Top Banner
Borinic Acid-Catalyzed Sulfation and Boronic Acid-Promoted Esterification of Carbohydrates by Yu Chen Lin A thesis submitted in conformity with the requirements for the degree of Master of Science Department of Chemistry University of Toronto © Copyright by Yu Chen Lin 2017
151

Borinic Acid-Catalyzed Sulfation and Boronic Acid-Promoted ......work on utilizing boronic acids as protective groups for the preparation of sugar fatty acid ester surfactants is also

Apr 25, 2020

Download

Documents

dariahiddleston
Welcome message from author
This document is posted to help you gain knowledge. Please leave a comment to let me know what you think about it! Share it to your friends and learn new things together.
Transcript
  • Borinic Acid-Catalyzed Sulfation and Boronic Acid-Promoted Esterification of Carbohydrates

    by

    Yu Chen Lin

    A thesis submitted in conformity with the requirements for the degree of Master of Science

    Department of Chemistry University of Toronto

    © Copyright by Yu Chen Lin 2017

  • ii

    Borinic Acid-Catalyzed Sulfation and Boronic Acid-Promoted

    Esterification of Carbohydrates

    Yu Chen Lin

    Master of Science

    Department of Chemistry University of Toronto

    2017

    Abstract

    Carbohydrates and their O-sulfates play important roles in biological functions, including cellular

    recognition and adhesion, neural processes, fibrosis, growth factor regulation, cancer metastasis,

    and cellular entry of viruses. However, preparation of sulfated carbohydrates remains a synthetic

    challenge with conventional methods requiring lengthy protection and deprotection steps.

    Described herein is our work toward the development of a method for the regioselective sulfation

    of fully unprotected carbohydrates using a borinic acid catalyst. Via an activated 1,2-cis-borinate

    intermediate, our method was shown to be robust in the sulfation of a range of substrates, including

    the synthesis of a sulfated galactosylceramide found in mammalian nervous systems. In addition,

    work on utilizing boronic acids as protective groups for the preparation of sugar fatty acid ester

    surfactants is also discussed.

  • iii

    Acknowledgments

    I would like to thank my supervisor, Professor Mark Taylor, for the opportunity to join his lab,

    and his continued support in my studies, research, and pursuits.

    I am also very grateful to all the Taylor lab members whom I’ve had the pleasure of meeting and

    working together with. You have all made me feel so at home here, and although it has only been

    a year, the great memories I’ve had here will stay with me long after. Furthermore, I’d like to give

    a shout out to my friends in and around the Department for every laughter and drink we’ve shared.

    Lastly, I would like to thank my family for their unconditional love and support throughout the

    years and during my degree.

  • Table of Contents Acknowledgments .................................................................................................................... iii

    Table of Contents ..................................................................................................................... iv

    List of Tables ........................................................................................................................... vi

    List of Figures ......................................................................................................................... vii

    List of Abbreviations ............................................................................................................... ix

    Chapter 1 Introduction ........................................................................................................1

    1.1 Sulfated carbohydrates ...................................................................................................1

    1.2 Direct, regioselective synthesis of sulfated carbohydrates ............................................4

    1.3 Synthesis of sulfated carbohydrates via masked sulfates ............................................11

    1.4 Organoboron compounds in carbohydrate chemistry ..................................................18

    1.5 Scope of Thesis ............................................................................................................20

    Chapter 2 Regioselective, catalytic sulfation of unprotected carbohydrates ....................21

    2.1 Sulfation of carbohydrates with 2,2,2-trichloroethyl chlorosulfate .............................21

    2.2 Sulfation of carbohydrates with alkyl and aryl 1,2-dimethylimidazolium salt ............22

    2.3 Sulfation of carbohydrates with sulfur trioxide amine complexes ..............................29

    2.4 Summary and future work ...........................................................................................32

    2.5 Experimental ................................................................................................................33

    2.5.1. General Information .........................................................................................33

    2.5.2. General Procedure A ........................................................................................34

    2.5.3. Preparation of catalyst and carbohydrate substrates ........................................34

    2.5.4. Synthesis and characterization of compounds .................................................36

    Chapter 3 Boronic acid-promoted Fischer esterification ..................................................47

    3.1 Introduction ..................................................................................................................47

    3.2 Chemical synthesis of sugar fatty acid esters ..............................................................49

  • v

    3.3 Summary ......................................................................................................................51

    3.4 Experimental ................................................................................................................51

    3.4.1 General Information .........................................................................................51

    3.4.2 General Procedure B ........................................................................................52

    3.4.3 Synthesis and characterization of compounds .................................................52

    Appendices ...............................................................................................................................58

    A1.NMR spectra of reported compounds ..........................................................................58

  • vi

    List of Tables Table 01. Optimization of methyl 6-O-TBS-α-D-mannopyranoside sulfation with 2,2,2-TCE chlorosulfate ..................................................................................................................................22

    Table 02. Optimization of methyl α-L-rhamnopyranoside sulfation with TCE-sulfuryl 1,2- dimethylimidazolium triflate .........................................................................................................24

    Table 03. Solvent screen of methyl α-L-rhamnopyranoside sulfation with TCE-sulfuryl 1,2- dimethylimidazolium triflate .........................................................................................................25

    Table 04. Boronic acid/Lewis base co-catalyst for sulfation of methyl α-L-rhamnopyranoside ....26

    Table 05. Optimization of methyl α-L-rhamnopyranoside sulfation with arylsulfuryl 1,2-dimethylimidazolium triflate .........................................................................................................28

    Table 06. Optimization of methyl α-D-mannopyranoside sulfation with sulfur trioxide amine complex .........................................................................................................................................30

    Table 07. Optimization of n-octyl α-D-galactopyranoside sulfation with sulfur trioxide amine

    complex .........................................................................................................................................31

    Table 08. Substrate scope of carbohydrate sulfation with SO3-Me3N ...........................................32

    Table 09. Substrate scope for fatty acid and sugar alcohol esterification .......................................50

  • vii

    List of Figures Figure 01. Structure of select naturally occurring sulfated carbohydrates .......................................2

    Figure 02. Structure of the major tri-sulfated disaccharide repeat unit in heparin ...........................3

    Figure 03. Structure of heparin derivatives .....................................................................................4

    Figure 04. Conventional routes to access various patterns of carbohydrate sulfation ......................5

    Figure 05. Transient boronate ester protection in the regioselective sulfation of steroids ...............6

    Figure 06. Temperature-dependent regioselective sulfation of galactoside ....................................7

    Figure 07. Effect of SO3-amine complexes in the sulfation of trimethylsilyl cellulose ...................8

    Figure 08. Regioselective sulfation with dibutyltin oxide of 1,2-cis-diol and 1,2-cis-dioxy ...........9

    Figure 09. Regioselective sulfation with dibutyltin oxide of 1,3-cis-diol ......................................10

    Figure 10. Regioselective sulfation with dibutyltin oxide in the absence of 1,2-cis-diol ...............10

    Figure 11. Routes of nucleophilic attack on a carbohydrate sulfate diester ...................................11

    Figure 12. Preparation and unmasking of phenyl sulfate diesters ..................................................12

    Figure 13. Preparation and unmasking of alkyl sulfate diesters, a) sulfation with iBu/nP

    chlorosulfate, b) unmasking of iBu sulfate diesters, c) unmasking of nP sulfate diesters ...............13

    Figure 14. Preparation and unmasking of trifluoroethyl sulfate diesters .......................................14

    Figure 15. Stability of TFE-masked sulfate diester to further functionalizations ..........................15

    Figure 16. Deprotection of TFE-masked sulfate diesters ..............................................................16

    Figure 17. Preparation and unmasking of trichloroethyl sulfate diesters .......................................17

    Figure 18. Regioselective TCE-masked sulfation of unprotected carbohydrates ..........................18

    Figure 19. Tetracoordinate organoboron activation of diols for regioselective functionalization .19

  • viii

    Figure 20. Synthesis of trichloroethyl chlorosulfate .....................................................................21

    Figure 21. Synthesis of 2,2,2-trichloroethoxysulfuryl 1,2-dimethylimidazolium salts .................23

    Figure 22. Sulfation with pre-formed cyclic arylboronate of methyl α-L-rhamnopyranoside .......26

    Figure 23. Preparation of alkyl- and arylsulfuryl 1,2-dimethylimidazolium triflate .....................27

    Figure 24. Decomposition study of 2.08, 2.18, 2.19 in CD3CN; A: 2.08 in 1 equiv. 1,2-

    dimethylimidazole; B: 2.08 in 1 equiv. DIPEA; C: 2.18 in 1 equiv. DIPEA; D: 2.18 in 1 equiv. 3

    MS-dried DIPEA; E: 2.19 in 1 equiv. DIPEA; F: 2.19 in 1 equiv. 3 MS-dried DIPEA .........29

    Figure 25. Structure of select fatty acids and sugar alcohols .........................................................47

    Figure 26. Preparation of sugar fatty acid esters with lipases ........................................................49

  • List of Abbreviations

    Ac acetyl

    app apparent Bn benzyl

    br broad

    Bu butyl

    Bz benzoyl calcd. calculated

    Cp cyclopentadienyl

    CSA (1S)-(+)-10-camphorsulphonic acid d doublet

    DAST (diethylamino)sulfur trifluoride

    DBU 1,8-diazabicyclo[5.4.0]undec-7-ene

    DCM dichloromethane

    dec. decomposed

    DIPEA N-N-diisopropylethylamine DMA N-N-dimethylacetamide

    DMAP 4-(dimethylamino)pyridine

    DMF N-N-dimethylformamide DMPU 1,3-dimethyl-3,4,5,6-tetrahydro-2(1H)-pyrimidinone equiv equivalents

    ESI electrospray ionization

    Et ethyl GML glycerol monolaurate h hour

    hept heptet

    HIV human immunodeficiency virus

    HMDS bis(trimethylsilyl)amide

  • x

    HMPA hexamethylphosphoramide

    HRMS high-resolution mass spectra iBu isobutyl

    Im imidazole

    iPr isopropyl

    IR infrared m multiplet (NMR), medium (IR)

    m meta

    Me methyl

    min minute

    MP 4-methoxyphenyl

    Ms methanesulfonyl

    MS molecular sieve NBS N-bromosuccinimide NMP N-methylpyrrolidine

    NMR nuclear magnetic resonance

    nP neopentyl o ortho p pentet

    p para

    Ph phenyl

    Piv pivaloyl

    PMB para-methoxybenzyl

    PMP 1,2,2,6,6-pentamethylpiperidine

    ppm parts per million q quartet rt room temperature s singlet (NMR), strong (IR)

    SFAE sugar fatty acid ester

    SIV simian immunodeficiency virus

  • xi

    t triplet

    TBAF tetrabutylammonium fluoride

    TBS tert-butyldimethylsilyl tBu tert-butyl

    TCE 2,2,2-trichloroethyl

    TDS thexyl-dimethylsilyl

    Tf trifluoromethanesulfonyl

    TFA trifluoroacetic acid

    TFE 2,2,2-trifluoroethyl

    THF tetrahydrofuran

    TMS trimethylsilyl

    TREAT-HF triethylamine trihydrofluoride UV ultraviolet w weak

  • 1

    Chapter 1 Introduction

    Carbohydrates present a wealth of diverse structures derived from joining together

    monosaccharides, primarily in their pyranose or furanose forms. These structures can undergo

    further functionalization such as alkylation, macrocyclization, phosphorylation, and sulfation.

    Their structural diversity is mirrored by their broad range of biological functions and properties,

    from energy storage as starch and structural support as cellulose, to cellular recognition and other

    intercellular functions as glycolipids and glycoproteins.

    The monosaccharides that make up complex carbohydrate structures contain several hydroxyl

    groups that differ in their stereochemical arrangements. Site selectivity among similar hydroxyl

    groups presents a difficult challenge for chemical modification. However, certain distinguishing

    features among monosaccharides have been exploited. Reactivity differences among amines,

    primary and secondary hydroxyl groups allow for site-selective modifications. cis-Diols present

    in mannose and galactose can be used to distinguish between two secondary hydroxyl groups.

    Differences between axial and equatorial hydroxyl groups can be used to influence stereochemical

    outcomes.

    1.1 Sulfated carbohydrates Sulfated carbohydrates are common in nature and play key roles in biological functions (Figure

    01). Dermatan sulfate 1.01, found in skin and blood vessels, plays a role in coagulation and

    fibrosis.1 Heparan sulfate 1.02, found on all cell surfaces, is a proteoglycan that functions in cell

    adhesion, blood coagulation, and growth factor regulation.1 Sulfated sialyl-Lewisx 1.03, found on

    cell surfaces, binds preferentially to lymphocyte cell-adhesion molecule L-selectin2 and has been

    1 Capila, I.; Linhardt, R. J. Angew. Chem. Int. Ed. 2002, 41, 390-412.

    2 Julien, S.; Ivetic, A.; Grigoriadis, A.; QiZe, D.; Burford, B.; Sproviero, D.; Picco, G.; Gillett, C.; Papp, S. L.;

    Schaffer, L. Cancer Res. 2011, 71, 7683-7693.

  • 2

    shown to play a role in bladder urothelial carcinoma metastasis.3 Sulfated galactosylceramide 1.04,

    found in the myelin sheath of nerve cells, play various functions in the nervous system. Abnormal

    expression of these sulfoglycolipids is associated with neurological disorders such as Alzheimer’s

    and Parkinson’s diseases.4 These sulfoglycolipids are also involved in the progression of other

    illnesses such as diabetes mellitus and the cellular entry of HIV-1.5

    Figure 01. Structure of select naturally occurring sulfated carbohydrates.

    Perhaps the most well-known sulfated carbohydrates belong to the heparin polysaccharide family

    (Figure 02). Discovered in 1916, naturally occurring heparin contains on average 25 units of the

    disaccharide 1.05, giving a molecular weight of 5–40 kDa.1 It is the biological macromolecule

    with the highest density of negative charge due to its sulfate and carboxylate groups. Heparin

    3 Taga, M.; Hoshino, H.; Low, S.; Imamura, Y.; Ito, H.; Yokoyama, O.; Kobayashi, M. Urol. Oncol.: Semin. Orig.

    Invest. 2015, 33, 496.e1-496.e9. 4 Eckhardt, M. Mol. Neurobiol. 2008, 37, 93-103.

    5 Compostella, F.; Panza, L.; Ronchetti, F. C. R. Chim. 2012, 15, 37-45.

    OO O

    OH-O3SO

    NHAc

    O OHO

    OSO3-HOOC

    Dermatan sulfate 1.01

    O-O3SO

    O

    OSO3-

    O

    AcHNO

    OHO

    OSO3-HOOC

    OHO

    OH

    AcHN

    OO

    OH

    HOOCOO

    Heparan sulfate 1.02

    OAcHN

    COOH

    OH

    OH

    HOHO

    OO O

    OSO3-OH

    OH

    OO OH

    OH

    NHAc

    OOH

    OHOH

    6-Sulfo sialyl-Lewisx 1.03

    OO

    -O3SO

    OHOH

    OHC13H27

    OH

    HN fatty acid

    O

    Sulfated galactosylceramide 1.04

  • 3

    initiates the blood coagulation cascade by binding to enzyme antithrombin III, which then

    inactivates thrombin and other proteases.1

    Figure 02. Structure of the major tri-sulfated disaccharide repeat unit in heparin.

    First approved in 1939, heparin sodium salt is an anticoagulant drug administered through

    intravenous catheter or as an injection. It remains one of the oldest pharmaceutical drugs that is

    still in use. Pharmaceutically relevant effects aside from anticoagulation include anti-inflammatory

    properties for the treatment of ulcerative colitis and relief of obstructive pulmonary diseases like

    asthma.6 Most pharmaceutical-grade heparin on the market is isolated from animal tissues,

    particularly from porcine intestine or bovine lungs. However, the sulfation pattern of livestock-

    grown heparin is variable and difficult to control, and the extracted heparin is prone to viral and

    bacterial contamination. Recently, mammalian cell production of heparin using the Chinese

    hamster ovary cell lines have been shown to be a safer and more robust alternative.6,7

    Due to the success of heparin as a drug, a number of derivatives and low-molecular weight

    analogues have been studied. Fondaparinux 1.06, a pentasaccharide marketed by

    GlaxoSmithKline, was approved in 2001 as an anticoagulant (Figure 03).8 Fondaparinux has

    advantages over heparin in that the former has a longer half-life, thus requiring a lower dosage.

    6 Oduah, E. I.; Linhardt, R. J.; Sharfstein, S. T. Pharmaceuticals 2016, 9, 38.

    7 Baik, J. Y.; Gasimli, L.; Yang, B.; Datta, P.; Zhang, F.; Glass, C. A.; Esko, J. D.; Linhardt, R. J.; Sharfstein, S. T.

    Metab. Eng. 2012, 14, 81-90. 8 Bauer, K. A.; Hawkins, D. W.; Peters, P. C.; Petitou, M.; Herbert, J. M.; Boeckel, C. A. A.; Meuleman, D. G.

    Cardiovasc. Ther. 2002, 20, 37-52.

    O

    O

    OSO3-

    O

    -O3SHN

    O OHO

    OSO3--OOCHO

    Heparin 1.05

  • 4

    Pentosan polysulfate 1.07, a plant-derived oligosaccharide, exhibits anti-HIV activity.9 Heparin

    tetrasaccharide 1.08, with increased oral bioavailability, has anti-allergic activity and is being

    studied for the treatment of asthma.10

    Figure 03. Structure of heparin derivatives.

    1.2 Direct, regioselective synthesis of sulfated carbohydrates Synthesis of sulfated carbohydrates remains a challenge despite their long history of biologically

    relevant properties. Sulfation is commonly carried out with sulfur trioxide-amine complexes,

    together forming a Lewis acid-base adduct. These complexes are easier to handle than liquid sulfur

    trioxide, which requires distillation prior to use. The relative reactivities of the SO3 complexes

    9 Baba, M.; Nakajima, M.; Schols, D.; Pauwels, R.; Balzarini, J.; De Clercq, E. Antiviral Res. 1988, 9, 335-343. 10

    Ahmed, T.; Smith, G.; Abraham, W. M. Pulm. Pharmacol. Ther. 2013, 26, 180-188.

    O

    OSO3-

    O

    -O3SHNHO

    O

    OSO3-OH

    -OOCOH

    O

    O

    OSO3-O

    -OOCOH

    OOH

    OHOSO3-

    Heparin tetrasaccharide 1.08

    O

    OSO3-O

    OSO3-

    n

    Pentosan polysulfate 1.07

    O

    O

    OSO3-

    HO

    -O3SHNHO

    O O-OOC

    OH

    HOO

    O

    OSO3-

    -O3SHN-O3SO

    O

    -OOC OSO3-HO

    O

    OMe

    OSO3-

    O

    -O3SHNHO

    Fondaparinux 1.06

  • 5

    generally vary inversely with the strength of the Lewis base component. Common complexes used

    are listed in order of decreasing basicity: Me3N ≈ Et3N > pyridine > DMF.11

    Conventional routes to sulfated carbohydrates consist of numerous protection and deprotection

    steps. As exemplified in Figure 04, the synthesis of galactopyranoside monosulfate requires

    multiple orthogonal protecting groups, functional group manipulation, and selective deprotection

    to reveal the free hydroxyl at the position of interest for sulfation.12

    Figure 04. Conventional routes to access various patterns of carbohydrate sulfation.12

    To overcome lengthy protection/deprotection steps, methods for direct and regioselective sulfation

    have been studied.13 Boronate ester protection was utilized by McLeod to transiently mask the cis-

    diols of steroid 1.09 for selective functionalization at the remaining hydroxyl group.14 The

    11

    Gilbert, E. E. Chem. Rev. 1962, 62, 549-589. 12 Marinier, A.; Martel, A.; Banville, J.; Bachand, C.; Remillard, R.; Lapointe, P.; Turmel, B.; Menard, M.; Harte, W. E.; Wright, J. J. K. J. Med. Chem. 1997, 40, 3234-3247. 13 Al-Horani, R. A.; Desai, U. R. Tetrahedron 2010, 66, 2907-2918. 14

    Hungerford, N. L.; McKinney, A. R.; Stenhouse, A. M.; McLeod, M. D. Org. Biomol. Chem. 2006, 4, 3951-3959.

    OHO OR

    OHHO

    OH

    OO OR

    OHO

    OH

    OO OR

    OSO3-O

    OH

    OO OR

    OPivO

    OH

    OO OR

    OPivO

    OSO3-

    OHO OR

    OPivAcO

    OPiv

    SO3-amine

    O-O3SO OR

    OPivAcO

    OPiv

    OBzO OR

    OBz

    OO

    Ph

    OBzO OR

    OPivOH

    OBz

    SO3-amine

    OBzO OR

    OPiv-O3SO

    OBz

    SO3-amineSO3-amine

    6-O-monosulfate 2-O-monosulfate 3-O-monosulfate 4-O-monosulfate

  • 6

    boronate ester of 1.10 can then be cleaved to reveal the cis-diol in 1.11 for selective sulfation to

    give 1.12 (Figure 05). Although this route still requires protecting groups, it reduces the number

    of separate purification steps.

    Figure 05. Transient boronate ester protection in the regioselective sulfation of steroids.14

    A temperature-dependent regioselective sulfation of galactoside 1.14 was described by Kondo

    (Figure 06).15 Sulfation performed at room temperature with SO3-pyridine yielded the 3,4-O-bis-

    sulfate 1.15 while at 0 oC, the reaction afforded only the 4-O-sulfate 1.16 without observable 3-O-

    or bis-sulfate. The 3-O-sulfate 1.18, however, could only be prepared after first protecting the C4-

    hydroxyl group. This method was not shown to be generalizable to other sugar moieties or to be

    applicable in the presence of exposed primary or C2-hydroxyl groups.

    15 Tsukida, T.; Yoshida, M.; Kurokawa, K.; Nakai, Y.; Achiha, T.; Kiyoi, T.; Kondo, H. J. Org. Chem. 1997, 62, 6876-6881.

    HO

    OH

    OH

    H

    1) PhB(OH)2, DMF/CH2Cl22) TBS-Cl, imidazole

    TBSOH

    H2O2, aq. NaOH

    THF73% over 3 steps

    OBO

    Ph

    TBSO

    OH

    OH

    H

    SO3-pyridineDMF, pyridine

    61%

    TBSO

    OH

    OSO3- +Na

    H

    80% AcOH/H2O

    69%HO

    OH

    OSO3- +Na

    H

    1.09 1.10

    1.11

    1.121.13

  • 7

    Figure 06. Temperature-dependent regioselective sulfation of galactoside.15

    Following up on previous reports of sulfate insertion into O-Si bonds,16,17 Richter showed that the

    Lewis base used in the sulfate complex can be tuned to influence the regiochemical outcome

    (Figure 07).18 Trimethylsilyl cellulose 1.19 with SO3-DMF preferentially yields the 6-O-sulfate

    1.20 while SO3-Et3N preferentially yields the 2-O-sulfate 1.21. The authors rationalize this effect

    by stating that the electron-donation of Et3N polarizes the O-S bond of its SO3 complex, promoting

    insertion at the more polarized O-Si bond of the O-2 position. This electron-donating effect is

    absent in the DMF complex, which favors the more sterically accessible position at O-6.

    16 Stein, A.; Wagenknecht, W.; Philipp, B.; Klemm, D.; Schnabelrauch, M., German Patent DD 299313, 1989. 17 Wagenknecht, W.; Nehls, I.; Stein, A.; Klemm, D.; Philipp, B. Acta Polym. 1992, 43, 266-269. 18 Richter, A.; Klemm, D. Cellulose 2003, 10, 133-138.

    OHO OR

    OBzOH

    OBz

    SO3-pyridineDMF, rt

    73%

    SO3-pyridineDMF, 0 oC

    73%

    OHO OR

    OBzAcO

    OBz

    SO3-pyridineDMF, 0 oC

    90%

    O-O3SO OR

    OBzAcO

    OBz

    OHO OR

    OBz-O3SO

    OBz

    O-O3SO OR

    OBz-O3SO

    OBz

    1.14

    1.17

    1.15 1.16 1.18

  • 8

    Figure 07. Effect of SO3-amine complexes in the sulfation of trimethylsilyl cellulose.18

    Dibutyltin oxide was employed by Flitsch in 1994 to facilitate the regioselective sulfation of 1.22

    (Figure 08).19,20 The dibutylstannylene acetal 1.23 was first formed at the cis-diol group with

    super-stoichiometric amount of dibutyltin oxide, followed by a solvent-switch and addition of SO3-

    Me3N complex to afford the product of sulfation at the more sterically accessible position of the

    19

    Guilbert, B.; Davis, N. J.; Flitsch, S. L. Tetrahedron Lett. 1994, 35, 6563-6566. 20

    Guilbert, B.; Davis, N. J.; Pearce, M.; Aplin, R. T.; Flitsch, S. L. Tetrahedron: Asymmetry 1994, 5, 2163-2178.

    OTMSO O

    OTMS

    OTMS

    SO3-DMFTHF

    SO3-Et3NTHF

    OTMSO O

    O

    OTMS

    SO

    OO

    SiO

    TMSO O

    OTMS

    OSO

    OO Si

    NaOHMeOH

    NaOHMeOH

    OHO O

    OSO3-

    OH

    OHO O

    OH

    OSO3-

    major product, 1.20 major product, 1.21

    1.19

  • 9

    1,2-cis-diol. The galactosylceramide glycolipid 1.24, found in mammalian nervous systems, was

    synthesized in excellent yield. This strategy was also shown for disaccharides, including a

    lactoside 1.25 to give the 3′-O-sulfate 1.26 as the major product with 10% of the 3′,6′-O-bis-sulfate

    byproduct 1.27. In the absence of Bu2SnO, this reaction gave no observable 1.26. Maltosides 1.28

    protected at the primary hydroxyl groups that do not possess a cis-diol were selectively 2′-O-

    sulfated to give 1.29 in decent yield. The authors rationalize that the regioselectivity is due to either

    the higher reactivity of the 2′-hydroxyl group or the C1′-C2′ cis-dioxy configuration. However,

    they did not perform the control reaction to show that the maltoside sulfation is only regioselective

    in the presence of Bu2SnO.

    Figure 08. Regioselective sulfation with dibutyltin oxide of 1,2-cis-diol and 1,2-cis-dioxy.20

    In 2004, Gelb adapted the method for the synthesis of 1.32, a fluorescent probe used in the assay

    for screening newborns against Hunter syndrome, a disease caused by the deficiency of iduronate-

    2-sulfatase (Figure 09).21 The dibutyltin oxide, in this case, binds to the 1,3-cis-diol of iduronic

    ester 1.30, delivering the sulfate to the more nucleophilic O-2 affording 1.31.13

    21

    Blanchard, S.; Turecek, F.; Gelb, M. H. Carbohydr. Res. 2009, 344, 1032-1033.

    OHO OR

    OHOH

    OH

    Bu2SnO (1.5 equiv)

    MeOHreflux, 2 h

    ORO

    O

    OHO

    OH

    SnBu

    BuSO3-Me3N (2 equiv)

    THFrt, 4 h

    1.2497%

    O-O3SO OR

    OHOH

    OH

    R =HN

    OH

    C13H27

    O

    C8H17

    1.22 1.23

    OHO

    O

    OHHO

    OH

    OOH

    OH

    OHSPh

    1) Bu2SnO (1 equiv) MeOH, reflux, 2 h

    2) SO3-Me3N (2 equiv) dioxane, rt, 30 h

    ORO

    O

    OR′HO

    OH

    OOH

    OH

    OHSPh

    1.25 1.26 R = SO3-, R′ = H1.27 R = SO3-, R′ = SO3-

    76%10%

    OOH

    OTBS

    OHO-allylO

    OOH

    OH

    OO

    Ph1) Bu2SnO (1 equiv) MeOH, reflux, 2 h

    2) SO3-Me3N (2 equiv) dioxane, rt, 93 h

    OOH

    OTBS

    OHO-allylO

    OOH

    -O3SO

    OO

    Ph

    1.28 1.2956%

  • 10

    Figure 09. Regioselective sulfation with dibutyltin oxide of 1,3-cis-diol.21

    In 2009, Kosma demonstrated the regioselective sulfation of xylopyranosides and xylotriosides.22

    Low-molecular weight xylans and their sulfates have been shown to possess antithrombin23 and

    antiasthmatic24 activities. With the dibutyltin oxide strategy, 1.33 and 1.35 gave the terminal 4-O-

    sulfated products 1.34 and 1.36, respectively (Figure 10).22 However, this selectivity for substrates

    without the presence of a cis-diol was not explained, and the remaining mass balance of the

    sulfation of 1.35 was unaccounted for.

    Figure 10. Regioselective sulfation with dibutyltin oxide in the absence of 1,2-cis-diol.22

    Although activation with dibutyltin oxide has shown to be highly selective for sulfating the

    equatorial hydroxyl of a cis-diol, it requires—at minimum—a stoichiometric amount of tin, long

    reaction times (93 h to afford 1.29), and a two-step activation/sulfation process. That being said,

    22

    Abad-Romero, B.; Mereiter, K.; Sixta, H.; Hofinger, A.; Kosma, P. Carbohydr. Res. 2009, 344, 21-28. 23 Yamagaki, T.; Tsuji, Y.; Maeda, M.; Nakanishi, H. Biosci. Biotechnol. Biochem. 1997, 61, 1281-1285. 24 Kuszmann, J.; Medgyes, G.; Boros, S. Carbohydr. Res. 2005, 340, 1739-1749.

    O

    OHOH

    O

    MeOOCOH O O

    O

    NH

    HNt-BuO

    O 1) Bu2SnO (1.5 equiv) MeOH, reflux, 40 min

    2) SO3-Me3N (1.5 equiv) DMF, 55 oC, 24 h O

    OSO3-OH

    O

    MeOOCOH O O

    O

    NH

    HNt-BuO

    O

    O

    OSO3-OH

    O

    -OOCOH O O

    O

    NH

    HNt-BuO

    ONaOH

    H2O

    1.3261% over 2 steps

    1.30 1.31

    HO OHO

    OHOMe

    1) Bu2SnO (1.08 equiv) toluene, reflux, 15 h

    2) SO3-Me3N (1.1 equiv) THF, rt, 48 h

    -O3SOO

    HOOH

    OMe

    O OHO

    OHHO O

    HOOH

    O OHO

    OHOMe

    1) Bu2SnO (4 equiv) toluene, reflux, 15 h

    2) SO3-Me3N (3.2 equiv) THF, rt, 72 h

    O OHO

    OH

    -O3SOO

    HOOH

    O OHO

    OHOMe

    1.33 1.3474%

    1.35 1.3627%

  • 11

    direct sulfation of carbohydrates presents a concise and efficient alternative to traditional

    protecting group-based methods. However, this strategy is often only applicable as the final step

    in a complex synthesis. The highly polar sulfated products are insoluble in organic solvents,

    making them difficult to purify and manipulate for subsequent elaborations. To this end, masked

    sulfates have been devised.

    1.3 Synthesis of sulfated carbohydrates via masked sulfates Masked sulfates, or protected sulfate esters, provide a way to address the issues of sulfated

    products, allowing for chemical transformations of the molecule elsewhere. Following appropriate

    transformations, the sulfate can be unmasked to reveal the desired product. These masked sulfates

    must be stable to conventional purification techniques and survive a wide range of reaction

    conditions including, in particular, the deprotection conditions of other functional groups on the

    molecule.

    Sulfate diesters have generally been used, providing the added benefit of a spectroscopic handle

    that is compatible with common characterization techniques. Sulfate diesters are prone to

    nucleophilic attack and substitution at one of three possible sites (Figure 11).25 Substitution via

    Route A leading to desulfation is generally slow, particularly on a carbohydrate’s secondary

    hydroxyl group. Design of sulfate diesters would ideally disfavor attack on the sulfur center (Route

    B), or premature deprotection of the ester (Route C).

    Figure 11. Routes of nucleophilic attack on a carbohydrate sulfate diester.25

    25

    Proud, A. D.; Prodger, J. C.; Flitsch, S. L. Tetrahedron Lett. 1997, 38, 7243-7246.

    SO O

    O OR carbohydrate

    NuA

    BC

  • 12

    Perlin first introduced phenyl chlorosulfate as a masked sulfating group for carbohydrates.26 The

    stable phenyl chlorosulfate is synthesized from phenol, sodium hydroxide, and sulfuryl chloride

    in high yields. After reaction with 1.37 (Figure 12), unmasking of the sulfate diester 1.38 was

    achieved by catalytic hydrogenolysis of the phenyl group to a cyclohexyl group, followed by alkyl

    fission to give 1.39. Although yields for unmasking the phenyl group were not reported, the authors

    stated that about 10% desulfation had occurred. The phenyl-masked sulfate is stable to the

    TFA/CHCl3 conditions used to deprotect the 5,6-O-isopropylidene group, and also stable to 1:1

    Ac2O:H2SO4, which was used to deprotect both isopropylidene groups. However, this method is

    not compatible with functional groups sensitive to base and hydrogenolysis.

    Figure 12. Preparation and unmasking of phenyl sulfate diesters.26

    Alkyl-masked sulfating groups have been described by Widlanski with particular emphasis on

    isobutyl (iBu) and neopentyl (nP) chlorosulfates.27 The carbohydrate 1.37 was first stirred in

    NaHMDS, followed by the addition of the masked chlorosulfate to give sulfated products 1.40 and

    1.41 (Figure 13a). Sulfation with iBu chlorosulfate was slower and required large excess of the

    chlorosulfate, while sulfation with nP chlorosulfate required DMPU as a co-solvent. The stabilities

    of the sulfates were studied using phenyl iBu or nP sulfate diesters as model substrates. The iBu

    sulfate diester degraded completely in 6% piperidine after 24 h, but was relatively stable in 50%

    TFA (7.1% degradation after 24 h). No degradation of the nP sulfate diester was observed in 20%

    piperidine, versus 6.9% degradation in 50% TFA after 24 h. Both sulfate diesters were stable under

    26

    Penney, C. L.; Perlin, A. S. Carbohydr. Res. 1981, 93, 241-246. 27

    Simpson, L. S.; Widlanski, T. S. J. Am. Chem. Soc. 2006, 128, 1605-1610.

    OOH

    OO

    OO

    1) NaH, THF, 30 min

    2)SO

    OClPhO , 20 h

    OO

    OO

    OO

    SOPh

    O O

    1.3875%

    K2CO3, PtO2, H2

    EtOH, H2O, 20 h

    OOSO3-

    OO

    OO

    1.37 1.39

  • 13

    conditions used to deprotect benzyl and isopropylidene groups (hydrogenolysis with Pd/C and H2,

    and acidification with aq. H2SO4/THF)

    Figure 13. Preparation and unmasking of alkyl sulfate diesters, a) sulfation with iBu/nP

    chlorosulfate, b) unmasking of iBu sulfate diesters, c) unmasking of nP sulfate diesters.27

    Unmasking of the diester was examined through Route C (Figure 11) via the attack of a small

    nucleophile on the alkyl group to reveal the sulfated carbohydrate. iBu sulfate diester 1.42 was

    unmasked cleanly in the presence of NaI at elevated temperatures to give 1.43 (Figure 13b). nP

    sulfate diester 1.44, however, was more difficult to unmask due to its increased bulk. Under the

    same deprotection conditions for the iBu sulfate diester, the nP sulfate diester 1.44 yielded no

    OOH

    OO

    OO

    1) NaHMDS, THF

    2) ClSO2OiBu (5–10 equiv), −15 oCor

    ClSO2OnP (1 equiv), DMPU, −75oC

    OO

    OO

    OO

    SOR

    O O

    1.40 R = iBu 95%1.41 R = nP 95%

    OiBu-O3SO

    OH

    HO

    OH OH

    NaI

    acetone, 55 oCO

    -O3SO

    OH

    HO

    OH OH1.4397%

    a)

    b)

    c)

    NaI

    acetone, 55 oC

    OOSO3-nP

    OO

    OO

    no reaction

    OnP-O3SO

    OH

    HO

    OH OH

    O

    OH

    HO

    OH OHN3

    OOSO3-nP

    OO

    OO

    OOSO3-

    OO

    OO

    NaN3

    DMF, 70 oC

    NaN3

    DMF, 70 oC

    1.3998%

    1.37

    1.42

    1.44

    1.44

    1.45 1.46

  • 14

    reaction (Figure 13c). NaN3 was found to be able to unmask the nP sulfate diester of the

    glucofuranose sulfate 1.44 to give 1.39, but led to the azide substitution product of the

    glucopyranose sulfate 1.45 to give 1.46. Although the method shown by Widlanski were not

    generally applicable and showed limited scope, it does offer a masked-sulfate approach with the

    potential for alkyl group manipulation to tune the reactivity of the masked sulfate.

    Trihaloethyl-masked sulfates have also long been considered as an alternative to aryl-masked

    sulfates. Trifluoroethyl (TFE) and trichloroethyl (TCE) are thought to hinder Routes B and C

    (Figure 11) compared to alkyl and aryl masking groups due to both steric and electronic reasons.

    TCE esters have been used in the protection of phosphate and carboxyl groups, and removed

    selectively by Zn/AcOH. Flitsch sought to extend trihaloethyl masking to the sulfation of

    carbohydrates.25 Focusing initially on TCE-masked sulfates, the authors were unable to sulfate

    carbohydrates in sufficiently high yields for further studies, citing steric hindrance of the trichloro

    group. Switching direction to TFE-masked sulfates, the authors were unable to introduce the TFE-

    sulfate in one step from the TFE-chlorosulfate. Instead, the carbohydrate 1.47 had to be first

    sulfated with an SO3-amine complex, followed by TFE protection with 2,2,2-trifluorodiazoethane

    in moderate yields (Figure 14). The TFE-sulfate diester 1.48 was stable to conditions including

    hydrogenation, 20% TFA/EtOH, and NaOMe/MeOH used for isopropylidene deprotection. Harsh

    conditions (reflux in t-BuOK/t-BuOH) were required for attack on the sulfur center and elimination

    of TFE to give the unmasked sulfate 1.49. However, deprotection of TFE-sulfate diester on O-2

    or O-3 positions resulted in sulfate migration to free, adjacent hydroxyl groups. Apart from

    difficult deprotection, this masked sulfate strategy was not widely adopted in carbohydrate

    synthesis due to the need for a two-step sulfation process and the potentially explosive nature of

    trifluorodiazoethane.

    Figure 14. Preparation and unmasking of trifluoroethyl sulfate diesters.25

    1) SO3-pyridine, MeCN, 80 oC

    2) F3C N2citric acid, MeCN, rt

    1.4851%

    OO

    O

    OHO

    O

    OO

    O

    OSO3TFEO

    O OHO

    OSO3-HO

    OH OH

    1) 4:1 TFA:EtOH rt, 2 h, 97%

    2) t-BuOK, t-BuOH, reflux, 96%

    1.47 1.49

  • 15

    Linhardt then showed the versatility and stability of TFE-masked sulfate diesters in subsequent

    functionalizations. Sulfate diesters 1.50 and 1.52 were stable to either TBAF in AcOH or TREAT-

    HF (Figure 15), thus demonstrating orthogonality to silyl protective groups.28 The masked sulfate

    diesters were also stable to fluorination and glycosylation conditions, affording TFE-masked

    sulfate disaccharides 1.54 and 1.56 in decent yields.

    Figure 15. Stability of TFE-masked sulfate diester to further functionalizations.28

    28

    Karst, N. A.; Islam, T. F.; Linhardt, R. J. Org. Lett. 2003, 5, 4839-4842.

    OBzO OTDS

    OSO3TFE

    O

    N3

    OBzO

    OSO3TFE

    O

    N3

    1) TBAF, AcOH, THF, −25 oC

    2) DAST, DCM, , −30 oC

    OBzO OTDS

    OSO3TFEBnO

    N3

    F

    1) TREAT-HF

    2) DAST, DCM, −30 oC

    1.5166%

    1.5363%

    OBzO

    OSO3TFEBnO

    N3 F

    OBzO

    OSO3TFEBnO

    N3F

    +O

    O

    O

    OHO

    OAgClO4, Cp2ZrCl2

    4 Å MS, DCMO

    O

    O

    OO

    O

    OBzO

    OSO3TFEBnO

    N3

    1.5464%

    OO

    O

    OHO

    O+

    BF3 Et2O

    4 Å MS, DCM

    OBzO

    OBnTFEO3SO

    N3O

    NH

    Cl3CO

    O

    O

    OO

    O

    OBzO

    OBnTFEO3SO

    N3

    1.5649%

    1.50

    1.52

    1.53 1.47

    1.55 1.47

    OCl

    OCl

  • 16

    Deprotection of TFE-masked sulfate diesters proved to be quite challenging.29 Previously reported

    t-BuOK deprotection conditions were effective on monosaccharide 1.57 to give 1.58, but led to

    the decomposition of disaccharide 1.59 (Figure 16). However, NaOMe/MeOH was found to be

    able to unmask the diester of disaccharide 1.59 in good yields to give 1.60 with minimal

    decomposition. Deprotection of TFE-masked bis-sulfate disaccharide 1.61 required a two step

    process. Unmasking with NaOMe/MeOH gave 1.62 in excellent yield, but the harsher t-BuOK

    reagent used in the second step gave 1.63 in only moderate yield, along with disaccharide

    decomposition, desilylation, and desulfonation.

    Figure 16. Deprotection of TFE-masked sulfate diesters.29

    In 2004, Taylor re-visited Flitsch’s attempts to develop a method for TCE-masked sulfate diesters

    and succeeded in sulfating aryl alcohols with TCE chlorosulfate.30 In follow-up reports of

    extending this methodology to the sulfation of carbohydrate 1.47, the authors managed to

    synthesize the sulfate diester 1.64 in approximately 50% yield, with the majority of the remaining

    mass balance being the chlorosugar byproduct 1.65 (Figure 17). This result, in contrary to Flitsch’s

    29

    Karst, N. A.; Islam, T. F.; Avci, F. Y.; Linhardt, R. J. Tetrahedron Lett. 2004, 45, 6433-6437. 30

    Liu, Y.; Lien, I. F. F.; Ruttgaizer, S.; Dove, P.; Taylor, S. D. Org. Lett. 2004, 6, 209-212.

    OBnO OMP

    OSO3TFE

    OOPh t-BuOK, t-BuOH

    OO

    O

    OO

    O

    OBzO

    OBnTFEO3SO

    N3 NaOMe, MeOH

    OBnO OMP

    OSO3-O

    OPh

    OBzO

    OBnTFEO3SO

    N3

    OBnO OMP

    OTBS

    O

    OSO3TFE

    NaOMe, MeOHO

    HO

    OBn-O3SO

    N3

    OBnO OMP

    OTBS

    O

    OSO3TFE

    t-BuOK, t-BuOHO

    HO

    OBn-O3SO

    N3

    OBnO OH

    OTBS

    O

    OSO3-

    OO

    O

    OO

    O

    OHO

    OBn-O3SO

    N3

    1.5882%

    1.6070%

    1.6290%

    1.6350%

    1.57

    1.59

    1.61

  • 17

    report in 1997,25 showed that TCE-sulfated carbohydrates could indeed be prepared in good yields.

    Encouraged by this finding, Taylor designed a sulfating reagent that did not release an effective

    nucleophilic species. With the optimized 1,2-dimethylimidazolium triflate salt 1.67 in hand, the

    authors achieved TFE-masked sulfation of carbohydrate 1.66 to give 1.68 in high yields, as well

    as deprotection of 1.68 under mild conditions with Zn or Pd/C and ammonium formate to give

    1.58 (Figure 17).31 The sulfating reagent 1.67 was stable to prolonged storage at room temperature

    and the TCE-masked sulfate diester 1.68 was stable to a range of reaction conditions:

    NaOMe/MeOH and ZnCl2/AcOH/Ac2O for debenzylation and deacetylation, NBS in

    acetone/water, DBU, and acidic conditions for benzylidene opening.

    Figure 17. Preparation and unmasking of trichloroethyl sulfate diesters.31,32

    Masked sulfation provides products that can be manipulated for further synthetic transformations.

    However, despite these recent advances in the field of masked sulfation of carbohydrates, there

    remains a need for a reliable method toward the regioselective installation of a masked sulfate

    group, particularly on fully unprotected carbohydrates. In 2016, Kaji and Makino33 bridged that

    gap by using the transient boronate ester protection for the regioselective sulfation of steroids

    31

    Ingram, L. J.; Desoky, A.; Ali, A. M.; Taylor, S. D. J. Org. Chem. 2009, 74, 6479-6485. 32

    Ingram, L. J.; Taylor, S. D. Angew. Chem. Int. Ed. 2006, 45, 3503-3506. 33

    Fukuhara, K.; Shimada, N.; Nishino, T.; Kaji, E.; Makino, K. Eur. J. Org. Chem. 2016, 2016, 902-905.

    OO

    O

    OHO

    O

    SO

    OClTCEO

    AgCN, Et3N, DMAP, THF

    OO

    O

    OSO3TCEO

    O

    OO

    O

    ClO

    O

    1.64~ 50%

    +

    OBnO OMP

    OH

    OOPh

    SO

    OTCEO N N

    OTf

    1,2-dimethylimidazoleDCM, 24 h

    OBnO OMP

    OSO3TCE

    OOPh

    1.6896%

    Zn or Pd/CHCO2NH4

    MeOH

    OBnO OMP

    OSO3-O

    OPh

    1.58with Zn: 94%

    1.47 1.65

    1.66

    1.67

  • 18

    described previously14 and the TCE-masked sulfate developed by Taylor.31 In one pot, unprotected

    carbohydrates 1.69 with 1,2-cis- or 4,6-diols are protected by phenylboronic acid, followed by

    TCE-masked sulfation at the remaining hydroxyl group of 1.70, and transesterification of the

    boronate ester with pinacol to reveal the TCE-masked sulfate carbohydrate 1.71 in good yields

    (Figure 18).

    Figure 18. Regioselective TCE-masked sulfation of unprotected carbohydrates.33

    1.4 Organoboron compounds in carbohydrate chemistry As discussed previously, organoboron compounds have been used in carbohydrate chemistry for

    functional group protection14,33 among other purposes.34 In addition, organoboron compounds in

    their tetracoordinate state have been used for the activation of diols. As first described by Aoyama,

    the phenylboronate ester of the 1,2-cis-diol of methyl α-L-fucopyranoside 1.72 can be activated

    by an amine base to give the tetracoordinate complex 1.73 (Figure 19).35 It can then undergo

    regioselective alkylation at the more accessible position to give 1.74. Other regioselective

    transformations following this strategy have also been explored, including the glycosidation of

    34

    McClary, C. A.; Taylor, M. S. Carbohydr. Res. 2013, 381, 112-122. 35

    Oshima, K.; Kitazono, E.-I.; Aoyama, Y. Tetrahedron Lett. 1997, 38, 5001-5004.

    O

    OR(HO)n

    PhB(OH)2

    DCMrt, 24–48 h

    O

    OR(HO)n

    OOB

    Ph

    1)

    SO

    OTCEO N N

    OTf

    1,2-dimethylimidazole4 Å MS, THF or DCM20–24 h, 0 oC to rt

    2) pinacol, DCM

    O

    OR(HO)n

    TCEO3SO

    Product:

    OTCEO3SO O

    OHHO

    OH

    OHO OMe

    OH

    HO

    TCEO3SO

    O

    OH

    TCEO3SO

    OMe

    OH

    76% 74% 66% 72%

    1.691.70 1.71

    OOSO3TCE

    OH

    OMe

    OH

  • 19

    unprotected carbohydrate 1.72, via the tetracoordinate complex 1.75, to give the disaccharide 1.76

    (Figure 19).36

    Figure 19. Tetracoordinate organoboron activation of diols for regioselective functionalization.34

    Expanding on the reactivity of tetracoordinate boronates, our group has developed catalytic

    versions employing borinic acid derivatives, avoiding the need for complexation with a Lewis

    36

    Oshima, K.; Aoyama, Y. J. Am. Chem. Soc. 1999, 121, 2315-2316.

    OOH

    OH

    OMe

    OH

    1) PhB(OH)2

    2) Ag2O, Et3N benzene, reflux

    OOH

    O

    OMe

    O

    B PhEt3N

    n-BuI OOH

    OH

    OMe

    On-Bu

    1.7450%

    OOH

    OH

    OMe

    OH Ag2CO3Et4N+ I-, 4 Å MS

    OOH

    O

    OMe

    O

    B

    OOH

    OH

    OMe

    O

    1.7693%

    O BOH

    O

    OAcO

    Br

    OAc

    AcO

    AcO

    OAcO

    OAc

    AcO

    AcO

    OHO

    OMe

    OTBSHO

    HO

    BO

    H2NPh

    Ph(10 mol%)

    BzCl, iPr2NEt

    MeCNO

    O

    OMe

    OTBSO

    HO

    BPhPh

    BzO NHBz

    OBzO

    OMe

    OTBSHO

    HO

    1.7995%

    1.721.73

    1.72

    1.75

    1.77 1.78

  • 20

    base. Catalytic, regioselective acylation,37 sulfonylation,38 alkylation,39 and glycosylation40 of

    carbohydrates have since been accomplished. An example with 1.77 is shown in Figure 19,

    forming the tetracoordinate complex 1.78, to give the regioselectively benzoylated product 1.79.

    1.5 Scope of Thesis Carbohydrates and their O-sulfated derivatives have long been known to play important roles in

    biology. With a growing number of carbohydrate drugs being approved in recent years, it follows

    that concise and elegant approaches to the synthesis of carbohydrate derivatives, including O-

    sulfates, are also increasingly needed. Furthermore, organoboron compounds have shown to be

    powerful in imparting regiocontrol in carbohydrate functionalization, both as protecting and

    activating groups.

    The remainder of this thesis discusses my research on expanding the scope of borinic acid catalysis

    and boronic ester protection in carbohydrate synthesis. Chapter 2 outlines the development of a

    method for the regioselective sulfation of unprotected sugars. In contrast to previously described

    sulfation methods, this synthesis employs catalytic borinic acid to effect direct sulfation with high

    selectivity. Chapter 3 describes the use of boronic acids in Fischer esterification between sugar

    alcohols and fatty acids for the preparation of surfactants. The esterification project outlined in this

    last chapter was conducted in conjunction with a fellow laboratory group member, Sanjay Manhas,

    and individual contributions to the joint work are delineated therein.

    37 Lee, D.; Taylor, M. S. J. Am. Chem. Soc. 2011, 133, 3724-3727. 38

    Lee, D.; Williamson, C. L.; Chan, L.; Taylor, M. S. J. Am. Chem. Soc. 2012, 134, 8260-8267. 39

    Chan, L.; Taylor, M. S. Org. Lett. 2011, 13, 3090-3093. 40

    Gouliaras, C.; Lee, D.; Chan, L.; Taylor, M. S. J. Am. Chem. Soc. 2011, 133, 13926-13929.

  • 21

    Chapter 2 Regioselective, catalytic sulfation of unprotected carbohydrates

    As discussed in Chapter 1, although regioselective methods using stoichiometric amounts of

    additives have been developed for direct sulfation, they are only suitable as the final step in a

    synthesis due to the highly polar nature of the products. Masked sulfates, with their various

    protective ester groups, address some of the issues of sulfated products but no regioselective

    strategies have been developed for sulfation. My research began by employing the catalytic

    method developed in our lab using borinic acid-activation of cis-diols to prepare TCE-masked

    sulfates of carbohydrates.

    2.1 Sulfation of carbohydrates with 2,2,2-trichloroethyl chlorosulfate

    TCE chlorosulfate 2.02 was prepared according to literature41 in 84% yield from sulfuryl chloride

    2.01 and 2,2,2-trichloroethanol (Figure 20).

    Figure 20. Synthesis of trichloroethyl chlorosulfate.

    Initial attempts for TCE-masked sulfation of 6-O-TBS-α-D-mannopyranoside 2.03 with 2.02,

    catalyzed by the borinic ester pre-catalyst 2.04 yielded the cyclic sulfate 2.05 in low yields (Table

    01). After the initial sulfation reaction, the sulfate diester presumably undergoes a rapid cyclization

    with the adjacent free hydroxyl group on the carbohydrate. The 2,2,2-trichloroethoxide generated

    in situ as a result of the cyclization may explain the intermolecular TBS migration product 2.06

    observed under 1.5 equivalents of DIPEA or Et3N (Entries 1 and 5). However, the sugar without a

    TBS-group was not able to be isolated cleanly for characterization. Furthermore, the strongly basic

    41

    Pitts, A. K.; O'Hara, F.; Snell, R. H.; Gaunt, M. J. Angew. Chem. Int. Ed. 2015, 54, 5451-5455.

    SO

    Cl ClO

    pyridine (1 equiv)Et2O (0.5 M)

    1.5 h at −20 oC, then 0.5 h at rt

    SO

    O ClOCl3C

    (1 equiv)2.01

    2,2,2-trichloroethanol (1 equiv)

    2.0284%

  • 22

    ethoxide could cause decomposition of the sulfating reagent 2.02, leading to low yields of sulfated

    products. Decreasing the reaction time (Entry 2), using DCM as solvent (Entry 3), or using 1.1

    equivalents of DIPEA (Entry 4) also gave cyclic sulfates with none of the desired, uncyclized

    product observed. A preliminary base screen with Et3N and pyridine (Entries 5, 6) did not yield

    any productive results. 1,2,2,6,6-Pentamethylpiperidine (PMP) was used as a bulky base (Entry 7)

    in hopes of minimizing cyclic sulfate formation, but the TCE-masked sulfate diester was not

    observed.

    Table 01. Optimization of methyl 6-O-TBS-α-D-mannopyranoside sulfation with 2,2,2-TCE

    chlorosulfate.

    2.2 Sulfation of carbohydrates with alkyl and aryl 1,2-dimethylimidazolium salt

    Following Taylor’s success using the sulfating reagent in the form of an imidazolium salt,31 we

    synthesized the 2,2,2-trichloroethoxysulfuryl 1,2-dimethylimidazolium salts from the reaction of

    2.02 with 2-methylimidazole to give 2.07, followed by methylation to give the

    dimethylimidazolium salts (Figure 21). MeOTf gave high yields of 2.08, but methylation with MeI

    or Me3OBF4 gave low conversion to 2.09 and 2.10, respectively, and were not pursued further.

    OHO

    HO

    OH

    OMe

    TBSO

    Base (1.5 equiv)2.04 (10 mol %)MeCN (0.2 M)

    24 h, rt2.03

    OHOO

    O

    OMe

    OTBSSOO

    OTBSOO

    O

    OMe

    OTBSSOO

    +

    BaseEntry

    2.05 2.06

    2.03 Yield (%)a 2.04 Yield (%)a

    DIPEADIPEA DIPEA DIPEAEt3N

    pyridine 1,2,2,6,6-pentamethylpiperidine

    4010636

    2919

    5---1--

    12b3c4d567

    aCrude 1H NMR yields were determined based on tetramethylsilane as an internal standard. b6 hour reaction time. cDCM used as solvent. d1.1 equivalents of 1.58 and DIPEA used.

    2.042-aminoethyl phenylborinate

    OB

    NH2

    Ph

    Ph(1.5 equiv)

    SO

    O ClOCl3C

  • 23

    Figure 21. Synthesis of 2,2,2-trichloroethoxysulfuryl 1,2-dimethylimidazolium salts.

    We then investigated TCE-masked sulfation with the imidazolium triflate salt using methyl α-L-

    rhamnopyranoside as a model substrate for easier characterization compared to the TBS-protected

    sugars and to avoid the possibility of silyl migration (Table 02). Conditions A with a tertiary amine

    base for substrate/catalyst binding in MeCN were based on previously optimized conditions for

    catalyst activity in our group.37 Conditions B with 1,2-dimethylimidazole as base in DCM at 0 oC

    to room temperature were based on Taylor’s optimized conditions for TCE-sulfation.31 TCE-

    masked sulfated rhamnopyranosides were observed, and a clear pattern of regioselectivity became

    evident. With Conditions A, there is a slight preference for the 3-O-sulfated product 2.12 which

    presumably goes through a tetracoordinate borinate complex, delivering the TCE-sulfate to the

    more accessible, equatorial position at O-3. However, with Conditions B, background reaction in

    the absence of catalyst was very prominent, with a slight preference for the 2-O-sulfated product

    2.13. The catalyst did not appear to have a significant effect for Conditions B, affording neither an

    increase in yield nor influence on regioselectivity. This may be due to the inability of 1,2-

    dimethylimidazole to promote effective substrate binding to the borinic acid catalyst.

    SO

    OClTCEO

    NHN

    SO

    OTCEO N N

    MeOTf (1 equiv)SO

    OTCEO N N

    OTf

    2.0776%

    2.0882%

    2.02

    (3.6 equiv)

    THF (0.025 M)1 h at 0 oC,

    then 1 h at rt

    Et2O (0.2 M)0 oC, 3 h

    MeI (1 equiv)

    THF (0.2 M)0 oC to rt, 4 d

    Me3OBF4 (1 equiv)

    THF (0.2 M)0 oC, 24 h

    SO

    OTCEO N N

    I

    2.09low conversion

    SO

    OTCEO N N

    BF4

    2.10low conversion

  • 24

    Table 02. Optimization of methyl α-L-rhamnopyranoside sulfation with TCE-sulfuryl 1,2-

    dimethylimidazolium triflate.

    Proceeding with the result that gave the most promising regioselectivity (Table 02, Entry 4,

    Condition A), a solvent screen quickly revealed that amide solvents were essential (Table 03).

    DMF and DMA gave the 3-O-sulfate 2.12 in improved regioselectivity (Entries 4, 5). A control

    reaction in DMA without catalyst gave trace yields (Entry 6), showing that the borinic acid is

    crucial. However, the switch to NMP did not prove to be productive (Entry 7). The use of solvent

    mixtures, DMA/MeCN and DMA/H2O, also did not prove to be beneficial (Entries 8, 10). In an

    attempt to improve yield, we increased the reaction temperature to 50 oC but were met with

    decreased yield, although complete regiocontrol (Entry 9). This result was puzzling at the time,

    and will be re-visited later on in the chapter.

    OHOHO

    OH

    OMe

    Conditions A or Conditions B

    (1.5 equiv)SO

    O NOCl3C N

    OTf

    OHOTCEO3SO OH

    OMe

    OHOHO

    OSO3TCE

    OMe

    +

    Equiv. of BaseEntry Yield (%)a with

    Conditions AYield (%)a with Conditions B

    123456

    2.11 2.12 2.13

    15243937354

    2.12:2.13 Ratiowith Conditions A

    2.12:2.13 Ratiowith Conditions BCatalyst (XX mol%)

    no catalyst2.04 (10 mol%)2.14 (5 mol%)

    2.15 (10 mol%)no catalyst

    2.04 (10 mol%)

    1:11:1

    1.5:11.8:1

    1:13:1

    605450--

    41

    1:21:21:2.3

    1:1.5

    2.04

    OB

    NH2

    Ph

    Ph

    2.14

    O BBPh

    Ph

    Ph

    Ph

    2.15

    B

    O

    OH

    1.51.51.51.50.20.2

    --

    aCrude 1H NMR yields were determined based on tetramethylsilane as an internal standard. Product ratio determined by 1H NMR. Conditions A: catalyst (XX mol%), DIPEA (YY equiv), MeCN (0.2 M), 24 h, rt. Conditions B: catalyst (XX mol%), 1,2-dimethylimidazole (YY equiv), DCM (0.2 M), 24 h, 0 oC to rt.

  • 25

    Table 03. Solvent screen of methyl α-L-rhamnopyranoside sulfation with TCE-sulfuryl 1,2-

    dimethylimidazolium triflate.

    After further optimization of the reaction conditions (base, catalyst, various methods of reagent

    addition) in DMA, we were unable to raise the yield of the desired 3-O-TCE-sulfated product 2.12

    above 50% and achieved only modest regioselectivity. Turning to alternate modes of activation,

    we looked at complexation induced activation with boronic acid derivatives. The pre-formed cyclic

    arylboronate of methyl α-L-rhamnopyranoside 2.14 was hypothesized to be activated by an

    external Lewis base (Et3N) through a tetracoordinate intermediate for sulfation (Figure 22).

    However, the cyclic boronate acted as a protective group instead to give the 4-O-sulfate, followed

    by hydrolysis of the boronate group during work-up to give 2.15. This showed that the competing

    background sulfation reaction occurs very rapidly. In the presence of the relatively unreactive 4-

    hydroxyl group, the direct sulfation reaction is more favorable than sulfation through the activated

    boronate.

    OHOHO

    OH

    OMe

    2.15 (10 mol%)DIPEA (1.5 equiv)

    Solvent (0.2 M)24 h, rt

    (1.5 equiv)SO

    O NOCl3C N

    OTf

    OHOTCEO3SO OH

    OMe

    OHOHO

    OSO3TCE

    OMe

    +

    2.11 2.12 2.13

    Entry Yield (%)a

    123456b789c10

    3745541541

    tracetrace3520-

    2.12:2.13 RatioSolvent

    MeCNDCMTHFDMFDMADMANMP

    DMA/MeCN (1:1)DMA/MeCN (1:1)DMA/H2O (20:1)

    1.8:11.2:11.2:14.3:14.3:1

    1:11.6:1

    >20:1

    aCrude 1H NMR yields were determined based on tetramethylsilane as an internal standard. Product ratio determined by 1H NMR. bNo catalyst 2.15 used. cReaction performed at 50 oC.

    -

    -

  • 26

    Figure 22. Sulfation with pre-formed cyclic arylboronate of methyl α-L-rhamnopyranoside.

    Revisiting previous work in our group of using electron-deficient boronic acids and Lewis base

    co-catalyst system for the silylation of pyranosides,42 we sought to adapt those conditions for

    regioselective sulfation. However, a brief screen of phosphine oxide and phosphoramide Lewis

    bases with 3,5-bis(trifluoromethyl)phenylboronic acid yielded no regioselectivity, though the 4-

    O-sulfate was not isolated in this case (Table 04).

    Table 04. Boronic acid/Lewis base co-catalyst for sulfation of methyl α-L-rhamnopyranoside.

    42

    Lee, D.; Taylor, M. S. Org. Biomol. Chem. 2013, 11, 5409-5412.

    OHOO

    O

    OMe

    B

    F3C

    OOHO

    OH

    OMe

    SO

    OOTCEEt3N (6 equiv)MeCN (0.2 M)

    24 h, rt;aqueous work-up

    2.1528%

    2.14

    (1.5 equiv)SO

    O NOCl3C N

    OTf

    OHOHO

    OH

    OMe2.08 (1.5 equiv)

    Lewis Base (20 mol%)DIPEA (1.5 equiv)

    MeCN (0.2 M)24 h, rt

    B(OH)2

    F3C

    F3C (20 mol%)

    1234

    1 : 1.21.1 : 1

    1 : 1.11.2 : 1

    35304133

    n-Bu3P=OPh2MeP=O

    Ph3P=OHMPA

    OHOTCEO3SO OH

    OMe

    OHOHO

    OSO3TCE

    OMe

    +

    2.12 2.132.11

    Entry Yield (%)a 2.12:2.13 RatioLewis Base

    aCrude 1H NMR yields were determined based on tetramethylsilane as an internal standard. Product ratio determined by 1H NMR.

  • 27

    To evaluate the effects of the sulfate ester substituent on reactivity and selectivity, we attempted

    to synthesize a series of alkyl- and arylsulfuryl 1,2-dimethylimidazolium triflate (Figure 23). The

    reaction of sulfuryl chloride with phenol and para-methoxyphenol gave 2.14 and 2.15,

    respectively, in good yields. However, the same reaction with benzyl alcohol, para-methoxybenzyl

    alcohol, or sec-butyl alcohol gave products in trace yields or decomposed upon work-up or storage

    under air at room temperature overnight. Previously attempts to prepare benzyl chlorosulfates were

    reported to lead to the formation of polybenzyl,43 and thus benzyl and alkyl chlorosulfates were

    not pursued further. Reaction of 2.14 and 2.15 with 2-methylimidazole gave 2.16 and 2.17,

    respectively, and methylation gave the triflate salts 2.18 and 2.19, respectively, in good overall

    yields.

    Figure 23. Preparation of alkyl- and arylsulfuryl 1,2-dimethylimidazolium triflate.

    Sulfation with 2.18 and 2.19 gave markedly better regioselectivity than with TCE-sulfuryl 1,2-

    dimethylimidazolium triflate (Table 05). Amide solvents such as DMF and DMA were, again,

    particularly effective (Entries 4-7 and 10) and the presence of catalyst was crucial (Entries 1, 3, 8).

    The optimal conditions achieved gave 52% yield with 13:1 selectivity of 3-O-sulfate:2-O-sulfate

    (Entry 6). However, this yield could not be improved further.

    43

    Gibbons, R. A.; Gibbons, M. N.; Wolfrom, M. L. J. Am. Chem. Soc. 1955, 77, 6374-6374.

    SO

    Cl ClO

    pyridine, Et2OSO

    O ClO(1 equiv)

    2.14 R = Ph2.15 R = p-OMePh R = Bn R = PMB R = sec-Bu

    R-OHR

    54%94%

    decompositiondecomposition

    trace

    NHN

    SO

    OO N N S

    O

    OO N N

    OTf(3.6 equiv)

    R R

    2.16 R = Ph2.17 R = p-OMePh

    2.18 R = Ph2.19 R = p-OMePh

    82%31%

    73%82%

    MeOTf(1 equiv)

    THF Et2O

  • 28

    Table 05. Optimization of methyl α-L-rhamnopyranoside sulfation with arylsulfuryl 1,2-

    dimethylimidazolium triflate.

    A closer examination of the sulfating reagents revealed that 2.08, 2.18, and 2.19 underwent

    appreciable base-mediated hydrolysis in deuterated acetonitrile, as monitored over time by 1H

    NMR spectroscopy (Figure 24). The triflate salts, in the absence of base, were stable in solution

    (50%

    decomposition after 24 h. 2.18 with 1 equiv. of DIPEA (Figure 24, C) showed >60%

    decomposition, while only 50% decomposition with 1 equiv. of DIPEA dried overnight with 3

    molecular sieves (Figure 24, D). The greater disparity after 24 h was seen with 2.19 in the presence

    of DIPEA (Figure 24, E, 52% decomposition) and 3 MS-dried DIPEA (Figure 24, F, 28%

    decomposition). This base-mediated hydrolysis of the sulfuryl imidazolium triflates explains our

    (1.5 equiv)

    Catalyst (10 mol%)DIPEA (1.5 equiv)

    Solvent (0.2 M)24 h, rt

    OHOHO

    OH

    OMeSO

    O NO N

    OTf

    R = H, 2.18123456b7R = p-OMe, 2.198910

    no catalyst2.15

    no catalyst2.152.152.152.15

    no catalyst2.15 2.15

    MeCNMeCNDMFDMFDMA

    DMA/MeCN (1:1)DMA/MeCN (1:1)

    MeCNMeCNDMA

    2:18:1

    >20:1>20:1

    13:1>20:1

    2:18:1

    >20:1

    29490

    30365248

    344820

    R

    aCrude 1H NMR yields were determined based on tetramethylsilane as an internal standard. Product ratio determined by 1H NMR. b3 equiv of 2.18 and DIPEA used instead.

    Entry Yield (%)a 2.12:2.13 RatioCatalyst Solvent

    OHOTCEO3SO OH

    OMe

    OHOHO

    OSO3TCE

    OMe

    +

    2.12 2.132.11

    -

  • 29

    inability to increase sulfation yields by heating the reaction or increasing the concentration—both

    changes presumably accelerated the decomposition of sulfating reagents to a greater extent than

    they increased the rate of sulfation. Although rigorously drying the reaction mixture would, in

    theory, prevent hydrolysis and decomposition of the sulfating reagent, we decided to place this

    portion of the project on hold and focus instead on direct, regioselective sulfation without masking

    groups.

    Figure 24. Decomposition study of 2.08, 2.18, 2.19 in CD3CN; A: 2.08 in 1 equiv. 1,2-

    dimethylimidazole; B: 2.08 in 1 equiv. DIPEA; C: 2.18 in 1 equiv. DIPEA; D: 2.18 in 1 equiv. 3

    MS-dried DIPEA; E: 2.19 in 1 equiv. DIPEA; F: 2.19 in 1 equiv. 3 MS-dried DIPEA.

    2.3 Sulfation of carbohydrates with sulfur trioxide amine complexes We began studying direct sulfation using methyl α-D-mannopyranoside 2.20 as a model substrate

    (Table 06). In the absence of catalyst (Entry 1), the reaction mixture included unreacted starting

    material and a distribution of O-sulfate and bis-sulfates. A brief catalyst screen (Entries 2-4)

    revealed that the reaction is highly influenced by catalyst, with the product being predominantly

    the 3-O-sulfate 2.21. This is particularly promising given the presence of a more reactive primary

    hydroxyl group in the substrate. Replacing SO3-pyridine with SO3-Me3N (Entry 5) diminished

    conversion but afforded a cleaner reaction. Increasing the temperature to 60 oC (Entry 6) and the

  • 30

    equivalence of sulfating reagent (Entry 7) gave increased yield while maintaining the

    regioselectivity of sulfation. Further increasing the amount of sulfating reagent (Entry 8) was not

    beneficial and afforded more over-sulfated products than the desired mono-3-O-sulfate (3-

    OSO3:over-sulfation 3:1).

    Table 06. Optimization of methyl α-D-mannopyranoside sulfation with sulfur trioxide amine

    complex.

    Further optimization was performed on n-octyl α-D-galactopyranoside to assess the efficiency of

    sulfation on a more soluble substrate with the standard conditions set out previously (Table 06,

    Entry 7). As discussed by Gilbert in 1962,11 the reactivities of sulfur trioxide amine complexes

    vary inversely with the strength of the Lewis base component. A survey of conventional sulfur

    trioxide amine complexes in our reaction, however, afforded the opposite trend (Table 07). SO3-

    Et3N (Entry 2) gave higher conversion but the regioselectivity of sulfation suffered drastically (3-

    OSO3:over-sulfation 0.6:1). Weaker Lewis base complexes (Entry 3, 4) gave low or no conversion.

    Sulfating Reagent (XX equiv)

    Catalyst (10 mol%)DIPEA

    MeCN (0.2 M)3 h, Temperature

    OHOHO

    OH

    OMe

    HO

    OHO-O3SO

    OH

    OMe

    HO

    aCrude 1H NMR yields were determined based on 1,3,5-trimethoxybenzene as an internal standard.

    Entry Yield (%)aCatalystSulfating Reagent (XX equiv) Temperature (oC)

    12345678

    SO3-pyridine (1.5 equiv)SO3-pyridine (1.5 equiv)SO3-pyridine (1.5 equiv)SO3-pyridine (1.5 equiv)SO3-Me3N (1.5 equiv)SO3-Me3N (1.5 equiv)SO3-Me3N (3 equiv)

    SO3-Me3N (4.5 equiv)

    no catalyst2.142.152.162.152.152.152.15

    4040404040606060

    2831616627498067

    2.14

    O BBPh

    Ph

    Ph

    Ph

    2.15

    B

    O

    OH2.22

    B

    S

    OH

    2.20 2.21

    Conversion (%)a

    7948827927499089

    Equiv. of DIPEA

    1.51.51.51.51.51.53

    4.5

  • 31

    To test the hypothesis that a highly Lewis basic additive could accelerate the sulfation,

    quinuclidine was added to the reaction with either SO3-Me3N (Entry 5) or SO3-DMF (Entry 6);

    however, both changes were not fruitful.

    Table 07. Optimization of n-octyl α-D-galactopyranoside sulfation with sulfur trioxide amine

    complex.

    With the optimized conditions in hand, we began to explore the substrate scope (Table 08). The

    sulfated products as a trialkylamine salt were exchanged with DOWEX Na+-resin to give the final

    product as a sodium salt. Substrates without primary hydroxyl groups gave sulfated products in

    high yields (2.25, 2.26). Sulfation could also be performed in good yields in the presence of

    primary hydroxyl groups for a wide range of manno- and galacto-derived pyranosides. No

    significant difference in reactivity between α- and β-anomers of methyl-D-galactopyranoside was

    observed (2.29 and 2.30). A known sulfated glycolipid 2.34 found in the mammalian nervous

    system was also synthesized in good yield.

    SO3-Me3N (3 equiv)

    2.15 (10 mol%)DIPEA (3 equiv)MeCN (0.2 M)

    3 h, 60 oC

    OHO

    O-Octyl

    OHOH

    OH

    O-O3SO

    O-Octyl

    OHOH

    OH

    Entry Yield (%)aChanges to Above Conditions

    123456

    -SO3-Et3N (3 equiv) instead of SO3-Me3N

    SO3-pyridine (3 equiv) instead of SO3-Me3NSO3-DMF (3 equiv) instead of SO3-Me3N

    quinuclidine (3 equiv) as additiveSO3-DMF (3 equiv) instead of SO3-Me3N, quinuclidine (3 equiv) as additive

    603617010

    aCrude 1H NMR yields were determined based on 1,3,5-trimethoxybenzene as an internal standard.

    2.23 2.24

    Conversion (%)a

    79>9551010

  • 32

    Table 08. Substrate scope of carbohydrate sulfation with SO3-Me3N.

    2.4 Summary and future work In summary, a catalytic, regioselective sulfation method was outlined that tolerated a range of fully

    unprotected carbohydrates. Our approach is similar to the one described by Flitsch in 199419,20 in

    that both strategies utilize cis-diol activation to selectively functionalize at the equatorial hydroxyl

    group. However, our method uses benign, organoboron compound in catalytic quantities as

    opposed to toxic, organotin reagent in stoichiometric amounts. Furthermore, our one-step

    O

    OR(HO)n

    SO3-Me3N (3 equiv)

    2.15 (10 mol%)DIPEA (3 equiv)MeCN (0.2 M)

    3 h, 60 oC

    O

    OR(HO)n

    Na+ -O3SO

    OOH

    OH

    OMe

    OSO3- +Na

    O

    OH

    HO

    OMe

    Na+ -O3SO

    OHONa+ -O3SO

    OH

    OMe

    HO

    OHONa+ -O3SO

    OH

    SPh

    HO

    ONa+ -O3SO

    OMe

    OHOH

    OH

    ONa+ -O3SO OMe

    OHOH

    OH

    O

    Na+ -O3SO SiPr

    OHOH

    OH

    ONa+ -O3SO

    O

    OHOH

    OH

    O

    Na+ -O3SO

    OHOH

    AcHNOMe

    OO

    Na+ -O3SO

    OHOH

    OHC13H27

    OH

    HN C15H31

    O

    >99%2.25

    72%2.26

    65%2.27

    54%2.28

    53%2.29

    57%2.30

    57%2.31

    54%2.32

    46%2.33

    42%a2.34

    Reactions were performed at 0.1 mmol scale of substrate. Isolated yields are reported. aReaction was performed at 0.005 mmol scale of substrate with 6 equiv. SO3-Me3N, 40 mol% 2.15, 6 equiv. DIPEA, and 0.5 M MeCN.

  • 33

    sulfation/activation is achieved in drastically shorter reaction time compared to those reported

    using the two-step process with dibutyltin oxide. There is still work to be done to understand the

    mechanism of sulfation and the reactivity of the sulfur trioxide amine complexes, particularly why

    the trend observed with various Lewis bases of the complex (Table 07) is opposite to that

    previously reported.11

    2.5 Experimental

    2.5.1. General Information

    All reactions were performed using a Teflon-coated magnetic stir bar under argon. All solvents

    used were dried using the Pure Solv-MD solvent purification system (Innovative Technology) or

    previously dried overnight with 3 molecular sieves. All reagents and carbohydrates, unless

    otherwise stated, were purchased from Sigma-Aldrich or Carbosynth Ltd. Flash column

    chromatography was performed using silica gel (60 , 230-400 mesh) (SiliCycle). Analytical

    thin-plate chromatography was performed on aluminum-backed silica gel 60 F254 plates (EMD

    Milipore) and visualized under UV or with aqueous basic permanganate stain.

    1H, 13C and 2D nuclear magnetic resonance (NMR) spectra were acquired on the Agilent DD2 600

    MHz or Agilent DD2 500 MHz, both equipped with a OneNMR probe. Chemical shifts (δ) are

    reported in parts per million (ppm), calibrated to the residual protium in the deuterated solvent.

    Spectral features are tabulated as follows: chemical shift (δ, ppm); multiplicity (app = apparent, s

    = singlet, d = doublet, t = triplet, q = quartet, p = pentet, hept = heptet, m = multiplet, where the

    range of chemical shift is given); number of protons, coupling constants (J, Hz); assignment.

    Infrared (IR) spectra were acquired on the Fourier-transform Spectrum 100 spectrometer

    (PerkinElmer) equipped with a single-bounce diamond/ZnSe ATR accessory. Spectral features are

    tabulated as follows: wavenumber (cm-1); intensity (s = strong, m = medium, w = weak, br =

    broad). High-resolution mass spectra (HRMS) were acquired on the Agilent 6538 UHD Q-TOF

    for electrospray ionization, negative mode (ESI−). Optical rotations were acquired on the

    AUTOPOL IV (Rudolph Research Analytical) in a 0.6 dm polarimeter sample cell, at 589 nm

    wavelength, at 20 oC, and the sample concentrations are reported in g per 100 mL in methanol.

    Melting points were acquired on the Mel-Temp II (Laboratory Devices Inc.), and reported as a

  • 34

    range of melting or decomposing (denoted by dec.).

    Presence of a single sulfate group is confirmed by HRMS. Assignment of sulfation position is

    based on change in NMR chemical shift between starting carbohydrate and sulfated product (1H:

    approx. +0.8 ppm, 13C: approx. +7 ppm).

    2.5.2. General Procedure A

    To a 2-dram vial equipped with a magnetic stir bar was added the carbohydrate (0.1 mmol, 1

    equiv), sulfur trioxide trimethylamine complex (42 mg, 0.3 mmol, 3 equiv), and 2.15 (2 mg, 0.01

    mmol, 0.1 equiv). The reaction vial was capped with a septum and purged with argon. Acetonitrile

    (0.5 mL, 0.2 M) was added to the vial, followed by N,N-diisopropylethylamine (0.06 mL, 0.3

    mmol, 3 equiv). The septum was quickly replaced with a screw cap, sealed with Teflon tape, and

    the reaction was stirred at 60 oC for 3 hours. The mixture was then quenched with MeOH and the

    solvent was removed by rotary evaporation. The crude mixture was purified by flash

    chromatography on silica gel (2% to 15% MeOH in DCM). Fractions containing the product were

    combined and stirred with Dowex 50WX2 Na+-form (50–100 mesh) for 30 min. The resulting

    mixture was dried, filtered through Celite with DCM, then MeOH. The MeOH fraction was

    collected and dried to give the product as a solid.

    2.5.3. Preparation of catalyst and carbohydrate substrates

    10H-Dibenzo[b,e][1,4]oxaborinin-10-ol (2.15):

    B

    O

    OH

    O 1) n-BuLi, diphenyl ether, THF

    2) tributyl borate3) 4N HCl

  • 35

    10H-Dibenzo[b,e][1,4]oxaborinin-10-ol was prepared according to literature procedure44 from

    diphenyl ether and tributyl borate (Sigma-Aldrich). Spectral features are in agreement with those

    previously reported.

    Methyl-2-acetamido-2-deoxy-α-D-galactopyranoside:

    Methyl-2-acetamido-2-deoxy-α-D-galactopyranoside was prepared according to an adapted

    literature procedure45 from N-acetyl-D-galactosamine (Sigma-Aldrich). The ⍺-anomer was separated cleanly by flash chromatography on silica gel to give the desired product. Spectral

    features are in agreement with those previously reported.46

    β-D-galactosyl N-palmitoyl-D-erythro-sphingosine:

    44

    Dimitrijević, E.; Taylor, M. S. Chem. Sci. 2013, 4, 3298-3303. 45

    Liang, H.; Grindley, T. B. J. Carbohydr. Chem. 2004, 23, 71-82. 46

    Grönberg, G.; Nilsson, U.; Bock, K.; Magnusson, G. Carbohydr. Res. 1994, 257, 35-54.

    OHO

    OHOH

    AcHNOMe

    OHO

    OHOH

    AcHN OH

    Amberlyst H+ resin

    MeOH, reflux

    OPMBO

    OPMBPMBO

    PMBO OH

    1) Ms2O, PMP, CH2Cl2

    2)

    HOHN

    OH

    C13H27C15H31

    O

    (Ph2B)2O, CH2Cl2

    OO

    PMBO

    OPMBPMBO

    PMBOC13H27

    OH

    HN C15H31

    O

    3) CF3COOH, anisole, CH2Cl2

  • 36

    β-D-galactosyl N-palmitoyl-D-erythro-sphingosine was prepared according to a literature

    procedure47 from 2,3,4,6-tetra-O-4-methoxybenzyl-D-galactopyranose and N-palmitoyl-D-

    sphingosine (Sigma-Aldrich). Spectral features are in agreement with those previously reported.

    2.5.4. Synthesis and characterization of compounds

    2,2,2-Trichloroethoxysulfuryl chloride (2.02):

    2.02 was prepared according to a literature procedure.31

    Spectral features are in agreement with those previously

    reported.

    Methyl 6-O-tert-butyldimethylsilyl-α-D-mannopyranoside 2,3-cyclic sulfate (2.05):

    1H NMR (399 MHz, CDCl3) δ 5.00 (app s, 1H, H-1), 4.94

    (dd, J = 5.3, 0.9 Hz, 1H, H-2), 4.89 (dd, J = 7.8, 5.3 Hz, 1H,

    H-3), 4.25 (dd, J = 9.8, 7.8 Hz, 1H, H-4), 3.96 (dd, J = 10.6,

    4.9 Hz, 1H, H-6a), 3.88 (dd, J = 10.6, 5.7 Hz, 1H, H-6b), 3.65

    (dddd, J = 9.8, 5.6, 4.9, 0.6 Hz, 1H, H-5), 3.42 (s, 3H, 1-

    OCH3), 0.91 (s, 9H, TBS), 0.12 (d, J = 3.2 Hz, 6H, TBS); 13C

    NMR (100 MHz, CDCl3) δ 95.4 (C-1), 85.3 (C-3), 78.9 (C-

    2), 69.1 (C-4), 67.8 (C-5), 64.0 (C-6), 55.3 (1-OCH3), 25.7 (TBS), 18.2 (TBS), -5.5 (TBS), -5.6

    (TBS).

    47

    D’Angelo, K. A.; Taylor, M. S. Chem. Commun. 2017, 53, 5978-5980.

    OHOO

    O

    OMe

    OTBSSOO

    Chemical Formula: C13H26O8SSiMolecular Weight: 370.4880

    SO

    O ClOCl3C

    Chemical Formula: C2H2Cl4O3SMolecular Weight: 247.8950

  • 37

    Methyl 4,6-bis-O-(tert-butyldimethylsilyl)-α-D-mannopyranoside 2,3-cyclic sulfate (2.06):

    1H NMR (400 MHz, CDCl3) δ 4.99 (d, J = 0.8 Hz, 1H, H-1),

    4.94 (dd, J = 5.2, 0.9 Hz, 1H, H-2), 4.80 (dd, J = 7.7, 5.3 Hz,

    1H, H-3), 4.19 (dd, J = 9.9, 7.8 Hz, 1H, H-4), 3.88 (dd, J =

    11.5, 2.0 Hz, 1H, H-6a), 3.82 (dd, J = 11.5, 4.8 Hz, 1H, H-

    6b), 3.55 (dddd, J = 9.9, 4.8, 2.1, 0.6 Hz, 1H, H-5), 3.40 (s,

    3H), 0.90 (s, 9H), 0.89 (s, 9H), 0.15 (d, J = 11.1 Hz, 6H), 0.08

    (d, J = 3.3 Hz, 6H); 13C NMR (100 MHz, CDCl3) δ 95.3 (C-

    1), 87.0 (C-3), 79.7 (C-2), 70.6 (C-5), 67.2 (C-4), 61.7 (C-6), 54.9 (1-OCH3), 25.8 (TBS), 25.7

    (TBS), 18.3 (TBS), 18.0 (TBS), -4.7 (TBS), -5.2 (TBS), -5.3 (TBS), -5.4 (TBS).

    1-(2,2,2-Trichloroethoxysulfuryl)-2-methyl imidazole (2.07):

    2.07 was prepared according to a literature procedure.31

    Spectral features are in agreement with those previously

    reported.

    1-(2,2,2-Trichloroethoxysulfuryl)-2,3-dimethylimidazolium triflate (2.08):

    2.08 was prepared according to a literature procedure.31

    Spectral features are in agreement with those

    previously reported.

    SO

    OO N N

    OTfCl3C

    Chemical Formula: C8F3H10Cl3N2O6S2Molecular Weight: 308.5775

    SO

    OO N N

    Cl3C

    Chemical Formula: C6H7Cl3N2O3SMolecular Weight: 293.5430

    OTBSOO

    O

    OMe

    OTBSSOO

    Chemical Formula: C19H40O8SSi2Molecular Weight: 484.7510

  • 38

    Methyl 3-trichloroethylsulfo-α-L-rhamnopyranoside (2.12):

    1H NMR (500 MHz, CDCl3) δ 4.85 (q, J = 10.8 Hz, 2H,

    CH2CCl3), 4.85 (dd, J = 9.2, 3.1 Hz, 1H, H-3), 4.71 (d, J =

    2.0 Hz, 1H, H-1), 4.32 (dd, J = 3.2, 2.0 Hz, 1H, H-2), 3.78

    (app t, J = 9.3 Hz, 1H, H-4), 3.75–3.68 (m, 1H, H-5), 3.39 (s,

    3H, 1-OCH3), 1.38 (d, J = 6.1 Hz, 3H, 5-CH3); 13C NMR (100

    MHz, CDCl3) δ 100.6 (C-1), 86.8 (C-3), 80.0 (CH2), 70.7 (C-

    4), 69.3 (C-2), 68.3 (C-5), 55.3 (1-OCH3), 17.7 (5-CH3).

    Methyl 2-trichloroethylsulfo-α-L-rhamnopyranoside (2.13):

    1H NMR (399 MHz, CDCl3) δ 4.97 (d, J = 10.7 Hz, 1H,

    CH2CCl3), 4.93 (d, J = 1.7 Hz, 1H, H-1), 4.89–4.82 (m, 1H,

    H-2), 4.78 (s, 1H, CH2CCl3), 4.01 (dd, J = 9.5, 3.3 Hz, 1H,

    H-3), 3.71–3.61 (m, 1H, H-5), 3.48 (app t, J = 9.5 Hz, 1H,

    H-4), 3.40 (s, 3H, 1-OCH3), 1.34 (d, J = 6.2 Hz, 3H, 5-CH3); 13C NMR (101 MHz, CDCl3) δ 97.7, 79.8, 77.9, 72.9, 69.4,

    68.0, 55.3, 17.4.

    Benzenesulfonyl chloride (2.14):

    2.14 was prepared according to a literature procedure.48

    Spectral features are in agreement with those previously

    reported.

    48 DeBergh, J. R.; Niljianskul, N.; Buchwald, S. L. J. Am. Chem. Soc. 2013, 135, 10638-10641.

    OHOO

    OH

    OMe

    SO

    OO

    Cl3C

    Chemical Formula: C9H15Cl3O8SMolecular Weight: 389.6210

    OHOHO

    O

    OMe

    SO

    O

    OCl3C

    Chemical Formula: C9H15Cl3O8SMolecular Weight: 389.6210

    SO

    OO

    Cl

    Chemical Formula: C6H5ClO3SMolecular Weight: 192.6130

  • 39

    4-Methoxybenzenesulfonyl chloride (2.15):

    2.15 was prepared according to a literature procedure.48

    Spectral features are in agreement with those previously

    reported.

    1H NMR (400 MHz, CDCl3) δ 7.25 (d, J = 9.2 Hz, 2H, Ph),

    6.89 (d, J = 9.2 Hz, 2H, Ph), 3.77 (s, 3H, CH3); 13C NMR (101

    MHz, CDCl3) δ 159.5, 143.5, 122.8, 115.1, 55.8.

    2-Methyl-1-(phenylsulfonyl) imidazole (2.16):

    2.16 was prepared according to a literature procedure.49

    Spectral features are in agreement with those previously

    reported.

    1-[(p-Methoxyphenyl)sulfonyl]-2-methyl imidazole (2.17):

    2.17 was prepared according to an adapted literature

    procedure.49 Spectral features are in agreement with those

    previously reported.50

    49 Desoky, A. Y.; Hendel, J.; Ingram, L.; Taylor, S. D. Tetrahedron 2011, 67, 1281-1287. 50

    Reuillon, T.; Bertoli, A.; Griffin, R. J.; Miller, D. C.; Golding, B. T. Org. Biomol. Chem. 2012, 10, 7610-7617.

    SO

    OO

    Cl

    MeO

    Chemical Formula: C7H7ClO4SMolecular Weight: 222.6390

    SO

    O NO N

    Chemical Formula: C10H10N2O3SMolecular Weight: 238.2610

    SO

    O NO N

    MeO

    Chemical Formula: C11H12N2O4SMolecular Weight: 268.2870

  • 40

    1,2-Dimethyl-3-(phenylsulfonyl)-imidazolium triflate (2.18):

    2.18 was prepared according to a literature procedure.49

    Spectral features are in agreement with those previously

    reported.

    1-[(p-Methoxyphenyl)sulfonyl]-2,3-dimethylimidazolium triflate (2.19):

    2.19 was prepared according to an adapted literature

    procedure.49

    1H NMR (400 MHz, CDCl3) δ 7.51 (d, J = 2.4 Hz, 1H,

    Im), 7.33 (d, J = 2.4 Hz, 1H, Im), 7.21 (d, J = 9.2 Hz, 2H,

    Ph), 6.96 (d, J = 9.2 Hz, 2H, Ph), 4.09 (s, 3H, 3-CH3),

    3.80 (s, 3H, OCH3), 2.93 (s, 3H, 2-CH3); 13C NMR (101

    MHz, CDCl3) δ 160.0, 147.9, 142.5, 123.9, 122.5, 120.6, 115.9, 55.8 (OCH3), 37.3 (3-CH3), 11.9

    (2-CH3).

    Methyl α-L-fucopyranoside 3-(sodium sulfate) (2.25):

    2.25 was prepared according to General Procedure A from

    methyl α-L-fucopyranoside (18 mg, 0.1 mmol, 1 equiv) and

    purified by flash chromatography on silica gel (5% to 15%

    MeOH in DCM) to give a beige solid in >95% yield (28 mg).

    SO

    O NO N

    OTf

    MeO

    Chemical Formula: C13F3H15N2O7S2Molecular Weight: 283.3215

    SO

    O NO N

    OTf

    Chemical Formula: C12F3H13N2O6S2Molecular Weight: 253.2955

    OOH

    OH

    OMe

    OSO3- +Na

    Chemical Formula: C7H13NaO8SMolecular Weight: 280.2228

  • 41

    [α]#$% = −20.2° (c 0.1, MeOH) (lit. 51: [α]&'( = −157.4° (c 1.3, H2O)); m.p.: 212–214 (dec.) (lit.51, monohydrate: 161–162 oC); 1H NMR (600 MHz, MeOD-d4) δ 4.68 (d, J = 3.9 Hz, 1H, H-

    1), 4.47 (dd, J = 10.2, 3.2 Hz, 1H, H-3), 4.07 (dd, J = 3.2, 1.2 Hz, 1H, H-4), 3.97–3.92 (m, 2H, H-

    2, H-5), 3.39 (s, 3H, 1-OCH3), 1.23 (d, J = 6.5 Hz, 3H, 5-CH3); 13C NMR (151 MHz, MeOD-d4)

    δ 100.0 (C-1), 78.0 (C-3), 70.3 (C-4), 66.6 (C-2), 65.8 (C-5), 54.1 (1-OCH3), 15.1 (5-CH3); IR

    (thin film, cm-1): 3438 (br, m), 2946 (w), 1650 (br, m), 1216 (s), 1046 (s), 987 (s), 839 (s), 748

    (m); HRMS (ESI−): calcd for C7H13O8S [M − Na]− 257.0337 m/z; found 257.0341 m/z.

    Methyl α-L-rhamnopyranoside 3-(sodium sulfate) (2.26):

    2.26 was prepared according to General Procedure A from

    methyl α-L-rhamnopyranoside (18 mg, 0.1 mmol, 1 equiv)

    and purified by flash chromatography on silica gel (5% to

    15% MeOH in DCM) to give a white solid in 79% yield (22

    mg).

    [α]#$% = −94.3° (c 0.27, MeOH); m.p.: 201–202 (dec.); 1H NMR (600 MHz, MeOD-d4) δ 4.58 (d, J = 1.8 Hz, 1H, H-1), 4.42 (dd, J = 9.5, 3.3 Hz, 1H, H-3), 4.16 (dd, J = 3.3, 1.8 Hz, 1H, H-2),

    3.65–3.60 (m, 1H, H-5), 3.55 (app t, J = 9.5 Hz, 1H, H-4), 3.36 (s, 3H,1-OCH3