk D. B. A. Gut 1983, 24,

1152-1160.e) Corring, T. Reprod. Nutr. Dev. 1980, 20, 1217-1235.

17. Ao, Z.; Quezada-Calvillo, R.; Sim, L.; Nichols, B. L.; Rose, D. R.; Sterchi,

E. E.; Hamaker, B. R. FEBS Lett. 2007a, 581, 2381-2388.

18. Sim, L.; Jayakanthan, J.; Mohan, S.; Nasi, R.; Johnston, B.D.; Pinto, B. M.;

Rose, D. R. Biochemistry 2010, 49, 443-451.

19. Sim, L.; Willemsma, C.; Mohan, S.; Naim, H. Y.; Pinto, B. M.; Rose, D. R.

J. Biol. Chem. 2010, 285, 17763-17770.

20. Sim, L.; Quezada-Calvillo, R.; Sterchi, E. E.; Nichols B. L.; Rose, D. R. J.

Mol. Biol. 2008, 375, 782-792.

21. Auricchio, S.; Rubino, A.; Murset, G. Pediatrics 1965, 35, 944-954.

22. Enattah, N. S.; Sahi, T.; Savilahti, E.; Terwilliger, J. D.; Peltonen, L.;

Jarvela, I. Nature Genet. 2002, 30, 233–237.

23. Savilahti, E.; Launiala, K.; Kuitunen, P. Arch. Dis. Childhood 1983, 58,

246–252.

39

24. Lebenthal, E.; Khin-Muang, U.; Zheng B. J. Pediatr. 1994, 124, 541–546.

25. Robayo-Torres, C. C.; Opekun, A. R.; Quezada-Calvillo, R.; Villa, X.;

Smith, E. O.; Navarrete, M.; Baker, S. S.; Nichols, B. L. J. Pediatr.

Gastroenterol. Nutr. 2009, 48, 412–418.

26. Sinnott, M. L. Chem. Rev. 1990, 90, 1171-1202.

27. Koshland, D. E. Biol. Rev. 1953, 28, 416-436.

28. McCarter, J. D.; Withers, S. G. Curr. Opin. Struct. Biol. 1994, 4, 885-892.

29. Davies, G.; Henrissat, B. Structure 1995, 3, 853-859.

30. Wang, Q. P.; Graham, R. W.; Trimbur, D.; Warren, R. A. J.; Withers, S. G.

J. Am. Chem. Soc. 1994, 116, 11594-11595.

31. Namchuk, M. N.; Withers, S. G. Biochemistry 1995, 34, 16194-16202.

32. McCarter, J. D.; Adam, M. J.; Withers, S. G. Biochem. J. 1992, 286, 721-

727.

33. Bennet, A. J.; Sinnott, M. L. J. Am. Chem. Soc. 1986, 108, 7287-7294.

34. Meister, A. Adv. Enzymol. 1975, 43, 219-410.

35. Sinnott, M. L. Chem. Rev. 1990, 90, 1171-1202.

36. Lillelund, V. H.; Jensen, H. H.; Liang, X.; Bols, M. Chem. Rev. 2002, 102,

515-553.

37. Holman, R. R.; Cull, C. A.; Turner, R. C. Diabetes Care 1999, 22, 960-964.

(b) Jacob, G. S. Curr. Opin. Struct. Biol. 1995, 5, 605-611. (d) Dwek, R. A.

Chem. Rev. 1996, 96, 683-720.

38. Ishida, N.; Kumagai, K.; Niida, T.; Tsuroka, T.; Yumoto, H. J. Antibiot.

1967, 20, 66-71.

39. Stutz, A. E. Wiley-VCH, New York, 1999, pp. 292-295 (b) Asano, N. Curr.

Top. Med. Chem., 2003, 3, 471-484.

40

40. Sears, P.; Wong, C. H. Angew. Chem. Int. Ed. 1999, 38, 2300-2324.

41. Heightman, T. D.; Vasella, A. T. Angew. Chem. Int. Ed. 1999, 38, 750-

770.

42. Inouye, S. E.; Tsuruoka, Y.; Ito, T.; Niida, T. Tetrahedron 1968, 24, 2125-

2144.

43. Pandey, G.; Kapur, M.; Khan, M. I.; Gaikwad, S. M. Org. Biomol. Chem.

2003, 1, 3321-3326.

44. Yu, C.; Asano, N.; Ikeda, K.; Wang, M.; Butters, T. D.; Wormald, M. R.;

Raymond, A. L. W.; Dwek, A.; Nash R. J.; Fleet, G. W. T. Chem.

Commun. 2004, 1936-1937.

45. Molyneux, R. J.; Benson, M.; Wong, R. Y.; Tropea, J. E.; Elbein, A. D. J.

Nat. Prod. 1988, 51, 1198-1206.

46. Tropea, J. E.; Molyneux, R. J.; Kaushal, G. P.; Pan, Y. T.; Mitchell, M.;

Elbein, A. D. Biochemistry 1989, 28, 2027-2034.

47. Vlietinck, A. J.; De Bruyne, T.; Apers, S.; Pieters, L. A. Planta Med.

1998, 64, 97-109.

48. Hohenschutz, L. D.; Bell, E. A.; Jewess, P. J.; Leworthy, D. P.; Pryce, R.

J.; Arnold, E.; Clardy, J. Phytochemistry 1981, 20, 811-814.(b) Nash, R. J.;

Fellows, L. E.; Dring, J. V.; Stirton, C. H.; Carter, D.; Hegarty, M. P.; Bell,

E. A. Phytochemistry, 1988, 27, 1403-1406.

49. Michael, J. P. Nat. Prod. Rep. 2007, 24, 191-222.

50. Tanaka, Y.; Kato, J.; Kohara, M.; Galinski, M. S. Antiviral Res., 2006, 72,

1-9.

51. Whitby, K.; Pierson, T. C.; Geiss, B.; Lane, K.; Engle, M.; Zhou, Y.; Doms,

R. W.; Diamond, M. S. J. Virol., 2005, 79, 8698-8706.

41

52. Ahmed, S. P.; Nash, R. J.; Bridges, C. G.; Taylor, D. L.; Kang, M. S.;

Porter, E. A.; Tyms, A. S. Biochem. Biophys. Res. Commun. 1995, 208,

267-273.

53. Bridges, C. G.; Brennan, T. M.; Taylor, D. L.; McPherson, M.; Tyms, A. S.

Antiviral Res. 1994, 25, 169-175.

54. Colegate, S. M.; Dorling, P. R.; Huxtable, C. R. Aust. J. Chem. 1979, 32,

2257-2264.

55. Molyneux, R. J.; Pan, Y. T.; Goldmann, A.; Tepfer, D. A.; Elbein, A. D.

Arch. Biochem. Biophys. 1993, 304, 81-88.

56. McCasland, G. E.; Futura, S.; Durham, L. J. J. Org. Chem., 1966, 31,

1516-1521. (b) McCasland, G. E.; Futura, S.; Durham, L. J. J. Org. Chem. 1968, 33, 2835-2841.

57. Miller, T. W.; Arison, B. H.; Albers-Schonberg, G. Biotechnol. Bioeng.

1973, 15, 1075-1080.

58. Yasuda, K.; Shimowada, K.; Uno, M.; Odaka, H.; Adachi, T.; Shihara, N.;

Suzuki, N.; Tamon, A.; Nagashima, K.; Hosokawa, M.; Tsuda, K.; Seino,

Y. Diabetes Res. Clin. Pract. 2003, 59, 113-122.

59. Vichayanrat, A.; Ploybutr, S.; Tunlakit, M.; Watanakejorn, P. Diabetes Res.

Clin. Pract. 2002, 55, 99-103.

60. For selected examples of marine α-glucosidase inhibitors, see: (a) Heo, S.

J.; Hwang, J. Y.; Choi, J. I.; Han, J. S.; Kim, H. J.; Jeon, Y. J. Eur. J.

Pharmacol. 2009, 615, 252-256. (b) Ingavat, N.; Dobereiner, J.;

Wiyakrutta, S.; Mahidol, C.; Ruchirawat, S.; Kittakoop, P. J. Nat. Prod.

2009, 72, 2049-2059. (c) Saludes, J. P.; Lievens, S. C.; Molinski, T. F. J.

Nat. Prod. 2007, 70, 436-638. (d) Nakao, Y.; Uehara, T.; Matsunaga, S.;

Fusetani N.; Van Soest, R. W. J. Nat. Prod. 2002, 65, 922-924.

42

61. Takada, K.; Uehara, T.; Nakao, Y.; Matsunaga, S.; Van Soest R. W. M.;

Fusetani, N. J. Am. Chem. Soc. 2004, 126, 187-193.

62. Nakao, Y.; Maki, T.; Matsunaga, S.; Van Soest R. W. M.; Fusetani, N.

Tetrahedron 2000, 56, 8977-8987.

63. Kajimoto, T.; Liu, K. K. C.; Pederson, R. L.; Zhong, Z. Y.; Ichikawa, Y.;

Porco, J. A.; Wong, C. H. J. Am. Chem. Soc. 1991, 113, 6187–6196.

64. Svansson, L.; Johnston, B. D.; Gu, J. H.; Patrick, B.; Pinto, B. M. J. Am.

Chem. Soc. 2000, 122, 10769-10775.

65. Johnson, M. A.; Jensen, M. T.; Svensson, B.; Pinto, B. M. J. Am. Chem.

Soc. 2003, 125, 5663-5670.

66. (a) Kumar, N. S.; Pinto. B. M. J. Org. Chem. 2006, 71, 1262-1264. (b)

Kumar, N. S.; Pinto. B. M. J. Org. Chem. 2006, 71, 2935-2943. (c) Kumar,

N. S.; Pinto. B. M. Carbohydr. Res. 2006, 341, 1685-1691.

67. Matsuda, H.; Morikawa, T.; Yoshikawa, M. Pure Appl. Chem. 2002, 74,

1301-1308.

68. Yoshikawa, M.; Murakami, T.; Shimada, H.; Matsuda, H.; Yamahara, J.;

Tanabe, G.; Muraoka, O. Tetrahedron Lett. 1997, 38, 8367-8370.

69. Matsuda, H.; Li, Y. H.; Murakami, T.; Matsumura, N.; Yamahara, J.;

Yoshikawa, M. Chem. Pharm. Bull. 1998, 46, 1399-1403.

70. Muraoka, O.; Xie, W. J.; Tanabe, G.; Amer, M. F. A.; Minematsu, T.;

Yoshikawa, M. Tetrahedron Lett. 2008, 49, 7315-7317.

71. Li, Y.; Huang T. H. W.; Yamahara, J. Life Sci. 2008, 82, 1045-1049.

72. Shimoda, H.; Fujimura, T.; Makino, K.; Yoshijima, K.; Naitoh, K.; Ihota, H.;

Miwa, Y. J. Food Hyg. Soc. Jpn. 1999, 40, 198-205.

73. Jayakanthan, K.; Mohan, S.; Pinto, B. M. J. Am. Chem. Soc. 2009, 131,

5621-5626.

43

74. Ozaki, S.; Oe, H.; Kitamura, S. J. Nat. Prod. 2008, 71, 981-984.

75. Yoshikawa, M.; Xu, F.; Nakamura, S.; Wang, T.; Matsuda, H.; Tanabe, G.;

Muraoka, O. Heterocycles 2008, 75, 1397-1405.

76. Minami, Y.; Kuriyama, C.; Ikeda, K.; Kato, A.; Takebayashi, K.; Adachi, I.;

Fleet, G. W. J.; Kettawan, A.; Okamoto, T.; Asano, N. Bioorg. Med.

Chem. 2008, 16, 2734-2740.

77. Tanabe, G.; Xie, W.; Ogawa, A.; Cao, C.; Minematsu, T.; Yoshikawa, M.;

Muraoka, O. Bioorg. Med. Chem. Lett. 2009, 19, 2195-2198.

78. Asada, M.; Kawahara, Y.; Kitamura, S. US Patent 0037870, 2007.

79. Johnston, B. D.; Jensen, H. H.; Pinto, B. M. J. Org. Chem. 2006, 71, 1111-

1118.

80. Matsuda, H.; Yoshikawa, M.; Morikawa, T.; Tanabe, G.; Muraoka, O. J.

Trad. Med. 2005, 22, 14-153.

44

CHAPTER 2: SYNTHESIS OF A BIOLOGICALLY ACTIVE ISOMER OF KOTALANOL, A NATURALLY-OCCURRING

GLUCOSIDASE INHIBITOR

This Chapter comprises the manuscript “Synthesis of a biologically

active isomer of kotalanol, a naturally-occurring glucosidase inhibitor”

which was published in Bioorganic and Medicinal Chemistry (2010, 18, 2829-

2835).

Razieh Eskandari,a Kumarasamy Jayakanthan,a Douglas A. Kuntz,b David R.

Rose,b B. Mario Pintoa

aDepartment of Chemistry, Simon Fraser University, Burnaby, British Columbia,

Canada V5A 1S6

b Department of Biology, University of Waterloo, Waterloo, Ontario,

Canada N2L 3G1

45

Kotalanol is one of the active principles in the aqueous extracts of Salacia

reticulata, a climbing tree shrub native to Sri Lanka and Southern India, used in

the treatment of diabetes. In this Chapter, kotalanol and its 6'-epimer were

synthesized and their glucosidase inhibitory activities against N-terminal maltase-

glucoamylase (ntMGAM) were studied. The thesis author performed all the

experimental synthetic work and the characterization of the compounds and

wrote and edited the manuscript with Dr. B. Mario Pinto. Dr. Jayakanthan

Kumarasamy assisted in the synthetic design. Drs. Douglas A. Kuntz and David

R. Rose performed the enzyme inhibition studies.

Graphical abstract:

O O

OSO3OMOM

OMOM

OMOMS

OPMBPMBO

PMBOS

OHHO

OSO3 OH

OH OH

HO

OHOH

O O

OS O

OOOMOM

OMOM

OMOMS

OPMBPMBO

PMBO +

46

2.1 Keywords

Kotalanol, isomer of kotalanol, D-mannitol, cyclic sulfate, glucosidase

inhibitors, human maltase-glucoamylase, type-2 diabetes.

2.2 Abstract

The syntheses of an isomer of kotalanol, a naturally-occurring glucosidase

inhibitor, and of kotalanol itself are described. The target compounds were

synthesized by nucleophilic attack of PMB-protected 1,4-anhydro-4-thio-D-

arabinitol at the least hindered carbon atom of two 1,3-cyclic sulfates, which were

synthesized from D-mannose. Methoxymethyl ether and isopropylidene were

chosen as protecting groups. The latter group was critical to ensure the facile

deprotection of the coupled products in a one-step sequence to yield kotalanol

and its isomer. The stereoisomer of kotalanol, with the opposite stereochemistry

at the C-6’ stereogenic centre, inhibited the N-terminal catalytic domain of

intestinal human maltase-glucoamylase (ntMGAM) with a Ki value of 0.20 ± 0.02

µM; this compares to a Ki value for kotalanol of 0.19 ± 0.03 µM. The results

indicate that the configuration at C-6’ is inconsequential for inhibitory activity

against this enzyme.

2.3 Introduction

Glycosidases are enzymes that are involved in the catabolism of

glycoproteins and glycoconjugates and the biosynthesis of oligosaccharides.

Disruption in regulation of glycosidases can lead to diseases.1,2 Over the years,

glycosidase inhibitors have received considerable attention in the field of

47

chemical and medicinal research3 because of their effects on quality control,

maturation, transport, secretion of glycoproteins, and cell-cell or cell-virus

recognition processes. This principle has potential for many therapeutic

applications, such as in the treatment of diabetes, cancer and viral infections.1

Bioactive components isolated from medicinal plants that are used in

traditional medicine or folk medicine often provide the lead structures for modern

drug-discovery programs. For example, the large woody climbing plant Salacia

reticulata, known as Kothalahimbutu in Singhalese, is used in traditional medicine

in Sri Lanka and Southern India for treatment of type-2 diabetes.4,5 A person

suffering from diabetes was advised to drink water stored overnight in a mug

carved from Kothalahimbutu wood.6 Several potent glucosidase inhibitors have

been isolated from the water soluble fraction of this plant extract and also other

plants that belong to the Salacia genus such as Salacia chinensis, Salacia

prinoides, and Salacia oblonga which explain, at least in part, the antidiabetic

property of the aqueous extract of this plant.7-9 All these compounds share a

common structural motif that comprises a 1,4-anhydro-4-thio-D-arabinitol and a

polyhydroxylated side chain. So far, five components have been isolated, namely

salaprinol (1),9 salacinol (2),7 ponkoranol (3),9 kotalanol (4),8 and de-O-sulfonated

kotalanol (5)10 (Figure 2-1). The absolute stereostructure for these compounds,

except salacinol, was not determined at the time of isolation, but synthetic work

has led to their stereochemical structure elucidation.11,12

48

S

OHHO

OSO3 OH

OH OH

HO

OHOH

S

OHHO

OSO3 OH

OH OH

HO

OH

S

OHHO

OSO3

OH OH

HOS

OHHO

OSO3

OH

HO

Kotalanol (4)

Salacinol (2) Ponkoranol (3)Salaprinol (1)

S

OHHO

OH OH

OH OH

HO

OHOH

CH3OSO3

De-O-sulfonated kotalanol (5)

Figure 2-1: Components isolated from Salacia species.

The synthesis of kotalanol 4 and its stereoisomer 6 (Figure 2-2) are of

interest here.

S

OHHO

OSO3 OH

OH OH

HO

OHOH

6

Figure 2-2: Kotalanol stereoisomer.

49

Our first attempt employed the reaction of the cyclic sulfates 8 in the

coupling reaction (Scheme 2-1).12 However, attempts to remove the methylene

acetal in the coupled products required forcing conditions and resulted in de-O-

sulfonation (Scheme 2-1).12 We have also reported a successful synthesis of

kotalanol using a cyclic sulfate derived from a naturally-occurring heptitol, D-

perseitol (Scheme 2-2).12

S

OPMBPMBO

PMBO

7

HFIP, K2CO3

S

OPMBPMBO

PMBO

OO

OSO3OBn

OBn

OBn

9

O O

OS O

OO

OBnOBn

OBn

8

+

then MeOH

S

OHHO

OH OH

OH OH

HO

OHOH

CH3OSO3

5

1.0 M BCl3

Scheme 2-1: First attempted synthesis of kotalanol 4.

50

O O O

PMBO OPMB OPMB

O

PhS

O O10

7HFIP, K2CO3

OSO3 O

PMBO OPMB OPMB

O

Ph

S

OPMBPMBO

PMBO

80% TFA 4

11

D-perseitol

Scheme 2-2: Synthesis of kotalanol 4.

However, it was of interest to develop a synthesis of the isomer of

kotalanol 6 in view of the fact that the C-6’ stereoisomer of de-O-sulfonated

kotalanol was just as active an inhibitor as de-O-sulfonated kotalanol 5 itself

against a key intestinal enzyme, human maltase-glucoamylase.13 We report here

a general synthetic route to this isomer 6 and also an alternative synthesis of

kotalanol 4.

2.4 Results and discussion

We chose to replace the problematic methylene acetal group of compounds 8

with an isopropylidene acetal (compound 12) to ensure not only its facile removal

after the coupling reaction but also to maintain some rigidity in the cyclic sulfate.

We chose also to replace the benzyl ethers with methoxymethyl (MOM) ethers,

because the latter can survive the hydrogenolysis conditions required for removal

51

of the benzylidene acetal. The cyclic sulfate 12 could be synthesized from D-

mannitol as shown in the retrosynthetic analysis (Scheme 2-3).

12

O O

OS O OMOM

OMOM

OMOM

OO

O O

OO OMOM

Ph

D-Mannitol

13

14

O O

OO O

Ph

O

Ph

Scheme 2-3: Retrosynthetic analysis.

Thus, the D-mannitol-derived diol 15,14 was protected as the acetonide to

give the C2-symmetric compound 14 in 73% yield. Mild hydrolysis of this

compound using catalytic PTSA in methanol effected the removal of one

benzylidene group to give the corresponding diol in 70% yield based on

recovered starting material. Selective protection of the primary hydroxyl group as

its TBDMS ether followed by sequential protection of the secondary hydroxyl

group as its MOM ether and removal of the TBDMS group with

tetrabutylammonium fluoride (TBAF) gave 17 in 73% yield over three steps.

Treatment of this alcohol with Dess-Martin periodinane provided the aldehyde

which was reacted with methyltriphenylphosphonium bromide to yield the olefinic

product 13 in 61% yield over two steps (Scheme 2-4).

52

O O

OH O

OH

O

Ph

Ph

O O

OO O

Ph

O1. DMP, pTSA 1. pTSA, MeOH

1. TBSCl, Imid.2. MOMCl, iPr2NEt

O O

OO OMOM

Ph

OH

3. TBAF, THF2. CH3PPh3Br, n-BuLi1. Dess-Martin periodinane

15 14

17

(70%)

(73%)(61%)

Ph

(73%)

O O

OO OMOM

Ph 13

O O

OO OMOM

Ph

OH

OH

18

12 3 4 5 6

7

O O

OO OMOM

Ph

OH

OH

19

123456

7+

dihydroxylation(80%)

OsO4

AD-mix-α

AD-mix-β

18 192.6 1.0

3.3

3.5

1.0

1.0

O O

OO OH

Ph

OH

16

1

234

5

67

Scheme 2-4: Synthesis of the diols 18 and 19.

With compound 13 in hand, our next goal was to introduce the two

hydroxyl groups. OsO4-catalyzed dihydroxylation of 13 afforded compound 18

(Scheme 2-4) as the major product with a diastereomeric ratio of 18:19 of 2.6:1.

Kishi's rule predicts that the relative stereochemistry between the pre-existing

hydroxyl group and the adjacent newly-introduced hydroxyl group in the major

product should be erythro.15 This result is also analogous to that obtained for

dihydroxylation of a corresponding methylene acetal.12

53

Interestingly, AD-mix-α and AD-mix-β also afforded compound 18 as a

major product, with a diastereomeric ratio of 3.3:1 and 3.5:1 (determined by 600

MHz 1H NMR), respectively. The unsatisfactory selectivity can be explained by

the steric hindrance imposed by the bicyclic structure, observed previously with a

similar structure.16 The two isomers were separated by column chromatography

and each was converted into its cyclic sulfate 12 or 22 as follows. The hydroxyls

in 18 were protected with MOM groups and the product was subjected to

hydrogenolysis to effect removal of the benzylidene group and to yield the

corresponding diol 20 in 72% yield over 2 steps. The cyclic sulfate 12 was then

obtained by treatment of 20 with thionyl chloride in the presence of triethylamine

to give the mixture of diastereomeric sulfites, followed by their oxidation with

sodium periodate and ruthenium (III) chloride as a catalyst. A similar sequence of

reactions with the diol 19 yielded the cyclic sulfate 22 (Scheme 2-5).

O O

OS O

OO

OMOMOMOM

OMOM

2. Pd(OH)2, MeOH1. SOCl2, Et3N, CH2Cl22. NaIO4, RuCl3 CCl4:CH3CN

O O

HOHO OMOM

OMOM

OMOM

21 22

(75%)(76%)

1. MOMCl, i-Pr2NEt19

O O

OS O

OO

OMOMOMOM

OMOM

2. Pd(OH)2, MeOH

1. SOCl2, Et3N, CH2Cl22. NaIO4, RuCl3 CCl4:CH3CN

O O

HOHO OMOM

OMOM

OMOM18

20 12

(72%)(70%)

1. MOMCl, i-Pr2NEt

Scheme 2-5: Synthesis of the cyclic sulfates 12 and 22.

54

The target compounds were prepared by opening of the cyclic sulfates 12

and 22 by nucleophilic attack of the sulfur atom in 2,3,5-tri-O-p-methoxybenzyl-

1,4-anhydro-4-thio-D-arabinitol 7.11 Reactions were carried out at 72 °C in

1,1,1,3,3,3-hexafluoro-2-propanol (HFIP) containing K2CO317 for 6 days to give

the sulfonium salts 23 and 24 in 65 and 57% yield, respectively. Finally,

deprotection of the coupled products 24 and 23 using aqueous 30%

trifluoroacetic acid (TFA) at 50 °C gave the desired compounds 4 and 6 in 91 and

93% yields, respectively (Scheme 2-6).

12, HFIP,K2CO3 65%

7

22, HFIP,K2CO3 57%

O O

OSO3 OMOM

OMOM

OMOMS

OPMBPMBO

PMBO

O O

OSO3OMOM

OMOM

OMOMS

OPMBPMBO

PMBO

24

23

30% TFA

30% TFAS

OHHO

OSO3 OH

OH OH

HO

OHOH

S

OHHO

OSO3 OH

OH OH

HO

OHOH

4

6

93%

91%

1' 2'

1' 2'

3'

3'

4'

4'5'

5'

6'

6'

Scheme 2-6: Coupling reactions.

Comparison of the 1H and 13C NMR spectra of kotalanol 4 with those

reported13 revealed identical data and served, therefore, to confirm the

stereochemistry at C-6', and, by inference, the stereochemistry at C-2 in each of

18 and 19.

55

Finally, the inhibitory activities of compounds 4 and 6 were examined

against the N-terminal catalytic domain of recombinant human maltase-

glucoamylase (ntMGAM), a critical intestinal glucosidase for processing starch-

derived oligosaccharides into glucose. The stereoisomer of kotalanol 4 inhibited

ntMGAM with a Ki value of 0.20 ± 0.02 µM; this compares to a Ki value for

kotalanol of 0.19 ± 0.03 µM,18 and Ki values of 0.10 ± 0.02 µM and 0.13 ± 0.02

µM for other stereoisomers of 4 with opposite configurations at C-5’ or both C-5’

and C-6’, respectively.16 It is clear, therefore, that the configurations at C-5’ and

C-6’ are not critical for dictating enzyme inhibitory activity against ntMGAM.

2.5 Experimental

2.5.1 General

Optical rotations were measured at 23 °C. 1H and 13C NMR spectra were

recorded at 600 and 150 MHz, respectively. All assignments were confirmed with

the aid of two-dimensional 1H, 1H (COSYDFTP) or 1H, 13C (INVBTP) experiments

using standard pulse programs. Column chromatography was performed with

Silica 60 (230-400 mesh). High resolution mass spectra were obtained by the

electrospray ionization method, using an Agilent 6210 TOF LC/MS high

resolution magnetic sector mass spectrometer.

2.5.2 Enzyme inhibition assays

Compounds 4 and 6 were tested for inhibition of ntMGAM, as previously

described.16

56

2.5.3 Compound characterization data

2.5.3.1 1,3,4,6-Di-O-benzylidene-2,5-O-isopropylidene-D-mannitol (14)

Compound 15 (9.30 g, 26.00 mmol) was dissolved in 2,2-di-

methoxypropane (150 mL), PTSA (1.50 g, 0.3 eq) was added, and the mixture

was rotated on a rotary evaporator at room temperature under reduced pressure

for 1 h. The reaction mixture was quenched by addition of Et3N to pH > 9. The

reaction mixture was concentrated under vacuum to give a white solid which was

dissolved in CHCl3 (200 mL) and washed with water (3×50 mL). The separated

organic layer was dried over Na2SO4, concentrated, and the residue was purified

by column chromatography with EtOAc/Hexanes (1:4) as eluent to afford 14 as a

white solid (7.55 g, 73%). Mp 160-162 °C; [α] 23D = -83°, (c = 1.1, CH2Cl2). 1H

NMR (CDCl3) δ 7.54-7.37 (10H, m, Ar), 5.54 (2H, s, 2CH-Ph), 4.24 (2H, dd, J1a,1b

= 10.8, J2,1 = 5.3 Hz, H-1), 3.95-3.91 (2H, m, H-6a, H-5), 3.84-3.80 (2H, m, H-3,

H-4), 3.74 (2H, t, J1,2 = J2,3 = J5,6b = J6a,6b = 10.5, H-2, H-6b), 1.42 (6H, s, 2Me).

13C NMR (CDCl3) δ 137.5 (CMe2), 129.9-126.2 (m, Ar), 100.7 (CH-Ph), 82.2 (C-3,

C-4), 69.4 (C-1, C-6), 61.7 (C-2, C-5), 24.4 (2Me). HRMS Calcd for C23H27O6

(M+H): 399.1802. Found: 399.1809.

2.5.3.2 1,3-O-Benzylidene-2,5-O-isopropylidene-D-mannitol (16)

To a solution of compound 14 (7.50 g, 18.84 mmol) in MeOH (300 mL),

was added PTSA (300 mg), and the reaction was stirred at room temperature for

30 min. The reaction mixture was then quenched by addition of Et3N to pH > 9,

and the solvent were removed under vacuum to give a solid. The solid was

dissolved in CH2Cl2 (100 mL) and washed with water (50 mL). The organic

57

solution was dried (Na2SO4), concentrated, and the crude product was purified

through a silica column with EtOAc/Hexanes (1:1) as eluent to yield 16 as a foam

(4.1g, 70%). [α] 23D = -15°, (c = 1 , CH2Cl2). 1H NMR (CDCl3) δ 7.42-7.30 (5H, m,

Ar), 5.40 (1H, s, CH-Ph), 4.12 (1H, dd, J1a,1b = 10.8, J1a,2 = 5.5 Hz, H-1a), 3.81

(1H, dd, J6a,6b = 10.9, J6a,5 = 4.3 Hz, H-6a), 3.76-3.72 (2H, m, H-3, H-5), 3.66 (1H,

m, H-6b), 3.60-3.53 (2H, m, H-4, H-1b), 3.43 (1H, t, J1,2 = 8.9 Hz, H-2), 2.23 (2H,

br, 2OH), 1.30 (6H, s, 2Me). 13C NMR (CDCl3) δ 137.3 (CMe2), 129.3- 101.7 (m,

Ar), 101.1 (CH-Ph), 85.2 (C-2), 73.9 (C-4), 70.3 (C-5), 69.3 (C-1), 63.6 (C-6),

61.2 (C-3), 24.8, 24.6 (2Me). HRMS Calcd for C16H23O6 (M+H): 311.1489. Found:

311.1487.

2.5.3.3 1,3-O-Benzylidene-2,5-O-isopropylidene-4-O-methoxymethyl-D-mannitol (17)

To a solution of 16 (6.80 g, 21.93 mmol) in DMF (125 mL) was added

imidazole (4.47g, 65.81 mmol). The reaction was cooled in an ice bath,

TBDMSCl (3.79 g, 24.13 mmol) was added portionwise, and the mixture was

stirred at 0 °C under N2 for 2 h. The reaction was quenched by the addition of ice-

water, and the reaction mixture was extracted with Et2O (3×75 mL). The

combined organic solvents were dried (Na2SO4) and concentrated to give the

crude product which was used directly in the next step without further purification.

The crude product was dissolved in DMF (60 mL), and i-Pr2NEt (26 mL, 150.75

mmol) and MOMCl (5.7 mL, 75.38 mmol) were added. The reaction mixture was

heated at 60 °C overnight, then quenched with ice, and extracted with ether

(3×50 mL). The organic solution was dried (Na2SO4) and concentrated to give a

58

crude product. The crude residue was dissolved in THF (100 mL), TBAF (1.0 M

solution in THF, 13.8 mL, 24.12 mmol) was added, and the reaction mixture was

stirred at room temperature. After 4 h it was concentrated and the residue was

purified by flash chromatography (EtOAc/Hexanes (1:3)) to yield 17 as a white

solid (5.67 g, 73%). Mp 65-67 °C; [α] 23D = +22° (c = 1, MeOH). 1H NMR (CDCl3)

δ 7.49-6.37 (5H, m, Ar), 5.50 (1H, s, CH-Ph), 4.93, 4.73 (2H, 2d, JA,B = 6.4 Hz,

CH2OMe), 4.20 (1H, dd, J1a,1b = 10.9, J1a,2 = 5.5 Hz, H-1a), 3.89-3.79 (4H, m, H-

2, H-5, H-6a,b), 3.74 (1H, t, J3,4 = J5,4 = 8.1 Hz, H-4), 3.69-3.65 (2H, m, H-1b, H-

3), 3.40 (3H, s, OMe), 2.69 (1H, t, J6,OH = 8.5 Hz, OH), 1.41, 1.38 (6H, 2s, 2Me).

13C NMR (CDCl3) δ 137.5 (CMe2), 128.9-101.5 (m, Ar), 100.9 (CH-Ph), 98.6

(CH2-OMe) 85.3 (C-3), 78.2 (C-4), 70.4 (C-5), 69.5 (C-1), 63.1 (C-6), 61.3 (C-2),

56.4 (OMe), 24.7, 24.4 (2Me). HRMS Calcd for C18H27O7 (M+H): 355.1751.

Found: 355.1741.

2.5.3.4 5,7-O-Benzylidene-1,2-dideoxy-3,6-O-isopropylidene-4-O-methoxymethyl-D-manno-hep-1-enitol (13)

Compound 17 (2.60 g, 7.34 mmol) was dissolved in CH2Cl2 (50 mL) and

NaHCO3 (2.77 g, 33.03 mmol) and Dess Martin periodinane (3.73 g, 8.81 mmol)

were added. The reaction mixture was stirred for 2h at room temprature, diluted

with ether (100 mL), and poured into saturated aqueous NaHCO3 (100 mL)

containing a seven fold excess of Na2S2O3. The mixture was stirred to dissolve

the solid, and the ether layer was separated and dried over Na2SO4. The ether

was removed to give the aldehyde that was further dried under high vacuum for 1

h. Methyltriphenylphosphonium bromide (2.99 g, 8.80 mmol) in dry THF (15 mL),

59

was cooled to -78 °C and n-BuLi (n-hexane solution, 14.67 mmol) was added

dropwise under N2. The reaction mixture was stirred at the same temperature for

1h, and a solution of the previously made aldehyde in THF (10 mL) was added.

The resulting mixture was allowed to warm to rt and was stirred overnight. The

reaction was quenched by the addition of acetone (1.5 mL), and the mixture was

extracted with ether (3×100 mL). The combined organic layers were washed with

brine, dried (Na2SO4), and concentrated in vacuo. Chromatographic purification

of the crude product (EtOAc/Hexanes (1:10)) gave 13 as a foam (1.56 g, 61%).

[α] 23D = +4° (c = 0.5, CH2Cl2). 1H NMR (CDCl3) δ 7.50-7.36 (5H, m, Ar), 6.05 (1H,

ddd, J5,6 = 6.1, J6,7b = 10.5, J6,7a = 16.6 Hz, H-6), 5.51 (1H, s, CH-Ph), 5.39 (1H,

ddd, J7b,7a = 17.1, J6,7a = 3.3, J5,7a = 1.5 Hz, H-7a), 5.36 (1H, ddd, J7a,7b = 10.7,

J6,7b = 3.1, J5,7b = 1.5 Hz, H-7b), 5.27, 5.26 (2H, 2d, JA,B = 6.25 Hz, CH2OMe),

4.25 (1H, m, H-5), 4.20 (1H, dd, J1a,1b = 10.8, J1a,2 = 5.4 Hz, H-1a), 3.90 (1H, dt,

J2,3 = 5.4, J2,1 = 9.9 Hz, H-2), 3.68 (2H, m, H-3, H-1b), 3.56 (1H, dd, J3,4 = 8.1 ,

J4,5 = 9.7 Hz, H-4), 3.33 (3H, s, OMe), 1.40, 1.37 (6H, 2s, 2Me). 13C NMR

(CDCl3) δ 137.6 (CMe2), 136.2 (C-6), 128.9-101.3 (m, Ar), 116.8 (C-7), 100.7

(CH-Ph), 97.9 (CH2OMe), 85.5 (C-3), 80.2 (C-4), 71.1 (C-5), 69.6 (C-1), 61.4 (C-

2), 56.4 (OMe), 24.8,24.1 (2Me). HRMS Calcd for C19H26NaO6 (M+Na):

373.1622. Found: 373.1606.

2.5.3.5 1,3-O-Benzylidene-2,5-O-isopropylidene-4-O-methoxymethyl-D-glycero-D-manno-heptitol (18)

To a solution of 13 (2.00 g, 5.71 mmol) in acetone:water (9:1, 6 mL) at rt

were added NMO (N-methylmorpholine-N-oxide) (735 mg, 6.28 mmol) and OsO4

60

(40 mg, 2.5 wt % solution in 2-methyl-2-propanol). The reaction mixture was

stirred at rt for 48 h before it was quenched with a saturated solution of NaHSO3.

After being stirred for an additional 15 min the reaction mixture was extracted

with ethyl acetate and the organic layer was washed with water and brine, dried

(Na2SO4), and concentrated in vacuo. The crude material was purified by column

chromatography on silica gel (MeOH/CH2Cl2 (1:100)) to give 18 (1.27 g, 58%)

and 19 (0.48 g, 22%) as foams. [α] 23D = +5.8° (c = 4.6, MeOH). 1H NMR (MeOD)

δ 7.49-7.36 (5H, m, Ar), 5.54 (1H, s, CH-Ph), 4.82 (1H, s, CH2OMe ), 4.13 (1H,

dd, br, H-1a), 4.00 (1H, br, q, H-6), 3.87-3.77 (3H, m, H-4, H-5, H-2), 3.68-3.55

(4H, H-1b, H-3, H-7a, H-7b), 3.32 (3H, s, OMe), 1.39, 1.34 (6H, 2s, 2Me). 13C

NMR (MeOD) δ 138.0 (CMe2), 128.4-101.1 (m, Ar), 100.8 (CH-Ph), 97.7

(CH2OMe), 85.3 (C-4), 77.1 (C-2), 69.2 (C-6), 69.1 (C-5), 69.0 (C-1), 62.3 (C-7),

61.1 (C-3), 55.3 (OMe), 23.5, 23.4 (2Me). HRMS Calcd for C19H29O8 (M+H):

385.1857. Found: 385.1875.

2.5.3.6 5,7-O-Benzylidene-3,6-O-isopropylidene-4-O-methoxymethyl-D-glycero-D-galacto-heptitol (19)

[α] 23D = -20° (c = 0.1, MeOH). 1H NMR (MeOD) δ 7.48-7.34 (5H, m, Ar),

5.51(1H, s, CH-Ph), 4.49, 4.47 (2H, 2d, JA,B = 6.2 Hz, CH2OMe), 4.13 (1H, dd,

J7a,7b = 10.7, J6,7b = 5.4 Hz, H-7a), 4.08 (1H, m, H-2), 3.95 (1H, dd, J3,4 = 9.7, J5,4

= 2.8 Hz, H-4), 3.85 (1H, dd, J1a,1b = 11.4, J2,1a = 3.6 Hz, H-1a), 3.78 (1H, dt, J6,7

= 9.9, J5,6 = 5.4 Hz, H-6), 3.67-3.60 (4H, m, H-5, H-7b, H-1b, H-3), 3.35 (3H, s,

OMe), 1.37, 1.36 (6H, 2s, 2Me). 13C NMR (MeOD) δ 137.9 (CMe2), 128.5 -101.3

(m, Ar), 100.6 (CH-Ph), 97.7 (CH2OMe), 86.0 (C-5), 78.2 (C-3), 72.1 (C-4), 71.3

61

(C-2), 69.1 (C-7), 61.3 (C-1), 61.0 (C-6), 55.6 (OMe), 23.6, 23.4 (2Me). HRMS

Calcd for C19H29O8 (M+H): 385.1857. Found: 385.1865.

2.5.3.7 2,5-O-Isopropylidene-4,6,7-tri-O–methoxymethyl-D-glycero-D-manno-heptitol (20)

Compound 18 (580 mg, 1.51 mmol), was dissolved in DMF (20 mL) and i-

Pr2NEt (4.21 mL, 24.16 mmol) and MOMCl (0.9 mL, 12.08 mmol) were added.

The reaction mixture was heated at 60 °C for 2 h, then quenched with ice, and

extracted with ether (3×30 mL). The organic solution was dried (Na2SO4) and

concentrated to give a crude product that was further dried under high vacuum

for 1 h. The crude product was dissolved in MeOH (50 mL) and the solution was

stirred with Pd(OH)2 20 wt% on carbon (520 mg) under 100 Psi of H2 for 1h. The

catalyst was removed by filtration through a bed of Celite, then washed with

methanol. The solvents were removed under reduced pressure and the residue

was purified by flash column chromatography (EtOAc/Hexanes (1.5:1)) to give 20

as a colorless syrup (420 mg, 72%). [α] 23D = +48.0° (c = 0.1, MeOH). 1H NMR

(MeOD) δ 4.90-4.63 (6H, m, 3CH2OMe), 4.20 (1H, dd, br, H-6), 3.95 (1H, d, br,

J4,5 = 8.6 Hz, H-5), 3.86-380 (2H, m, H-1a, H-7a), 3.68-3.58 (3H, m, H-2, H-7b,

H-1b), 3.45, 3.42, 3.36 (9H, 3s, 3OMe), 3.34 (2H, m, H-4, H-3), 1.35 (6H, s,

2Me). 13C NMR (CDCl3) δ 100.7 (CMe2), 98.4, 96.3, 95.5 (3CH2OMe), 83.9 (C-4),

75.0 (C-6), 74.9 (C-3), 71.2 (C-2), 70.5 (C-5), 66.2 (C-7), 62.5 (C-1), 55.3, 54.5,

54.1 (3OMe), 22.6, 22.4 (2Me). HRMS Calcd for C16H33O10 (M+H): 385.2068.

Found: 385.2083.

62

2.5.3.8 3,6-O-Isopropylidene-1,2,4-tri-O–methoxymethyl-D-glycero-D-galacto-heptitol (21)

Compound 21 was obtained as a colorless syrup (285 mg, 75%) from 19

(380 mg, 1 mmol) using the same procedure that was used to obtain 20. [α] 23D = -

30° (c = 0.4, MeOH). 1H NMR (MeOD) δ 4.84-4.61 (6H, m, 3CH2OMe), 4.08 (1H,

ddd, J3,2 = 1.3 , J2,1a = 5.6, J2,1b = 7.2 Hz, H-2), 3.86-3.84 (2H, m, H-7a, H-3),

3.74 (1H, dd, J1a,1b = 9.5, J1a,2 = 5.6 Hz, H-1a), 3.69 (1H, ddd, J6,5 = 2.9, J6,7b

=6.8, J6,7a = 9.8 Hz, H-6), 3.60-3.55 (2H, m, H-1b, H-7b), 3.45(1H, m, H-5), 3.44,

3.38, 3.35 (9H, 3s, 3OMe), 3.34 (1H, m, H-4), 1.36, 1.32 (6H, 2s, 2Me). 13C NMR

(CDCl3) δ 100.1 (CMe2), 97.8, 96.8, 95.9 (3CH2OMe), 83.3 (C-5), 75.1 (C-2),

73.8 (C-4), 70.3 (C-6), 67.8 (C-3), 65.9 (C-1), 61.8 (C-7), 54.3, 54.1, 53.7

(3OMe), 22.9, 22.8 (2Me). HRMS Calcd for C16H33O10 (M+H): 385.2068. Found:

385.2067.

2.5.3.9 2,5-O-Isopropylidene-4,6,7-tri-O–methoxymethyl-D-glycero-D-manno-heptitol-1,3-cyclic sulfate (12)

A mixture of 20 (400 mg, 1.04 mmol) and Et3N (0.57 mL, 4.16 mmol) in

CH2Cl2 (10 mL) was stirred in an ice bath. Thionyl chloride (0.12 mL, 1.56 mmol)

in CH2Cl2 (2 mL) was then added dropwise over 15 min, and the mixture was

stirred for an additional 30 min. The mixture was poured into ice-cold water and

extracted with CH2Cl2 (3×30 mL). The combined organic layers were washed

with brine and dried over Na2SO4. The solvent was removed under reduced

pressure and the residue was dried under high vacuum for 1 h. The

diasteromeric mixture of cyclic sulfites was dissolved in a mixture of CH3CN:CCl4

(1:1, 25 mL) and sodium periodate (333 mg, 1.56 mmol) and RuCl3 (10 mg) were

63

added, followed by water (2 mL). The reaction mixture was stirred for 2 h at rt,

then filtered through a bed of Celite, and washed with ethyl acetate. The volatile

solvents were removed, and the aqueous solution was extracted with EtOAc

(2×30 mL). The combined organic layers were washed with brine, dried over

Na2SO4, concentrated under reduced pressure, and the residue purified by flash

column chromatograghy (EtOAc/Hexanes (1:2)) to give 12 as a colorless syrup

(325 mg, 70%). [α] 23D = +1.2° (c = 0.85, CH2Cl2). 1H NMR (CDCl3) δ 4.77-4.65

(6H, m, 3CH2OMe), 4.62 (1H, t, J2,3 = J4,3 = 9.1 Hz, H-3), 4.54 (1H, t, J1a,1b = J2,1a

= 11.1 Hz, H-1a), 4.37 (1H, dd, J2,1a = 5.4, J1a,1b = 11.1 Hz, H-1b), 4.16 (2H, m,

H-2, H-6), 3.98 (1H, d, J4,5 = 9.8 Hz, H-5), 3.80 (1H, dd, J6,7a = 4.8, J7a,7b = 10.8

Hz, H-7a), 3.75 (1H, t, J3,4 = J4,5 = 8.6 Hz, H-4), 3.64 (1H, t, J7a,7b = J6,7b = 8.9 Hz,

H-7b), 3.44, 3.41, 3.39 (9H, 3s, 3OMe), 1.38, 1.36 (6H, 2s, 2Me). 13C NMR

(CDCl3) δ 102.3 (CMe2), 97.9, 96.7, 96.1 (3CH2OMe), 89.2 (C-3), 76.9 (C-4),

74.6 (C-6), 72.2 (C-1), 71.0 (C-5), 66.8 (C-5), 66.8 (C-7), 56.5 (C-2), 56.6, 55.7,

55.3 (3OMe), 24.4, 23.9 (2Me). HRMS Calcd for C16H31O12S (M+H): 447.1531.

Found: 447.1516.

2.5.3.10 3,6-O-Isopropylidene-1,2,4-tri-O–methoxymethyl-D-glycero-D-galacto-heptitol-5,7-cyclic sulfate (22)

Compound 22 was obtained as a colorless syrup (220 mg, 76%) from 21

(250 mg, 0.65 mmol) using the same procedure that was used to obtain 12. [α] 23D

= -32° (c = 0.46, CH2Cl2). 1H NMR (CDCl3) δ 4.83-4.63 (6H, m, 3CH2OMe), 4.70

(1H, m, H-5), 4.55 (1H, t, J7a,7b = J6,7a = 11.1 Hz, H-7a), 4.39 (1H, dd, J6,7a = 4.9,

J7a,7b = 10.7 Hz, H-7b), 4.24 (1H, td, J5,6 = 5.7, J6,7 = 10.5 Hz, H-6), 4.09 (1H,

64

ddd, J1a,2 = 6.9, J1b,2 = 5.3, J3,2 = 1.4 Hz, H-2), 3.97 (1H, dd, J4,3 = 10.0, J3,2 = 1.5

Hz, H-3), 3.89 (1H, dd, J3,4 = 10.0, J5,4 = 7.7 Hz, H-4), 3.80 (1H, dd, J2,1a = 5.4,

J1a,1b = 9.8 Hz, H-1a), 3.58 (1H, t, J1a,1b = J2,1b = 9.2 Hz, H-1b), 3.45, 3.41, 3.39

(9H, 3s, 3OMe), 1.43, 1.37 (6H, 2s, 2Me). 13C NMR (CDCl3) δ 102.32 (CMe2),

98.1, 98.0, 97.9 (3CH2OMe), 89.6 (C-5), 76.6 (C-4), 75.1(C-2), 72.0 (C-7), 68.9

(C-3), 66.2 (C-1), 59.5 (C-6), 56.4, 56.0, 55.7 (3OMe), 24.8, 23.8 (2Me). HRMS

Calcd for C16H30NaO12S (M+Na): 470.1383. Found: 470.1399.

2.5.3.11 2,3,5-Tri-O–p–methoxybenzyl-1,4-dideoxy-1,4-[[2S,3S,4R,5R,6R]-2,5-isopropylidene-4,6,7-tri-O–methoxymethyl-3-(sulfooxy)heptyl]-(R)-epi-sulfoniumylidine-D-arabinitol inner salt (23)

The cyclic sulfate 12 (260 mg, 0.58 mmol) and the thiosugar 7 (360 mg,

0.70 mmol) were dissolved in HFIP (1.5 mL), containing anhydrous K2CO3 (10

mg). The mixture was stirred in a sealed reaction vessel in an oil bath at 72 °C

for 6 days. The progress of the reaction was followed by TLC analysis

(developing solvent EtOAc:MeOH, 10:1). The mixture was cooled, then diluted

with EtOAc and evaporated to give a syrupy residue. Purification by column

chromatograghy (EtOAc/MeOH 99:1) gave the sulfonium salt 23 as a syrup (360

mg, 65%). [α] 23D = +62° (c = 0.85, CH2Cl2). 1H NMR (acetone-d6) δ 7.32-

6.91(12H, m, Ar), 5.12-4.52 (12H, m, 3CH2OMe, 3CH2-Ph), 4.69 (1H, m, H-2),

4.55 (1H, m, H-3), 4.39-4.30 (4H, m, H-1'a, H-2', H-3', H-6'), 4.08 (1-H, t, J3,4 =

J5,4 = 7.4 Hz, H-4), 4.02-3.90 (4H, m, H-1a, H-1'b, H-5'), 3.85-3.78 (3H, m, H-5a,

H-7'a, H-1b), 3.82 (9H, s, 3Ph-OMe), 3.60 (1H, t, J7'a,7'b = J6',7'b = 9.1 Hz, H-7'b),

3.42 (1H, m, H-4'), 3.39, 3.36, 3.33 (9H, 3s, 3CH2OMe) 1.37, 1.32 (6H, 2s, 2Me).

13C NMR (acetone-d6) δ 159.8-129 (m, Ar), 101.6 (CMe2), 98.7, 96.5, 95.2

65

(3CH2OMe), 83.5 (C-3), 81.2 (C-2), 79.7 (C-2'), 78.6 (C-4'), 74.0 (C-6'), 72.7,

71.6, 71.4 (3CH2Ph), 71.3 (C-5'), 66.9 (C-7'), 66.6 (C-3'), 66.5 (C-5), 65.1 (C-4),

55.9-54.2 (6OMe), 51.5 (C-1'), 47.4 (C-1), 24.4, 23.5 (2Me). HRMS Calcd for

C45H65O18S2 (M+H): 957.3607. Found: 957.3604.

2.5.3.12 1,4-Dideoxy-1,4-[[2S,3S,4R,5R,6R-2,4,5,6,7-pentahydroxy-3-(sulfooxy)heptyl]-(R)-epi-sulfoniumylidine]-D-arabinitol inner salt (6)

The protected sulfonium salt 23 (150 mg, 0.16 mmol) was dissolved in

30% aqueous solution of TFA (25 mL) and the mixture was stirred at 50 °C for

5h. The solvent was removed under reduced pressure and the residue was

dissolved in water (5 mL) and washed with CH2Cl2 (3×5 mL). The water layer

was evaporated to give the crude product that was purified on silica gel with

EtOAc/MeOH/H2O 6:3:1 (v/v) as eluent to give compound 6 as a colorless solid

(61 mg, 93%). Mp 82-84 °C [α] 23D = +5.5° (c = 0.55, CH2Cl2). 1H NMR (D2O) δ

4.67 (1H, dd, J1a,2 = 3.7, J1b,2 = 7.4 Hz, H-2), 4.56 (1H, d, J2',3' = 8.2 Hz, H-3'),

4.39 (1H, t, J2,3 = J3,4 = 3.1 Hz, H-3), 4.35 (1H, dt, J2’,3’ = 3.3, J2’,1’ = 7.8 Hz, H-2'),

4.02 (3H, m, H-5a, H-1'a, H-4), 3.91-3.83 ( 5H, m, H-6', H-5', H-5b, H-4', H-1’b),

3.81 (2H, d, J1,2 = 3.9 Hz, H-1a,b), 3.71 (1H, dd, J7'a,7'b = 3.2, J7'b,6' = 11.9 Hz, H-

7'b), 3.62 (1H, dd, J7'b,7'a = 7.8, J7'a,6' = 11.6 Hz, H-7'a). 13C NMR (D2O) δ 78.3 (C-

3'), 77.7 (C-3), 76.7 (C-2), 72.9 (C-6'), 70.7 (C-5'), 70.0 (C-4), 69.1 (C-4'), 66.0

(C-2'), 61.6 (C-7'), 59.2 (C-5), 50.7 (C-1'), 47.7 (C-1). HRMS Calcd for

C12H25O12S2 (M+H): 425.0782. Found: 425.0778.

66

2.5.3.13 1,4-Dideoxy-1,4-[[2S,3S,4R,5R,6S-2,4,5,6,7–pentahydroxy-3-(sulfooxy)heptyl]-(R-)epi-sulfoniumylidine]-D-arabinitol inner salt (4)

A mixture of the thiosugar 7 (100 mg, 0.224 mmol) and the cyclic sulfate

22 (137 mg, 0.269 mmol) in HFIP (1 mL) containing K2CO3 (5 mg) was placed in

a sealed reaction vessel and heated at 72 °C with stirring for 6 days. The

progress of the reaction was followed by TLC analysis (developing solvent

EtOAc:MeOH, 10:1). The mixture was cooled, then diluted with EtOAc and

evaporated to give a syrupy residue. Purification by column chromatography

(EtOAc/MeOH 95:5) gave the protected sulfonium salt as a foam (120 mg, 57%).

The protected sulfonium salt 24 (100 mg, 0.11 mmol) was dissolved in 30%

aqueous TFA (10 mL) and stirred at 50 °C for 5h. The solvents were removed

under reduced pressure and the residue was dissolved in water (5 mL) and

washed with CH2Cl2 (3×5 mL). The water layer was evaporated to give the crude

product that was purified on silica gel column with EtOAc/MeOH/H2O 6:3:1 (v/v)

as eluent to give compound 4 as a colorless solid (40 mg, 91%).12

2.6 Acknowledgments

We are grateful to the Canadian Institutes for Health Research (FRN79400) and

the Heart and Stroke Foundation of Ontario (NA-6305) for financial support.

67

2.7 References

1. Dwek, R. A. Chem. Rev. 1996, 96, 683-720.

2. Varki, A.; Cummings, R.; Esko, J.; Freeze, H.; Hart, G.; Marth, J.

Essentials of Glycobiology; Cold Spring Harbor Laboratory Press: Cold

Spring Harbor, NY, 1999.

3. For a recent review on glycosidase inhibitors: de Melo, E. B.; Gomes, A.

D.; Carvalho, I. Tetrahedron 2006, 62, 10277-10302.

4. Chandrasena, J. P. C. The Chemistry and Pharmacology of Ceylon and

Indian medicinal plants, H&C Press, Colombo, Sri Lanka, 1935.

5. Jayaweera, D. M. A. Medicinal Plants Used in Ceylon-Part 1, National

Science Council of Sri Lanka: Colombo, 1981, p. 77.

6. Vaidyartanam, P. S. In Indian Medicinal Plants: a compendium of 500

species, Warrier, P. K.; Nambiar, V. P. K.; Ramankutty, C.; Eds. Orient

londman: India 1993, pp. 47-48.

7. Yoshikawa, M.; Murakami, T.; Shimada, H.; Matsuda, H.; Yomahara, J.;

Tanabe, G.; Muraoka, O. Tetrahedron Lett. 1997, 38, 8367-8370.

8. Yoshikawa, M.; Murakami, T.; Yashiro, K.; Matsuda, H.; Chem. Pharm.

Bull. 1998, 46, 1339-1340.

9. Yoshikawa, M.; Xu, F.; Nakamura, S.; Wang, T.; Matsuda, H.; Tanabe, G.;

Muraoka, O. Heterocycles 2008, 75, 1397-1405.

10. Muraoka, O.; Xie, W.; Tanabe, G.; Amer, M. F. A.; Minematsu, T.;

Yoshikawa, M. Tetrahedron Lett. 2008, 49, 7315-7317.

11. For reviews see: a) Mohan, S.; Pinto, B. M. Carbohydr. Res. 2007, 342,

1551-1580. b) Mohan, S.; Pinto, B. M. Collect. Czech. Chem. Commun.

2009, 74, 1117-1136.

68

12. Jayakanthan, K.; Mohan, S.; Pinto, B. M. J. Am. Chem. Soc. 2009, 131,

5621-5626.

13. Mohan, S.; Jayakanthan, K.; Nasi, R.; Kuntz D. A.; Rose, D. R.; Pinto, B.

M. Organic Lett. 2010, 12, 1088-1091.

14. Baggett, N.; Stribblehill, P. J. Chem. Soc., Perkin Trans. 1, 1977, 1123-

1126.

15. Harris, J. M.; Keranen, M. D.; O'Doherty, G. A. J. Org. Chem. 1999, 64,

2982-2983.

16. Nasi, R.; Patrick, B. O.; Sim, L.; Rose, D. R.; Pinto, B. M. J. Org. Chem.

2008, 73, 6172-6181.

17. Ghavami, A.; Sadalapure, K. S.; Johnston, B. D.; Lobera, M.; Snider, B.

B.; Pinto, B. M. Synlett 2003, 9, 1259-1262.

18. Sim, L.; Jayakanthan, J.; Mohan, S.; Nasi, R.; Johnston, B.D.; Pinto, B. M.;

Rose, D. R. Biochemistry 2010, 49, 443-451.

69

2.8 Supporting Information

Synthesis of a biologically active isomer of kotalanol, a naturally-

occurring glucosidase inhibitor

Razieh Eskandari, Kumarasamy Jayakanthan, Douglas A. Kuntz, David R.

Rose, B. Mario Pinto

70

ppm (t1)50100

ppm (t1)2.03.04.05.06.07.0

O O

OO O

Ph

O

14Ph

O O

OO O

Ph

O

14Ph

71

ppm (t1)2.03.04.05.06.07.0

O O

OO OH

Ph

OH

16

ppm (t1)50100

O O

OO OH

Ph

OH

16

72

ppm (t1)50100

ppm (t1)2.03.04.05.06.07.0

O O

OO OMOM

Ph

OH

17

O O

OO OMOM

Ph

OH

17

73

ppm (t1)50100

ppm (t1)2.03.04.05.06.07.0

O O

OO OMOM

Ph 13

O O

OO OMOM

Ph 13

74

ppm (t1)50100

ppm (t1)2.03.04.05.06.07.0

O O

OO OMOM

Ph

OH

OH

18

12 3 4 5 6

7

O O

OO OMOM

Ph

OH

OH

18

12 3 4 5 6

7

75

ppm (t1)50100

ppm (t1)2.03.04.05.06.07.0

O O

OO OMOM

Ph

OH

OH

19

123456

7

O O

OO OMOM

Ph

OH

OH

19

123456

7

76

ppm (t1)30405060708090100

ppm (t1)1.502.002.503.003.504.004.50

O O

HOHO OMOM

OMOM

OMOM

20

O O

HOHO OMOM

OMOM

OMOM

20

77

ppm (t1)30405060708090100

ppm (t1)1.502.002.503.003.504.004.50

O O

OS O

OOOMOM

OMOM

OMOM

12

O O

OS O

OOOMOM

OMOM

OMOM

12

78

ppm (t1)30405060708090100

ppm (t1)1.502.002.503.003.504.004.50

O O

HOHO OMOM

OMOM

OMOM

21

O O

HOHO OMOM

OMOM

OMOM

21

79

ppm (t1)30405060708090100

ppm (t1)1.502.002.503.003.504.004.50

O O

OS O

OOOMOM

OMOM

OMOM

22

O O

OS O

OOOMOM

OMOM

OMOM

22

80

ppm (t1)2.03.04.05.06.07.0

ppm (t1)50100150

O O

OSO3OMOM

OMOM

OMOMS

OPMBPMBO

PMBO

23

O O

OSO3OMOM

OMOM

OMOMS

OPMBPMBO

PMBO

23

81

ppm (t1)50.055.060.065.070.075.0

ppm (t1)4.004.50

S

OHHO

OSO3 OH

OH OH

HO

OHOH

6

S

OHHO

OSO3 OH

OH OH

HO

OHOH

6

82

CHAPTER 3: POTENT GLUCOSIDASE INHIBITORS: DE-O-SULFONATED PONKORANOL AND ITS

STEREOISOMER

This Chapter comprises the manuscript “Potent glucosidase inhibitors:

de-O-sulfonated ponkoranol and its stereoisomer” which was published in

Organic Letters (2010, 12, 1632-1635).

Razieh Eskandari,a Douglas A. Kuntz,b David R. Rose,b B. Mario Pintoa

aDepartment of Chemistry, Simon Fraser University, Burnaby, British Columbia,

Canada V5A 1S6

b Department of Biology, University of Waterloo, Waterloo, Ontario,

Canada N2L 3G1

83

Ponkoranol is a naturally-occurring glucosidase inhibitor isolated from the

plant Salacia reticulata. The compound comprises a sulfonium ion with an

internal sulfate counter ion. Since de-O-sulfonated compounds of naturally-

occurring sulfonium-ion glucosidase inhibitors isolated from Salacia species show

better inhibitory activities against α-glucosidases, we report in this Chapter, an

efficient synthetic route to de-O-sulfonated ponkoranol and its stereoisomer, and

show that these compounds are potent glucosidase inhibitors that inhibit a key

human intestinal glucosidase, the N-terminal domain of maltase glucoamylase.

The thesis author performed all the experimental synthetic work and the

characterization of the compounds, and wrote and edited the manuscript with Dr.

B. Mario Pinto. Drs. Douglas A. Kuntz and David R. Rose performed the enzyme

inhibition studies.

Graphical abstract:

O

HOOH

OBnTsO

S

OHHO

HO

OH

OH

OH

OH

ClS

OBnBnO

BnO+

R1 = OH, R2 = HR1 = H, R2 = OH

R2R1

8. R1 = OH, R2 = H9. R1 = H, R2 = OH

R2 R1

84

3.1 Keywords

Ponkoranol, sulfonium-ion, de-O-sulfonated ponkoranol, glucosidase

inhibitors.

3.2 Abstract

Ponkoranol, a glucosidase inhibitor isolated from the plant Salacia

reticulata, comprises a sulfonium-ion with an internal sulfate counterion. An

efficient synthetic route to de-O-sulfonated ponkoranol and its 5’-stereoisomer is

reported, and it is shown that these compounds are potent glucosidase inhibitors

that inhibit a key intestinal human glucosidase, the N-terminal catalytic domain of

maltase-glucoamylase, with Ki values of 43 ± 3 and 15 ± 1 nM, respectively.

3.3 Introduction

Compounds isolated from medicinal plants can provide the lead structures

for drug development programs.1,2 For example, the aqueous extracts of the

roots and stems of the large woody climbing plant Salacia reticulata, known as

Kothalahimbutu in Singhalese, have been used in the Ayurvedic system of Indian

medicine in Sri Lanka and Southern India for the treatment of type-2 diabetes.3,4

Several glucosidase inhibitors have been isolated from the water-soluble fraction

of this plant extract and also other plants that belong to the Salacia genus such

as Salacia chinensis, Salacia prinoides, and Salacia oblonga which explain, at

least in part, the antidiabetic property of the aqueous extracts of these plants.5-7

Thus far, six components have been isolated from the plant S. reticulate, namely

salaprinol (1), 7 salacinol (2), 6 ponkoranol (3), 7 kotalanol (4),5 de-O-sulfonated

85

kotalanol (5),8 and de-O-sulfonated salacinol (6)9 (Figure 3-1), all of which

possess a common structural motif that comprises a 1,4-anhydro-4-thio-D-

arabinitol and a polyhydroxylated side chain.

Figure 3-1: Components isolated from Salacia species.

We have carried out extensive research on the synthesis of higher

homologues of salacinol (2) which has led to the stereochemical structure

elucidation of compounds 3-5.10-12 Interestingly, ponkoranol (3), the recently

isolated7 six-carbon-chain homologue of salacinol, was synthesized by us several

years earlier with the expectation that it would be an effective glucosidase

inhibitor.13

S

OHHO

OSO3 OH

OH OH

HO

OHOH

S

OHHO

OSO3 OH

OH OH

HO

OH

S

OHHO

OSO3

OH OH

HOS

OHHO

OSO3

OH

HO

Kotalanol (4)

Salacinol (2) Ponkoranol (3)Salaprinol (1)

S

OHHO

OH OH

OH OH

HO

OHOH

CH3OSO3

De-O-sulfonated kotalanol (5)

S

OHHO

OH

OH OH

HOHCO2

De-O-sulfonated salacinol (6)

86

Recently, Minami et al . 9 reported the isolation of a thiosugar sulfonium-

alkoxide inner salt (7), neosalacinol (Figure 3-2), from S. reticulata; however,

Tanabe et al.14 have shown that this compound is de-O-sulfonated salacinol (6).

Figure 3-2: Proposed structure of neosalacinol.

Comparison of the inhibitory activities of de-O-sulfonated salacinol (6) vs

salacinol (2) and de-O-sulfonated kotalanol (5) vs kotalanol (4) against rat

intestinal α-glucosidases (maltase, sucrase, and isomaltase) revealed that the

desulfonated analogues were either equivalent or better inhibitors than the parent

compounds.7,15,16 Furthermore, we have shown recently that de-O-sulfonated

kotalanol (5) (Ki = 0.03 ± 0.01 µM) is more potent an inhibitor of the N-terminal

catalytic domain of human intestinal maltase-glucoamylase (ntMGAM) than

kotalanol (4) itself (Ki = 0.19 ± 0.03 µM) (Table 3-1).17

S

OHHO

O

OH OH

HO

7

87

Table 3-1: Experimentally determined Ki values (nM)a of compounds 2-5, 8 and 9.

Inhibitor Ki (nM)

2

190 ± 2019

3

170 ± 3019

4

190 ± 3017

5

30 ± 1017

8

43 ± 1

9

15 ± 1

a Analysis of MGAM inhibition was performed using maltose as the substrate.

S

OHHO

OSO3 OH

OH OH

HO

OHOH

S

OHHO

OH OH

OH OH

HO

OHOH

CH3OSO3

S

OHHO

HO

OH

OH

OH

OH

OHCl

S

OHHO

HO

OH

OH

OH

OH

OHCl

S

OHHO

HO

OH

OH

OSO3

OH

OH

S

OHHO

HO

OH

OSO3

OH

88

In view of these findings, it was of interest to question whether de-O-

sulfonated ponkoranol 8 and its 5’-stereoisomer 9 (Figure 3-3) would be more

potent inhibitors than ponkoranol itself.

Figure 3-3: De-O-sulfonated ponkoranol and its 5’-stereoisomer.

Our previous work with kotalanol analogues had suggested that the

configuration at C-6’ was not critical for inhibitory activity.17,18 We report here an

efficient synthetic route to de-O-sulfonated ponkoranol 8 and its 5’-stereoisomer

9, and show that they are very potent inhibitors of the amino terminal catalytic

domain of human maltase-glucoamylase (ntMGAM).19

3.4 Results and discussion

The sulfonium-ions A could be synthesized by alkylation of an

appropriately protected 1,4-anhydro-4-thio-D-arabinitol B at the ring sulfur atom

with agent C. The desired stereochemistry at C-5' could be readily obtained by

choice of either D-glucose or D-mannose as starting material (Scheme 3-1).

S

OHHO

OH OH

OH OH

HO

OH

S

OHHO

OH OH

OH OH

HO

OH

Cl Cl

8 9

1'2'

3'4'

5'6'

1'2'

3'4'

5'6'

89

Scheme 3-1: Retrosynthetic analysis.

Initially, the S-alkylation of thioarabinitol 1220 with methyl 6-iodo-β-D-

glucopyranoside 1121 in CH3CN using AgBF4 at 65 °C was examined, based on

the procedure that has been reported for S-alkylation with simple alkyl halides

(Scheme 3-2).22 No product formation and decomposition of the starting material

were observed by TLC; the reaction in 1,1,1,3,3,3-hexafluoroisopropyl alcohol

(HFIP)23 as a solvent was also unsuccessful.

Scheme 3-2: First attempted synthesis of 8.

O

HOOH

OPL

S

OPPO

POS

OHHO

HO

OH

OH

OH

OH

L

L = leaving groupP = protecting group

+

A B C

R2 R1

R1R2

R1 = OH, R2 = H R1 = H, R2 = OH

8. R1 = OH, R2 = H, L = Cl9. R1 = H, R2 = OH, L = Cl

O

HOOH

OH

OMeI

AgBF4, CH3CN, 65oC

11

No reaction

PPh3, I2, imidazole,THF, reflux, 2 h

S

BnO OBn

BnO

12

O

HOOH

OH

OMeHO

10

90

In contrast, the coupling reaction with the p-toluenesulfonyl ester 1324 in

HFIP at 70 °C proceeded smoothly and yielded the sulfonium ion 14 (Scheme 3-

3). The benzyl groups of compound 14 were removed by treatment with boron

trichloride at -78 °C in CH2Cl2. However, attempts to hydrolyze the methyl

glycoside 15 with 2 M HCl were not successful, and decomposition of the product

was observed.

Scheme 3-3: Second attempted synthesis of 8.

Therefore, a benzyl glycoside was chosen as a protecting group at the

anomeric position to ensure its facile removal after the coupling reaction. Thus,

benzyl 6-O-p-toluenesulfonyl-α-D-gluco 17 or manno- pyranoside 20 were readily

prepared from D-glucose and D-mannose, respectively, according to literature

procedures.25-27 The thioether 12 was reacted with 17 in HFIP containing K2CO323

to give the protected sulfonium-ion 18 in 52% yield (Scheme 3-4). The benzyl

O

HOOH

OH

OMeTsO

HFIP, 70 oC,4 d, 45%

O

HOOH

OH

OMeS

BnO OBn

BnO

OTs

1314

15

S

BnO OBn

BnO

1. BCl3, CH2Cl2,, -78 oC

2. 2M HCl, -70oC, 20 h

12

O

HOOH

OH

OMeS

HO OH

HO

OTs

91

groups were then removed by treatment with boron trichloride at -78 °C in

CH2Cl2. During the course of deprotection, the p-toluenesulfonate counterion was

partially exchanged with chloride ion. Similar results were observed in previous

work from our laboratory.22 Hence, after removal of the benzyl groups, the

product was subsequently treated with Amberlyst A-26 resin (chloride form) to

completely exchange the p-toluenesulfonate counterion with chloride ion. Finally,

the crude product was reduced with NaBH4 to provide the desired de-O-

sulfonated ponkoranol 8 in 48% yield over three steps (Scheme 3-4).

Scheme 3-4: Synthesis of compound 8.

The other diastereomer was obtained similarly. Thus, compound 20 was

reacted with the thioether 12 to give the protected sulfonium-ion 21 in 47% yield

which was converted, as before, to the desired compound 9 in 41% yield over

three steps (Scheme 3-5).

1.BnOH, p-TSOH,70 oC, 4 h, 36%2.TsCl, Pyridine,-10 oC, 4 h, 60%

O

HOOH

OH

OBnTsO

HFIP, 70 oC, 4 d, 52%

O

HOOH

OH

OBnS

BnO OBn

BnO

1. BCl3, CH2Cl2,

-78 oC, 6 h

2. Amberlyst A-26,

H2O, 3 h

3. NaBH4, H2O, 3 h

S

OHHO

HO

OH

OH

OH

OH

OHOTs Cl

17

18 8

1' 1' 2'2'

3'

3'4'

4'

5'5'6'

6'

48%

S

BnO OBn

BnOO

HOOH

OH

OHHO

16

12

92

Scheme 3-5: Synthesis of compound 9.

The absolute stereochemistry at the stereogenic sulfur center in 18 and 21

was established by means of 1D-NOESY experiments (Figure 3-4) which showed

H-4 to H-6' correlations, implying that these atoms are syn-facial with respect to

the sulfonium salt ring.

Figure 3-4: 1D-NOESY correlations of selected protons in compounds 18 and 21.

1.BnOH, HCl,rt, 1 h, 35%2.TsCl, Pyridine,-10 oC, 15 h, 97%

O

HOOH

OH

OBnTsO

HFIP, 70 oC,4 d, 47%

O

HOOH

OH

OBnS

BnO OBn

BnOS

OHHO

HO

OH

OH

OH

OH

OHOTs Cl

20

21 9

1'

2'3'

3'

4'4'

5'5'6'

6'1' 2'

41%

S

BnO OBn

BnOO

HOOH

OH

OHHO

1. BCl3, CH2Cl2,

-78 oC, 6 h

2. Amberlyst A-26,

H2O, 3 h

3. NaBH4, H2O, 3 h

19

12

O

OH

OH

OHBnO

S

BnO OBn

18

H

H2C

4

BnO

O

OH

OH

OHBnO

S

BnO OBn

21

H

H2C

4

BnO

93

Finally, we comment on the inhibitory activities of compounds 8 and 9

against the N-terminus of recombinant human maltase-glucoamylase

(ntMGAM),19 a critical intestinal glucosidase for postamylase processing of

starch-derived oligosaccharides into glucose. The de-O-sulfonated ponkoranol 8

and its 5’-stereoisomer 9 inhibited ntMGAM with Ki values of 43 ± 3 and 15 ± 1

nM, respectively, both significantly lower than that (170 ± 30)19 for ponkoranol (3)

itself (Table 3-1). Thus, it would appear that de-O-sulfonation is beneficial. We

have attributed this fact previously to alleviation of steric compression of the

sulfate anion in a hydrophobic pocket within the active site of ntMGAM.17 The Ki

values for 8 and 9 compare to a Ki value for de-O-sulfonated kotalanol of 30 ± 1

nM.17 It would appear, therefore, that the configuration at C-5’ is not critical for

dictating enzyme inhibitory activity against ntMGAM and, furthermore, that

extension of the acyclic carbon chain beyond six carbons is not essential. We

note that 9 is the most potent compound to date in this class of molecules.

3.5 Experimental

3.5.1 General methods

Optical rotations were measured at 23 °C. 1H and 13C NMR spectra were

recorded at 600 and 150 MHz, respectively. All assignments were confirmed with

the aid of two-dimensional 1H, 1H (COSYDFTP) or 1H, 13C (INVBTP) experiments

using standard pulse programs. Column chromatography was performed with

Silica 60 (230-400 mesh). Reverse column chromatography was performed with

Silica C-18 cartridges. High resolution mass spectra were obtained by the

94

electrospray ionization method, using an Agilent 6210 TOF LC/MS high

resolution magnetic sector mass spectrometer.

3.5.2 Compound characterization data

3.5.2.1 Methyl 6-deoxy-6-[2,3,5-tri-O-benzyl-1,4-dideoxy-episulfoniummylidene-D-arabinitol]-α-D-glycopyranoside-p-toluenesulfonate (14)

The mixture of the thioether 1220 (270 mg, 0.64 mmol) and methyl 6-O-p-

toluenesulfonyl-β-D-glucopyranoside 1324 (200 mg, 0.57 mmol) in HFIP (1 mL)

were placed in a sealed tube in an oil bath at 70 oC for 4 days. The mixture was

cooled, then diluted with EtOAc, and evaporated to give a syrupy residue.

Purification by column chromatography (CHCl3/MeOH 10:1) gave the sulfonium

salt 14 as a syrup (150 mg, 45%). [α] 23D = + 19 (c = 1.4, MeOH). 1H NMR (MeOD)

δ 7.72-7.20 (19H, m, Ar), 4.69 (2H, m, H-1', H-3), 4.67-4.50 (6H, m, 3CH2-Ph),

4.47 (1H, br, H-2), 4.40 (1H, dd, J3,4 = 5.9, J4,5 = 10 Hz, H-4 ), 4.13 (1H, d, H-1a),

3.96 (2H, m, H-6'a , H-5'), 3.84 (2H, m, H-5a, H1b), 3.74 (2H, m, H-6'b, H-5b),

3.62 (1H, t, J2',3' = J3',4' = 8.9 Hz, H-3'), 3.39 (1H, dd, J1',2' = 3.7, J2',3' = 9.5 Hz, H-

2'), 3.36 (3H, s, OMe), 3.27 (1H, t, J4',5' = J3',4' = 8.9 Hz, H-4'), 2.35 (3H, s, Me).

13C NMR (MeOD) δ 142.3-125.6 (m, Ar), 100.5 (C-1'), 83.3 (C-3), 83.2 (C-2),

73.1, 72.0, 71.9 (3CH2-Ph), 73.0 (C-3'), 72.9 (C-4'), 71.7 (C-2'), 68.3 (C-5'), 66.9

(C-4), 66.4 (C-5), 55.1 (OMe), 48.6.0 (C-1), 48.2 (C-6'), 19.9 (Me). HRMS Calcd

for C33H41O8S (M+.): 599.2542. Found: 599.2564.

95

3.5.2.2 Attempted hydrolysis of the methyl glycoside (15)

To a solution of the sulfonium-ion 15 (150 mg, 0.274 mmol) in CH2Cl2 (5

mL) at -78 oC under N2 atmosphere was added BCl3 (1M, in CH2Cl2, 2.20 mmol).

The reaction mixture was stirred at the same temperature for 30 min, and then

allowed to warm to 5 oC for 6 h. The reaction was quenched by addition of MeOH

(5 mL), the solvents were removed, and the residue was co-evaporated with

MeOH (2 × 5 mL) to give a solid residue. The crude product was dissolved in 2 M

HCl (5 mL) and the solution was stirred at 70 °C for 20 h. The progress of the

reaction was monitored by 1H NMR spectroscopy which showed decomposition

of the starting material without hydrolysis of the methyl group.

3.5.2.3 Benzyl 6-deoxy-6-[2,3,5-tri-O-benzyl-1,4-dideoxy-episulfoniummylidene-D-arabinitol]-α-D-glycopyranoside-p-toluenesulfonate (18)

Benzyl 6-O-p-toluene-sulfonyl-β-D-glucopyranoside 17 25,27 (470 mg, 1.11

mmol) and the thioether 12 (558 mg, 1.33 mmol) were dissolved in HFIP (1.5

mL), containing anhydrous K2CO3 (10 mg). The mixture was stirred in a sealed

reaction vessel in an oil bath at 70 oC for 4 days. The mixture was cooled, then

diluted with EtOAc, and evaporated to give a syrupy residue. Purification by

column chromatograghy (EtOAc/MeOH 92:8) gave the sulfonium salt 18 as a

syrup (388 mg, 52%). [α] 23D = + 16 (c = 0.8, MeOH). 1H NMR (MeOD) δ 7.67-7.17

(24H, m, Ar), 4.87 (1H, d, J1',2' = 3.6 Hz, H-1'), 4.63 (1H, m, H-3), 4.64-4.46 (8H,

m, 4CH2-Ph), 4.41 (1H, br, H-2), 4.28 (1H, dd, J3,4 = 5.7 J4,5 = 9.4 Hz, H-4 ), 3.93

(2H, m, H-1a, H-5'), 3.81 (1H, dd, J6'a,5' = 3.1, J6'a,6'b = 13.2 Hz, H-6'a), 3.77 (1H,

dd, J5a,4 = 6.0, J5a,5b = 10.5 Hz, H-5a), 3.72 (1H, dd, J1,2 = 3.6, J1a,1b = 13.3 Hz,

96

H-1b), 3.67-3.62 (3H, m, H-5b, H-6b, H-3'), 3.35 (1H, dd, J1',2' = 3.6, J2',3' = 9.8

Hz, H-2') 3.21 (1H, t, J4',5' = J3',4' = 8.9 Hz, H-4'), 2.32 (3H, s, Me). 13C NMR

(MeOD) δ 142.2-125.6 (m, Ar), 99.3 (C-1'), 83.2 (C-3), 83.0 (C-2), 73.1, 72.0,

71.9, 70.7 (4CH2-Ph), 73.0 (C-4'), 72.9 (C-3'), 71.7 (C-2'), 68.8 (C-5'), 66.9 (C-4),

66.5 (C-5), 49.0 (C-1), 48.2 (C-6'), 19.9 (Me). HRMS Calcd for C39H45O8S (M+.):

673.2830. Found: 673.2831.

3.5.2.4 1,4-Dideoxy-1,4-[[2S, 3S, 4R, 5S]-2,3,4,5,6-pentahydroxy-hexyl]-(R)-epi-sulfoniumylidine]-D-arabinitol chloride (8)

Compound 18 (300 mg, 0.36 mmol) was dissolved in CH2Cl2 (25 mL), the

mixture was cooled to -78 oC, and BCl3 (1M solution in CH2Cl2, 3.56 mmol) was

added under N2. The reaction mixture was stirred at the same temperature for 30

min, and then allowed to warm to 5 oC for 6 h. The reaction was quenched by

addition of MeOH (5 mL), the solvents were removed, and the residue was co-

evaporated with MeOH (2 × 5 mL). The crude residue was dissolved in H2O (10

mL), Amberlyst A-26 resin (200 mg) was added, and the reaction mixture was

stirred at room temprature for 3 h. Filtration through cotton, followed by solvent

removal gave the crude hemiacetal. The crude product was dissolved in water (8

mL), and the solution was stirred at room temperature while NaBH4 (67 mg, 1.78

mmol) was added in small portions over 30 min. Stirring was continued for

another 3 h and the mixture was acidified to pH < 4 by dropwise addition of 2 M

HCl. The mixture was evaporated to dryness and the residue was co-evaporated

with anhydrous MeOH (3 × 30 mL). Treatment of the solid residue with 50%

EtOAc:MeOH (5-10 mL) resulted in precipitation of most of the borate salt.

97

Filtration through cotton, followed by solvent removal gave the crude compound.

The residue was purified by reverse phase column chromatography (MeOH/H2O

(2:100)) to give 8 as a colorless solid (60 mg, 48%). [α] 23D = + 4o, (c = 0.5 , H2O).

1H NMR (D2O) δ 4.64 (1H, m, H-2), 4.35 (1H, t, br, H-3), 4.14 (1H, td, J1',2' = 9.1,

J2',3' = 3.0 Hz, H-2'), 4.02 (2H, m, H-5a, H-4), 3.87-3.77 (4H, m, H-5b, H-1'a, H-

1a, H-1b), 3.72-3.66 (3H, m, H-4', H-5', H-1'b), 3.62 (2H, m, H-6'a, H-3'), 3.50

(1H, dd, J6'a,6,b = 11.7, J5',6'b = 5.6 Hz, H-6'b). 13C NMR (D2O) δ 77.5 (C-3), 76.9

(C-2), 73.1 (C-3'), 72.7 (C-5'), 70.0 (C-4), 69.3 (C-4'), 67.4 (C-2'), 62.2 (C-6'),

59.2 (C-5), 50.0 (C-1'), 48.2 (C-1). HRMS Calcd for C11H23O8S (M+.): 315.1108.

Found: 315.1117.

3.5.2.5 Benzyl 6-deoxy-6-[2,3,5-tri-O-benzyl-1,4-dideoxy-episulfoniummylidene-D-arabinitol]-α-D-mannopyranoside-p-toluenesulfonate (21)

Reaction of the thioether 12 (590 mg, 1.41 mmol) with benzyl 6-O-p-

toluenesulfonyl-β-D-mannopyranoside 2026 (500 mg, 1.18 mmol) in HFIP (1.5

mL), containing anhydrous K2CO3 (10 mg) at 70 oC for 4 days gave the sulfonium

salt 21 as a foam (370 mg, 47%) after purification by column chromatography

(EtOAc/MeOH (92:8)). [α] 23D = +8o, (c = 0.5, MeOH). 1H NMR (MeOD) δ 7.73-7.23

(24H, m, Ar), 4.87 (1H, m, H-1'), 4.70 (1H, m, H-2), 4.69-4.52 (8H, m, 4CH2-Ph),

4.49 (1H, m, H-3), 4.31 (1H, t, J3,4 = J4,5 = 9.6 Hz, H-4 ), 4.04 (1H, d, br, J1,2 =

13.1 Hz, H-1a), 3.94 (2H, m, H-6'a, H-4'), 3.90-3.85 (3H, m, H-2', H-1b, H-5a),

3.79-3.73 (3H, m, H-6'b, H-5b, H-3'), 3.61 (1H, t, J4',5' = J5',6' = 9.3 Hz, H-5'), 2.38

(3H, s, Me). 13C NMR (MeOD) δ 142.2-125.6 (m, Ar), 100.5 (C-1'), 83.3 (C-2),

82.9 (C-3), 73.1, 72.0, 71.8, 70.1(4CH2-Ph), 70.6 (C-2'), 70.4 (C-3'), 66.7 (C-4),

98

66.5 (C-5), 48.8 (C-1), 48.2 (C-6'), 19.9 (Me). HRMS Calcd for C39H45O8S (M+.):

673.2830. Found: 673.2828.

3.5.2.6 1,4-Dideoxy-1,4-[[2S, 3S, 4R, 5R]-2,3,4,5,6-pentahydroxy-hexyl]-(R)-epi-sulfoniumylidine]-D-arabinitol chloride (9)

Compound 9 was obtained as a colorless solid (51 mg, 41%) from 21 (300

mg, 0.36 mmol) using the same procedure that was used to obtain 8. [α] 23D = +

11o, (c = 0.3 , H2O). 1H NMR (D2O) δ 4.65 (1H, d, br, H-2), 4.35 (1H, t, br, H-3),

4.12 (1H, td, J1',2' = 9.1, J2',3' = 3.0 Hz, H-2'), 4.02 (2H, m, H-5a, H-4), 3.89-3.60

(9H, m, H-1'a, H-5b, H-1a, H-1b, H-3', H-6'a, H-1'b, H-4', H-5'), 3.55 (1H, dd,

J6'a,6,b = 11.7, J5',6'b = 5.8 Hz, H-6'b). 13C NMR (D2O) δ 77.5 (C-3), 76.9 (C-2),

71.5 (C-3'), 70.5 (C-5'), 70.0 (C-4), 68.8 (C-4'), 67.3 (C-2'), 63.0 (C-6'), 59.2 (C-

5), 50.4 (C-1'), 48.1 (C-1). HRMS Calcd for C11H23O8S (M+.): 315.1108. Found:

315.1122.

3.6 Acknowledgments

We are grateful to the Canadian Institutes for Health Research

(FRN79400) and the Heart and Stroke Foundation of Ontario (NA-6305) for

financial support.

3.7 References

1. Harvey, A. L. Natural products in drug discovery. Drug Discov. Today

2008, 13, 894-901.

2. Butler, M. S. Natural products to drugs: natural product-derived

compounds in clinical trials. Nat. Prod. Rep. 2008, 25, 475-516.

99

3. Chandrasena, J. P. C. The Chemistry and Pharmacology of Ceylon and

Indian Medicinal Plants; H&C Press: Colombo, Sri Lanka, 1935.

4. Jayaweera, D. M. A. Medicinal Plants Used in Ceylon-Part 1: National

Science Council of Sri Lanka: Colombo, 1981.

5. Matsuda, H.; Li, Y. H.; Murakami, T.; Matsumura, N.; Yamahara, J.;

Yoshikawa, M. Chem. Pharm. Bull. 1998, 46, 1399-1403.

6. Yoshikawa, M.; Murakami, T.; Shimada, H.; Matsuda, H.; Yamahara, J.;

Tanabe, G.; Muraoka, O. Tetrahedron Lett. 1997, 38, 8367-8370.

7. Yoshikawa, M.; Xu, F. M.; Nakamura, S.; Wang, T.; Matsuda, H.; Tanabe,

G.; Muraoka, O. Heterocycles 2008, 75, 1397-1405.

8. Muraoka, O.; Xie, W. J.; Tanabe, G.; Amer, M. F. A.; Minematsu, T.;

Yoshikawa, M. Tetrahedron Lett. 2008, 49, 7315-7317.

9. Minami, Y.; Kurlyarna, C.; Ikeda, K.; Kato, A.; Takebayashi, K.; Adachi, I.;

Fleet, G. W. J.; Kettawan, A.; Karnoto, T.; Asano, N. Bioorg. Med. Chem. 2008, 16, 2734-2740.

10. Jayakanthan, K.; Mohan, S.; Pinto, B. M. J. Am. Chem. Soc. 2009, 131,

5621-5626.

11. Mohan, S.; Pinto, B. M. Carbohydr. Res. 2007, 342, 1551-1580.

12. Mohan, S.; Pinto, B. M. Collect. Czech. Chem. Commun. 2009, 74, 1117-

1136.

13. Johnston, B. D.; Jensen, H. H.; Pinto, B. M. J. Org. Chem. 2006, 71, 1111-

1118.

14. Tanabe, G.; Xie, W. J.; Ogawa, A.; Cao, C. N.; Minematsu, T.; Yoshikawa,

M.; Muraoka, O. Bioorg. Med. Chem. Lett. 2009, 19, 2195-2198.

15. Matsuda, H.; Yoshikawa, M.; Murakami, T.; Tanabe, G.; Muraoka, O. J.

Trad. Med. 2005, 22, 145-153.

100

16. Ozaki, S.; Oae, H.; Kitamura, S. J. Nat. Prod. 2008, 71, 981-984.

17. Sim, L.; Jayakanthan, K.; Mohan, S.; Nasi, R.; Johnston, B. D.; Pinto, B.

M.; Rose, D. R. Biochemistry 2010, 49, 443-451.

18. Nasi, R.; Patrick, B. O.; Sim, L.; Rose, D. R.; Pinto, B. M. J. Org. Chem.

2008, 73, 6172-6181.

19. Rossi, E. J.; Sim, L.; Kuntz, D. A.; Hahn, D.; Johnston, B. D.; Ghavami, A.;

Szczepina, M. G.; Kumar, N. S.; Sterchi, E. E.; Nichols, B. L.; Pinto, B. M.;

Rose, D. R. FEBS J. 2006, 273, 2673-2683.

20. Satoh, H.; Yoshimura, Y.; Sakata, S.; Miura, S.; Machida, H.; Matsuda, A.

Bioorg. Med. Chem. Lett. 1998, 8, 989-992.

21. Skaanderup, P. R.; Poulsen, C. S.; Hyldtoft, L.; Jorgensen, M. R.;

Madsen, R. Synthesis 2002, 1721-1727.

22. Sankar, M.; Sim, L.; David, R. R.; Pinto, B. M. Carbohydr. Res. 2007, 342,

901-912.

23. Ghavami, A.; Sadalapure, K. S.; Johnston, B. D.; Lobera, M.; Snider, B.

B.; Pinto, B. M. Synlett 2003, 1259-1262

24. Pellowska-Januszek, L.; Dmochowska, B.; Skorupa, E.; Chojnacki, J.;

Wojnowski, W.; Wiśniewski, A. P. Carbohydr. Res. 2004, 339, 1537-1544.

25. Andreana, P. R.; Sanders, T.; Janczuk, A . ; Warrick, J. I.; Wang, P. G.

Tetrahedron Lett. 2002, 43, 6525-6528.

26. Branchaud, B. P.; Meier, M. S. J. Org. Chem. 1989, 54, 1320-1326.

27. Koto, S.; Inada, S.; Yoshida, T.; Toyama, M.; Zen, S. Can. J. Chem. 1981,

59, 255-259.

101

3.8 Supporting Information

Potent glucosidase inhibitors: de-O-sulfonated ponkoranol and its stereoisomer

Razieh Eskandari, Douglas A. Kuntz, David R. Rose, B. Mario Pinto

102

ppm (t1)3.04.05.06.07.0

3.33

1.00

0.99

2.35

2.02

2.08

1.31

1.03

1.22

2.83

4.27

2.20

0.97

4.49

13.7

7

3.93

O

HOOH

OH

OMeS

BnO OBn

BnO

OTs

14600 MHz, MeOD

ppm (t1)50100

O

HOOH

OH

OMeS

BnO OBn

BnO

OTs

14150 MHz, MeOD

103

ppm (t1)3.04.05.06.07.01.

071.

04

3.03

1.00

1.95

23.1

3

5.15

1.20

1.68

2.20

1.02

1.01

2.05

1.20

3.16

0.98

1.17

O

HOOH

OH

OBnS

BnO OBn

BnO

OTs

18

1'

2'3'4'5'6'

600 MHz, MeOD

ppm (t1)4.004.50

H-6’a H-4

H-6’b

O

HOOH

OH

OBnS

BnO OBn

BnO

OTs

18

1'

2'3'4'5'6'

150 MHz, MeOD

104

ppm (t1)3.504.004.50

2.22

3.66

2.46

1.57

1.00

1.09

4.54

ppm (t1)50.055.060.065.070.075.0

S

OHHO

HO

OH

OH

OH

OH

OHCl

8

1' 2'3'

4'5'

6'

150 MHz, D2O

S

OHHO

HO

OH

OH

OH

OH

OHCl

8

1' 2'3'

4'5'

6'

600 MHz, D2O

105

ppm (t1)3.04.05.06.07.0

3.00

2.21

1.20

3.89

4.19

2.60

1.24

1.20

5.40

3.37

4.98

25.0

6

O

HOOH

OH

OBnS

BnO OBn

BnO

OTs

21

1'2'3'4'

5'6'

600 MHz, MeOD

ppm (t1)50100

O

HOOH

OH

OBnS

BnO OBn

BnO

OTs

21

1'2'3'4'

5'6'

150 MHz, MeOD

106

ppm (t1)3.504.004.50

0.83

0.88

1.00

1.80

1.23

9.03

S

OHHO

HO

OH

OH

OH

OH

OHCl

9

3'4'

5'6'1' 2'

600 MHz, D2O

ppm (t1)4.004.50

H-4H-6’a

H-6’b

O

HOOH

OH

OBnS

BnO OBn

BnO

OTs

21

1'2'3'4'

5'6'

500 MHz, MeOD

107

ppm (t1)45.050.055.060.065.070.075.080.0

S

OHHO

HO

OH

OH

OH

OH

OHCl

9

3'4'

5'6'1' 2'

150 MHz, D2O

108

CHAPTER 4: THE EFFECT OF HETEROATOM SUBSTITUTION OF SULFUR FOR SELENIUM IN GLUCOSIDASE INHIBITORS ON INTESTINAL α-

GLUCOSIDASE ACTIVITIES

This Chapter comprises the manuscript “The effect of heteroatom

substitution of sulfur for selenium in glucosidase inhibitors on intestinal α-

glucosidase activities” which was published in Journal of Chemical Society

Chemical Communications (2011, 47, 9134-9136).

Razieh Eskandari,a Kyra Jones, b David R. Rose,b B. Mario Pintoa

aDepartment of Chemistry, Simon Fraser University, Burnaby, British Columbia,

Canada V5A 1S6

b Department of Biology, University of Waterloo, Waterloo, Ontario,

Canada N2L 3G1

109

After we had successfully synthesized de-O-sulfonated ponkoranol and its

5'-epimer, we turned our attention to the structure-activity relationship (SAR)

studies of ponkoranol. In this Chapter, we describe the effect of selenium

substitution on the inhibitory activities of de-O-sulfonated ponkoranol and its 5'-

epimer. Thus, the selenium analogues of de-O-sulfonated ponkoranol were

synthesized and their glucosidase inhibitory activities against four human

intestinal glucosidase enzymes: maltase-glucoamylase MGAM (ntMGAM,

ctMGAM) and sucrase-isomaltase (ntSI, ctSI) were studied and compared with

the activities of the previously synthesized (Chapter 3) de-O-sulfonated

ponkoranol and its 5'-epimer. The thesis author performed all the experimental

synthetic work and the characterization of the compounds, and wrote and edited

the manuscript with Dr. B. Mario. Pinto. Mrs. Kyra Jones and Dr. David R. Rose

performed the enzyme inhibition studies.

Graphical abstract:

O

HOOH

OBnTsO Se

OHHO

HO

OH

OH

OH

OH

ClSe

OBnBnO

BnO+

R1 = OH, R2 = HR1 = H, R2 = OH

R2R1

7. R1 = OH, R2 = H8. R1 = H, R2 = OH

R2 R1

110

4.1 Keywords

Selenium analogue, maltase-glucoamylase, sucrase-isomaltase, de-O-

sulfonated ponkoranol.

4.2 Abstract

The synthesis of selenium analogues of de-O-sulfonated ponkoranol, a

naturally occuring sulfonium-ion glucosidase inhibitor isolated from Salacia

reticulata, and their evaluation as glucosidase inhibitors against two recombinant

intestinal enzymes maltase-glucoamylase (MGAM) and sucrase-isomaltase (SI)

are described.

4.3 Introduction

The roots and stems extract of Salacia reticulata, a large woody climbing

plant found in Sri Lanka and southern India, have traditionally been used for

treatment of type-2 diabetes.1 Human clinical trials on patients with type-2

diabetes, conducted with extract of Salacia reticulata, have shown effective

treatment of type-2 diabetes with minimal side effects.2 Our efforts in recent years

have focused on this novel class of α-glucosidase inhibitors comprising

sulfonium-ions as putative mimics of the oxacarbenium ion intermediates in

glucosidase-mediated hydrolysis reactions.3 Thus, we have described the

synthesis and stereochemical structure elucidation of the naturally-occurring

glucosidase inhibitors salacinol (1),4 ponkoranol (2),5 kotalanol (3),6 de-O-

sulfonated kotalanol (4)6 (Figure 4-1) from the plant extract of Salacia reticulata.

111

Interestingly, one of our synthetic compounds, de-O-sulfonated ponkoranol (5),7

(Figure 4-1), was isolated recently from the same plant.8

Figure 4-1: Components isolated from Salacia species.

The antidiabetic property of these herbal extract can be attributed, at least

in part, to the inhibition of intestinal α-glucosidases by these sulfonium-ion

components.9-11 Crystallographic studies of the N-terminal domain of human

intestinal maltase-glucoamylase (ntMGAM), a family 31 glycoside hydrolase

(GH31),12 with several candidate inhibitors suggested that the permanent positive

charge of these entities may mimic the oxacarbenium-ion like transition state of

the glucosidase-mediated hydrolysis reactions.13 We also suggested that the ring

conformation adopted by the thiocyclitol in kotalanol (3), a 3T2 conformation,

S

OHHO

OSO3 OH

OH OH

HO

OHOH

S

OHHO

OSO3 OH

OH OH

HO

OH

S

OHHO

OSO3

OH OH

HO

Kotalanol (3)Salacinol (1) Ponkoranol (2)

S

OHHO

OH OH

OH OH

HO

OHOH

CH3OSO3

De-O-sulfonated kotalanol (4) De-O-sulfonated ponkoranoll (5)

S

OHHO

OH OH

OH OH

HO

OH

Cl

112

closely resembles the proposed 4H3 conformation of the oxacarbenium-ion like

transition state.13

We have also reported the synthesis of the 5′-stereoisomer of de-O-

sulfonated ponkoranol 6 (Figure 4-2), and have shown that it is a potent (Ki = 15

nM) inhibitor of ntMGAM.7

Figure 4-2: Structure of the 5′-stereoisomer of de-O-sulfonated ponkoranol.

As part of our continuing interest in evaluating the effect of heteroatom

substitution in the sugar ring on glucosidase inhibitory activity, we have also

synthesized selenium congeners of several candidates as potential glucosidase

inhibitors. 14-16

Comparison of the proposed intermediate in the glucosidase-catalyzed

reaction and the selenonium ion, inferred from the X-ray crystal structures of

complexes of kotalanol, miglitol, and acarbose with ntMGAM,13 shows that the

selenonium ion in a 3T2 conformation, superimposes well on the 4H3

conformation of the oxacarbenium ion (Figure 4-3).

S

OHHO

OH OH

OH OH

HO

OH

Cl

6

113

Figure 4-3: Superimposition of the ring carbon atoms of the proposed intermediate in glucosidase-catalyzed reactions (in green) and the selenonium ion (in blue).

We now report the synthesis and enzyme inhibitory activity of analogues

of 5 and 6, in which the sulfur atom has been replaced by the heavier cognate

atom selenium to give 7 and 8, respectively (Figure 4-4).

Figure 4-4: Selenium analogues of de-O-sulfonated ponkoranol and its 5’-stereoisomer.

4.4 Results and discussion

Retrosynthetic analysis indicated that 7 or its analogue 8 could be

obtained by alkylation of an appropriately protected 1,4-anhydro-4-seleno-D-

arabinitol B at the ring heteroatom with agent C (Scheme 4-1).7

Se

OHHO

OH OH

OH OH

HO

OH

Se

OHHO

OH OH

OH OH

HO

OH

Cl Cl

7 8

1'2'

3'4'

5'6'

1'2'

3'4'

5' 6'

OHO

OH

OH

HO

4H3

1 2

3

4

5

O

HO

OH

OH

HO 123

4

5

SeHO OH

HO

3T2

1

2

3

4

Half chair

Twist

RSe

OHHO

HO

R

5

66

1

23

45 C3

C1 C2

C3

C4

C1

C4

C6

C5

C5 C2

114

Scheme 4-1: Retrosynthetic analysis.

The required benzyl (Bn)-protected D-selenoarabinitol (10)14 and benzyl 6-

O-p-toluenesulfonyl-α-D-gluco 9 or manno pyranoside 127 were obtained by

literature methods. The selenonium salts 11 and 13 were synthesized by

alkylation of 10 with 9 or 12 (1.2 equiv) in hexafluoroisopropanol (HFIP),

containing K2CO3, to give 11 and 13 in 55% and 45%, yield, respectively

(Scheme 4-2). In the case of compound 11, just one isomer was obtained at the

stereogenic selenium atom, whereas in the case of compound 13, a mixture of

isomers was obtained; 1H NMR spectroscopy showed the presence of two

isomers in a ratio of 10:1. The benzyl protecting groups were removed with boron

trichloride at -78 °C to give the desired sulfonium salts. During the course of

deprotection, some of the tosylate counterion was exchanged with the chloride

ion. Hence, the deprotected sulfonium salts were treated with Amberlyst A-26

(chloride form) to completely exchange the tosylate counterion with chloride ion.7

Finally, the products were reduced with NaBH4 to provide the desired selenium

analogues of de-O-sulfonated ponkoranol 7 and its 5’ epimer 8 in 52% and 45%

yield, respectively, over three steps (Scheme 4-2).

O

HOOH

OPL

Se

OPPO

POSe

OHHO

HO

OH

OH

OH

OH

L

L = leaving groupP = protecting group

+

A B C

R2 R1

R1R2

R1 = OH, R2 = H R1 = H, R2 = OH

7. R1 = OH, R2 = H, L = Cl8. R1 = H, R2 = OH, L = Cl

115

Scheme 4-2: Synthesis of compounds 7 and 8.

The stereochemistry at the selenium centre for the major isomer of 13 was

confirmed with the aid of a 2D-NOESY experiment, which showed a correlation

between H-4 and H-6′a, thus indicating that these atoms are syn-facial with

respect to the sulfonium salt ring (Figure 4-5). In the case of 11, the absolute

O

HOOH

OH

OBnTsO

HFIP, 70 oC, 4 d, 55%

O

HOOH

OH

OBnSe

BnO OBn

BnO

1. BCl3, CH2Cl2,

-78 oC, 6 h

2. Amberlyst A-26,

H2O, 3 h

3. NaBH4, H2O, 3 h

Se

OHHO

HO

OH

OH

OH

OH

OH

OTs

Cl

9 11

7

1'

1' 2'

2'

3'

3'

4'

4'

5'

5'

6'

6'

52%

Se

BnO OBn

BnO

10

O

HOOH

OH

OBnTsO

HFIP, 70 oC,4 d, 45%

O

HOOH

OH

OBnSe

BnO OBn

BnO

Se

OHHO

HO

OH

OH

OH

OH

OH

OTs

Cl

12 13

8

1'

2'3'

3'

4'

4'

5'

5'

6'

6'1' 2'

45%

Se

BnO OBn

BnO

1. BCl3, CH2Cl2,

-78 oC, 6 h

2. Amberlyst A-26,

H2O, 3 h

3. NaBH4, H2O, 3 h

10

d.r, ~ 10:1

d.r, ~ 10:1

116

configuration at the selenium centre was assigned by analogy since a NOESY

experiment was not possible owing to overlapping signals for H-4 and H-6′. This

assignment is consistent with our previous work.3

Figure 4-5: 2D-NOESY correlations of selected protons in compound 13.

Finally, we comment on the inhibitory activities of the compounds

synthesized in this study and previous study7 against the two recombinant

intestinal enzymes, maltase-glucoamylase (MGAM) and sucrase-isomaltase (SI),

critical for postamylase processing of starch-derived oligosaccharides into

glucose. Each of these enzymes exhibits two separate catalytic units (by gene

duplication) at their C- and N-terminal domains, resulting in ntMGAM, ctMGAM,

ntSI, and ctSI enzyme activities.17-22 There are also various alternative splicing

patterns of ctMGAM in mammals; two spliceforms from mice are discussed in this

paper: ctMGAM-N2 and ctMGAM-N20.23,24 The experimentally determined

inhibition constants (Kis) for the competitive inhibitors de-O-sulfonated

ponkoranol (5) and its synthetic analogues (6-8) are listed in Table 4-1.

O

OH

OH

OHBnO

Se

BnO OBn

13

H

H2C

4

BnO

117

Table 4-1: Experimentally determined Ki values (µM)a of compounds 5-8.

a Analysis of inhibition was performed using maltose as the substrate.

Comparison of the data in Table 4-1 indicates that substitution of the ring

sulfur atom in de-O-sulfonated ponkoranol (5) and its 5’ epimer (6) by selenium

results in a 13-fold and 23-fold improvement, respectively, in inhibition of ntSI.

NtSI demonstrates the broadest substrate specificity of all four subunits and

hydrolyses both α (1-4) and α (1-6) linkages.25 Similarly, substitution of sulfur by

selenium leads to an increase in inhibitory activity against ctSI (cf. 5 and 6 vs. 7

and 8). In contrast, little discrimination is observed between the congeners for

inhibition of ntMGAM or ctMGAM-N20. Interestingly, compounds 5-8 differentiate

between the spliceforms of ctMGAM, and none shows inhibition of ctMGAM-N2;

the reason for this selectivty must await structural analysis.

ctMGAM-N2 ctMGAM-N20 ntMGAM ctSI ntSI

5 No Inhibition 0.096 ± 0.015 0.043 ± 0.0017 0.103 ± 0.037 0.302 ± 0.123

6 No Inhibition 0.138 ± 0.068 0.015 ± 0.0017 0.132 ± 0.047 0.138 ± 0.012

7 No Inhibition 0.047 ± 0.014 0.038 ± 0.008 0.018 ± 0.004 0.013 ± 0.008

8 No Inhibition 0.041 ±0.027 0.025 ± 0.014 0.019 ±0.010 0.010 ± 0.002

118

4.5 Experimental

4.5.1 General methods

Optical rotations were measured at 23 °C. 1H and 13C NMR spectra were

recorded at 600 and 150 MHz, respectively. All assignments were confirmed with

the aid of two-dimensional 1H, 1H (COSYDFTP) or 1H, 13C (INVBTP) experiments

using standard pulse programs. Column chromatography was performed with

Silica 60 (230-400 mesh). High resolution mass spectra were obtained by the

electrospray ionization method, using an Agilent 6210 TOF LC/MS high

resolution magnetic sector mass spectrometer.

4.5.2 Enzyme kinetics

Kinetic parameters were determined by measuring the amount of glucose

produced upon the addition of enzyme at increasing maltose concentrations

(from 0.5 to 30 mM) in the presence of increasing inhibitor concentration (0-200

nM) by a two-step glucose oxidase assay in a 96-well plate. The enzyme was

allowed to act on the maltose substrate in the presence of inhibitor for 45 minutes

at 37°C. The reactions were then quenched with Tris-HCl to a final concentration

of 1M. Glucose oxidase reagent (Sigma-Aldrich) was then added to each well

(125µl) and the reactions were allowed to develop for 30 minutes at 37°C.

Reactions were performed in quadruplicate and absorbance measured at 405 nM

by a SpectraMax 190 Plate Reader (Molecular Devices). Absorbance readings

were averaged to give the final value, which was compared to a glucose standard

curve to determine the amount of glucose released by the enzyme from the

substrate. The program KaleidaGraph 4.1 was used to fit the data to the

119

Michaelis-Menten equation and estimate Km, Kmobs (Km in the presence of the

inhibitor) and Vmax of the catalytic subunits. Ki values for each inhibitor were

determined using the equation Ki = [I]/(Kmobs/Km)-1). The Ki values reported for

each inhibitor were determined by averaging the Ki values from three different

inhibitor concentrations. The weight of compounds 8 was adjusted for the

presence of major isomer. The data was also plotted on Lineweaver-Burk plots to

verify that the inhibitors were acting as competitive inhibitors. The methods used

for kinetic assays were reported previously.25

4.5.3 Compound characterization data

4.5.3.1 Benzyl 6-deoxy-6-[2,3,5-tri-O-benzyl-1,4-dideoxy-(R)-epi-seleniumylidene-D-arabinitol]-β-D-glycopyranoside-p-toluenesulfonate (11).

Benzyl 6-O-p-toluenesulfonyl-β-D-glucopyranoside 97 (310 mg, 0.73 mmol)

and 1,4-dideoxy-2,3,5-tri-O-benzyl-1,4-anhydro-4-seleno-D-arabinitol 1014 (409

mg, 0.88 mmol) were dissolved in HFIP (1.5 mL), containing anhydrous K2CO3

(10 mg). The mixture was stirred in a sealed reaction vessel in an oil bath at 65-

70 oC for 4 days. The mixture was cooled, then diluted with EtOAc, and

evaporated to give a syrupy residue. Purification by column chromatography

(CHCl3/MeOH 95:5) gave the sulfonium salt 11 as a white amorphous solid (358

mg, 55%). [α] 23D = + 22o. 1H NMR (MeOD) δ 7.74-7.24 (24H, m, Ar), 4.78 (1H,

dd, J1,2 = 7.8, J2,3 = 4.2 Hz, H-2), 4.68-4.49 (8H, m, 4CH2-Ph), 4.55 (1H, m, H-3),

4.48 (1H, m, H-4), 4.41 (1H, d, J1′,2′ = 7.8 Hz, H-1′), 3.98 (1H, m, H-6′a), 3.91 (1H,

m, H-1a), 3.84 (1H, dd, J5a,4 = 6.3, J5a,5b = 10.3 Hz, H-5a), 3.75 (3H, m, H-6′b, H-

3′, H-5b), 3.70 (1H, dd, J1,2 = 2.8, J1a,1b = 12.5 Hz, H-1b), 3.40 (1H, t, J4′,5′ =

120

J6′,5′ = 9.1 Hz, H-5′), 3.28 (2H, m, H-2′, H-4′), 2.38 (3H, s, Me). 13C NMR (MeOD)

δ 142.2-125.5 (m, Ar), 103.0 (C-1′), 83.8 (C-2), 83.5 (C-3), 75.8 (C-5′), 73.4 (C-

4′), 73.3 (C-2′), 73.1, 72.0, 71.7, 71.3 (4CH2-Ph), 71.8 (C-3′), 66.5 (C-5), 66.2 (C-

4), 45.5 (C-1), 45.2 (C-6′), 19.8 (Me). HRMS Calcd for C39H45O8Se (M+.):

721.2278. Found: 721.2279.

4.5.3.2 1,4-Dideoxy-1,4-[[2S, 3S, 4R, 5S]-2,3,4,5,6-pentahydroxyhexyl]-(R/S)-epi-seleniumylidine]-D-arabinitol chloride (7).

Compound 11 (200 mg, 0.22 mmol) was dissolved in CH2Cl2 (20 mL), the

mixture was cooled to -78 oC, and BCl3 (1M solution in CH2Cl2, 3.6 mmol) was

added under N2. The reaction mixture was stirred at the same temperature for 30

min, and then allowed to warm to -5 oC for 6 h. The reaction was cooled to -78 oC

and quenched by addition of MeOH (5 mL), the solvents were removed, and the

residue was co-evaporated with MeOH (2 × 5 mL). The crude residue was

dissolved in H2O (10 mL), Amberlyst A-26 resin (200 mg) was added, and the

reaction mixture was stirred at room temperature for 3 h. Filtration through

cotton, followed by solvent removal gave the crude hemiacetal. The crude

product was dissolved in water (8 mL), and the solution was stirred at room

temperature while NaBH4 (34 mg, 0.9 mmol) was added in small portions over 30

min. Stirring was continued for another 3 h and the mixture was acidified to pH <

4 by dropwise addition of 2M HCl. The mixture was evaporated to dryness and

the residue was co-evaporated with anhydrous MeOH (3 × 30 mL). Treatment of

the solid residue with 50% EtOAc:MeOH (5-10 mL) resulted in precipitation of

most of the borate salt. Filtration through cotton, followed by solvent removal

121

gave the crude compound. The residue was purified by crystallization with

minimum amount of MeOH to give 7 as a colorless solid (46 mg, 52%).

[α] 23D = + 12.8 o, (c = 0.5 , H2O). 1H NMR (D2O) δ 4.76 (1H, dd, J1,2 = J2,3

= 3.7 Hz, H-2), 4.46 (1H, t, br, H-3), 4.20 (1H, m, H-2′), 4.15 (1H, m, H-4), 4.05

(1H, dd, J4,5a = 12.2, J5a,5b = 4.8 Hz, 5a) 3.92-3.87 (2H, m, H-5b, H-1′a), 3.78-

3.74 (5H, m, H-5′, H-4′, H-1′b, H-1a, H-1b), 3.69-3.65 (2H, m, H-3′, H-6′a ), 3.55

(1H, dd, J6′a,6,b = 10.7, J5′,6′b = 5.9 Hz, H-6′b). 13C NMR (D2O) δ 77.8 (C-3), 77.2

(C-2), 73.2 (C-3′), 72.2 (C-5′), 69.1 (C-4), 69.0 (C-4′), 67.0 (C-2′), 61.8(C-6′),

558.9 (C-5), 47.0 (C-1′), 44.7 (C-1). HRMS Calcd for C11H23O8Se (M+.):

363.0551. Found: 363.0559.

4.5.3.3 Benzyl 6-deoxy-6-[2,3,5-tri-O-benzyl-1,4-dideoxy-(R)-epi-seleniumylidene-D-arabinitol]-α-D-mannopyranoside-p-toluenesulfonate (13).

Reaction of 1,4-dideoxy-2,3,5-tri-O-benzyl-1,4-anhydro-4-seleno-D-

arabinitol 10 (660 mg, 1.4 mmol) with benzyl 6-O-p-toluenesulfonyl-β-D-

mannopyranoside 13 (500 mg, 1.2 mmol) in HFIP (1.5 mL), containing

anhydrous K2CO3 (10 mg) at 65-70 oC for 4 days gave the selenonium salt 13

as a foam (473 mg, 45%) after purification by column chromatography

(CHCl3/MeOH (95:5)). Analysis by NMR showed that the product was a mixture

of two isomers (~10:1) at the stereogenic selenium centre. [α] 23D = + 13.6 o Data

for the major diastereomer: 1H NMR (MeOD) δ 7.74-7.24 (24H, m, Ar), 4.86 (1H,

m, H-1′), 4.81 (1H, m, H-2), 4.71-4.50 (8H, m, 4CH2-Ph), 4.58 (1H, m, H-3), 4.42

(1H, dd, J3,4 = 6.8, J4,5 = 9.4 Hz, H-4 ), 4.03 (1H, d, J1,2 = 12.8 Hz, H-1a), 3.94

(2H, m, H-6′a, H-4′), 3.88 (1H, dd, J1′,2′ =2.0 J2′,3′ = 2.7 Hz, H-2′) 3.83 (1H, dd, J4,5

122

= 6.7, J5a,5b = 10.3 Hz, H-5a), 3.78-3.73 (4H, m, H-1b, H-5b, H-3′, H-6b), 3.59

(1H, t, J4′,5′ = J5′,6′ = 9.3 Hz, H-5′), 2.39 (3H, s, Me). 13C NMR (MeOD) δ 141.7-

125.1 (m, Ar), 100.0 (C-1′), 83.7 (C-2), 83.2 (C-3), 72.6, 71.5, 71.2, 69.5 (4CH2-

Ph), 70.2 (C-2′), 70.0(C-5′), 69.9 (C-3′), 68.9 (C-4′), 66.1 (C-5),65.7 (C-4), 45.9

(C-1), 45.6 (C-6′), 19.4 (Me). HRMS Calcd for C39H45O8Se (M+.): 721.2278.

Found: 721.2278.

4.5.3.4 1,4-Dideoxy-1,4-[[2S, 3S, 4R, 5R]-2,3,4,5,6-pentahydroxy-hexyl]-(R/S)-epi-seleniumylidene]-D-arabinitol chloride (8).

Compound 8 was obtained as a colorless solid (40 mg, 45%) from 13 (200

mg, 0.22 mmol) using the same procedure that was used to obtain 7. [α] 23D = + 4o.

1H NMR (D2O) δ 4.85 (1H, dd, J1,2 =7.6, J2,3 = 3.85, H-2), 4.54 (1H, t, J3,4 = J2,3 =

3.5 Hz,, H-3), 4.30-4.23 (2H, m, H-2′, H-4), 4.13 (1H, dd, J4,5a = 12.4, J5a,5b = 5.2

Hz, 5a) 4.02-3.97 (2H, m, H-1′a,H-5b), 3.94-3.82 (5H, m, H-6′a, H-3′, H-1′b, H-

1a, H-1b), 3.78-3.73 (2H, m, H-4′, H-5′b) 3.68 (1H, dd, J6′a,6,b = 11.7, J5′,6′b = 5.6

Hz, H-6′b). 13C NMR (D2O) δ 78.3 (C-3), 77.9 (C-2), 72.1 (C-3′), 70.7(C-5′), 69.6

(C-4), 69.2 (C-4′), 67.5 (C-2′), 63.1 (C-6′), 59.4 (C-5), 48.1 (C-1′), 45.3 (C-1).

HRMS Calcd for C11H23O8Se (M+.): 363.0553. Found: 363.0544.

4.6 Acknowledgments

We thank Buford L. Nichols, Roberto Quezada-Calvillo and Hassan Naim

for reagents for recombinant expression, and Lyann Sim for MGAM and SI

enzyme purification. Work by K.J. was funded by CIHR and the Canadian

Digestive Health Foundation (CDHF). We also thank the Canadian Institutes of

123

Health Research and the Heart and Stroke Foundation of Ontario (NA-6305) for

financial support.

4.7 References

1. (a) J. P. C. Chandrasena, The Chemistry and Pharmacology of Ceylon

and Indian Medicinal Plants; H&C press: Colombo, Sri Lanka, 1935. (b) D.

M. A. Jayaweera, Medicinal Plants Used in Ceylon-Part 1;

NationalScience Council of Sri Lanka: Colombo, 1981; p 77. (c) P. S.

Vaidyartanam, In Indian Medicinal Plants: a compendium of 500 species,

P. K. Warrier V. P. K. Nambiar and C. Ramankutty, Eds.; Orient Longman:

Madras, India, 1993, pp. 47-48.

2. M. H. S. Jayawardena, N. M. W. de Alwis, V. Hettigoda and D. J. S.

Fernando, J. Ethnopharmacol. 2005, 97, 215–218.

3. For reviews see: a) S. Mohan and B. M. Pinto, Carbohydr. Res. 2007, 342,

1551-1580. b) S. Mohan and B. M. Pinto, Collect. Czech. Chem. Commun.

2009, 74, 1117-1136. c) S. Mohan and B. M. Pinto, Nat. Prod. Rep. 2010,

27, 481-488.

4. A. Ghavami, B. D. Johnston and B. M. Pinto, J. Org. Chem. 2001, 66,

2312–2317.

5. B. D. Johnston, H. H. Jensen and B. M. Pinto, J. Org. Chem. 2006, 71,

1111–1118.

6. K. Jayakanthan, S. Mohan and B. M. Pinto, J. Am. Chem. Soc. 2009, 131,

5621-5626.

7. R. Eskandari, D. A. Kuntz, D. R. Rose and B. M. Pinto, Org. Lett. 2010, 12, 1632-1635.

124

8. W. Xie, G. Tanabe, J. Akaki, T. Morikawa, K. Ninomiya, T. Minematsu, M.

Yoshikawa, X. Wu and O. Muraoka, Bioorg & Med. Chem. 2011, 19, 2015-

2022.

9. M. Yoshikawa, T. Murakami, H. Shimada, H. Matsuda, J. Yamahara, G.

Tanabe and O. Muraoka, Tetrahedron Lett. 1997, 38, 8367–8370.

10. M. Yoshikawa, T. Murakami, K. Yashiro and H. Matsuda, Chem. Pharm.

Bull. 1998, 46, 1339–1340.

11. S. Ozaki, H. Oe and S. Kitamura, J. Nat. Prod. 2008, 71, 981–984. (b) O.

Muraoka, W. Xie, G. Tanabe, M. F. A. Amer, T. Minematsu and M.

Yoshikawa, Tetrahedron Lett. 2008, 49, 7315–7317.

12. L. Sim, R. Quezada-Calvillo, E. E. Sterchi, B. L. Nichols and D. R. Rose, J.

Mol. Biol. 2008, 375, 782-792.

13. L. Sim, K. Jayakanthan, S. Mohan, R. Nasi, B. D. Johnston, B. M. Pinto,

and D. R. Rose, Biochemistry 2010, 49, 443–451.

14. B. D. Johnston, A. Ghavami, M. T. Jensen, B. Svensson and B. M. Pinto,

J. Am. Chem. Soc. 2002, 124, 8245–8250.

15. H. Liu, R. Nasi, K. Jayakanthan, L. Sim, H. Heipel, D. R. Rose and B.M.

Pinto, J. Org. Chem. 2007, 72, 6562–6572.

16. S. Mohan, K. Jayakanthan, R. Nasi, D. A. Kuntz, D. R. Rose and B. M.

Pinto, Org. Lett. 2010, 12, 1088.

17. B. L. Nichols, J. Eldering, S. E. Avery, D. Hahn, A. Quaroni and E. E.

Sterchi, J. Biol. Chem. 1998, 273, 3076-3081.

18. R. Quezada-Calvillo, L. Sim, Z. Ao, B. R. Hamaker, A. Quaroni, G. D.

Brayer, E. E. Sterchi, C. C. Robayo-Torres, D. R. Rose and B. L. Nichols,

J. Nutr. 2008, 138, 685-692.

125

19. B. L. Nichols, S. E. Avery, P. Sen, D. M. Swallow, D. Hahn and E. E.

Sterchi, PNAS 2003, 100, 1432-1437.

20. H. Heymann, D. Breitmeier and S. Günther, Biol. Chem. Hoppe-Seyler

1995, 376, 249-253.

21. H. Heymann and S. Günther, Biol. Chem. Hoppe-Seyler 1994, 375, 451-

455.

22. C. Robayo-Torres, R. Quezada-Calvillo and B. L. Nichols, Clin.

Gastroenterol. Hepatol. 2006, 4, 276-287.

23. Naumoff, D. G. Mol. Biol. (Mosk). 2007, 41, 1056-1068.

24. K. Jones, L. Sim, S. Mohan, K. Jayakanthan, H. Liu, S. Avery, H. H.

Naim, R. Quezada-Calvillo, B. L. Nichols, B. M. Pinto and D. R. Rose,

Bioorg. Med. Chem. 2011, 19, 3929-3934.

25. L. Sim, C. Willemsma, S. Mohan, H. Y. Naim, B.M. Pinto and D. R. Rose,

J. Biol. Chem. 2010, 285, 17763-17770.

126

4.8 Supporting Information

The effect of heteroatom substitution of sulfur for selenium

in glucosidase inhibitors on intestinal α-glucosidase activities

Razieh Eskandari, Kyra Jones, David R. Rose, B. Mario Pinto

127

ppm (t1)2.03.04.05.06.07.08.0

ppm (t1)50100

O

HOOH

OH

OBnSe

BnO OBn

BnO

OTs

11

1'

2'3'4'

5'6'

O

HOOH

OH

OBnSe

BnO OBn

BnO

OTs

11

1'

2'3'4'

5'6'

128

ppm (t1)3.504.004.50

Se

OHHO

HO

OH

OH

OH

OH

OHCl

7

1' 2'3'

4'5'

6'

ppm (t1)45.050.055.060.065.070.075.080.0

Se

OHHO

HO

OH

OH

OH

OH

OHCl

7

1' 2'3'

4'5'

6'

129

ppm (t1)2.03.04.05.06.07.08.0

O

HOOH

OH

OBnSe

BnO OBn

BnO

OTs

13

1'

2'3'4'

5'6'

ppm (t1)50100150

O

HOOH

OH

OBnSe

BnO OBn

BnO

OTs

13

1'

2'3'4'

5'6'

130

H-4 H-6’a H-6’b

O

OH

OH

OHBnO

Se

BnO OBn

13

H

H2C

4

BnO

ppm (t1)4.004.505.00

Se

OHHO

HO

OH

OH

OH

OH

OHCl

8

3'4'

5'6'1' 2'

131

Representative Lineweaver-Burk plot of ntSI inhibited by 5 at concentrations of 0 nM, 75 nM, 125 nM, and 200 nM.

ppm (t1)4050607080

Se

OHHO

HO

OH

OH

OH

OH

OHCl

8

3'4'

5'6'1' 2'

132

CHAPTER 5: PROBING THE ACTIVE-SITE REQUIREMENTS OF HUMAN INTESTINAL N-TERMINAL

MALTASE-GLUCOAMYLASE: THE EFFECT OF REPLACING THE SULFATE MOIETY BY A METHYL

ETHER IN PONKORANOL, A NATURALLY-OCCURRING α-GLUCOSIDASE INHIBITOR

This Chapter comprises the manuscript “Probing the active-site requirements of human intestinal N-terminal maltase-glucoamylase: The effect of

replacing the sulfate moiety by a methyl ether in ponkoranol, a naturally-

occurring α-glucosidase inhibitor” which was published in Bioorganic and

Medicinal Chemistry Letters (2010, 20, 5686-5689).

Razieh Eskandari,a Kyra Jones,b David R. Rose,b B. Mario Pintoa

aDepartment of Chemistry, Simon Fraser University, Burnaby, British Columbia,

Canada V5A 1S6

b Department of Biology, University of Waterloo, Waterloo, Ontario,

Canada N2L 3G1

133

The crystal structures of N-terminal maltase-glucoamylase (ntMGAM) in

complex with kotalanol and de-O-sulfonated kotalanol have shown that removing

the sulfate group enhances the inhibitory activity presumably due to alleviation of

interaction between the polar sulfate group and hydrophobic groups in the active

site. In this Chapter, we set out to probe the active-site requirements of ntMGAM

using 3’-O-methylponkoranol. We describe here the synthesis of this compound

and the results obtained from the evaluation of its inhibitory activity against N-

terminal maltase-glucoamylase (ntMGAM). The thesis author performed all the

experimental synthetic work and the characterization of the compound, and wrote

and edited the manuscript with Dr. B. Mario Pinto. Mrs. Kyra Jones and Dr. David

R. Rose performed the enzyme inhibition studies.

Graphical abstract:

O

MeOO

OBnTsO S

OHHO

HO

OH

OH

OMe

OH

ClS

OBnBnO

BnO+ O

OH

MeO

OMe

134

5.1 Keywords

α-Glucosidase inhibitors, 3’-O-methylponkoranol, maltase-glucoamylase,

thiosugar, sulfonium-ion.

5.2 Abstract

Ponkoranol is a naturally-occurring glucosidase inhibitor isolated from the

plant Salacia reticulata. The compound comprises a sulfonium-ion with an

internal sulfate counter ion. We report here an efficient synthetic route to 3'-O-

methylponkoranol to test the hypothesis that occupation of a hydrophobic pocket

by a methyl group instead of the polar sulfate ion within the active site of human

N-terminal maltase-glucoamylase would be beneficial. The synthetic strategy

relies on the nucleophilic attack of 2,3,5-tri-O-benzyl-1,4-anhydro-4-thio-D-

arabinitol at the C-6 position of benzyl 6-O-p-toluenesulfonyl β-D-

glucopyranoside, followed by deprotection using boron trichloride and reduction

with sodium borohydride. The target compound inhibited the N-terminal catalytic

domain of intestinal human maltase-glucoamylase (ntMGAM) with a Ki value of

0.50 ± 0.04 µM, higher than those of de-O-sulfonated ponkoranol (Ki = 43 ± 3

nM), or its 5’-stereoisomer (Ki = 15 ± 1 nM). We conclude that the interaction of

the methyl group with hydrophobic residues in the active site is not as beneficial

to inhibition of ntMGAM as the other interactions of the polyhydroxylated chain

with active-site residues.

135

5.3 Introduction

Glycosidase inhibitors have many potential therapeutic applications

because glycosidase enzyme-catalyzed hydrolysis of complex carbohydrates is

biologically widespread, and has been implicated in several disease states.1,2 For

example, inhibition of starch-hydrolyzing enzymes such as pancreatic α-amylase

and intestinal α-glucosidases that leads to a delay in digestion of ingested

carbohydrates is one of the therapeutic approaches for the treatment of type-2

diabetes.3,4 Bioactive components isolated from medicinal plants often provide

the lead structures for drug development programs.5,6 For example, a relatively

new and interesting class of inhibitors is the sulfonium-ion containing inhibitors,

which were first isolated from Salacia reticulata, a plant that is used in traditional

Ayurvedic medicine in Sri Lanka and South India for treating type-2 diabetes.7-9

The active components in Salacia reticulata were found to include salaprinol

(1),10 salacinol (2),11 ponkoranol (3), 10 kotalanol (4),12 de-O-sulfonated kotalanol

(5),13 and de-O-sulfonated salacinol (6)14 (Figure 5-1), whose structures comprise

a 1,4-anhydro-4-thio-D-arabinitol core and polyhydroxylated acyclic chain (Figure

5-1).

136

Figure 5-1: Components isolated from Salacia species.

Comparison of the inhibitory activities against the human N-terminal

catalytic domain of maltase-glucoamylase (ntMGAM) of de-O-sulfonated

kotalanol (5) and some of its stereoisomers vs kotalanol (4) and the

corresponding sulfated stereoisomers, respectively, revealed that the de-O-

sulfonated analogues were more potent inhibitors than the parent

compounds.15,16 Furthermore, we have shown recently that de-O-sulfonated

ponkoranol (7) (Ki = 0.043 ± 0.01 µM) and its 5’-stereoisomer (8) (Ki = 0.015 ±

0.01 µM) are more potent inhibitors of ntMGAM than ponkoranol (3) itself (Ki =

0.17 ± 0.03 µM) (Figure 5-2).17

S

OHHO

OSO3 OH

OH OH

HO

OHOH

S

OHHO

OSO3 OH

OH OH

HO

OH

S

OHHO

OSO3

OH OH

HOS

OHHO

OSO3

OH

HO

Kotalanol (4)

Salacinol (2) Ponkoranol (3)Salaprinol (1)

S

OHHO

OH OH

OH OH

HO

OHOH

CH3OSO3

De-O-sulfonated kotalanol (5)

S

OHHO

OH

OH OH

HOHCO2

De-O-sulfonated salacinol (6)

137

Figure 5-2: De-O-sulfonated ponkoranol and its 5’-stereoisomer.

Our previous X-ray crystallographic studies of ntMGAM in complex with

kotalanol (4) and de-O-sulfonated kotalanol (5) had indicated that removal of the

sulfate group affects the conformation of the rest of the polyhydroxylated chain.18

We concluded that although the stereoconfiguration at C3’ does not affect

inhibitory activity, the proximity of the sulfate group to the large hydrophobic

groups (Y299, W406 and F575) likely restricts its conformational freedom.

Therefore, by removing the sulfate group, the positional constraint imposed by

the bulky hydrophobic residues surrounding the C3’ group is relieved, allowing

the rest of the polyhydroxylated chain to make optimal contacts with the ntMGAM

active site (Figure 5-3).18

S

OHHO

OH OH

OH OH

HO

OH

S

OHHO

OH OH

OH OH

HO

OH

Cl Cl

7 8

1'2'

3'4'

5'6'

1'2'

3'4'

5' 6'

138

Figure 5-3: Effect of removing the sulfate group. Superposition of kotalanol (4) (orange) and de-O-sulfonated kotalanol (5) (purple) structures. Double-headed arrows show the proximities of the sulfate group to the surrounding hydrophobic residues Y299, W406 and F575. (Reproduced with permission from Biochemistry, 2010, 49, 443–451. Copyright American Chemical Society).

In view of these findings, it was of interest to question whether replacing

the sulfate group with a hydrophobic methyl ether in ponkoranol (9) (Figure 5-3)

would increase its inhibitory properties through hydrophobic interactions in the

site compared to de-O-sulfonated ponkoranol 7.

Figure 5-4: 3’-O-Methylponkoranol.

S

OHHO

OMe OH

OH OH

HO

OH

Cl

9

1'2'

3'4'

5'6'

139

5.4 Results and discussion

We report here an efficient synthetic route to 3’-O-methylponkoranol (9).

Our synthetic strategy involved the alkylation of an appropriate protected

anhydrothioarabinitol B at the ring sulfur atom with agent C. Agent C could be

obtained by methylation of protected D-glucose at C-4’(Scheme 5-1).

Scheme 5-1: Retrosynthetic analysis.

Thus, benzyl 2,3-di-O-benzyl-4-O-methyl-6-O-tosyl-β-D-glucopyranoside

(11) was obtained from 1019 by treatment with a mixture of methyl iodide in

aqueous sodium hydroxide and dimethyl sulfoxide (50% w/w) 20 to afford 11. Our

initial attempt at the coupling reaction employed 11 with the anhydrothioarabinitol

(12)21 in 1,1,1,3,3,3-hexafluoroisopropyl alcohol (HFIP) at 70 °C, as in our

previous work (Scheme 2).22 No product formation and decomposition of the

starting material were observed by TLC; the coupling reaction of benzyl 4-O-

methyl-6-O-tosyl-β-D-glucopyranoside (13) with the anhydrothioarabinitol (12)

was also unsuccessful, the same result being observed (Scheme 5-2).

O

MeOOP

OPL

S

OPPO

POS

OHHO

HO

OH

OH

OMe

OH

L

L = leaving groupP = protecting group

+

A B C

OHOP

O

HOOP

OPL

OP

140

Scheme 5-2: First attempted synthesis of 9.

Next, a trifluoromethanesulfonyl (OTf) group was chosen as a leaving

group to ensure milder conditions for the coupling reaction. Thus, the primary

hydroxyl group in the diol 1419 was protected as its TBDMS ether followed by

sequential methylation of the secondary hydroxyl group to give 15 in 53% over

two steps (Scheme 5-3). Removal of the TBDMS group using

tetrabutylammonium fluoride (TBAF) gave the alcohol which was treated with

trifluoromethanesulfonyl anhydride to yield the glycoside 16 in 76% yield over two

steps. The coupling reaction of the OBn-protected thioether (12) with benzyl 2,3-

di-O-benzyl-4-O-methyl-6-O-trifluoromethanesulfonyl-α-D-glucopyranoside 16

was carried out in dry CH2Cl2 at room temperature to give the corresponding

O

MeOOBn

OBn

OBnTsO

HFIP, 70oC

11

Starting material decompositon

MeI, NaOH, DMSO, rt. 85%

S

BnO OBn

BnO

12

O

HOOBn

OBn

OBnTsO

10

O

MeOOH

OH

OBnTsO

HFIP, 70oC

13

S

BnO OBn

BnO

12 Starting material decompositon

141

protected sulfonium-ion 17 as a 4:1 mixture of diastereomers at the stereogenic

sulfur center. However, attempts to separate the mixture of diastereomers, even

after deprotection and reduction, were unsuccessful.

Scheme 5-3: Second attempted synthesis of 9.

In another attempt, the tosylate 18, containing a butane diacetal (BDA)

protecting group, was chosen as the coupling partner. As shown in Scheme 5-4,

the tosylate 18 was synthesized in two steps from benzyl 2,3-O-[(2R,3R)-2,3-

dimethoxybutane-2,3-diyl]-β-D-glucopyranoside 19 which was, in turn, prepared

from D-glucose according to literature procedures.23 Thus, compound 19 was

treated with p-toluenesulfonyl chloride to afford the p-toluenesulfonyl ester 20 in

1.TBDMSCl,Imid,2.MeI, NaOHDMSO, rt, 53%

O

MeOOBn

OBn

OBnTBSO

CH2Cl2, rt, 90%

O

MeOOBn

OBn

OBnS

BnO OBn

BnO

S

OHHO

HO

OH

OH

OMe

OH

OHOTfCl

15

179

1'

2'3'

3'

4'

4'5'5'6'

6'1' 2'

57%

S

BnO OBn

BnO

O

HOOBn

OBn

OBnHO

1. BCl3, CH2Cl2,

-78 oC, 6 h

2. Amberlyst A-26,H2O, 3 h3. NaBH4, H2O, 3 h

14

1. TBAF, THF2. TfO2, Pyridine, CH2Cl2, 76%

O

MeOOBn

OBn

OBnTfO

16

12

142

75%, which was methylated with methyl iodide to give 18 in 91% yield. The

coupling reactions of the OBn-protected 1,4-anhydro-4-thio-D-arabinitol (12) with

the p-toluenesulfonyl ester 18 were carried out in HFIP to give the protected

sulfonium-ion 21 as a single diastereomer in 70% yield (Scheme 5-4).

Deprotection of the coupled product 21 was carried out in a two-step

procedure. The benzyl group was first removed by treatment with boron

trichloride at -78 °C in CH2Cl2 ,followed by sequential BDA deprotection with 80%

TFA. During the course of benzyl group deprotection with BCl3, the p-

toluenesulfonate counterion was partially exchanged with chloride ion. Similar

results were observed in our previous work.18 Hence, after deprotection, the

product was treated with Amberlyst A-26 resin (chloride form) to completely

exchange the p-toluenesulfonate counterion with chloride ion. Finally, the crude

product was reduced with NaBH4 to provide the desired 3'-O-methylponkoranol 9

in 51% yield over three steps (Scheme 5-4).

143

Scheme 5-4: Synthesis of compound 9.

The absolute stereochemistry at the stereogenic sulfur center in 9 was

established by means of 1D-NOESY experiments (Figure 5-5). Correlation

between H-1’ and H-4 and also a correlation between H-2’ and H-4 confirmed the

anti relationship between the alkyl side chain and the C-4 substituent on the

anhydroarabinitol moiety in compound 9.

Figure 5-5: 1D-NOESY correlations of selected protons in compound 9.

HFIP, 70 oC, 70%

S

BnO OBn

BnOS

OHHO

HO

OH

OH

OMe

OH

OHOTs Cl

21 9

3'4'

5'6'1' 2'

51%

S

BnO OBn

BnO

O

HOO

O

OBnHO

1. BCl3, CH2Cl2,

-78 oC, 6 h

2. 80% TFA

3. Amberlyst A-26,4. NaBH4, H2O, 3 h

19

OMe

MeO

O

MeOO

O

OBnTsO

OMe

MeO

TsCl, Pyridine

1' 2'

3'4'5'

6'

O

OHO

O

BnOOMe

OMe

18

MeI, NaOH, DMSO, 91%

O

HOO

O

OBnTsO

20

OMe

MeO

75%

S

OHHOHO

OH

OH

OMe

OH

OHCl

9

3'4'

5'6'1' 2'

H4

H

144

Finally, the inhibitory activity of compound 9 was examined against the N-

terminal catalytic domain of recombinant human maltase-glucoamylase

(ntMGAM), a critical intestinal glucosidase for processing starch-derived

oligosaccharides into glucose. The 3’-O-methylponkoranol 9 inhibited ntMGAM

with a Ki value of 0.50 ± 0.04 µM; by comparison, de-O-sulfonated ponkoranol 7

and its 5’-stereoisomer 8 inhibited ntMGAM with Ki values of 43 ± 3 and 15 ± 1

nM, respectively.17. We conclude, therefore, that the hydrophobic interactions

between the methyl group and the hydrophobic residues Y299, W406 and F575

in the active site are not as optimal as the the interactions of the latter groups

with the rest of polyhydroxylated chain in the absence of the methyl ether; a

similar situation is also observed in the binding of the sulfated compound,

ponkoranol (3) (Ki = 0.17 ± 0.03 µM).

5.5 Experimental

5.5.1 General methods

Optical rotations were measured at 23 °C. 1H and 13C NMR spectra were

recorded at 600 and 150 MHz, respectively. All assignments were confirmed with

the aid of two-dimensional 1H, 1H (COSYDFTP) or 1H, 13C (INVBTP) experiments

using standard pulse programs. Column chromatography was performed with

Silica 60 (230-400 mesh). Reverse column chromatography was performed with

Silica C-18 cartridges. High resolution mass spectra were obtained by the

electrospray ionization method, using an Agilent 6210 TOF LC/MS high

resolution magnetic sector mass spectrometer.

145

5.5.2 Enzyme kinetics

Activity of recombinant N-terminal domain of maltase-glucoamylase

(ntMGAM) was determined using the glucose oxidase assay17 to follow the

production of glucose from maltose upon addition of the enzyme (0.8 nM). A no-

inhibitor control and five different inhibitor concentrations were used in

combination with 7 different maltose concentrations (ranging from 1.5 to 24 mM).

A reaction time of 60 min at 37 ºC was employed. Reactions were linear within

this time frame. Values of Ki and standard deviations were determined by the

program GraFit 4.0.14 (Erithacus Software) which employs nonlinear fitting of the

data for each inhibitor concentration to the Michaelis-Menten equation.

5.5.3 Compound characterization data

5.5.3.1 General Procedure for Methylation

To a vigorously stirred solution of the glucopyranoside in dimethyl

sulfoxide (DMSO) was added 50% aqueous sodium hydroxide (1.7 equiv.) to

form a gel-like suspension. Methyl iodide (1.5 equiv), was added dropwise in one

portion, and the mixture was further stirred for 2-15 min, then quenched with 2:1

ether:water. The organic layer was separated and the aqueous layer was

extracted with ether. The organic solution was dried (Na2SO4) and concentrated

to give the crude product.

5.5.3.2 Benzyl 2,3-di-O-benzyl-4-O-methyl-6-O-p-toluenesulfonyl-β-D-glucopyranoside 11

Reaction of 1019 (150 mg, 0.25 mmol) with methyl iodide (0.026 mL) in

presence of 50% aqueous sodium hydroxide (6 drops) in DMSO (3.0 mL) for 5

146

min at room temperature gave compound 11 as a white, amorphous foam (130

mg, 85%). [α] 23D = + 3 (c = 1, CH2Cl2). 1H NMR (CDCl3) δ 7.76-7.18 (19H, m, Ar),

4.82, 4.78, 4.75, 4.65, 4.59, 4.48 (6H, 6d, JA,B = 12.1 Hz, 3CH2Ph), 4.33 (1H, d,

J1,2 = 7.7 Hz, H-1), 4.23 (1H, dd, J6a,6b = 2.3, J5,6a = 10.7 Hz, H-6a), 4.16 (1H, dd,

J6a,6b = 5.3, J5,6b = 10.5 Hz, H-6b), 3.44 (1H, t, J2,3 = J3,4 = 9.6 Hz, H-3 ), 3.40

(3H, s, OMe), 3.31 (2H, m, H-2, H-5), 3.08 (1H, dd, J3,4 = 8.9, J4,5 = 10.1 Hz, H-

4), 2.34 (3H, s, Me). 13C NMR (CDCl3) δ 144.9-127.7 (m, Ar), 102.1 (C-1), 84.3

(C-3), 81.8 (C-2), 79.1 (C-4), 75.5, 74.9, 71.0 (3 CH2Ph), 72.8 (C-5), 68.7 (C-6),

60.8 (OMe). HRMS Calcd for C35H39O8S (M+H): 619.236. Found: 619.2363.

5.5.3.3 Benzyl 4-O-methyl-6-O-p-toluenesulfonyl-β-D-glucopyranoside 13

Compound 20 (report in this paper) (200 mg, 0.36 mmol) was dissolved in

80% TFA (5 mL) and stirred at room temperature for 1 h. The solvents were

removed under reduced pressure, and the residue was purified by flash

chromatography (EtOAc/Hexanes (1:2)) to yield 13 as a white foam (150 mg,

95%). [α] 23D = - 22.8 (c =0.4, MeOH). 1H NMR (CDCl3) δ 7.84-7.28 (9H, m, Ar),

4.75, 4.74 (2H, 2d, JA,B = 11.5 Hz, CH2Ph), 4.26 (3H, m, H-1, H-6a, H6b), 3.49

(3H, s, OMe), 3.40 (2H, m, H-3, H-5), 3.19 (1H, t, J2,3 = J1,2 = 8.5 Hz, H-2 ), 3.04

(1H, t, J4,5 = J3,4 = 9.0 Hz, H-4 ), 2.42 (3H, s, Me). 13C NMR (CDCl3) δ 145.2-

127.3 (m, Ar), 101.7 (C-1), 79.0 (C-4), 76.6 (C-3), 73.6 (C-2), 72.5 (C-5), 70.3

(CH2Ph), 68.9 (C-6), 59.5 (OMe), 120.2 (Me). HRMS Calcd for C12H26NaO8S

(M+Na): 463.1248. Found: 463.1242.

147

5.5.3.4 Benzyl 2,3-di-O-benzyl-6-O- tert-butyldimethylsilyl-4-O-methyl-6-O-p-toluenesulfonyl-α- D-glucopyranoside 15

To a solution of 1419 (500 mg, 1.1 mmol) in DMF (15 mL) was added

imidazole (235 mg, 3.4 mmol). The reaction was cooled in an ice bath, TBDMSCl

(190 mg, 1.2 mmol) was added portionwise, and the mixture was stirred at 0 °C

under N2 for 2 h. The reaction was quenched by the addition of ice-water, and the

reaction mixture was extracted with Et2O (3×30 mL). The combined organic

solvents were dried (Na2SO4), and concentrated to give the crude product which

was used directly in the next step without further purification. The residue was

dissolved in DMSO (10 mL) and methyl iodide (0.12 mL) and 50% aqueous

sodium hydroxide (12 drops) were added. The reaction mixture was stirred for 5

min at room temperature to yield 15 as a syrup (340 mg, 53%). [α] 23D = + 98 (c =

1, CH2Cl2). 1H NMR (CDCl3) δ 7.35-7.17 (15H, m, Ar), 4.89, 4.76, 4.64, 4.60,

4.49, 4.47 (6H, 6d, JA,B = 12.1 Hz, 3CH2Ph), 4.75 (1H, d, J1,2 = 2.8 Hz, H-1), 3.89

(1H, t, J2,3 = J3,4 = 9.3 Hz, H-3), 3.72 (1H, dd, J6a,6b = 4.3, J5,6a = 11.1 Hz, H-6a),

3.67 (1H, dd, J6a,6b = 1.8, J5,6b = 11.1 Hz, H-6b), 3.54 (1H, ddd, J4,5 = 1.8, J6a,5 =

4.3, J6b,5 = 5.8 Hz, H-5), 3.50 (3H, s, OMe), 3.40 (1H, dd, J2,3 = 3.6, J1,2 = 10.1

Hz, H-2 ), 3.21 (1H, t, J3,4 = J4,5 = 9.4 Hz, H-4), 0.84 (9H, s, 3Me), 0.01, 0.00 (6H,

2s, 2Me). 13C NMR (CDCl3) δ 139.0-127.6 (m, Ar), 95.2 (C-1), 82.2 (C-3), 80.1

(C-2), 79.5 (C-4), 75.8, 73.0, 68.7 (3 CH2Ph), 71.9 (C-5), 62.2 (C-6), 60.8 (OMe),

26.0, -5.1, -5.3 (5Me). HRMS Calcd for C34H46NaO6Si (M+Na): 601.2956. Found:

601.2954.

148

5.5.3.5 Benzyl 2,3-di-O-benzyl-4-O-methyl-6-O-tert-butyldimethylsilyl-α-D-glucopyranoside 16

To a solution of 15 (600 mg, 1.04 mmol) in THF (50 mL), TBAF (1.0 M

solution in THF, 1.2 mL, 1.2 mmol) was added, and the reaction mixture was

stirred at room temperature. After 2 h it was concentrated and further dried under

high vacuum for 1 h. The crude product was dissolved in CH2Cl2 (25 mL), and

pyridine (0.1 mL, 1 eq) was added. The mixture was cooled to -10 °C, and Tf2O

(0.2 mL, 1.1 eq) was added under N2. After 5-10 min it was concentrated and the

residue was purified by flash chromatography (EtOAc/Hexanes (1:5)) to yield 16

as a white foam (620 mg, 76%). [α] 23D = +75 (c = 0.1, CH2Cl2). 1H NMR (CDCl3) δ

7.40-7.25 (15H, m, Ar), 4.98, 4.77 (2H, 2d, JA,B = 11.1 Hz, CH2Ph), 4.84 (1H, d,

J1,2 = 3.6 Hz, H-1), 4.68-4.49 (6H, m, 2CH2Ph, H-6a, H-6b), 3.95 (1H, t, J2,3 = J3,4

= 9.2 Hz, H-3 ), 3.80 (1H, ddd, J4,5 = 1.8, J6a,5 = 4.6, J6b,5 = 6.6 Hz, H-5), 3.55

(3H, s, OMe), 3.48 (1H, dd, J1,2 = 3.5, J2,3 = 9.4 Hz, H-2), 3.17 (1H, dd, J3,4 = 8.8,

J4,5 = 10.1 Hz, H-4). 13C NMR (CDCl3) δ 138.5 (CF3), 137.9-127.7 (m, Ar), 95.6

(C-1), 81.7 (C-3), 79.6 (C-2), 78.7 (C-4), 75.6, 73.1, 69.7 (3 CH2Ph), 74.8 (C-6),

68.5 (C-5), 60.9 (OMe). HRMS Calcd for C29H31F3NaO8S (M+Na): 619.1584.

Found: 619.1585.

5.5.3.6 Benzyl 2,3-O-[(2R,3R)-2,3-dimethoxybutane-2,3-diyl]-6-O-p-toluenesulfonyl -β-D-glucopyranoside 20

To a solution of 1923 (634 mg, 1.65 mmol) in pyridine (10 mL) was added

p-toluenesulfonyl chloride (TsCl) (380 mg, 3.27 mmol), and the mixture was

stirred at room temperature overnight. Another portion of TsCl (35 mg, 0.16

mmol) was added, the mixture was stirred for 4h, and the pyridine was removed

149

by distillation. The residue was purified by flash chromatography

(EtOAc/Hexanes (1:2)) to yield 20 as a white foam (660 mg, 75%). [α] 23D = - 208

(c = 1, CH2Cl2). 1H NMR (CDCl3) δ 7.85-7.32 (9H, m, Ar), 4.86, 4.63 (2H, 2d, JA,B

= 12.1 Hz, CH2Ph), 4.58 (1H, d, J1,2 = 8.0 Hz, H-1), 4.38 (1H, dd, J6a,6b = 1.7, J5,6a

= 10.9 Hz, H-6a), 4.28 (1H, dd, J6a,6b = 5.5, J5,6b = 10.7 Hz, H-6b), 3.66 (2H, m,

H-3, H-4), 3.55 (2H, m, H-2, H-5), 3.29 (6H, s, OMe), 2.43 (3H, s, Me), 1.97 (1H,

br, OH), 1.33, 1.34 (6H, 2s, 2Me). 13C NMR (CDCl3) δ 144.9-127.4 (m, Ar), 99.8

(C-1), 99.6, 99.5 (2MeCOMe), 74.1 (C-5), 72.4 (C-3), 70.7 (CH2Ph), 69.0 (C-2),

68.7 (C-6), 67.3 (C-4), 48,0 47.9 (2OMe), 21.6 (OMe), 17.6 (2Me). HRMS Calcd

for C26H34NaO10S (M+Na): 561.1765. Found: 561.1772.

5.5.3.7 Benzyl 2,3-O-[(2R,3R)-2,3-dimethoxybutane-2,3-diyl]-4-O-methyl-6-O-p-toluenesulfonyl-β-D-glucopyranoside 18

Reaction of 20 (710 mg, 1.3 mmol) with methyl iodide (0.13 mL) in

presence of 50% aqueous sodium hydroxide (25 drops) in DMSO (5.0 mL) for 3

min at room temperature gave compound 18 as a white, amorphous foam (660

mg, 91%). [α] 23D = -111 (c =1, CH2Cl2). 1H NMR (CDCl3) δ 7.84-7.29 (9H, m, Ar),

4.82, 4.61 (2H, 2d, JA,B = 12.7 Hz, CH2Ph), 4.51 (1H, d, J1,2 = 7.8 Hz, H-1), 4.31

(1H, dd, J6a,5 = 2.0, J6b,6b = 10.6 Hz, H-6a), 4.25 (1H, dd, J6b,5 = 5.4, J6b,6b = 10.6

Hz, H-6b), 3.78 (1H, t, J2,3 = J3,4 = 9.9 Hz, H-3 ), 3.56 (1H, dd, J1,2 = 8.0, J2,3 =

10.1 Hz, H-2), 3.55 (3H, s, OMe), 3.45 (1H, ddd, J4,5 = 2.0, J6a,5 = 5.4, J6b,5 = 7.6

Hz, H-5), 3.31, 3.30 (6H, 2s, 2OMe), 3.27 (1H, dd, J3,4 = 9.2, J4,5 = 18.6 Hz, H-4),

2.42 (3H, s, OMe), 1.34, 1.33 (6H, 2s, 2Me). 13C NMR (CDCl3) δ 144.8-127.4 (m,

Ar), 99.6 (C-1), 99.4 (MeCOMe), 75.6 (C-4), 73.7 (C-3), 73.4 (C-5), 70.6

150

(CH2Ph), 69.1 (C-2), 68.6 (C-6), 60.6 (OMe), 48,0 47.9 (2OMe), 21.6 (OMe),

17.8, 17.7 (2Me). HRMS Calcd for C27H36NaO10S (M+Na): 575.1921. Found:

575.1931.

5.5.3.8 Benzyl 2,3-O-[(2R,3R)-2,3-dimethoxybutane-2,3-diyl]-4-O-methyl-6-deoxy-6-[2,3,5-tri-O-benzyl-1,4-dideoxy-episulfoniumylidene-D-arabinitol]-β-D-glucopyranoside-p-toluenesulfonate 21

The mixture of the thioether 1221 (540 mg, 1.3 mmol) and benzyl 2,3-O-

[(2R,3R)-2,3-dimethoxybutane-2,3-diyl]-4−Ο−methyl-6−Ο−p-toluenesulfonyl−β-D-

glucopyranoside 20 (235 mg, 0.43 mmol) in HFIP (1 mL) were placed in a sealed

tube in an oil bath at 70 °C for 6 days. The mixture was cooled, then diluted with

EtOAc, and evaporated to give a syrupy residue. Purification by column

chromatography (EtOAc/MeOH 20:1) gave the sulfonium salt 21 as a syrup (310

mg, 70% based on 20). [α] 23D = -22.5 (c = 0.1, CH2Cl2). 1H NMR (CDCl3) δ 7.78-

7.01 (24H, m, Ar), 4.69, 4.56 (2H, 2d, JA,B = 12.1 Hz, CH2Ph), 4.49-4.35 (6H, m,

3 CH2Ph), 4.45 (1H, m, H-1’), 4.44 (1H, m, H-3), 4.41 (1H, m, H-4), 4.28 (1H, br,

H-2), 4.13 (1H, t, J6a’,5 = J6b’,6b’ = 11.2 Hz, H-6a’), 4.01 (1H, dd, J6b’,5 = 7.7, J6’b,6’b

= 13.3 Hz, H-6b’), 3.83 (1H, dd, J4,5a =7.6 J5a,5b = 10.0 Hz, H-5a ), 3.79-3.69 (5H,

m, H-1a, H-1b, H-5b, H-5’, H-3’), 3.49 (3H, s, OMe), 3.47 (1H, dd, J1’,2’ = 2.1,

J2’,3’ = 5.1 Hz, H-2’), 3.23, 3.19 (6H, 2s, 2OMe), 3.18 (1H, m, H-4’), 2.23 (3H, s,

OMe), 1.25, 1.24 (6H, 2s, 2Me). 13C NMR (CDCl3) δ 144.1-126.2 (m, Ar), 100.7

(C-1’), 99.5, 99.4 (2MeCOMe), 83.1 (C-2), 82.7 (C-3), 78.0 (C-4’), 73.5, 72.2,

71.9, 71.6 (4CH2Ph), 73.1 (C-3’), 71.8 (C-5’), 68.9 (C-2’), 66.9 (C-5), 65.3 (C-4),

60.7 (OMe), 48.1 (C-1) 48,0 47.9 (2OMe), 47.6 (C-6’), 21.2 (Me), 17.7, 17.5

(2Me). HRMS Calcd for C46H57O10S (M+.): 801.3667. Found: 801.3676.

151

5.5.3.9 1,4-Dideoxy-1,4-[[2S, 3S, 4R, 5S]-2,4,5,6-tetrahydroxy-3-(methyl) hexyl]-(R)-epi-sulfoniumylidine]-D-arabinitol chloride 9

Compound 21 (150 mg, 0.15 mmol) was dissolved in CH2Cl2 (15 mL), the

mixture was cooled to -78 °C, and BCl3 (1M solution in CH2Cl2, 2.3 mmol) was

added under N2. The reaction mixture was stirred at the same temperature for 30

min, and then allowed to warm to 5 °C for 6 h. The reaction was quenched by

addition of MeOH (5 mL), the solvents were removed, and the residue was co-

evaporated with MeOH (2 × 5 mL). The crude residue was dissolved in 80% TFA

(5 mL) and stirred at room temperature for 1h. The solvents were removed under

reduced pressure, the residue was dissolved in water (5 mL), and washed with

CH2Cl2 (3×5 mL). The water layer was evaporated to give the crude product. The

residue was dissolved in H2O (10 mL), Amberlyst A-26 resin (200 mg) was

added, and the reaction mixture was stirred at room temperature for 2 h.

Filtration through cotton, followed by solvent removal gave the crude hemiacetal.

The crude product was dissolved in water (8 mL), and the solution was stirred at

room temperature while NaBH4 (23 mg, 0.6 mmol) was added in small portions

over 30 min. Stirring was continued for another 3 h and the mixture was acidified

to pH < 4 by dropwise addition of 2M HCl. The mixture was evaporated to

dryness and the residue was co-evaporated with anhydrous MeOH (3 × 20 mL).

Treatment of the solid residue with 50% EtOAc:MeOH (5-10 mL) resulted in

precipitation of most of the borate salt. Filtration through cotton, followed by

solvent removal gave the crude compound. The residue was purified by reverse

phase column chromatography (H2O) to give 9 as a colorless solid (29 mg, 51%).

[α] 23D = + 3.3, (c = 0.6, H2O). 1H NMR (D2O) δ 4.66 (1H, m, H-2), 4.36 (1H, t, br,

152

H-3), 4.25 (1H, td, J1',2' = 3.5, J2',3' = 8.9 Hz, H-2'), 4.05 (1H, dd, J4,5a = 4.9, J5a,5b

= 12.0 Hz, H-5a), 3.97 (1H, m, H-4), 3.87-3.73 (5H, m, H-5b, H-1a, H-1b, H-1'a,

H1'b), 3.70-3.63 (3H, m, H-4', H-5', H-6'a), 3.54 (1H, dd, J5',6'b = 6.6, J6'a,6'b = 11.7

Hz, H-6'b), 3.46 (3H, s, OMe), 3.43 (1H, m, H-3'). 13C NMR (D2O) δ 83.4 (C-3'),

77.5 (C-3), 76.9 (C-2), 71.9 (C-5'), 70.2 (C-4'), 69.9 (C-4), 67.9 (C-2'), 62.5 (C-6'),

60.6 (OMe), 59.2 (C-5), 49.3 (C-1'), 48.2 (C-1). HRMS Calcd for C12H25O8S

(M+.): 329.1265. Found: 329.126.

5.6 Acknowledgments

We are grateful to the Canadian Institutes for Health Research and the

Heart and Stroke Foundation of Ontario grant # NA-6305 for financial support of

this work.

5.7 References

1. Lillelund, V. H.; Jensen, H. H.; Liang, X.; Bols, M. Chem. Rev. 2002, 102,

515.

2. Gloster, T. M. Davies, G. J. Org. Biomol. Chem., 2010, 8, 305.

3. Holman, R. R.; Cull, C. A.; Turner, R. C. Diabetes Care 1999, 22, 960.

4. Jacob, G. S. Curr. Opin. Struct. Biol. 1995, 5, 605.

5. Harvey, A. L. Drug Discov. Today 2008, 13, 894.

6. Butler, M. S. Nat. Prod. Rep. 2008, 25, 475.

7. Chandrasena, J. P. C. The Chemistry and Pharmacology of Ceylon and

Indian Medicinal Plants; H&C Press: Colombo, Sri Lanka, 1935.

153

8. Jayaweera, D. M. A. Medicinal Plants Used in Ceylon-Part 1, National

Science Council of Sri Lanka: Colombo, 1981.

9. Matsuda, H.; Yoshikawa, M.; Murakami, T.; Tanabe, G.; Muraoka, O. J.

Trad. Med. 2005, 22, 145.

10. Yoshikawa, M.; Xu, F. M.; Nakamura, S.; Wang, T.; Matsuda, H.; Tanabe,

G.; Muraoka, O. Heterocycles 2008, 75, 1397.

11. Yoshikawa, M.; Murakami, T.; Shimada, H.; Matsuda, H.; Yamahara, J.;

Tanabe, G.; Muraoka, O. Tetrahedron Lett. 1997, 38, 8367-8370.

12. Matsuda, H.; Li, Y. H.; Murakami, T.; Matsumura, N.; Yamahara, J.;

Yoshikawa, M. Chem. & Pharm. Bull. 1998, 46, 1399.

13. Muraoka, O.; Xie, W. J.; Tanabe, G.; Amer, M. F. A.; Minematsu, T.;

Yoshikawa, M. Tetrahedron Lett. 2008, 49, 7315.

14. Minami, Y.; Kurlyarna, C.; Ikeda, K.; Kato, A.; Takebayashi, K.; Adachi, I.;

Fleet, G. W. J.; Kettawan, A.; Karnoto, T.; Asano, N. Bioog. & Med. Chem.

2008, 16, 2734.

15. Jayakanthan, K.; Mohan, S.; Pinto, B. M. J. Am. Chem. Soc. 2009, 131,

5621.

16. Mohan, S.; Jayakanthan, K.; Nasi, R.; Kuntz, D. A.; Rose, D. R.; Pinto, B.

M. Org. Lett., 2010, 12, 1088.

17. Eskandari, R.; Kuntz, D. A.; Rose, D. R.; Pinto, B. M. Org. Lett. 2010, 12,

1632.

18. Sim, L.; Jayakanthan, K.; Mohan, S.; Nasi, R.; Johnston, B. D.; Pinto, B.

M.; Rose, D. R. Biochemistry 2010, 49, 443.

19. Shinkiti, K.; Kusunoki, A.; Hirooka, M. Bull. Chem. Soc. Jpn. 2000, 73,

967.

154

20. Wang, H.; Sun, L.; Glazebnik, S.; Zhao, K. Tetrahedron Lett. 1995, 36,

2953.

21. Satoh, H.; Yoshimura, Y.; Sakata, S.; Miura, S.; Machida, H.; Matsuda, A.

Bioorg & Med. Chem. Lett. 1998, 8, 989.

22. Ghavami, A.; Sadalapure, K. S.; Johnston, B. D.; Lobera, M.; Snider, B.

B.; Pinto, B. M. Synlett 2003, 1259.

23. Liu, H.; Nasi, R.; Jayakanthan, K.; Sim, L.; Heipel, H.; Rose, D. R.; Pinto,

B. M. J. Org. Chem. 2007, 72, 6562.

155

5.8 Supporting Information

Probing the active-site requirements of human intestinal N-terminal

maltase-glucoamylase: The effect of replacing the sulfate moiety by a

methyl ether in ponkoranol, a naturally-occurring α-glucosidase inhibitor

Razieh Eskandari, Kyra Jones, David R. Rose, B. Mario Pinto

156

ppm (t1)3.04.05.06.07.0

ppm (t1)50100

O

MeOOBn

OBn

OBnTsO

11

O

MeOOBn

OBn

OBnTsO

11

157

ppm (t1)3.04.05.06.07.0

ppm (t1)50100150

O

MeOOH

OH

OBnTsO

13

O

MeOOH

OH

OBnTsO

13

158

ppm (t1)0.01.02.03.04.05.06.07.0

ppm (t1)050100

O

MeOOBn

OBn

OBnTBSO

15

O

MeOOBn

OBn

OBnTBSO

15

159

ppm (t1)4.05.06.07.0

ppm (t1)60708090100110120130140

O

MeOOBn

OBn

OBnTfO

16

O

MeOOBn

OBn

OBnTfO

16

160

ppm (t1)1.02.03.04.05.06.07.0

ppm (t1)50100

O

HOO

O

OBnTsO

OMe

MeO20

O

HOO

O

OBnTsO

OMe

MeO20

161

ppm (t1)2.03.04.05.06.07.08.0

ppm (t1)50100

O

MeOO

O

OBnTsO

OMe

MeO

18

O

MeOO

O

OBnTsO

OMe

MeO

18

162

ppm (t1)2.03.04.05.06.07.0

ppm (t1)50100

S

BnO OBn

BnO

OTs

1' 2'

3'4'5'

6'

O

OHO

O

BnO OMe

OMe

21

S

BnO OBn

BnO

OTs

1' 2'

3'4'5'

6'

O

OHO

O

BnO OMe

OMe

21

163

ppm (t1)3.504.004.50

ppm (t1)50607080

S

OHHO

HO

OH

OH

OMe

OH

OHCl

3'4'

5'6'1' 2'

9

S

OHHO

HO

OH

OH

OMe

OH

OHCl

3'4'

5'6'1' 2'

9

164

ppm (t1)3.803.904.004.104.204.30

H-4H-2’ H-1’

S

OHHOHO

OH

OH

OMe

OH

OHCl

9

3'4'

5'6'1' 2'

H4

H

165

CHAPTER 6: SELECTIVITY OF 3'-O-METHYLPONKORANOL FOR INHIBITION OF N- AND C-

TERMINAL MALTASE-GLUCOAMYLASE AND SUCRASE-ISOMALTASE, POTENTIAL THERAPEUTICS FOR DIGESTIVE DISORDERS OR THEIR SEQUELAE

This Chapter comprises the manuscript “Selectivity of 3'-O-methylponkoranol for inhibition of N- and C-terminal maltase-glucoamylase and sucrase-

isomaltase, potential theraputics for digestive disorders or their sequelae”

which was published in Bioorganic and Medicinal Chemistry Letters (2011, 21,

6491-6494).

Razieh Eskandari,a Kyra Jones,b David R. Rose,b B. Mario Pintoa

aDepartment of Chemistry, Simon Fraser University, Burnaby, British Columbia,

Canada V5A 1S6

b Department of Biology, University of Waterloo, Waterloo, Ontario,

Canada N2L 3G1

166

In this Chapter, the selectivity of 3'-O-methylponkoranol for the different

intestinal enzymes N- and C- terminal maltase-glucoamylase and sucrase-

isomaltase is reported. The thesis author wrote and edited the manuscript with

Dr. B. Mario Pinto. Mrs. Kyra Jones and Dr. David R. Rose performed the

enzyme inhibition studies.

Graphical abstract:

Starch

167

6.1 Keywords

α-Glucosidase inhibitors, 3'-O-methylponkoranol, maltase-glucoamylase,

sucrase-isomaltase, selective inhibition, therapeutics of digestive disorders.

6.2 Abstract

Human maltase-glucoamylase (MGAM) and sucrase-isomaltase (SI) are

two human intestinal glucosidases responsible for the final step of starch

hydrolysis.

MGAM and SI are anchored to the small intestinal brush border epithelial

cells and contain two catalytic N-terminal and C-terminal subunits. In this study,

we report the inhibition profile of 3'-O-methylponkoranol for the individual

recombinant N and C terminal enzymes and compare the inhibitory activities of

this compound with de-O-sulfonated ponkoranol. We show that 3'-O-

methylponkoranol inhibits the different subunits to different extents, with

extraordinary selectivity for C-terminal SI (Ki = 7 ± 2 nM). The enzymes

themselves could serve as therapeutic targets for the treatment of digestive

disorders or their sequelae.

6.3 Introduction

Starches constitute the main source of energy for adult humans and have

particular nutritional importance in children. Ingestion of starch in the human diet

has received particular attention in recent years due to their nutritional value. The

imbalance between energy consumption and dietary intake of energy-rich foods

has a direct relationship to the development of energy metabolism diseases,

168

such as obesity or diabetes.1-3 Such adverse consequences have led to the study

of enzyme/substrate or enzyme/inhibitor interactions in the starch digestion

process.

Starch is present in two main molecular structures: mostly linear α-1,4-D-

glucopyranose linked polymers known as amyloses, and those with a mixture of

α-1,4- and α-1,6-branched α-D-glucopyranose linkages, known as amylopectins

(Figure 6-1).

Figure 6-1: Components of starch: amylose and amylopectin.

The manner in which mammals derive glucose molecules from starch

involves multiple enzymes.4-8 Two luminal α-1,4 endoglucosidases, namely

salivary and pancreatic amylases, hydrolyze linear, unbranched glucose

sequences present in amylose and amylopectin molecules into soluble glucose

Amylose

AmylopectinO

O

OH

HOHO

OH O

O

HOOH

OH

1

4

6

α-1,4 linkage

O

OH

HOHO

OH

1 α-1,6 linkage

O

O

OH

HOHO

OH O

OH

HOOH

OH

1

12

34 5

6

α-1,4 linkage

169

oligomers with both linear and branched structures,9-12 and the membrane bound

enzymes maltase-glucoamylase (MGAM) and sucrase-isomaltase (SI) are then

responsible for the hydrolysis of these oligomers into glucose.13,14 The enzymes

each contain two catalytic subunits: an N-terminal subunit (ntMGAM and ntSI)

near the membrane-bound end of the enzyme and a C-terminal luminal subunit

(ctMGAM and ctSI) (Figure 6-2).15

Figure 6-2: Schematic diagram of MGAM and SI indicating their hydrolytic activities.

Since MGAM and SI genes arose from the duplication of an ancestral

gene,16 the N-terminal catalytic subunits of MGAM and SI are more similar to one

another in sequence, as are the C-terminal domains (60% sequence identity),

than the N- and C-terminal subunits within the same protein.17 There are also

multiple spliceforms of C-terminal MGAM in mammals, two of which are studied

in this paper (ctMGAM-N2 and ctMGAM-N20).17,18

From a therapeutic point of view, it is important to understand the

thermodynamics and kinetics of action of the enzymes involved in the starch

digestion pathway, which will in turn inform the treatment of digestive disorders

and their sequelae. The individual roles of the four catalytic subunits comprising

170

the human intestinal MGAM and SI complexes in the digestion of food starches,

have not been investigated in depth, until recently. Compounds are being

developed to regulate the activities of the catalytic subunits independently. Small

modifications in a compound can have profound effects on its inhibitory

properties and ability to inhibit one catalytic subunit selectively in comparison to

the other enzymes. The characterization of these enzymes, using glucosidase

inhibitors, for example, is an essential prelude to their use in the treatment of

digestive disorders and their sequelae. Certain clinical indications might also

benefit from specific inhibition of one particular enzyme, but not others, to reduce

unwanted side-effects.

The crystal structure of human ntMGAM in complex with inhibitors such as

kotalanol (1) and de-O-sulfonated kotalanol (2)19 (Figure 6-3) had indicated that

removal of the sulfate group did not significantly affect the conformation of the

rest of the polyhydroxylated chain,20 although the C3'-sulfate anion seemed to be

constrained by the hydrophobic residues of the enzyme, and made no hydrogen

bonding interactions with other residues. Removal of the sulfate group allowed

the rest of the polyhydroxylated chain to make optimal contacts with the ntMGAM

active site,20 and resulted in an approximately 7-fold decrease in Ki values.19 In

order to further probe the accommodation by this hydrophobic pocket of other

groups, we synthesized 3'-O-methylponkoranol (3). Disappointingly, the Ki value

(0.50 ± 0.04 µM) was higher than that of 2, indicating that the hydrophobic

interactions between the methyl group and the hydrophobic residues in the active

171

site of ntMGAM are not as optimal as the interactions of the latter groups with the

rest of polyhydroxylated chain in the absence of the methyl ether.21

Very recently Tanabe et al. have synthesized four 3'-O-alkylated salacinol

derivatives (4a–4d), in which the 3'-sulfate moiety was replaced by OCH3,

OC2H5, OC13H27, or OCH2Ph groups, and their inhibitory activities were evaluated

against several disaccharidases (Figure 6-3).22

Figure 6-3: Sulfonium-ion α-glucosidase inhibitors 1-7.

Compared to salacinol (5) and de-O-sulfonated salacinol (6) all the

molecules synthesized (4a-4d) showed equal or higher inhibitory activities (Table

6-1). Compound 4b had the highest inhibitory activity against sucrase and

compound 4d was the most potent inhibitor of maltase and isomaltase among

this type of molecules. The enhanced inhibitory activity of 4d was attributed to

S

OHHO

OSO3 OH

OH OH

HO

OHOH

S

OHHO

OSO3

OH OH

HO

Kotalanol (1)

Salacinol (5)

S

OHHO

OH OH

OH OH

HO

OHOH

CH3OSO3

De-O-sulfonated kotalanol (2)

S

OHHO

OR

OH OH

HOX

3'-O-Alkylated salacinol R-4a: Me, 4b: Et, 4c: C13H27, 4d: Bn

S

OHHO

OMe OH

OH OH

HO

OH

3'-O-Methyl ponkoranol (3)

Cl

S

OHHO

OH

OH OH

HO

De-O-sulfonated salacinol (6)

HCO2 S

OHHO

OH OH

OH OH

HO

OH

Cl

De-O-sulfonated ponkoranol (7)

172

π/π or CH/π interactions of the phenyl ring at 3' with surrounding aromatic

residues of the active site in the enzyme. It is clear from this work and our

previous work that subtle structural modifications of the ligands can differentiate

between the active sites of the various glucosidase enzymes.

Table 6-1: IC50 (µM) against disaccharidases.

In the present work, we report the glucosidase inhibitory activity of 3'-O-

methylponkoranol (3) against recombinant human maltase-glucoamylase

(ntMGAM and ctMGAM) and sucrase-isomaltase (ntSI and ctSI), in comparison

with the parent compound, de-O-sulfonated ponkoranol (7) (Figure 6-3).24

Compound Sucrase Maltase Isomaltase

4a 0.46 5.3 0.39

4b 0.12 1.7 0.27

4c 1.3 1.0 0.95

4d 0.32 0.44 0.14

5 1.619 5.223 1.323

6 1.319 8.023 0.323

173

6.4 Results and discussion

Enzyme kinetics. Kinetic parameters of the four domains were determined

by measuring the amount of glucose produced upon the addition of enzyme at

increasing maltose concentrations (from 0.5 to 30 mM) in the presence of

increasing inhibitor concentration (0-200 nM) by a two-step glucose oxidase

assay. The methods used for kinetic assays were reported previously.25,26

Reactions were performed in quadruplicate and absorbance measured at 405 nM

with a SpectraMax 190 Plate Reader (Molecular Devices). Absorbance readings

were averaged to give the final value, which was compared to a glucose standard

curve to determine the amount of glucose released by the enzyme from the

substrate. The program KaleidaGraph4.1 was used to fit the data to the

Michaelis-Menten equation and estimate Km, Kmobs (Km in the presence of the

inhibitor) and Vmax of the catalytic subunits. Ki values for each inhibitor were

determined using the equation Ki = [I]/(Kmobs/Km)-1). The Ki values reported for

each inhibitor were determined by averaging the Ki values from three different

inhibitor concentrations. The data were also plotted on Lineweaver-Burk plots to

verify that the inhibitors were acting as competitive inhibitors (Figure 6-4).

174

Figure 6-4: Representative Lineweaver-Burk plot of ctMGAM-N2 inhibited by 3 at concentrations of 0 nM, 75 nM, 125 nM, and 200 nM.

Kinetic Analysis. Inhibition constants (Ki), calculated for each catalytic

subunit with compound 3, as described above, are shown in Table 6-2.

Compound 3 was a potent inhibitor of all the catalytic subunits. With regard to

MGAM, it was found to inhibit both ctMGAM spliceforms similarly, demonstrating

inhibitory activity in the nanomolar range for both ctMGAM-N2 (Ki = 0.060 ± 0.015

µM) and ctMGAM-N20 (Ki = 0.055 ± 0.014 µM). Unlike the other catalytic

subunits, compound 3 is a weak inhibitor of ntMGAM (Ki = 0.5 ± 0.04 µM);21 it is,

therefore, ten times more potent an inhibitor of ctMGAM-N2 and ctMGAM-N20

175

than ntMGAM. With respect to ctSI, Compound 3 exhibited remarkable inhibitory

activity in the nanomolar range (Ki = 0.007 ± 0.002 µM). This compound

demonstrates the most potency against ctSI, approximately seventy times more

active than with ntMGAM. Compound 3 inhibited ntSI in the nanomolar range as

well, with a Ki value of 0.035 ± 0.013 µM. Furthermore, there is some

differentiation between ctSI and ntSI as compound 3 is five times more potent

against ctSI compared to ntSI. These results demonstrate the variation in

biochemical and structural properties of the enzymes despite their similarity in

amino acid sequence.

Finally, we compare the inhibitory activities of compounds 3 and de-O-

sulfonated pokoranol (7) against ntMGAM, ctMGAM, ntSI, and ctSI. The results

(Table 6-2) indicate that addition of the methyl group at the 3' position of de-O-

sulfonated ponkoranol (7) results significant inhibition of ctMGAM-N2 (Ki = 0.060

± 0.015 µM), in contrast to 7 which does not inhibit ctMGAM-N2.24

Table 6-2: Experimentally determined Ki values (µM) of 3 and 7.a

a analysis of inhibition was performed with maltose as the substrate.

ctMGAM-N2 ctMGAM-N20 ntMGAM ctSI ntSI

3 0.060 ± 0.015 0.055 ± 0.014 0.50 ± 0.0421 0.007 ± 0.002 0.035 ± 0.013

724 No Inhibition 0.096 ± 0.015 0.043 ± 0.001 0.103 ± 0.037 0.302 ± 0.123

176

Both compounds 3 and 7 inhibited ctMGAM-N20 with similar Ki values.

However, an order of magnitude greater inhibition of ntSI by 3 over 7 was

observed. Strikingly, compound 3 inhibited ctSI with Ki values of 7 ± 2 nM,

significantly lower than that (103 ± 30)24 for de-O-sulfonated ponkoranol (7) itself

(Table 6-2). We note that 3 is the most potent compound against ctSI to date in

this class of molecules. We speculate that ctSI will have a hydrophobic pocket in

the catalytic site that better accommodates the methyl group and provides

favorable hydrophobic interactions, relative to the other enzymes in which steric

interactions between the methyl group and active-site residues might lead to

unfavorable contacts. Such unfavorable contacts are also seen in the binding of

the sulfated derivative, kotalanol (1) relative to its de-O-sulfonated analogue

(2).20

6.5 Acknowledgments

We thank Buford L. Nichols, Roberto Quezada-Calvillo and Hassan Naim

for reagents for recombinant expression, and Lyann Sim for MGAM and SI

enzyme purification. K. J. was supported by a scholarship from the Canadian

Institutes of Health Research (CIHR) and the Canadian Digestive Health

Foundation. We also thank CIHR (MOP111237), NSERC, and the Heart and

Stroke Foundation of Ontario (NA-6305) for financial support.

177

6.6 References

1. Peters, A.; Schweiger, U.; Pellerin, L.; Hubold, C.; Oltmanns, K. M.;

Conrad, M.; Schultes, B.; Born J.; Fehm, H. L. Neurosci. Biobehav. Rev.

2004, 28, 143.

2. Livesey, G. Proc. Nutr. Soc. 2005, 64, 105.

3. McCall, A. L. Eur. J. Pharmacol. 2004, 490,147.

4. Corring, T. Reprod. Nutr. Dev. 1980, 20, 1217.

5. Jones, B. J.; Brown, B. E.; Loran, J. S.; Edgerton, D.; Kennedy J. F.;

Stead, J. A.; Silk D. B. A. Gut 1983, 24, 1152.

6. Lee, P.C. J. Pediatr. Gastroenterol. Nutr. 1983, 2, S227.

7. Fujita, S.; Fuwa, H. J. Nutr. Sci. Vitaminol. (Tokyo) 1984, 30,135.

8. Gray, G. M. J. Nutr. 1992, 122, 172.

9. Gray, G. M. N. Engl. J. Med. 1975, 292, 1225.

10. Brayer, G. D.; Sidhu, G.; Maurus, R.; Rydberg, E. H.; Braun, C.; Wang, Y.;

Nguyen, N. T.; Overall, C. M.; Withers, S. G. Biochemistry 2000, 39, 4778.

11. Horvathova, V.; Janecek, S.; Sturdik, E. Gen. Physiol. Biophys. 2001, 20,

7.

12. Robyt, J.; French, D. J. Biol. Chem. 1970, 15, 3917.

13. Van Beers, E. H.; Buller, H. A.; Grand, R. J.; Einerhand, A. W.; Dekker, J.

Crit. Rev. Biochem. Mol. Boil. 1995, 30, 197.

14. Semenza, G. Ann. Rev. Cell Biol. 1986, 2, 255.

15. Ernst, H. A.; Leggio, L. L.; Willemoës, M.; Leonard, G.; Blum, P.; Larsen,

S. J. Mol. Biol. 2006, 358, 1106.

178

16. Quezada-Calvillo, R.; Sim, L.; Ao, Z.; Hamaker, B. R.; Quaroni, A.;

Brayer, G. D.; Sterchi, E. E.; Robayo-Torres, C. C.; Rose, D. R.; Nichols,

B. L. J. Nutr. 2008, 138, 685.

17. Jones, K.; Sim, L.; Mohan, S.; Jayakanthan, K.; Liu, H.; Avery, S.; Naim,

H. H.; Quezada-Calvillo, R.; Nichols, B. L.; Pinto B.M.; Rose, D. R. Bioorg.

Med. Chem. 2011, 19, 3929.

18. Naumoff, D. G. Mol. Biol. (Mosk). 2007, 41, 1056.

19. Jayakanthan, K.; Mohan, S.; Pinto, B. M. J. Am. Chem. Soc. 2009, 5621.

20. Sim, L.; Jayakanthan, K.; Mohan, S.; Nasi, R.; Johnston, B. D.; Pinto, B.

M.; Rose, D. R. Biochemistry 2010, 49, 443.

21. Eskandari, R.; Jones, K.; Rose, D. R.; Pinto, B. M. Bioorg. Med. Chem.

Lett. 2010, 20, 5686.

22. Tanabe, G.; Otani, T.; Cong, W.; Minematsu, T.; Ninomiya, K.; Yoshikawa,

M.; Muraoka, O. Bioorg. Med. Chem. Lett. 2011, 21, 3159.

23. Yoshikawa, M.; Xu, F.; Nakamura, S.; Wang, T.; Matsuda, H.; Tanabe, G.;

Muraoka, O. Heterocycles 2008, 75, 1397.

24. a) Eskandari, R.; Kuntz, D. A.; Rose, D. R.; Pinto, B. M. Org. Lett. 2010,

12, 1632; b) Eskandari, R.; Jones, K.; Rose, D. R.; Pinto, B. M. Chem.

Commun. 2011, 47, 9134.

25. Sim, L.; Willemsma, C.; Mohan, S.; Naim, H. Y.; Pinto, B.M.; Rose, D. R.

J. Biol. Chem. 2010, 285, 17763.

26. Nasi, R.; Sim, L.; Rose, D. R.; Pinto, B. M. J. Org. Chem. 2007, 72, 180.

179

CHAPTER 7: PROBING THE INTESTINAL α-GLUCOSIDASE ENZYME SPECIFICITIES OF STARCH-

DIGESTING MALTASE-GLUCOAMYLASE AND SUCRASE-ISOMALTASE: SYNTHESIS AND INHIBITORY PROPERTIES OF 3′- AND 5′- MALTOSE-EXTENDED DE-

O-SULFONATED PONKORANOL

This Chapter comprises the manuscript “Probing the intestinal α-glucosidase enzyme specificities of starch-digesting maltase-glucoamylase and

sucrase-isomaltase: Synthesis and inhibitory properties of 3′- and 5′- maltose-extended de-O-sulfonated ponkoranol” which was published in

Chemistry A European Journal (2011, 17, 14817-14825).

Razieh Eskandari,a Kyra Jones,b Kongara Ravinder Reddy, a Kumarasamy

Jayakanthan, a Marcia Chaudet, b David R. Rose,b B. Mario Pintoa

aDepartment of Chemistry, Simon Fraser University, Burnaby, British Columbia,

Canada V5A 1S6

b Department of Biology, University of Waterloo, Waterloo, Ontario,

Canada N2L 3G1

180

An approach to controlling blood glucose level in individuals with type-2

diabetes is to inhibit intestinal glucosidases using α-glucosidase inhibitors.

Maltase-glucoamylase and sucrase-isomaltase are two intestinal glucosidases

that can be targeted for this purpose. The design and synthesis of specific

domain inhibitors that exploit differences between MGAM and SI subsites will be

helpful to reduce unwanted side effects. The design of domain-specific inhibitors

will also aid in the characterization of these domains and help to better define the

complementary roles of MGAM and SI in the process of terminal starch digestion.

In this Chapter, C-3′- and C-5′-β-O-maltose-extended de-O-sulfonated

ponkoranol were synthesized to probe the intestinal α-glucosidase enzyme

specificities of MGAM and SI. We have evaluated the inhibitory activities of

these compounds against recombinant human maltase-glucoamylase (ntMGAM

and ctMGAM) and sucrase-isomaltase (ntSI and ctSI).

The thesis author performed all the experimental synthetic work and

characterization of the C-3′-β-maltose-extended de-O-sulfonated ponkoranol. Dr.

Kongara Ravinder Reddy performed part of the experimental synthetic work and

characterization of the C-5′-β-maltose-extended de-O-sulfonated ponkoranol

compound. Dr. Jayakanthan Kumarasamy assisted in the synthetic design and

part of the experimental synthetic work and characterization of the C-5′-β-O-

maltose-extended de-O-sulfonated ponkoranol compound. The thesis author

wrote and edited the manuscript with Dr. B. Mario Pinto. Mrs. Kyra Jones, Ms.

Marcia Chaudet and Dr. David R. Rose performed the enzyme inhibition studies.

181

Graphical abstract:

7.1 Keywords

Intestinal glucosidase inhibition, maltose-extended-de-O-sulfonated

ponkoranol, sulfonium-ions, glucosidase inhibitors.

7.2 Abstract

The synthesis and glucosidase inhibitory activities of two C-3′- and C-5′-β-

maltose-extended de-O-sulfonated analogues of the naturally-occurring

sulfonium-ion inhibitor, de-O-sulfonated ponkoranol, are described. The

compounds are designed to test the specificity towards four intestinal glycoside

hydrolase family 31 enzyme activities, responsible for the hydrolysis of terminal

starch products and sugars into glucose, in humans. The target sulfonium-ion

compounds were synthesized by means of nucleophilic attack of benzyl

protected 1,4-anhydro-4-thio-D-arabinitol at the C-6 position of 6-O-

trifluoromethanesulfonyl trisaccharides as alkylating agents. The alkylating

S

OHHO

OH O

OH OH

HO

OH

S

OHHO

OH

OH OH

HO

OH

O

O OH

OH

O OH

O

HOOH

HO

O OH

OH

O OH

O

HOOH

HO

Cl

HO HO

Cl

182

agents were synthesized from D-glucose via glycosylation at C-4 or C-2 with

maltosyl trichloroacetimidate. Deprotection of the coupled products using a two-

step sequence, followed by reduction, afforded the final compounds. Evaluation

of the target compounds for inhibition of the four glucosidase activities indicated

that selective inhibition of one enzyme over the others is possible.

7.3 Introduction

Type-2 (noninsulin-dependent) diabetes, is a metabolic disorder

characterized by elevated blood glucose levels with defects in insulin secretion,

insulin action or both. In the treatment of type-2 diabetes, controlling blood

glucose levels is critical. One strategy is to slow down the breakdown of ingested

carbohydrates and starches and thus delay glucose absorption by inhibiting the

enzymes that are involved in the breakdown of dietary starches and sugars into

glucose. In humans, six enzyme activities are involved in the complete digestion

of dietary starches and sugars into glucose. Two endohydrolases, salivary and

pancreatic α-amylases are responsible for digestion of starch into shorter linear

and branched dextrin chains and four exohydrolase α-glucosidase activities,

maltase-glucoamylase (MGAM) and sucrose-isomaltase (SI), are responsible for

the hydrolysis of terminal starch products and sugars into glucose.1,2 MGAM and

SI are anchored to the brush-border epithelial cells of the small intestine each

containing two catalytic subunits classified under glycosyl hydrolase family 31

(GH31): an N-terminal subunit (ntMGAM and ntSI) near the membrane bound

end of the enzyme and a C-terminal luminal subunit (ctMGAM and ctSI) (Figure

7-1).3

183

Figure 7-1: Schematic diagram of MGAM and SI indicating hydrolytic activity.

SI exhibits hydrolytic activity on branched α-1,6 linkages, complemented

by the hydrolytic activity of both SI and MGAM on α-1,4 linkages.4 These

complementary activities of the human enzymes permit digestion of starches of

plant origin comprising two-thirds of most diets; however, the main substrate of SI

is that with α-1,4 linkages.5

SI is more abundant than MGAM; however, to counteract this deficit in

abundance, MGAM displays higher hydrolytic activities.6-8 With respect to

similarity in sequence, SI and MGAM show 59% amino acid sequence similarity.

The catalytic subunits of MGAM and SI are 40-60% identical in amino acid

sequence.3 N-terminal catalytic subunits and the C-terminal catalytic subunits of

MGAM and SI are more closely related in sequence to one another than the N-

and C-terminal subunits within the same protein.9 There are also multiple

spliceforms of C-terminal MGAM in mammals, two of which are studied in this

paper (ctMGAM-N2 and ctMGAM-N20).9,10 In recent years, the aqueous extracts

of the plant Salacia reticulata found in Sri Lanka and Southern India, have been

used by patients as a remedy for the treatment of type-2 diabetes.11 The active

184

compounds of Salacia reticulata were found to be a novel class of sulfonium-ion

glucosidase inhibitors, including salaprinol (1),12b salacinol (2),13 ponkoranol

(3),12a,b kotalanol (4),14 de-O-sulfonated kotalanol (5),14b, 15 de-O-sulfonated

salacinol (6),16 de-O-sulfonated ponkoranol (7)17a,b and de-O-sulfonated

salaprinol (8)17b (Figure 7-2), whose structures comprise a 1,4-anhydro-4-thio-D-

arabinitol core and polyhydroxylated acyclic chain (Figure 7-2).

Figure 7-2: Components 1-8 isolated from Salacia species.

S

OHHO

OSO3 OH

OH OH

HO

OHOH

S

OHHO

OSO3 OH

OH OH

HO

OH

S

OHHO

OSO3

OH OH

HOS

OHHO

OSO3

OH

HO

Kotalanol (4)

Salacinol (2) Ponkoranol (3)Salaprinol (1)

S

OHHO

OH OH

OH OH

HO

OHOH

CH3OSO3

De-O-sulfonated Kotalanol (5)

S

OHHO

OH

OH OH

HOHCO2

De-O-sulfonated Salacinol (6)

S

OHHO

OH OH

OH OH

HO

OH

De-O-sulfonated Ponkoranol (7)

S

OHHO

OH

OH

HO

De-O-sulfonated Salaprinol (8)

ClCl

185

They have been shown to be stronger inhibitors of ntMGAM, with Ki values

in the sub micromolar range (i.e., 0.03-0.19 µM) compared to acarbose (Ki =

62±13 µM), the naturally-occurring α-glucosidase inhibitor currently in use for the

treatment of type-2 diabetes.18-23 Interestingly, de-O-sulfonated ponkoranol (7)

was one of our synthetic compounds,17a isolated recently from the same plant.17b

Acarbose (9) (Figure 7-3) has been shown to be an efficient inhibitor of α-

amylase with a reported Ki of 15 nM24 and of the C-terminal domain of MGAM-N2

(Ki = 0.009±0.002 µM),4, 9 but a weaker inhibitor of the N-terminal domain (Ki =

62±13 µM) (Table 7-1).4, 19

Figure 7-3: Structure of acarbose 9, an α-glucosidase inhibitor currently used in the treatment of type-2 diabetes.

The study of ntMGAM in complex with acarbose had indicated that

acarbose was bound to the ntMGAM active site primarily through side chain

interactions with its acarvosine unit and almost no interactions were made with its

other sugar rings. Since it is suggested that additional subsite interactions with

the acarbose sugar rings would significantly increase its inhibitory properties for

OHN

HO

OOH

H3C

O

OOH

HO

O

HO OH

HO

HOHO

OH

HO

HO

OH9

Acarvosine

Maltose

186

ctMGAM4, 25 it is of interest to examine whether glucose residues appended to the

polyhydroxylated carbon chain of salacinol-based compounds would lead to

differential inhibition of the four enzyme activities. We have shown recently that

extension of the acyclic carbon chain beyond six carbons in salacinol-based

compounds is not essential for activity; and furthermore that the de-O-sulfonated

analogues were more potent inhibitors than the parent compounds.17a Therefore,

de-O-sulfonated ponkoranol has been chosen for modification at C-3′-OH and C-

5′-OH of its side chain in our preliminary studies. We report here the synthesis of

C-3′-β-maltose-extended de-O-sulfonated ponkoranol analogue 10 and C-5′-β-

maltose extended-de-O-sulfonated ponkoranol 11 (Figure 7-4) to investigate the

individual roles of the four catalytic subunits comprising human intestinal MGAM

and SI. These candidates were chosen to test whether occupying the subsite

with glucosyl units would suffice, irrespective of their orientation (reducing or non-

reducing positional substitution) or α- or β-stereochemistry at the anomeric

centre. The β-isomers were chosen as initial candidates because their syntheses

were more tractable.

187

Figure 7-4: 3′-O-β-maltosyl-de-O-sulfonated ponkoranol 10 and 5′-O-β-maltosyl-de-O-sulfonated ponkoranol 11.

7.4 Results and discussion

Synthesis. Retrosynthetic analysis (Scheme 7-1) revealed that the target

molecules could be synthesized by alkylation at the sulfur atom of suitably

protected alkylating agents. The alkylating agents (B) could be obtained by

glycosylation of protected D-glucose (C) at C-4 or C-2 with maltosyl

trichloroacetimidate (12). The required maltosyl trichloroacetimidate donor (12)

could be prepared from maltose according to the literature procedure.26

S

OHHO

OH O

OH OH

HO

OH

S

OHHO

OH

OH OH

HO

OH

O

O OH

OH

O OH

O

HOOH

HO

O OH

OH

O OH

O

HOOH

HO

X

HO HO

X

10 11

188

Scheme 7-1: Retrosynthetic analysis.

To synthesize the desired β-(1→4) and β-(1→2)- maltosyl-linked

glucopyranosides, we chose to use benzyl 2,3-O-[(2R,3R)-2,3-dimethoxybutane-

2,3-diyl]-6-O-tert-butyldimethylsilyl-β-D-glucopyranoside (14) and benzyl 3,4-O-

[(2R,3R)- 2,3 -dimethoxybutane- 2,3 -diyl]- 6 -O- tert -butyldimethylsilyl- β - D -

glucopyranoside (19) as glycosyl acceptors. The acceptors 14 and 19 were

synthesized by selective protection of their primary hydroxyl groups as tert-

butyldimethylsilyl (TBDMS) ethers from butane-2,3-diacetal (BDA) protected

glucopyranosides 13 and 18 which were readily obtained from benzyl β-D-

S

OPPO

POS

OHHO

OH OH

HO

OH

X5'

O

R2O

R1O

PO

L

OP

OOAcO

OC(NH)CCl3AcO

OAc

D-Glucose Maltose

3'

OR2OR1

10. R1 = Maltose, R2 = H11. R1 = H, R2 = Maltose

R1 = Maltose, R2 = P R1 = P, R2 = Maltose

L = leaving groupP = protecting group

O

R2O

R1O

PO

OTBDS

OP

A B

C 12

OAcOAcO

AcO

OAc

189

glucopyranoside by a known procedure.27 The syntheses of the desired

trisaccharides 15 and 20 were achieved by treatment of maltosyl

trichloroacetimidate 12 with the acceptors 14 and 19 in dichloromethane, using a

catalytic amount of boron trifluoride etherate as promoter at room temperature to

give the compounds maltosyl-(1 4)-β- linked 6-O-tert-butyldimethylsilyl-D-

glucopyranoside 15 and maltosyl-(1 2)-β-linked 6-O-tert-butyldimethylsilyl-D-

glucopyranoside 20 in 45% and 52% yield, respectively. The O-deacetylation of

15 and 20 was carried out using NaOMe, followed by protection of the hydroxyl

groups as benzyl ethers to afford benzyl-protected maltosyl-(1 4)-β- linked 6-O-

tert-butyldimethylsilyl-D-glucopyranoside 16 and benzyl protected maltosyl-(1

2)-β- linked 6-O-tert-butyldimethylsilyl-D-glucopyranoside 21. 6- O- Desilylation

of trisaccharides 16 and 21 with TBAF in THF and subsequent triflation of the

free hydroxyl group afforded maltosyl-(1 4)-β- linked 6-O-

trifluoromethanesulfonyl-D-glucopyranoside 17 and maltosyl-(1 2)-β- linked 6-O-

trifluoromethanesulfonyl-D-glucopyranoside 22 in 41% and 86% yield,

respectively (Scheme 7-2).

190

Scheme 7-2: Synthesis of benzyl 6-O-trifluoromethanesulfonyl-D-glucopyranoside derivatives 17 and 22, with benzylated maltose units at C-2 or C-4.

OAcO

AcO

OOAc O

O

OAc

AcO

AcOHO

O

OO

HO

OBn

OMe

OMe

TBDMSCl, Imid. DMF, 0 oC-RT

HOO

OO

TBDMSO

OBn

OMe

OMe

12, BF3.Et2O, DCM,

4oA, 30 min.

O

OO

OTBDMS

OBn

OMe

OMe

OBnOBnO

OOBn O O

OBn

BnO

BnO

O

OO

TBDMSO

OBn

OMe

OMe

1. NaOMe, MeOH, RT, 24h

2. BnBr, NaH, DMF, RT, 2h,

1. TBAF, THF, 12h, RT

2. Tf2O, DCM, -10oC, pyridine

OBnO

BnO

OOBn O

O

OBn

BnO

BnO

O

OO

OTf

OBn

OMe

OMe

13 14 15

16 17

OAc

OBnOBn

O

OAcOAc

O

OAc

O OAc

AcOOAc

O

OH

OH

OBn

TBDMSCl, Imid. DMF, 0 oC-RT

12, BF3.Et2O, DCM,

4oA, 30 min.

1. NaOMe, MeOH,

RT, 24h

2. BnBr, NaH, DMF,

RT, 1h,

1. TBAF, THF,

4h, RT

2. Tf2O, DCM,

-10oC, pyridine

18

20 21 22

OO

OMe

OMe

O

OH

OTBDMS

OBn

19

OO

OMe

OMe

O

O

OTBDMS

OBnOO

OMe

OMe

O

OBnOBn

O

OBn

O OBn

BnOOBn

O

O

OTBDMS

OBnOO

OMe

OMe

O

OBnOBn

O

OBn

O OBn

BnOOBn

O

O

OTf

OBnOO

OMe

OMe

AcO BnO BnO

191

The coupling reaction of the benzyl-protected anhydrothioarabinitol 2328

with 6-O-trifluoromethanesulfonyl trisaccharides 17 and 22 as alkylating agents

was carried out in dichloromethane at room temperature to give the

corresponding sulfonium-ions 24 and 25, respectively, as a 5:1 and 2.8:1 mixture

of diastereomers at the stereogenic sulfur centre (Scheme 7-3).

Scheme 7-3: Coupling reactions to give sulfonium-ions.

SBnO

BnO BnO

OOBnO

OBn

OBnOBnO

BnOOBn

OBn

OOO

OOMe

OMe

SBnO

BnOBnO

OTf

17

S

OBnBnO

BnO

DCM, 24h., RT, 95%

O

O

OO

OMe

OMe

OBn

O

OOBn

OBn

BnO

O

OBnOBn

OBn

BnO

OTf

OBn

22

S

OBnBnO

BnO

DCM, 24h., RT, 90%

24 d.r, 5:1

25a:25bd.r, 2.8:1

192

The diastereoisomers of 24 and 25 were found to be inseparable, even by

HPLC, although repeated attempts to separate the mixture of diastereomers of

25 were partially successful. Prior to choosing 6-O-trifluoromethanesulfonyl

trisaccharides 17 and 22 as alkylating agents, the coupling reaction of 6-O-p-

toluenesulfonyl trisaccharides was carried out in 1,1,1,3,3,3-hexafluoro-2-

propanol (HFIP), based on the procedure that has been reported for synthesis of

3′-O-methylponkoranol (26) (Figure 7-5)29 but the reaction failed and no product

formation was observed. Temperature variation to -78 °C and -25 °C failed to

give a single isomer.

Figure 7-5: Structure of 3'-O-methylponkoranol 26.

The reaction was also carried out at 0 °C and reflux conditions in

dichloromethane but the same ratio was observed. Deprotection of the coupled

products 24 and 25 was carried out by a two-step procedure, first hydrogenolysis

of the benzylethers followed by BDA deprotection. The corresponding crude

materials were then treated with Amberlyst A-26 resin (chloride form) to

completely exchange their triflate counterions to chloride ions.29 Finally, the

corresponding hemiacetals were reduced with NaBH4 in water to provide the

desired C-3′-β-maltose-extended de-O-sulfonated ponkoranol 10 (dr, 5:1) and C-

S

OHHO

OMe OH

OH OH

HO

OH

Cl

26

193

5′-β- maltose-extended de-O-sulfonated ponkoranol 11 (dr, 2.8:1) in 70% and

64% yield, respectively, over four steps (Scheme 7-4).

Scheme 7-4: Synthesis of compounds 10 and 11.

The absolute stereochemistry at the stereogenic sulfur center for the major

isomer of 24 was established by means of a 2D-NOESY experiment (Figure 7-5).

A NOESY correlation was observed between the H-4 proton and H-6′a, implying

that these two hydrogens are syn-facial with respect to the sulfonium ring. In the

case of 25, the absolute configuration at the sulfonium centre for the major

isomer was assigned to be the same as in 24, since a NOESY experiment was

not possible owing to overlapping signals for H-4 and H-6′.

1. Pd(OH)2, 80% AcOH:H2O

2. 85% TFA3. Amberlyst A-26,H2O, 3 h4. NaBH4, 70%

24

1. Pd(OH)2, 80% AcOH:H2O

2. 85% TFA3. Amberlyst A-26,H2O, 3 h4. NaBH4, 64%

25S

OHHO

OH O

OH OH

HO

OH

S

OHHO

OH

OH OH

HO

OH

O

O OH

OH

O OH

O

HOOH

HO

O OH

OH

O OH

O

HOOH

HO

Cl

HO HO

Cl

10 11

194

Figure 7-6: NOESY correlations in compound 24.

Enzyme kinetics. Kinetic parameters were determined by measuring the

amount of glucose produced upon the addition of enzyme at increasing maltose

concentrations (from 0.5 to 30 mM) in the presence of increasing inhibitor

concentration (0-200 nM) by a two-step glucose oxidase assay in a 96-well plate.

The enzyme was allowed to act on the maltose substrate in the presence of

inhibitor for 45 minutes at 37°C. The reactions were then quenched with Tris-HCl

to a final concentration of 1M. Glucose oxidase reagent (Sigma-Aldrich) was then

added to each well (125µl) and the reactions were allowed to develop for 30

minutes at 37°C. Reactions were performed in quadruplicate and absorbance

measured at 405 nM by a SpectraMax 190 Plate Reader (Molecular Devices).

Absorbance readings were averaged to give the final value, which was compared

to a glucose standard curve to determine the amount of glucose released by the

enzyme from the substrate. The program KaleidaGraph4.1 was used to fit the

data to the Michaelis-Menton equation and estimate Km, Kmobs (Km in the

presence of the inhibitor) and Vmax of the catalytic subunits. Ki values for each

inhibitor were determined using the equation Ki = [I]/(Kmobs/Km)-1). The Ki values

OOBnO

OBn

OBnOBnO

BnOOBn

OBn

OOH2C

O

OOMe

OMe

SBnO

BnOOBn

OTf

OBn

24

H4

6'

195

reported for each inhibitor were determined by averaging the Ki values from three

different inhibitor concentrations. The weights of compounds 10 and 11 were

adjusted for the presence of major isomer. The data were also plotted on

Lineweaver-Burk plots to verify that the inhibitors were acting as competitive

inhibitors (see Supporting Information). The methods used for kinetic assays

were reported previously.30, 31

Kinetic Analysis. The inhibition constants (Ki) of acarbose,19, 25, 30 C-3′-β-

maltose-extended de-O-sulfonated ponkoranol 10 and C-5′-β-maltose-extended

de-O-sulfonated ponkoranol 11 against MGAM and SI were determined using the

glucose oxidase assay and maltose as a substrate. There are also multiple,

alternatively-spliced isoforms of ctMGAM; the two studied in this context, referred

to as ctMGAM-N2 and ct-MGAM-N20, occur in humans.9 The experimentally

determined inhibition constants for acarbose,19, 25, 30 10, and 11 are listed Table

7-1. The current anti-diabetic compound, acarbose, was found to be a

micromolar inhibitor of both ntMGAM (Ki = 62 ±13 µM)19 and ntSI (Ki = 14 ± 1

µM).30 In contrast, acarbose is a 1000-fold better inhibitor of ctMGAM-N2 (Ki =

0.009 ± 0.002 µM), ctMGAM-N20 (Ki = 0.028 ± 0.005 µM) and 100-fold better of

ctSI (Ki = 0.246 ± 0.005 µM). C-3′-β-maltose-extended de-O-sulfonated

ponkoranol 10, generally shows some selectivity for inhibiting N-terminal MGAM,

as compared to C-terminal enzymes: ntMGAM (Ki = 0.039 ± 0.025) over

ctMGAM-N2 (Ki = No Inhibition) and ctMGAM-N20 (Ki = 0.655 ± 0.063µM), ctSI

(Ki = 0.062 ± 0.005µM) and ntSI (Ki = 0.046 ± 0.018µM). Surprisingly, C-5′-β-

maltose-extended de-O-sulfonated ponkoranol 11, shows improved inhibition of

196

ntMGAM (Ki = 0.008 ± 0.002 µM), with little distinction between the other catalytic

enzyme subunits: ctMGAM-N20 (Ki = 0.067 ± 0.012 µM), ctSI (Ki = 0.045 ± 0.001

µM), ctMGAM-N2 (Ki = 0.077 ± 0.015), and ntSI (Ki 0.019 ± 0.008 µM).

In this study, we report the design, synthesis, and glucosidase inhibitory

activities of two ponkoranol-based compounds 10 and 11 against recombinant

human maltase-glucoamylase (ntMGAM and ctMGAM) and sucrase-isomaltase

(ntSI and ctSI). These compounds were intended to probe whether one could

differentiate between the different enzymes, i.e. to toggle between the different

enzymes even though they are not strictly analogous to acarbose. Kinetic

analysis confirmed the enzyme activity of the recombinant proteins, and inhibition

analysis confirmed classic competitive inhibition by the α-glucosidase inhibitors.

The results (Table 7-1) indicate that despite the overall similarity between the

subunits in terms of amino-acid sequence, they exhibit different biochemical and

structural properties. Compound 11 demonstrates the ability to inhibit all of the

catalytic subunits very well, especially ntMGAM. In fact, compound 11 is the most

potent inhibitor of ntMGAM to date. Compound 10 is also a good inhibitor of ctSI,

ntSI, and ntMGAM, with Ki values in the nanomolar range. Although elongation of

the scaffold does not result in a significant gain in binding energy, nonetheless it

does result in some interesting selectivities in inhibitory activities. A significant

finding is that 10 differentiates between the two C-terminal catalytic subunits.

Compound 10 is a very poor inhibitor of ctMGAM-N20 and shows no inhibition

against ctMGAM-N2. By confirming the higher potency of 10 against other

subunits as compared with ctMGAM, our results suggest that this inhibitor shows

197

specificity towards these different α-glucosidases. Since ctMGAM is expressed in

more than one spliceform,9 it provides further complexity to the system. It is

hypothesized that these units act in a complementary manner depending on the

organism’s nutritional sources and requirements, as well as its physical and

environmental conditions. Therefore, with compound 10, we are able to keep the

ctMGAM activity on and dampen the others, and we are in a position to test this

hypothesis. According to previous studies, acarbose is a very powerful inhibitor of

ctMGAM over ntMGAM and SI subunits.9 Therefore, we are now also in a

position to turn off the ctMGAM unit and study the effect of others subunits in

starch digestion. The results also demonstrate that the ability to selectively inhibit

one enzyme unit over the others can result from relatively small changes in the

structure of the compound and we are closer to being able to independently

toggle each subunit on and off. This observation is important clinically because

the design of α-glucosidase inhibitors for the treatment of type-2 diabetes might

require specificity for enzymes later in the starch digestion pathway in order to

reduce unwanted side-effects. The initial results of the inhibition assays are

promising; at this point, analysis of structure-activity relationships can only be

somewhat speculative. However, the very interesting and unprecedented enzyme

selectivity observed in this study set the stage for improvement of the specificity

and affinity of these compounds for their potential development as antidiabetic

agents, irrespective of whether or not the binding of these compounds occurs

through a mechanism analogous to that of acarbose. It is noteworthy that de-O-

sulfonated ponkoranol 7 (Figure 7-2) 32 and 3’-O-methylponkoranol (26) (Figure

198

7-5)33 also show an interesting selectivity profile for the different subunits (Table

7-1). Further confirmation of the importance of inhibitor structure and how it

affects binding in the ctMGAM active site will be possible with an analysis of the

atomic structure of the ctMGAM binding site in the presence of bound inhibitor.

Determination of this structural information will be a valuable tool in future design

and synthesis of α-glucosidase inhibitors effective against and specific to MGAM

and SI subunits. These inhibitors should be promising lead candidates as oral

agents for the treatment and prevention of type-2 diabetes.

Table 7-1: Comparison of inhibition profiles against MGAM and SI subunits, Ki (µM).

ctMGAM-N2 ctMGAM-N20 ntMGAM ctSI ntSI

732 no Inhibitiona 0.096 ± 0.015 0.043 ± 0.001 0.103 ± 0.037 0.302 ± 0.123

9 0.009±0.002 0.028±0.005 62±13 0.246±0.005 14±1

10 no Inhibitiona 0.655±0.063 0.039±0.025 0.062±0.005 0.046±0.018

11 0.077 ± 0.015 0.067±0.012 0.008±0.002 0.045±0.001 0.019±0.008

2633 0.060 ± 0.015 0.055 ± 0.014 0.50 ± 0.0429 0.007 ± 0.002 0.035 ± 0.013

a at 200 nM

199

7.5 Experimental

7.5.1 Compound characterization data

7.5.1.1 Benzyl 6-O-tert-butyldimethylsilyl-2,3-O-[(2R,3R)-2,3-dimethoxybutane-2,3-diyl]-β-D-glucopyranoside 14.

To a solution of 13 (500 mg, 1.3 mmol) in DMF (15 mL) was added

imidazole (270 mg, 3.9 mmol). The reaction was cooled in an ice bath, TBDMSCl

(226 mg, 1.4 mmol) was added portionwise, and the mixture was stirred at 0 °C

for 15 min and at room temperature 1 h. The reaction was quenched by the

addition of ice-water, and the reaction mixture was extracted with Et2O (3×30

mL). The combined organic solvents were washed with water (15 mL) and brine

(15 mL), dried (Na2SO4), and concentrated to give the crude. The crude was

purified by column chromatography (Hexanes/EtOAc 2:1) to afford 14 as foam

(564 mg, 87%). [α] 23D = - 124.5 (c = 2.3, CH2Cl2). 1H NMR (CDCl3) δ 7.28-7.15

(5H, m, Ar), 4.79, 4.55 (2H, 2d, JA,B = 12.3 Hz, CH2Ph), 4.51 (1H, d, J1,2 = 8.0 Hz,

H-1), 3.83 (1H, dd, J6a,6b = 5.5, J5,6a = 10.3 Hz, H-6a), 3.75 (1H, dd, J6a,6b = 5.8,

J5,6b = 10.7 Hz, H-6b), 3.63 (2H, m, H-3, H-4), 3.49 (1H, t, J1,2 = J2,3 = 8.5 Hz, H-

2 ), 3.30 (1H, ddd, J5,6a = 5.7, J5,6b = 8.6, J4,5 = 10.7 Hz, H-5), 3.21, 3.19 (6H, 2s,

2OMe), 3.06 (1H, br, OH), 1.25, 1.23 (6H, 2s, 2Me), 0.81 (9H, s, 3Me), 0.00 (6H,

s, 2Me). 13C NMR (CDCl3) δ 137.7-127.5 (m, Ar), 99.9 (C-1), 99.5, 99.4

(2MeOCMe), 74.9 (C-5), 72.5 (C-3), 70.8 (CH2Ph), 70.3 (C-4), 69.2 (C-2), 64.8

(C-6), 48.0, 47.9 (2OMe), 25.9 (3Me), 18.32 (CMe3), 17.7 (2Me), -5.4, -5.5 (2Me).

HRMS Calcd for C25H42NaO8Si (M+Na): 521.2541. Found: 521.2537.

200

7.5.1.2 Benzyl 6-O-tert-butyldimethylsilyl-3,4-O-[(2R,3R)-2,3-dimethoxybutane-2,3-diyl)]-β-D-glucopyranoside 19

The compound was obtained as white solid (654.4 mg, 82 %) from 18 (614

mg, 1.6 mmol) using the same procedure that was used to obtain 14. M.P. 145-

147 oC. [α] 23D = + 81.8 (c = 1.0, CH2Cl2). 1H NMR (CDCl3): δ 7.38−7.27 (5H, m,

Ar), 4.90 (1H, d, JAB = 11.4 Hz, CH2Ph), 4.60 (1H, d, JAB = 11.4 Hz, CH2Ph), 4.39

(1H, d, J1, 2 = 7.4 Hz, H-1), 3.90 (1H, dd, J6a, 6b = 11.6, J6a, 5 = 2.2 Hz, H-6a), 3.85

(1H, dd, J6b, 6a = 11.6, J6b, 5 = 3.8 Hz, H-6b), 3.77−3.68 (2H, m, H-3, H-4), 3.59

(1H, t, J1, 2 = J2, 3 = 9.6 Hz, H-2 ), 3.44 (1H, ddd, J5, 6a = 2.2, J5, 6b = 3.7, J5, 4 =

9.4 Hz, H-5), 3.29, 3.27 (6H, 2s, 2OMe), 2.53 (1H, br, OH), 1.33, 1.29 (6H, 2s,

2Me), 0.90 (9H, s, 3 Me), 0.10, 0.08 (6H, 2s, 2Me). 13C NMR (CDCl3) δ

137.0−127.9 (m, Ar), 101.9 (C-1), 99.5, 99.0 (2MeOCMe), 74.7 (C-5), 71.9 (C-3),

71.2 (C-2), 70.7 (CH2Ph), 65.1 (C-4), 61.3 (C-6), 48.0, 47.9 (2OMe), 25.8 (3 Me),

18.3 (CMe3), 17.6 (2 Me) -5.0, -5.4 (2Me). HRMS Calcd for C25H42NaO8Si

(M+Na): 521.2541. Found: 521.2546.

7.5.1.3 Benzyl 2,3,4,6-tetra-O-acetyl-α-D-glucopyranosyl-(1 4)-2,3,6-tri-O-acetyl-β-D-glucopyranosyl-(1 4)-6-O-tert-butyldimethylsilyl-2,3-O-[(2R,3R)-2,3-dimethoxybutane-2,3-diyl]-β-D-glucopyranoside 15

To a solution of the acceptor glycoside 14 (0.7 g, 1.4 mmol) and maltosyl

trichloroacetimidate 12 (1.3 g, 1.68 mmol) in dry CH2Cl2 (10 mL) under N2,

BF3.Et2O (0.02 mL, 0.2 eq) was added at room temperature. After stirring for 30

min, the mixture was quenched with NEt3 (0.2 eq.) with vigorous stirring. The

mixture was concentrated and purified by column chromatography

(Hexanes/EtOAc 1:1) to give the coupled product 15 as white foam (0.7g, 45%).

201

[α] 23D = - 9 (c = 0.7, CH2Cl2). 1H NMR (CDCl3) δ 7.39-7.28 (5H, m, Ar), 5.42 (1H,

d, J1′′,2′′ = 4.1 Hz, H-1′′), 5.38 (1H, t, J2′′,1′′ = J2′′,3′′ = 10.4 Hz, H-2′′), 5.19 (1H, t, J3′,4′

= J2′,3′ = 7.7 Hz, H-3′), 5.08 (1H, t, J3′′,4′′ = J4′′,5′′ = 9.9 Hz, H-4′′), 4.90, 4.66 (2H, 2d,

JA,B = 12.0 Hz, CH2Ph), 4.87 (1H, t, J3′′,4′′ = J2′′,3′′ = 4.1 Hz, H-3′′), 4.83 (2H, m, H-

1′, H-2′), 4.55 (1H, d, J1,2 = 7.8 Hz, H-1), 4.42 (1H, dd, J6′a,6′b = 3.1, J5′,6′a = 12.5

Hz, H-6′a), 4.29 (1H, dd, J6′a,6′b = 3.6, J5′,6′b = 12.2 Hz, H-6′b), 4.24 (1H, dd, J6′′a,6′′b

= 3.6, J5′′,6′′a = 12.5 Hz, H-6′′a), 4.06 (2H, m, H-6′′b, H-4′), 3.94 (1H, ddd, J5′′,6a′′ =

3.1, J5′′,6b′′ = 2.6, J4′′,5′′ = 10.0 Hz, H-5′′), 3.88 (1H, dd, J6a,6b = 1.5, J5,6a = 11.5 Hz,

H-6a), 3.82 (1H, t, J3,4 = J2,3 = 9.6 Hz, H-3), 3.71 (2H, m, H-4, H-6b), 3.67 (1H,

ddd, J5′,6a′ = 3.2, J5′,6b′ = 3.0, J4′,5′ = 9.5 Hz, H-5′), 3.59 (1H, dd, J1,2 =7.8, J2,3 =

10.0 Hz, H-2 ), 3.34 (1H, ddd, J5,6a = 1.5, J5,6b = 5.1, J4,5 = 6.8 Hz, H-5), 3.32,

3.30 (6H, 2s, 2OMe), 2.14, 2.12, 2.06, 2.05, 2.04, 2.02, 2.01 (21H, 7s, 7OAc),

1.34, 1.32 (6H, 2s, 2Me), 0.94 (9H, s, 3Me), 0.11, 0.10 (6H, 2s, 2Me). 13C NMR

(CDCl3) δ 170.4-169.3 (m, CO), 137.4-127.4 (m, Ar), 100.0 (C-1′), 99.4, 99.3

(2MeOCMe), 99.2 (C-1), 95.4 (C-1′′), 76.1 (C-5), 75.5 (C-3′), 75.4 (C-4), 72.8 (C-

2′), 72.3 (C-4′), 71.9 (C-5′), 71.4 (C-3), 70.3 (CH2Ph), 69.8 (C-3′′), 69.4 (C-2),

69.2 (C-2”), 68.3 (C-5”), 67.8 (C-4”), 62.9 (C-6′), 61.9 (C-6), 61.2 (C-6”), 47.8,

47.7 (2OMe), 25.8 (3Me), 20.7-20.4 (m, OAc), 18.22 (CMe3), 17.5, 17.4 (2Me), -

5.1, -5.3 (2Me). HRMS Calcd for C51H76NaO25Si (M+Na): 1139.4337. Found:

1139.4330.

202

7.5.1.4 Benzyl 2,3,4,6-tetra-O-acetyl-α-D-glucopyranosyl-(1 4)-2,3,6-tri-O-acetyl-β-D-glucopyranosyl-(1 2)-6-O-tert-butyldimethylsilyl-3,4-O-[(2R,3R)-2,3-dimethoxybutane-2,3-diyl]-β-D-glucopyranoside 20

The compound was obtained as white solid (1.7 g, 52 %) from 19 (1.5 g,

2.95 mmol) using the same procedure that was used to obtain 15. M.P. 95-98 oC.

[α] 23D = + 64.4 (c = 0.1, CH2Cl2). 1H NMR (CDCl3) δ 7.40−7.29 (5H, m, Ar), 5.37

(1H, d, J1′′, 2′′ = 4.0 Hz, H-1′′), 5.32 (1H, t, J3′′,2′′ = J3′′,4′′ = 9.5 Hz, H-3′′), 5.16 (1H,

t, J3′,2′ = J3′,4′ = 9.1 Hz, H-3′), 5.03 (1H, t, J4′′,3′′ = J4′′,5′′ = 10.1 Hz, H-4′′), 4.94−4.91

(2H, m, H-1′, CH2Ph), 4.85 (1H, dd, J2′′,1′′ = 7.5, J2′′,3′′ = 4.0 Hz, H-2′′), 4.81 (1H,

dd, J2′,1′ = 7.5, J2′,3′ = 2.1 Hz, H-2′), 4.62 (1H, d, JAB = 11.7 Hz, CH2Ph), 4.50 (1H,

d, J1, 2 = 6.9 Hz, H-1), 4.28 (1H, dd, J6′b,6′a = 12.2, J6′b,5′ = 2.6 Hz, H-6′b), 4.23 (1H,

dd, J6′′a, 6′′b = 12.6, J6′′a,5′′ = 3.8 Hz, H-6′′a), 4.08−3.96 (3H, m, H-6′a, H-6′′b, H-4′),

3.89 (1H, ddd, J5′′,4′ ′ = 10.1, J6′′a,5′′ = 3.3, J6′′b,5′′ = 2.3 Hz , H-5′′), 3.86−3.84 (2H, m,

H-6a, H-6b), 3.75−3.64 (3H, m, H-2, H-4, H-5), 3.49 (1H, ddd, J5′,4′ = 9.7, J5′, 6′a =

3.7, J5′, 6′b = 2.4, H-5′), 3.37 (1H, dd, J3,2 = 3.4, J3,4 = 1.8, H-3), 3.24, 3.20 (6H, 2s,

2OMe), 2.07−2.00 (21H, 7s, 7OAc), 1.26, 1.25 (6H, 2s, 2Me), 0.88 (9H, s, 3Me),

0.07, 0.05 (6H, 2s, 2Me). 13C NMR (CDCl3, 100MHz) δ 170.5−169.4 (m, CO),

137.6−127.4 (m, Ar), 101.8 (C-1), 99.9 (C-1′), 99.6, 99.4 (2MeOCMe), 95.6 (C-

1′′), 77.9 (C-2), 75.8 (C-3′), 74.4 (C-3), 73.3 (C-2′), 72.6 (C-4′), 72.2 (C-5′), 71.2

(C-5), 70.7 (CH2Ph), 69.9 (C-2′′), 69.4 (C-3′′), 68.4 (C-5′′), 68.0 (C-4′′), 65.2 (C-4),

63.0(C-6′), 61.4 (C-6′′), 61.3 (C-6), 48.0, 47.9 ( 2OMe), 25.8 (3Me), 21.0−20.6 (m,

OAc), 18.4 (CMe3), 17.6, 17.5 (2Me), -5.0, -5.4 (2Me). HRMS Calcd. for

C51H76NaO25Si (M+Na): 1139.4337.Found: 1139.4325.

203

7.5.1.5 Benzyl 2,3,4,6-tetra-O-benzyl-α-D-glucopyranosyl-(1 4)-2,3,6-tri-O-benzyl-β-D-glucopyranosyl-(1 4)-6-O-tert-butyldimethylsilyl-2,3-O-[(2R,3R)-2,3-dimethoxybutane-2,3-diyl]-β-D-glucopyranoside 16

To a solution of compound 15 (500 mg, 4.5 mmol) in MeOH (15 mL), was

added catalytic amount of 1M NaOMe and the reaction was stirred at room

temperature overnight. The solvent were removed under vacuum to give the

crude product which was used directly in the next step without further purification.

The crude product was dissolved in DMF (20 mL), and NaH (130 mg, 5.4 mmol)

and BnBr (0.5 mL, 4 mmol) were added at 0 °C. The reaction mixture was stirred

for 2 h at room temperature, then quenched with ice, and extracted with ether

(3×50 mL). The organic solution was washed with water (30 mL) and brine (30

ml), dried (Na2SO4) and concentrated and the residue was purified by flash

chromatography (EtOAc/Hexanes (1:5)) to yield 16 as white foam (475 mg, 73%).

[α] 23D = - 15 (c = 0.8, CH2Cl2). 1H NMR (CDCl3) δ 7.40-7.11 (40H, m, Ar), 5.73

(1H, d, J1′′,2′′ = 3.7 Hz, H-1′′), 4.96-4.52 (14H, m, 7CH2Ph), 4.63 (1H, m, H-1′),

4.60 (1H, m, H-1), 4.45, 4.26 (2H, 2d, JA,B = 12.1 Hz, CH2Ph), 4.18 (1H, t, J3′′,4′′ =

J4′′,5′′ = 8.9 Hz, H-4′′), 4.01 (1H, t, J3′,4′ = J4′,5′ = 9.5 Hz, H-4′), 3.95-3.77 (8H, m, H-

6a, H-3′′, H-4′, H-6′′a, H-3, H-6b, H-6b′′, H-3′), 3.66 (2H, m, H-5′, H-2), 3.54-3.49

(3H, m, H-5′′, H-6′a, H-2′′), 3.46 (1H, dd, J2′,1′ = 8.0 J2′,3′ = 9.0 Hz, H-2′), 3.38 (1H,

dd, J6′a,6′b = 1.8, J5′,6′b = 10.9 Hz, H-6′b), 3.33 (1H, ddd, J5,6a = 1.8, J5,6b = 3.5,

J4,5 = 5.3 Hz, H-5), 3.31, 3.28 (6H, 2s, 2OMe), 1.34, 1.32 (6H, 2s, 2Me), 1.32,

1.21 (6H, 2s, 2Me), 0.91 (9H, s, 3Me), 0.07, 0.06 (6H, 2s, 2Me). 13C NMR

(CDCl3) δ 138.3-126.1 (m, Ar), 101.3 (C-1′), 99.2 (C-1), 99.1, 99.0 (2MeOCMe),

96.2 (C-1′′), 84.6 (C-3′), 82.3 (C-2′), 81.6 (C-4′), 78.8 (C-5′′), 77.3 (C-5′), 76.3 (C-

204

5), 75.0, 74.5, 74.2, 73.6, 73.2, 73.0, 72.8, 69.8 (8CH2Ph), 74.4 (C-2”), 72.5 (C-

4), 71.6 (C-3′′), 70.4 (C-3), 70.3 (C-4”), 69.4 (C-2), 68.1 (C-6′′), 67.7 (C-6′), 61.2

(C-6), 47.5, 47.4 (2OMe), 25.5 (3Me), 17.93 (CMe3), 17.2, 17.1 (2Me), -5.4, -5.6

(2Me). HRMS Calcd for C86H104NaO18Si (M+Na): 1477.6941. Found: 1477.6941.

7.5.1.6 Benzyl 2,3,4,6-tetra-O-benzyl-α-D-glucopyranosyl-(1 4)-2,3,6-tri-O-benzyl-β-D-glucopyranosyl-(1 2)-6-O-tert-butyldimethylsilyl-3,4-O-[(2R,3R)-2,3-dimethoxybutane-2,3-diyl]-β-D-glucopyranoside 21

The compound was obtained as white solid (585 mg, 76 %) from 20 (592

mg, 0.53 mmol) using the same procedure that was used to obtain 16. M.P. 90-

92 oC. [α] 23D = + 76.8 (c = 1.0, CH2Cl2). 1H NMR (CDCl3) δ 7.43−7.17 (40H, m,

Ar), 5.77 (1H, d, J1′′,2′′ = 3.6 Hz, H-1′′), 5.05-4.84 (8H, m, H-1′, 7CH2Ph),

4.71−4.51 (9H, m, H-1, 8CH2Ph), 4.40 (1H, d, JA,B = 12.1 Hz, CH2Ph), 4.24 (1H,

t, J4′,3′ = J4′,5′ = 9.2 Hz, H-4′), 4.18 (1H, t, J4′′,3′′ = J4′′,5′′ = 8.9 Hz, H-4′′), 4.13−3.80

(8H, m, H-6a, H-6b, H-3, H-3′, H-3′′, H-4, H-6′a, H-6′′a), 3.76−3.66 (4H, m, H-6′′b,

H-6′b, H-2, H-5′′), 3.58−3.48 (4H, m, H-2′′, H-2′, H-5′, H-5), 3.35, 3.17 (6H, 2s,

2OMe), 1.33, 1.30 (6H, 2s, 2Me), 0.98 (9H, s, 3Me), 0.18, 0.16 (6H, 2s, 2Me).

13C NMR (CDCl3) δ 138.8−126.5 (m, Ar), 101.9 (C-1′), 101.2 (C-1), 99.5, 99.4

(2MeOCMe), 96.7 (C-1′′), 84.8 (C-5′′), 83.0 (C-2′), 81.9 (C-3′′), 79.2 (C-2′′), 77.7

(C - 4′′), 75.4, 74.9, 74.6 (3 CH2Ph), 74.4 (C-5, C-5′), 73.8, 73.4 (CH2Ph), 73.3

(C-3), 73.1 (CH2Ph), 72.3 (C-4′), 70.8 (C-3′), 70.9 (C-2), 69.7 (CH2Ph), 68.2 (C-

6′′), 68.1 (C-6′), 65.1 (C-4), 61.4 (C-6), 48.0 (2OMe), 25.8, 25.5 (3Me), 18.3

(CMe3), 17.8, 17.6 (2Me), -5.0, -5.4 (2Me). HRMS Calcd. for C86H104NaO18Si

(M+Na): 1477.6941. Found: 1476.6938.

205

7.5.1.7 Benzyl 2,3,4,6-tetra-O-benzyl-α-D-glucopyranosyl-(1 4)-2,3,6-tri-O-benzyl-β-D-glucopyranosyl-(1 4)-6-O-trifluoromethanesulfonyl-2,3-O-[(2R,3R)-2,3-dimethoxybutane-2,3-diyl]-β-D-glucopyranoside 17

To a solution of 16 (205 mg, 0.14 mmol) in THF (15 mL), TBAF (1.0 M

solution in THF, 0.28 mL, 0.28 mmol) was added, and the reaction mixture was

stirred at room temperature. After 12 h it was concentrated and further dried

under high vacuum for 1 h. The crude product was dissolved in CH2Cl2 (10 mL),

and pyridine (0.013 mL, 1.2 eq) was added. The mixture was cooled to -10 °C,

and Tf2O (0.036 mL, 1.5 eq) was added under N2. After 30 min the reaction was

quenched by the addition of cold saturated NaHCO3 (1.5 mL). The organic layers

were washed with 1N HCl (5 mL), water (5 mL) and brine (5 mL), dried (Na2SO4),

and concentrated in vacuo. Chromatographic purification of the crude product

(EtOAc/Hexanes (1:10)) gave 17 as a foam (85 mg, 41%). [α] 23D = - 5.5 (c = 0.2,

CH2Cl2). 1H NMR (CDCl3) δ 7.38-7.11 (40H, m, Ar), 5.62 (1H, d, J1′′,2′′ = 3.7 Hz,

H-1′′), 4.91- 4.32 (16H, m, 8CH2Ph), 4.85 (1H, m, H-6a), 4.60 (1H, m, H-6b), 4.59

(1H, m, H-1′), 4.53 (1H, m, H-1), 4.05 (2H, m, H-5′′, H5′), 3.91 (1H, t, J3,4 = J4,5 =

9.7 Hz, H-4), 3.85 (1H, t, J3′,4′ = J4′,5′ = 9.6 Hz, H-4′), 3.76 (4H, m, H-4′′, H-6′a,b,

H-3′), 3.66 (2H, m, H-2, H-3), 3.56-3.48 (4H, m, H-6′′a, H-5, H-2′, H-2′′), 3.42 (1H,

dd, J6′′a,6′′b = 1.8, J5′′,6′′b = 10.5 Hz, H-6′′b), 3.30, 3.29 (6H, 2s, 2OMe), 1.32, 1.23

(6H, 2s, 2Me). 13C NMR (CDCl3) δ 138.6-126.5 (m, Ar), 118.4 (m, CF3), 100.6 (C-

1′), 99.7, 99.5 (2MeOCMe), 99.4 (C-1), 96.9 (C-1′′), 84.9 (C-4′′), 82.2 (C-2′), 81.9

(C-4), 79.1 (C-5), 77.6 (C-3), 75.5, 75.0, 74.7, 74.3, 73.9, 73.4, 73.3, 73.2

(8CH2Ph), 74.6 (C-2′′), 74.5 (C-4′), 72.8 (C-3′), 72.6 (C-5′′), 71.0 (C-3′′), 70.7 (C-

6), 69.6 (C-2), 69.1 (C-6′), 68.4 (C-5′), 68.1 (C-6′′), 48.6, 47.9 (2OMe), 17.5

(2Me). HRMS Calcd for C81H89F3NaO20S (M+Na): 1493.5512. Found: 1493.5543.

206

7.5.1.8 Benzyl 2,3,4,6-tetra-O-benzyl-α-D-glucopyranosyl-(1 4)-2,3,6-tri-O-benzyl-β-D-glucopyranosyl- (1 2)-6-O-triflouromethanesulfonyl-3,4-O-[(2R,3R)-2,3-dimethoxybutane-2,3-diyl]-β-D-glucopyranoside 22

Compound was obtained as white solid (227 mg, 86 %) from 21 (262 mg,

0.18 mmol) using the same procedure that was used to obtain 17. M.P. 80-82 oC.

[α] 23D = + 51.4. (c = 0.9, CH2Cl2). 1H NMR (CDCl3): δ 7.34−7.10 (40H, m, Ar), 5.67

(1H, d, J1′′,2′′ = 3.7 Hz, H-1′′), 4.95−4.73 (9H, m, H-1′, H-6b, 7CH2Ph), 4.63−4.45

(10H, m, H-1, H-6a, 8CH2Ph), 4.33 (1H, d, JA,B = 12.1, CH2Ph), 4.14 (1H, t, J4′,3′ =

J4′,5′ = 9.2 Hz, H-4′), 4.05 (1H, t, J5,4 = J5, 6a = 9.3 Hz, H-5), 3.91−3.84 (2H, m, H-

3′′, H-4), 3.78−3.73 (4H, m, H-5′′, H-6′′a, H-3′, H-2), 3.70−3.64 (3H, m, H-3, H-4′′,

H-6′′b), 3.60 (1H, dd, J6′b,6′a =10.7 Hz, J6′b,5′ = 2.7 Hz, H-6′b), 3.50−3.42 (4H, m, H-

2′′, H-2′, H-5′, H-6′a), 3.24, 3.07 (6H, 2s, 2OMe), 1.26, 1.21 (6H, 2s, 2Me). 13C

NMR (CDCl3) δ 138.6-126.5 (m, Ar), 121.8, 119.6, 117.5, 115.4 (m, CF3), 102.0

(C-1′), 101.3(C-1), 99.8, 99.7 (2MeOCMe), 96.8 (C-1′′), 84.8 (C-5′′), 82.9 (C-2′),

82.0 (C-3′′), 79.2 (C-2′′), 77.7 (C-4′′), 75.5, 75.0 (CH2Ph), 74.7 (C-6, C-5), 74.5

(C-5′), 73.9, 73.8, 73.5, 73.3, 73.2 (6CH2Ph), 72.6 (C-4), 72.4 (C-4′), 71.0 (C-3′),

70.9 (C-2), 68.3 (C-6′′), 68.2 (C-6′), 65.3 (C-3), 48.1 (2OMe), 17.7,17.5 (2Me).

HRMS Calcd. for C81H89F3NaO20S (M+Na): 1493.5512. Found: 1493.5487.

7.5.1.9 Benzyl 2,3,4,6-tetra-O-benzyl-α-D-glucopyranosyl-(1 4)-2,3,6-tri-O-benzyl-β-D-glucopyranosyl-(1 4)-6-deoxy-6-[2,3,5-tri-O-benzyl-1,4-dideoxy-episulfoniumylidene-D-arabinitol]-2,3-O-[(2R,3R)-2,3-dimethoxybutane-2,3-diyl]-β-D-glucopyranoside trifluoromethanesulfonate 24

The mixture of the thioether 23 (64 mg, 0.15 mmol) and 17 (75 mg, 0.051

mmol) in CH2Cl2 (1 mL) were stirred overnight at room temperature for 24 h. The

mixture was concentrated and purified by column chromatography (CHCl3/MeOH

20:1) to give the sulfonium salt 24 as syrup (90 mg, 95%). Analysis by NMR

207

showed that the product was a mixture of two isomers (~5:1) at the stereogenic

sulfur center. The major component of the mixture was assigned to be the

diastereomer with a cis relationship between H-4 and H-6′ on the basis of

analysis of the NOESY spectrum.

Data for the major diastereomer (trans-24) follow. 1H NMR (CDCl3) δ 7.36-

7.11 (55H, m, Ar), 5.56 (1H, d, J1′′′,2′′′ = 3.6 Hz, H-1′") 4.90-4.36 (22H, m, CH2Ph),

4.67 (1H, m, H-1"), 4.56 (1H, m, H-1′), 4.46 (1H, m, H-2), 4.27 (1H, m, H-3), 4.11-

4.01 (4H, m, H-5′, H-4, H-4′, H-5′′′), 3.92 (1H, t, J3′′′,2′′′ = J3′′′,4′′′ = 9.5 Hz, H-3′′′),

3.83-3.70 (8H, m, H-1a, H-5′′, H-3′′, H-5a, H-5b, H-6′a, H-3, H-6′′′′a), 3.68-3.56

(6H, m, H-4′′′, H-6′b, H-4′′, H-6′′a, H-6′′′b, H-2′), 3.52-3.47 (3H, m, H-2′′′, H-2′′, H-

6′′b), 3.43 (1H, dd, J1a,1b = 4.0, J1b,2 = 13.4 Hz, H-1b), 3.31, 3.29 (6H, 2s, 2OMe),

1.34, 1.23 (6H, 2s, 2Me). 13C NMR (CDCl3) δ 138.6-126.6 (m, Ar), 120.8 (m,

CF3), 100.7 (C-1′), 100.3 (H-1′′), 99.7, 99.6 (2MeOCMe), 99.0 (C-1′′′), 84.6 (C-5′′),

82.4 (C-3), 82.3 (C-2), 81.8 (C-3′′′), 81.5 (C-2′′), 79.2 (C-2′′′), 77.6 (C-4′′′), 76.5 (H-

3′′), 75.5, 75.0, 74.0, 73.6, 73.5, 73.2, 73.0, 72.4, 72.1, 71.8 (m, CH2Ph), 74.5 (H-

2′), 73.3 (H-4′), 71.5 (C-5′), 71.1 (C-3′), 69.5 (C-4′′), 69.3 (C-5), 68.1 (C-6′′), 67.7

(C-5′′′), 67.1 (C-4), 66.4 (C-6′′′), 48,8 47.9 (2OMe), 47.2 (C-1), 46.9 (C-6′), 17.5,

17.4 (2Me). HRMS Calcd for C106H118O20S (M+H):1742.7932. Found: 1742.7932.

7.5.1.10 Benzyl 2,3,4,6-tetra-O-benzyl-α-D-glucopyranosyl-(1 4)-2,3,6-tri-O-benzyl-β-D-glucopyranosyl-(1 2)-6-deoxy-6-(2,3,5-tri-O-benzyl-1,4-dideoxy-episulfoniumylidene-D-arabinitol)-3,4-O-[(2R,3R)-2,3-dimethoxybutane-2,3-diyl]-β-D-glucopyranoside trifluoromethanesulfonate 25

The mixture of the thioether 23 (64 mg, 0.15 mmol) and 22 (80mg, 0.054

mmol) in CH2Cl2 (1 mL) were stirred overnight at room temperature for 24 h. The

208

mixture was concentrated and purified by column chromatography (CHCl3/MeOH

20:1) to give the sulfonium salts 25a and 25b as syrups (91.8 mg, 90%). Analysis

by crude NMR showed that the product was a mixture of two isomers (~2.8:1) at

the stereogenic sulfur center.

Data for the major diastereomer (25a): [α] 23D = + 62.5 (c = 0.2, CH2Cl2). 1H

NMR (CDCl3) δ 7.34−7.08 (55H, m, Ar), 5.64 (1H, d, J1′′′,2′′′ = 3.6 Hz, H-1′") ,

4.93−4.72 (8H, m, H-1′′, 7CH2Ph), 4.64 (1H, d, JA,B = 12.6, CH2Ph), 4.60−4.40

(15H, m, H-2, H-1′, 13CH2Ph), 4.39 (1H, brd, J = 7.5, H-4), 4.31−4.27 (2H, m, H-

3, CH2Ph), 4.09 (1H, t, J4′′′,3′′′ = J4′′′, 5′′′ = 9.2 Hz, H- 4′′), 3.99−3.92 (3H, m, H-2′, H-

6′b, H-1b), 3.90−3.85 (1H, m, H-5′, H-3′′′), 3.83−3.73 (2H, m, H-3′, H-1a),

3.77−3.71 (6H, H-5′′′, H-6′′′a, H-6′′′b, H-3′′, H-6′′a, H-6′a), 3.65−3.62 (2H, m, H-

6′′b, H-4′′′), 3.57 −3.54 (2H, m, H-4′, H-5a), 3.49−3.40 (4H, m, H-2′′′, H-2′′, H-5′′,

H-5b), 3.19, 3.06 (6H, 2s, 2OMe), 1.25, 1.20 (6H, 2s, 2Me). 13C NMR (CDCl3) δ

138.7− 126.5 (m, Ar), 121.9, 119.7 (CF3), 102.5 (C-1′), 102.1 (C-1′′), 100.1, 99.8

(2 MeOCMe), 96.8 (C-1′′′), 84.8 (C-5′′′, C-2′′), 82.7 (C-2, C-3), 82.0 (C-3′′′), 79.2

(C-2′′′), 77.7 (C-4′′′), 75.5, 75.0, 74.7(3CH2Ph), 74.5 (C-5′′), 74.4 (C-2′), 73.9,

73.7, 73.5, 73.4, 73.3 (5CH2Ph), 72.6 (C-4′′), 72.5, 72.2 (CH2Ph), 72.0 (C-3′),

71.3 (CH2Ph), 71.0 (C-3′′), 70.0 (C-5′), 68.8 (C-6′′), 68.2 (C-5, C-4′ ), 66.5 (C-4,

C-6′′′), 48.6 (OMe), 48.2 (C-1), 48.1 (OMe), 46.6 (C-6′), 17.7, 17.3 (2Me). HRMS

Calcd for C106H118O20S (M+H): 1742.7932. Found: 1742.7882.

Data for the minor diastereomer (25b): [α] 23D = - 125.0 (c = 0.04, CH2Cl2).

1H NMR (CDCl3) δ 7.38−7.12 (55H, m, Ar), 5.68 (1H, d, J1′′′,2′′′ = 3.6 Hz, H-1′"),

4.97−4.88 (4H, m, H-1′′, 3CH2Ph), 4.83−4.78 (5H, m, 5CH2Ph), 4.66 (1H, d, J1′,2′

209

= 7.5 Hz, H-1′), 4.63−4.46 (12H, m, H-3, 11CH2Ph), 4.48−4.46 (2H, m, 2CH2Ph),

4.44 (1H, m, H-2), 4.37−4.33 (2H, m, H-4, CH2Ph), 4.13 (1H, t, J4′′,3′′ = J4′′,5 ′′ = 9.2

Hz, H-4′′), 4.08 (1H, dd, J6′′a, 5′′ = 4.1, J6′′a, 6′′b = 11.7 Hz, H-6′′a), 3.98 (1H, dd, J2′, 1′

= 9.7 Hz, J2′, 3′ = 2.2 Hz, H-2′), 3.95−3.89 (4H, m, H-3′′′, H-5′, H-6′′b, H-6′a), 3.84

(1H, t, J3′,4′ = J3′,2′ = 9.6 Hz, H-3′), 3.82−3.76 (4H, m, H-5′′′,H-6′′′a, H-3′′, H-6′b),

3.75−3.70 (2H, m, H-6′′′b, H-1a), 3.67 (1H, t, J3′′′,4′′′ = J5′′′,4′′′ = 9.5 Hz, H-4′′′),

3.63−3.59 (2H, m, H-1b, H-5a), 3.55 (1H, t, J4′,3′ = J5′,4′ = 9.7 Hz, H-4′),

3.52−3.45 (4H, H-2′′′, H-2′′, H-5′′, H-5b), 3.07, 2.97 (6H, 2s, 2OMe ), 1.19, 1.06

(6H, 2s, 2Me). 13C NMR (CDCl3) δ 138.7-126.5 (m, Ar), 121.9, 119.8 (CF3), 102.8

(C-1′), 102.1 (C-1′′), 100.0, 99.7 (2 MeOCMe), 96.8 (C-1′′′), 84.8 (C-5′′′), 83.2 (C-

2), 83.5 (C-3), 82.7 (C-2′′), 82.0 (C-3′′′), 79.2 (C-2′′′), 77.7 (C-4′′′), 75.5, 75.0

(CH2Ph), 74.7 (C-2′), 74.6 (C-5′′), 73.8 - 72.9 (9CH2Ph), 72.6 (C-4′′), 71.9 (C-3′),

71.0 (C-3′′), 69.1(C-5′), 68.8 (C-6′′′), 68.2 (C-5), 67.1 (C-4′), 65.0 (C-6′′), 63.3 (C-

4), 48.0, 48.1 (2OMe), 46.5 (C-1), 38.6 (C-6′), 17.7, 17.3 (2Me). HRMS Calcd. for

C106H118O20S (M+H): 1742.7932. Found: 1742.7882.

7.5.1.11 1,4-Dideoxy-1,4-[[2S, 3S, 4R, 5S]-2,4,5,6-tetrahydroxy-3-(4-O-α-D-Glucopyranosyl-D-glucosyl) hexyl]-(R/S)-epi-sulfoniumylidine]-D-arabinitol chloride 10

To a solution of sulfonium salt 24 (85 mg, 0.045 mmol, 5:1 mixture of

isomers) in 85% AcOH:H2O (40 mL) was added Pd(OH)2, 20 % weight on carbon

(200 mg) and the mixture was stirred under 100 Psi of H2 for 24h. The catalyst

was removed by filtration through a bed of Celite, and then washed with water.

The solvents were removed under reduced pressure and the residue was

dissolved in 80% TFA (5 mL) and stirred at room temperature for 2h. The

210

solvents were removed under reduced pressure; the residue was dissolved in

water (5 mL), and washed with CH2Cl2 (2×5 mL). The water layer was

evaporated to give the crude product. The residue was dissolved in H2O (10 mL),

Amberlyst A-26 resin (100 mg) was added, and the reaction mixture was stirred

at room temperature for 2 h. Filtration through cotton, followed by solvent

removal gave the crude hemiacetal. The crude product was dissolved in water (5

mL), and the solution was stirred at room temperature while NaBH4 (7 mg, 0.18

mmol) was added. Stirring was continued for another 3 h and the mixture was

acidified to pH < 4 by dropwise addition of 2M HCl. The mixture was evaporated

to dryness and the residue was co-evaporated with anhydrous MeOH (3 × 10

mL). The residue was purified by crystallization with EtOAc:MeOH:H2O (10:3:1)

followed by reverse phase chromatography (H2O) to give 10 (d.r, 5:1) as a

colorless solid (21 mg, 70%).

Data for the major diastereomer 10 follow. 1H NMR (D2O) δ 5.34 (1H, d,

J1′′′,2′′′ = 3.7 Hz, H-1′′′), 4.68 (1H, m, H-2), 4.61 (1H, d, J1′′,2′′ = 7.6 Hz, H-1′′), 4.36

(2H, m, H-3, H-2′), 4.08 (1H, dd, J4,5a = 5.0, J5a,5b = 12.2 Hz, H-5a), 4.02 (2H, m,

H-1a, H-3′′′), 3.92 (1H, m, H-3′), 3.89-3.57 (16H, m, H-4′′, H-5b, H-1′a, H-1′b,

H1b, H-6′′′a, H-4, H-6′′′b, H-6′′a, H-6′′b, H-4′, H-3′′, H-6′a, H-5′′, H-5′, H-6′b), 3.51

(2H, m, H-4′′′, H-2′′′), 3.33 (2H, m, H-5′′′, H-2′′). 13C NMR (D2O) δ 102.9 (H-1′′),

99.5 (C-1′′′), 81.5 (C-3′), 77.3 (C-3), 76.9 (H-2), 66.2 (C-5′′), 76.0 (C-3′′), 74.4 (C-

4′′′), 73.1 (C-2′′), 72.7 (C-5′), 72.6 (C-4′), 71.9 (C-4), 71.5 (H-2′′′), 69.8 (H-3′′′),

69.3 (H-5′′′), 69.2 (H-2′′), 67.9 (C-2′), 62.3 (C-6′), 60.4 (C-6′′′), 60.3 (H-6′′), 59.2

211

(C-5), 49.2 (C-1), 48.2 (C-1′). HRMS Calcd for C23H43O18S (M+ ): 639.2165.

Found: 639.2152.

7.5.1.12 1,4-Dideoxy-1,4-[[2S, 3S, 4R, 5S]-2,3,4,6-tetrahydroxy-5-(4-O-α-D-Glucopyranosyl-D-glucosyl) hexyl]-(R/S)-epi-sulfoniumylidine]-D-arabinitol chloride 11

Compound was obtained as white solid (18 mg, 64 %, 2.8:1 mixture of

isomers) from mixture of 25a and 25b (79 mg, 0.042 mmol, 2.8:1 mixture of

isomers) using the same procedure that was used to obtain 10.

Data for the major isomer (11) : 1H NMR (D2O) : δ 5.33 (1H, d, J1′′′, 2′′′ = 2.6

Hz, H-1′′′), 4.70 (1H, m, H-2), 4.57 (1H, d, J1′′,2′′ = 7.9 Hz, H-1′′), 4.39 (1H, brs, H-

3), 4.19 (1H, brs, H-2′), 4.08−4.04 (2H, m, H-4, H-6′a), 3.98−3.50 (20H, m, H-6′′b,

H-1a, H-1b, H-5a, H-5b, H-1′a, H-3′,H-4′, H-6′b, H-6′′′a, H-6′′′b, H-3′′, H-6′′a, H-1′b,

H-5′, H-3′′′, H-5′′′, H-4′′, H-2′′′, H-5′′), 3.36−3.26 (2H, m, H-4′′′, H-2′′). 13C NMR

(D2O) δ 103.0 (C-1′′), 99.5 (C-1′′′), 83.1 (C-5′), 77.0 (C-3, C-2), 76.5 (C-3′, C-4′′),

74.5 (C-5′′), 73.2 (C-2′′), 72.8 (C-5′′′), 72.6 (C-3′′′), 72.5 (C-3′′), 71.6 (C-2′′′), 70.0

(C-4), 69.3 (C-4′′′), 68.5 (C-4′), 67.5 (C-2′), 61.1 (C-5), 60.6 (C-6′′), 60.4 (C-6′′′),

59.2 (C-6′), 50.1 (C-1′), 48.2(C-1) HRMS Calcd. for C23H43O18S (M+): 639.2165.

Found: 639.2160.

7.6 Acknowledgments

We thank Buford L. Nichols, Roberto Quezada-Calvillo and Hassan Naim

for reagents for recombinant expression, and Lyann Sim for MGAM and SI

enzyme purification. K. J. was supported by a scholarship from the Canadian

Institutes of Health Research (CIHR) and the Canadian Digestive Health

212

Foundation. We also thank the CIHR (111237) and the Heart and Stroke

Foundation of Ontario (NA-6305) for financial support.

7.7 References

1. E. H. Van Beers, H. A. Buller, R. J. Grand, A. W. Einerhand J. Dekker,

Crit. Rev. Biochem. Mol. Bio. 1995, 30, 197-262.

2. G. Semenza, Annu. Rev. Cell Biol. 1986, 2, 255-313.

3. H. A. Ernst, L. L. Leggio, M. Willemoës, G. Leonard, P. Blum, S. Larsen, J.

Mol. Biol. 2006, 358(4), 1106-1124.

4. R. Quezada-Calvillo, L. Sim, Z. Ao, B. R. Hamaker, A. Quaroni, G. D.

Brayer, E. E. Sterchi, C. C. Robayo-Torres, D. R. Rose, B. L. Nichols, J.

Nutr. 2008, 138, 685-692.

5. B. L. Nichols, S. Avery, P. Sen, D. M. Swallow, D. Hahn, E. Sterchi, Proc.

Natl. Acad. Sci. U S A 2003, 100(3), 1432-1437.

6. H. Heymann, D. Breitmeier, S. Günther, Biol. Chem. Hoppe-Seyler 1995,

376, 249-253.

7. H. Heymann, S. Günther, Biol. Chem. Hoppe-Seyler 1994, 375, 451-456.

8. C. Robayo-Torres, R. Quezada-Calvillo, B. L. Nichols, Clin. Gastroenterol.

Hepatol. 2006, 4, 276-287.

9. K. Jones, L. Sim, S. Mohan, K. Jayakanthan, H. Liu, S. Avery, H. H.

Naim, R. Quezada-Calvillo, B. L. Nichols, B. M. Pinto, D. R. Rose, Bioorg.

Med. Chem. 2011, 19, 3929-3934.

10. Naumoff, D. G. Mol. Biol. (Mosk). 2007, 41, 1056-1068.

213

11. a) J. P. C. Chandrasena, The Chemistry and Pharmacology of Ceylon

and Indian Medicinal Plants; H&C press: Colombo, Sri Lanka, 1935; b) D.

M. A. Jayaweera, Medicinal Plants Used in Ceylon-Part 1;

NationalScience Council of Sri Lanka: Colombo, 1981; p 77; c) P. S.

Vaidyartanam, In Indian Medicinal Plants: a compendium of 500 species,

P. K. Warrier V. P. K. Nambiar and C. Ramankutty, Eds.; Orient Longman:

Madras, India, 1993, pp. 47-48.

12. a) B. D. Johnston, H. H. Jensen, B. M. Pinto, J. Org. Chem. 2006, 71,

1111-1118; b) M. Yoshikawa, F. M. Xu, S. Nakamura, T. Wang, H.

Matsuda, G. Tanabe, O. Muraoka, Heterocycles 2008, 75, 1397-1405.

13. a) M. Yoshikawa, T. Murakami, H. Shimada, H. Matsuda, J. Yamahara, G.

Tanabe and O. Muraoka, Tetrahedron Lett. 1997, 38, 8367–8370; b) H.

Yuasa, J. Takada, H. Hashimoto, Tetrahedron Lett. 2000, 41, 6615-6618;

c) A. Ghavami, B. D. Johnston, B. M. Pinto, J. Org. Chem. 2001, 66,

2312-2317.

14. a) H. Matsuda, Y. H. Li, T. Murakami, N. Matsumura, J. Yamahara, M.

Yoshikawa, Chem. & Pharm. Bull. 1998, 46, 1399-1403; b) K.

Jayakanthan, S. Mohan, B. M. Pinto, J. Am. Chem. Soc. 2009, 131, 5621-

5626; c) R. Eskandari, K. Jayakanthan, D. A. Kuntz, D. R. Rose, B. M.

Pinto, Bioorg. Med. Chem. 2010, 18, 2829-2835.

15. a) S. Ozaki, H. Oe, S. Kitamura, J. Nat. Prod. 2008, 71, 981-984; b) H. Oe,

S. Ozaki, Biosci. Biotechnol. Biochem. 2008, 72, 1962-1964; c) O.

Muraoka, W. Xie, G. Tanabe, M. F. A. Amer, T. Minematsu, M.

Yoshikawa, Tetrahedron Lett. 2008, 49, 7315–7317.

16. a) Y. Minami, C. Kurlyarna, K. Ikeda, A. Kato, K. Takebayashi, I. Adachi,

G. W. J. Fleet, A. Kettawan, T. Karnoto, N. Asano, Bioorg. & Med. Chem.

2008, 16, 2734-2740; b) G. Tanabe, W. Xie, A. Ogawa, C. Cao, T.

Minematsu, M. Yoshikawa, O. Muraoka, Bioorg. Med. Chem. Lett. 2009,

19, 2195-2198.

214

17. a) R. Eskandari, D. A. Kuntz, D. R. Rose, B. M. Pinto, Org. Lett. 2010, 12,

1632-1635; b) W. Xie, G. Tanabe, J. Akaki, T. Morikawa, K. Ninomiya, T.

Minematsu, M. Yoshikawa, X. Wu, O. Muraoka, Bioorg. Med. Chem. 2011, 19, 2015-2022.

18. E. J. Rossi, L. Sim, D. A. Kuntz, D. Hahn, B. D. Johnston, A. Ghavami, M.

G. Szczepina, N. S. Kumar, E. E. Sterchi, B. L. Nichols, B. M. Pinto, D. R.

Rose, FEBS J. 2006, 273, 2673-2683.

19. L. Sim, K. Jayakanthan, S. Mohan, R. Nasi, B. D. Johnston, B. M. Pinto,

and D. R. Rose, Biochemistry 2010, 49, 443–451.

20. S. Mohan, B. M. Pinto, Carbohydr. Res. 2007, 342, 1551-1580.

21. S. Mohan, B. M. Pinto, Collect. Czech. Chem. Commun. 2009, 74, 1117-

1136.

22. S. Mohan, B. M. Pinto, Nat. Prod. Rep. 2010, 27, 481-488.

23. J. D. Wardrop, S. L. Waidyarachchi, Nat. Prod. Rep. 2010, 27, 1431-1468.

24. C. Li, A. Begum, S. Numao, K. H. Park, S. G. Withers, G. D. Brayer,

Biochem. 2005, 44(9), 3347-3357.

25. L. Sim, R. Quezada-Calvillo, E. E. Sterchi, B. L. Nichols, D. R. Rose, J.

Mol. Biol. 2008, 375, 782-792.

26. Y. Su, J. Xie, Y. Wang, X. Hu, X. Lin, Eur. J. Med. Chem. 2010, 45, 2713-2718.

27. H. Liu, R. Nasi, K. Jayakanthan, L. Sim, H. Heipel, D. R. Rose, B. M.

Pinto, J. Org. Chem. 2007, 72, 6562-6572.

28. H. Satoh, Y. Yoshimura, S. Sakata, S. Miura, H. Machida, A. Matsuda,

Bioorg. Med. Chem. Lett. 1998, 8, 989-992.

29. R. Eskandari, K. Jones, D. R. Rose, B. M. Pinto Bioorg. Med. Chem. Lett.

2010, 20, 5686-5689.

215

30. L. Sim, C. Willemsma, S. Mohan, H. Y. Naim, B.M. Pinto, D. R. Rose, J.

Biol. Chem. 2010, 285, 17763-17770.

31. R. Nasi, L. Sim, D. R. Rose, B. M. Pinto, J. Org. Chem. 2007, 72, 180-186

32. R. Eskandari, K. Jones, D. R. Rose, B. M. Pinto, Chem. Commun. 2011,

47, 9134-9136.

33. R. Eskandari, K. Jones, D. R. Rose, B. M. Pinto, Bioorg. Med. Chem. Lett.

2011, 21, 6491-6494.

216

7.8 Supporting Information

Probing the intestinal α-glucosidase enzyme specificities of starch-

digesting maltase-glucoamylase and sucrase-isomaltase: Synthesis and

inhibitory properties of 3′- and 5′- maltose-extended de-O-sulfonated

ponkoranol

Razieh Eskandari, Kyra Jones, Kongara Ravinder Reddy, Kumarasamy

Jayakanthan, Marcia Chaudet, David R. Rose, B. Mario Pinto

217

ppm (t1)050100

ppm (t1)0.01.02.03.04.05.06.07.0

O

OBn

O

O OMe

OMe

HO

600 MHz, CDCl314

OTBDMS

O

OBn

O

O OMe

OMe

HO

600 MHz, CDCl314

OTBDMS

218

ppm (t1)

050100

OO

O

OMe

OMe

OBnOH

19

OTBDMS

100 MHz, CDCl3

ppm (t1)0.01.02.03.04.05.06.07.0

400 MHz, CDCl3

OOO

TBDMSOOMe

OMeOBn

OH

17

OO

O

OMe

OMe

OBnOH

19

OTBDMS

400 MHz, CDCl3

219

ppm (t1)0.01.02.03.04.05.06.07.0

O

OBn

O

O OMe

OMe

OO

OAcAcO

O

OAcO

AcOAcO OAc

OTBDMS

OAc

15600 MHZ, CDCl3

ppm (t1)050100150

O

OBn

O

O OMe

OMe

OO

OAcAcO

O

OAcO

AcOAcO OAc

OTBDMS

OAc

15150 MHZ, CDCl3

220

ppm (t1)0.01.02.03.04.05.06.07.0

ppm (t1)050100150

OO

O

MeO

OMe

OBn

OOAc

O

OAc

OOAc

O

OAcOAc

20

OTBDMS

AcO

AcO

400 MHz, CDCl3

OO

O

MeO

OMe

OBn

OOAc

O

OAc

OOAc

O

OAcOAc

20

OTBDMS

AcO

AcO

100 MHz, CDCl3

221

ppm (t1)0.01.02.03.04.05.06.07.0

O

OBn

O

O OMe

OMe

OO

OBnBnO

O

OBnO

BnOBnO OBn

OTBDMS

OBn

16600 MHZ, CDCl3

ppm (t1)050100

O

OBn

O

O OMe

OMe

OO

OBnBnO

O

OBnO

BnOBnO OBn

OTBDMS

OBn

16150 MHz, CDCl3

222

ppm (t1)0.01.02.03.04.05.06.07.0

OO

O

MeO

OMe

OBn

OOBn

O

OBn

OOBn

O

OBnOBn

OTBDMS

BnO

BnO

21400 MHz, CDCl3

ppm (t1)050100

OO

O

MeO

OMe

OBn

OOBn

O

OBn

OOBn

O

OBnOBn

OTBDMS

BnO

BnO

21100 MHz, CDCl3

223

ppm (t1)2.03.04.05.06.07.0

ppm (t1)050100

O

OBn

O

O OMe

OMe

OO

OBnBnO

O

OBnO

BnOBnO OBn

OTf

OBn

17150 MHz, CDCl3

O

OBn

O

O OMe

OMe

OO

OBnBnO

O

OBnO

BnOBnO OBn

OTf

OBn

17600 MHz, CDCl3

224

ppm (t1)1.02.03.04.05.06.07.0

OO

O

MeO

OMe

OBn

OOBn

O

OBn

OOBn

O

OBnOBn

OTf

BnO

BnO

22600 MHz, CDCl3

ppm (t1)050100

OO

O

MeO

OMe

OBn

OOBn

O

OBn

OOBn

O

OBnOBn

OTf

BnO

BnO

22150 MHz, CDCl3

225

ppm (t1)2.03.04.05.06.07.0

SBnO

BnO OBn

OTf

24 d.r, 5:1

OOBn

O

O OMe

OMe

OO

OBnBnO

O

OBnO

BnOBnO OBn

OBn

600 MHz, CDCl3

ppm (t1)50100

SBnO

BnO OBn

OTf

24 d.r, 5:1

OOBn

O

O OMe

OMe

OO

OBnBnO

O

OBnO

BnOBnO OBn

OBn

150 MHz, CDCl3

226

ppm (t2)3.503.603.703.803.904.00

3.50

4.00

ppm (t1

compound 24 NOESY mix t, 500 ms

ppm (t1)1.02.03.04.05.06.07.0

SBnO

BnO BnO

OTf

OO

O

MeO

OMe

OBn

OOBn

O

OBn

OOBn

O

OBnOBn

BnO

BnO

25a600 MHz, CDCl3

H-4 H-6’a

227

ppm (t1)50100

SBnO

BnO BnO

OTf

OO

O

MeO

OMe

OBn

OOBn

O

OBn

OOBn

O

OBnOBn

BnO

BnO

25a150 MHz, CDCl3

ppm (t1)1.02.03.04.05.06.07.0

SBnO

BnO BnO

OTf

OO

O

MeO

OMe

OBn

OOBn

O

OBn

OOBn

O

OBnOBn

BnO

BnO

25b600 MHz, CDCl3

228

ppm (t1)050100

SBnO

BnO BnO

OTf

OO

O

MeO

OMe

OBn

OOBn

O

OBn

OOBn

O

OBnOBn

BnO

BnO

25b150 MHz, CDCl3

ppm (t1)3.504.004.505.005.50

S

OHHO

OH

OH OH

HO

OH

O

10

O OH

HO

O OH

O

HOOH

HO

Cl

HO

600 MHz, D2O

229

ppm (t1)5060708090100

S

OHHO

OH

OH OH

HO

OH

O

10

O OH

HO

O OH

O

HOOH

HO

Cl

HO

150 MHz, D2O

ppm (t1)3.504.004.505.005.50

S

OHHO

OH O

OH OH

HO

OH

11

O OH

HO

O OH

O

HOOH

HO

HO

Cl

600 MHz, D2O

230

Representative Lineweaver-Burk plot of ntMGAM inhibited by 11 at concentrations of 0 nm, 30 nM, 125 nM, and 200 nM.

ppm (t1)405060708090100110

S

OHHO

OH O

OH OH

HO

OH

11

O OH

HO

O OH

O

HOOH

HO

HO

Cl

150 MHz, D2O

231

CHAPTER 8: CONCLUSIONS AND FUTURE WORK

8.1 Conclusions

The work described in this thesis focused on the design and syntheses of

α-glucosidase inhibitors. Since the isolation of kotalanol in late 1990s, the

absolute stereostructure of kotalanol 1 (Figure 8-1) had not been determined.

Our group successfully established the absolute stereostructure of kotalanol (1)

and also completed its total synthesis and that of its de-O-sulfonated derivative 2

(Figure 8-1).1 It was interesting to note that de-O-sulfonated kotalanol (2) showed

higher α-glucosidase inhibitory activities than the original sulfate (1). In Chapter 2

of this thesis, an alternative route for the synthesis of kotalanol (1) as well as the

synthesis of its 6'-epimer (3) were shown.2 As part of these extended studies, the

syntheses of de-O-sulfonated ponkoranol (4) and its 5'-epimer (5)3 (Figure 8-1),

and their selenium analogues,4 (6 and 7, Figure 8-1) were completed by the

thesis author as shown in Chapters 3 and 4. Biological evaluation of compounds

4-7 as glucosidase inhibitors against two recombinant intestinal enzymes

maltase-glucoamylase (MGAM) and sucrase-isomaltase (SI) were also described

(Chapter 4).4 The de-O-sulfonated ponkoranol (4) was also used as a scaffold for

designing additional inhibitors because extension of the polyhydroxylated side

chain beyond six carbon was shown not to be beneficial.

The crystal structures of ntMGAM in complex with kotalanol (1) and de-O-

sulfonated kotalanol (2),5 obtained by our collaborator Dr. David R. Rose, showed

232

that the 3'-O-sulfate anion in kotalanol was surrounded by hydrophobic residues

in the enzyme active site, and made no hydrogen bonding interactions with other

residues. Therefore, 3'-O-methylponkoranol (8) (Figure 8-1) was synthesized

(Chapter 5) to see whether the methyl group in this compound would make

hydrophobic interactions within the ntMGAM active site and lead to a tighter

inhibitor.6 In Chapter 6, the inhibition profile of 3'-O-methylponkoranol (8) for the

individual recombinant N- and C- terminal MGAM and SI were reported.7 The

results indicated that the interaction of the methyl group with hydrophobic

residues in the active site is not beneficial to inhibition of ntMGAM but resulted in

significant inhibition of ctSI.

Figure 8-1: Structures of compounds 1-8.

To better define the individual roles of the MGAM and SI domains in the

process of terminal starch digestion, we proposed further modification to the

S

OHHO

OSO3 OH

OH OH

HO

OHOH

1

S

OHHO

OH OH

OH OH

HO

OHOH

CH3OSO3

2

S

OHHO

OSO3 OH

OH OH

HO

OHOH

3

S

OHHO

OH OH

OH OH

HO

OH

X: S=4X: Se=6

ClS

OHHO

OH

OH OH

HO

OH

X: S=5X: Se=7

ClOH S

OHHO

OMe OH

OH OH

HO

OH

8

Cl

233

ponkoranol scaffold to toggle their activities on and off with domain specific

inhibitors.

From the structural and biological studies of ntMGAM and ctMGAM, it was

postulated that ctMGAM might have an extended binding site compared to

ntMGAM, which favors binding of longer inhibitors such as acarbose (9, Figure 8-

2).8 Based on this difference we synthesized extended structures of de-O-

sulfonated ponkoranol (4) by linking maltose units at the C3’ (10) or C5’ (11)

position of the hydroxylated side chain; these additional sugars were intended to

provide further contacts with the subsites within the active site. The inhibitory

activities of these compounds against individual subunits of MGAM and SI

indicated that C-3′-β-maltose-extended de-O-sulfonated ponkoranol 10 showed

some selectivity for inhibiting ntMGAM compared to ctMGAM, ntSI, and ctSI; C-

5′-β-maltose-extended de-O-sulfonated ponkoranol 11 showed the ability to

inhibit all the subunits very well especially, ntMGAM.9

Examination of the enzyme inhibitory activities showed for the first time

that de-O-sulfonated ponkoranol (4), its 5'-epimer (5), and their selenium

analogues (6, 7) could differentiate between ctMGAM spilceforms. We also

showed that substitution of the ring sulfur atom by selenium leads to 13-fold and

23-fold improvement in inhibition of ntSI; and 6-fold and 7-fold improvement in

inhibition of ctSI (4 and 5 vs. 6 and 7, respectively). Interestingly, 3'-O-

methylponkoranol (8) could not differentiate between ctMGAM spliceforms and

inhibited both spliceforms; however, it showed extraordinary selectivity for ctSI (7

± 2 nM). The inhibitory activities of C-3′-β-maltose-extended de-O-sulfonated

234

ponkoranol 10 and 5′-β-maltose-extended de-O-sulfonated ponkoranol 11

indicated that 10 showed higher potency against other subunits as compared

with ctMGAM, whereas 11 demonstrated the ability to inhibit all subunits,

especially ntMGAM. A significant finding was that 10 also could differentiate

between ctMGAM spliceforms and showed selectivity towards the different

enzyme activities. Based on a previous study acarbose (9) was shown to be a

very powerful inhibitor of ctMGAM over ntMGAM and SI subunits. Therefore, we

are now in a position to turn on and off the ctMGAM unit and study the effect of

other subunits in starch digestion.

Ultimately, small structural changes in a compound can result in

significant changes in inhibitory activity and selective inhibition of one enzyme

over the others, despite the overall similarity between the subunits in term of

amino acid sequence.

235

Figure 8-2: Structures of compounds 9-11.

8.2 Future work

It was observed from kinetic studies, that maltose (1,4-glycosidic linkage)

could be hydrolyzed by both ntMGAM and ntSI, whereas isomaltose (1,6-

glycosidic linkage) could only be hydrolyzed efficiently by ntSI.10 With respect to

MGAM and SI roles in terminal starch digestion, the different substrate

specificities of ntMGAM and ntSI seem consistent. The redundancy in α-1,4

OHN

HO

OOH

H3C

O

OOH

HO

O

HO OH

HO

HOHO

OH

HO

HO

OH9

S

OHHO

OH O

OH OH

HO

OH

S

OHHO

OH

OH OH

HO

OH

O

O OH

HO

O OH

O

HOOH

HO

O OH

HO

O OH

O

HOOH

HO

X

HO HO

X

10 11

236

activity seen in both enzymes (also in ctMGAM and ctSI) is beneficial due to the

higher distribution of α-1,4 linkages (95%) in starch molecules. In contrast, α-1,6

linkages make up a small fraction of the linkages (5%) in starch molecules.

Thus, only one subunit (ntSI) is likely necessary for processing these structures.

Close inspection of the active site complexes of kotalanol with ntMGAM and ntSI,

indicates that it might be possible to increase the inhibitory potency towards ntSI

through the addition of a side chain with an isomaltose extension, as shown in

compound 12 (Figure 8-3). Hence, we propose compound 12 as a potential

inhibitor of ntSI, with selectivity over ntMGAM. Selective inhibitors of ntSI could

then be used to further our understanding of the critical role of different subunits

of SI and MGAM in the starch digestion process and the structural basis of

substrate specificity in these intestinal α-glucosidases.

Figure 8-3: Proposed selective inhibitors of ntSI.

For the synthesis of target compounds 12, we propose a synthetic route

as shown in Scheme 8-1.

S

OHHO

OH

HO

12HO

O

HO OH

O

HOO

OH

O

HO

OMe

OHOH

OH

237

Scheme 8-1: Proposed synthetic route for target compound 12.

O

O

O

OH

HOHO

OH

HOHO

OH

OMe

O

O

O

OH

BnOBnO

OBn

BnOBnO

OBn

OMe1. TrCl, Py.2. BnBr, NaH,3.70% AcOH,

CBr4, PPh3

O

O

O

Br

BnOBnO

OBn

BnOBnO

OBn

OMe

OOBnO

HO

OTBS

OMeO

MeO

, NaH

OOBnO

Isomaltose-O

OTBS

OMeO

MeO

1. TBAF2. TfO2

OOBnOIsomaltose-O

OTf

OMeO

MeO

S

BnO OBnBnO

2. BCl33. NaBH4

1.

Methyl maltoside

13

S

OHHO

OH

HO

12HO

O

HO OH

O

HOO

OH

O

HO

OMe

OHOH

HO

238

For the synthesis of the target compound 12, we propose a synthetic route

that takes advantage of the same key intermediate 139 from the synthesis of our

3'-O-maltose-extended de-O-sulfonated ponkoranol (see Scheme 8-1).

Since α-amylase starch-digesting enzymes in the small-intestine break

down starch to α-limit dextrins, SI and MGAM are exposed to α-limit dextrins as a

substrate rather than maltose. The true in vivo substrate composition of SI and

MGAM is likely to change throughout the entire length of the small intestine, as

starch is broken down gradually as it passes through. Therefore, an important

avenue to be explored is the role of our different inhibitors on MGAM and SI

enzyme activities using dextrin structures as substrates to determine how our

inhibitors inhibit real starch (α-limit dextrins) and whether they have preferences

for inhibition of one subunit over others. Finally, experiments in which the

different inhibitors are mixed with α-limit dextrins and the mixtures are used as

substrates for the different intestinal mucosal enzymes, ntMGAM, ctMGAM, ntSI,

and ctSI should be explored. The thesis that one should be able to control starch

digestion to affect slow release of glucose should be explored further.

239

8.3 References

1. Jayakanthan, K.; Mohan, S.; Pinto, B. M. J. Am. Chem. Soc. 2009, 131,

5621-5626.

2. Eskandari, R.; Jayakanthan, K.; Kuntz, D. A.; Rose, D. R.; Pinto B. M.

Bioorg. Med. Chem. 2010, 18, 2829-2835.

3. Eskandari, R.; Kuntz, D. A.; Rose, D. R.; Pinto, B. M. Org. Lett. 2010, 12,

1632-1635.

4. Eskandari, R.; Jones, K.; Rose, D. R.; Pinto, B. M. J. Chem. Soc., Chem.

Commun. 2011, 47, 9134-9136.

5. Sim, L.; Jayakanthan, K.; Mohan, S.; Nasi, R.; Johnston, B. D.; Pinto, B.

M.; Rose, D. R. Biochemistry 2009, 49, 443-451.

6. Eskandari, R.; Jones, K.; Rose, D. R.; Pinto B. M. Bioorg. Med. Chem.

Lett. 2010, 20, 5686-5689.

7. Eskandari, R.; Jones, K.; Rose, D. R.; Pinto, B. M. Bioorg. Med. Chem.

Lett. 2011, 21, 6491-6494.

8. Sim, L.; Quezada-Calvillo, R.; Sterchi, E. E.; Nichols, B. L.; Rose, D. R. J.

Mol. Biol. 2008, 375, 782-792.

9. Eskandari, R.; Jones, K.; Reddy, K. R.; Jayakanthan, K. Chaudet, M.;

Rose, D. R.; Pinto, B. M. Chem. Eur. J. 2011, 17, 14817-14825.

10. Sim, L.; Willemsma, C.; Mohan, S.; Naim, H. Y.; Pinto, B. M.; Rose, D. R.

J. Biol. Chem. 2010, 285, 17763-17770.