phase‑transfer and ion‑pairing catalysis of pentanidiums ... · bisguanidinium formed an ion...

36
This document is downloaded from DR‑NTU (https://dr.ntu.edu.sg) Nanyang Technological University, Singapore. Phase‑Transfer and Ion‑Pairing Catalysis of Pentanidiums and Bisguanidiniums Zong, Lili; Tan, Choon‑Hong 2017 Zong, L., & Tan, C.‑H. (2017). Phase‑Transfer and Ion‑Pairing Catalysis of Pentanidiums and Bisguanidiniums. Accounts of Chemical Research, 50(4), 842‑856. https://hdl.handle.net/10356/87360 https://doi.org/10.1021/acs.accounts.6b00604 © 2017 American Chemical Society. This is the author created version of a work that has been peer reviewed and accepted for publication by Accounts of Chemical Research, American Chemical Society. It incorporates referee’s comments but changes resulting from the publishing process, such as copyediting, structural formatting, may not be reflected in this document. The published version is available at: [http://dx.doi.org/10.1021/acs.accounts.6b00604]. Downloaded on 08 Jul 2021 05:08:56 SGT

Upload: others

Post on 17-Feb-2021

7 views

Category:

Documents


0 download

TRANSCRIPT

  • This document is downloaded from DR‑NTU (https://dr.ntu.edu.sg)Nanyang Technological University, Singapore.

    Phase‑Transfer and Ion‑Pairing Catalysis ofPentanidiums and Bisguanidiniums

    Zong, Lili; Tan, Choon‑Hong

    2017

    Zong, L., & Tan, C.‑H. (2017). Phase‑Transfer and Ion‑Pairing Catalysis of Pentanidiums andBisguanidiniums. Accounts of Chemical Research, 50(4), 842‑856.

    https://hdl.handle.net/10356/87360

    https://doi.org/10.1021/acs.accounts.6b00604

    © 2017 American Chemical Society. This is the author created version of a work that hasbeen peer reviewed and accepted for publication by Accounts of Chemical Research,American Chemical Society. It incorporates referee’s comments but changes resultingfrom the publishing process, such as copyediting, structural formatting, may not bereflected in this document. The published version is available at:[http://dx.doi.org/10.1021/acs.accounts.6b00604].

    Downloaded on 08 Jul 2021 05:08:56 SGT

  • 1

    Phase transfer and ion pairing catalysis of

    pentanidiums and bisguanidiniums

    Lili Zong and Choon-Hong Tan*

    Division of Chemistry and Biological Chemistry, School of Physical and Mathematical Sciences,

    Nanyang Technological University, 21 Nanyang Link, Singapore 637371

    CONSPECTUS: Catalysts accelerate biological processes and organic reactions in a controlled

    and selective fashion. There are continuing efforts in asymmetric catalysis to develop efficient

    catalysts with broad reaction scope and industrial practicability. Amongst the various modes of

    asymmetric catalysis, phase transfer catalysis has attracted intense interest due to its ease to scale

    up and low catalyst loading. Chiral quaternary ammonium and phosphonium salts are well-

    studied classes of chiral phase transfer catalysts and they typically are comprised of sp3-

    hybridized quaternary onium salts. In this Account, we describe our recent attempts to develop

    N-sp2 hybridized guanidinium-type salts as efficient phase transfer catalysts as well as ion pair

    catalysis based on N-sp2 hybridized bisguanidinium-type salts.

    The sp2-quaternized ammonium salts, pentanidiums, which contain five nitrogen atoms in

    conjugation, displayed remarkable phase transfer catalytic efficiency. We have shown that

    pentanidium can catalyze Michael additions of tert-butyl glycinate-benzophenone Schiff base

    with various ,-unsaturated acceptors, such as vinyl ketones, acrylates and chalcones in high

  • 2

    enantioselectivities. The structurally amendable pentanidium phase transfer catalysts supply

    diverse reactivity and selectivity to various other organic transformations, such as α-

    hydroxylation of 3-substituted-2-oxindoles, Michael addition of 3-alkyloxindoles with vinyl

    sulfone and alkylation reactions of sulfenate anions and dihydrocoumarins. Pentanidium salts are

    applicable to enantioselective transformations on a preparative scale at low catalyst loading,

    allowing for the synthesis of a broad range of enantiopure compounds. From computational and

    experimental results, we also proposed that halogenated pentanidium catalysts participated in

    halogen bonding and this contributed to the excellent stereocontrol in alkylation reactions.

    Subsequently, we articulated that chiral cations can direct functional anions besides basic anions

    in traditional Brønsted basic phase transfer reactions, including metal-centered anions. We

    identified dicationic bisguanidinium as an excellent ion pairing catalyst, first demonstrating that

    bisguanidinium formed an ion pair with permanganate and directed the anion in enantioselective

    dihydroxylation and oxohydroxylation of a,β-unsaturated esters. This initial success led us to

    explore chiral cationic ion pairing catalysis as a general mode of catalysis. This mode of catalysis

    is at the interphase between organocatalysis, phase transfer catalysis and organometallic

    catalysis. We then identified bisguanidinium diphosphatobisperoxotungstate and bisguanidinium

    dinuclear oxodiperoxomolybdosulfate ion pairs as the active catalysts in enantioselective

    sulfoxidations using aqueous H2O2 as oxidant. The structure of bisguanidinium dinuclear

    oxodiperoxomolybdosulfate ion pair was elucidated using single crystal X-ray analysis.

    Bisguanidinium-catalyzed sulfoxidations emerged as a practical methodology for the synthesis of

    enantioenriched sulfoxides including armodafinil and lansoprazole, which are commercial drugs.

    Finally, we are also able to show that pentanidium and bisguanidinium hypervalent silicates are

    intermediates in enantioselective alkylations using silylamide as Brønsted probase.

  • 3

    1. INTRODUCTION

    Guanidine, the functional group on the side chain of arginine, is a strong base and is protonated

    over a wide pH range. Thus, in physiological conditions, it usually exists as guanidinium. The

    guanidinium motif is known to play crucial roles in substrate recognition at the active site of

    enzymes through non-covalent interactions such as hydrogen bonding.1 Interactions between

    guanidinium and functional groups such as phosphates, carboxylates and enolates are also of

    considerable interest in bioorganic chemistry and host-guest macromolecular chemistry.2 Since

    the pioneering works of Chinchilla,3

    Lipton4

    and Corey5

    in the 1990s, numerous exciting

    outcomes have been achieved by utilizing chiral guanidines as general Brønsted base catalysts.6

    Our group has reported a variety of chiral guanidine-catalyzed asymmetric reactions and these

    results have been summarized in several reviews.7 This Account will emphasize our recent

    efforts in using guanidinium-type and bisguanidinium-type salts 1, 2 and 3 as phase transfer8 and

    ion pairing catalysts (Figure 1).

    Figure 1. Guanidinium-type salts as phase transfer and ion pairing catalysts.

    2. PENTANIDIUM-CATALZYED ENANTIOSELECTVE PHASE

    TRANSFER REACTIONS

    In our attempt to enhance the basicity of the guanidine catalyst 1, structurally novel pentanidine

    containing five nitrogen atoms in conjugation was designed. The hypothesis was that with

    NN

    N

    N N

    N

    R1

    R2

    R1

    R2

    Ph

    PhPh

    Ph

    Bisguanidinium

    R1, R2 = benzyl

    N

    N

    N

    R1

    Ph

    PhN

    N

    R2

    Ph

    Ph

    Pentanidium

    R1, R2 = alkyl, benzyl

    R1 R2

    ClNH

    N

    NH

    Bicyclic guanidinium

    I ClCl1.HI 2 3

  • 4

    increased conjugation, it might render pentanidine with increased basicity and less acidic

    nucleophiles can be activated. Although the pentanidine did not exhibit enhanced basicity,

    serendipitously, we found that its fully alkylated salt, pentanidium salts 2, is an excellent phase

    transfer catalyst (Figure 1). We prepared the pentanidium salts 2 from commercially available

    chiral diamines and tetra-methylated pentanidium 2a can be obtained in five steps with good

    chemical yield. Generally, pentanidium salts 2 can be synthesized through the coupling between

    imidazolinium chloride and a guanidine partner under basic conditions (Scheme 1). Single-

    crystal X-ray diffraction analysis of pentanidium 2a reveals that the five nitrogen atoms are not

    coplanar and the resulting two five-membered rings are in a twisted spatial arrangement (Figure

    2).9

    Scheme 1. Synthesis of pentanidium salts 2 through imidazolinium chloride.

    2 R1, R2

    2a R1 = R2 = Me

    R1 =

    tBu

    tBu

    R2 =

    X = H, F, Cl, Br, IX

    tBu

    tBu

    R1 =

    tBu

    tBu

    R2 =

    Cl

    2b

    2c-g

    2h-k

    X = Cl, Br, I, OMe

    R1 = R2 =

    tBu

    tBu

    X

    Imidazolinium chloride

    +MeCN reflux

    Et3N

    N

    N

    Cl

    R1

    Ph

    Ph

    R1

    ClHN N

    N

    R2

    Ph

    Ph

    R2

    N

    N

    N

    R1

    Ph

    PhN

    N

    R2

    Ph

    Ph

    Pentanidium

    R1, R2 = alkyl, benzyl

    R1 R2

    Cl 2

    Guanidine

  • 5

    Figure 2. Crystal structure of tetra-methylated pentanidium 2a.

    2.1 Pentanidium-catalyzed Michael addition of tert-butylglycinate Schiff base

    Scheme 2. Pentanidium 2a-catalyzed Michael addition of Schiff base.

    When we investigated the propensity for pentanidium to function as phase transfer catalyst, we

    found that pentanidium 2a is an excellent catalyst for the Michael addition of tert-butylglycinate-

    benzophenone Schiff base 4.9 Pentanidium 2a displayed excellent reactivity; with only 2.0 mol%

    of catalyst, the reaction with methyl vinyl ketone 5a was completed within 3 h, providing the

    Cs2CO3 (5.0 equiv.)

    mesitylene, -20 oC

    NPh2C OtBu

    O

    +

    R2

    O

    R1

    N

    CPh2

    tBuO2C

    HH

    4

    2a (2.0 mol%)

    R1 R2

    O

    Me

    O

    N

    CO2tBu

    Ph2C Et

    O

    N

    CO2tBu

    Ph2C n-Bu

    O

    N

    CO2tBu

    Ph2C

    Ph

    O

    N

    CO2tBu

    Ph2C OEt

    O

    N

    CO2tBu

    Ph2C OBn

    O

    N

    CO2tBu

    Ph2C

    ON

    CPh2

    tBuO2C

    HHO

    N

    CPh2

    tBuO2C

    HH

    5a-l6a-l

    6a, 86% yield, 91% ee 6b, 92% yield, 93% ee 6c, 97% yield, 93% ee

    F

    ON

    CPh2

    tBuO2C

    HH

    MeO

    ON

    CPh2

    tBuO2C

    HHO

    N

    CPh2

    tBuO2C

    HHO

    N

    CPh2

    tBuO2C

    HH

    6d, 50% yield, 88% ee 6e, 71% yield, 97% ee 6f, 80% yield, 96% ee

    6g, 98% yield, 92% ee 6h, 89% yield, 90% ee 6i, 89% yield, 85% ee

    6j, 95% yield, 90% ee 6k, 92% yield, 90% ee

    S

    6l, 71% yield, 87% ee

    O

  • 6

    adduct 6a in 86% yield and in 91% ee (Scheme 2). The generality of the method was

    successfully demonstrated using different Michael addition acceptors such as vinyl ketones,

    acrylates and chalcones. It was found that loading of the catalyst could be further lowered to 0.05

    mol% for a gram-scale reaction (Scheme 3). The Michael addition between Schiff base 4 and

    chalcone 5m proceeded with high stereoselectivity providing the adduct 6m in 90% ee and as a

    single diastereomer; it was transformed to pyrrolidine derivative 7 in simple steps.

    Scheme 3. Preparative-scale reaction with low catalyst loading.

    Figure 3. Pentanidium-catalyzed reactions using prochiral intermediates.

    The aforementioned result indicates that chiral pentanidium cation can discriminate efficiently

    between the two enantiotopic faces of prochiral ester enolate. Further investigations reveal that

    amide enolate, sulfenate anion and cyclic ester enolate are also suitable prochiral nucleophiles

    (Figure 3).

    NPh2C OtBu

    O

    +Cs2CO3 (2.5 equiv.)

    mesitylene, -20 oC

    24 h

    2a (0.05 mol%)

    5m, Ar = 4-ClC6H4 6m1.16 g, 87% yield dr> 99:1, 90% ee

    4 2.5 mmol

    NH

    Ph

    Ar

    CO2tBu

    7

    Ph

    O

    Ar Ph

    O

    Ar

    N

    CPh2

    tBuO2C

    HH

    NPh2C OtBu

    O

    NO

    R2

    R1

    ester enolate

    amide enolate

    Hydroxylation using molecular oxygen

    Michael addition to vinyl sulfone

    Michael addition to a,b-unsaturatedcarbonyl compounds

    cyclic ester enolate

    O O

    R1

    Alkylation using alkyl bromides

    S

    S

    O

    Alkylation to form sulfoxides

    sulfenate anion

    a)

    c)

    b)

    d)

  • 7

    2.2 Pentanidium-catalyzed -hydroxylation of 3-substituted-2-oxindoles

    The amide enolate (Figure 3b) generated from 3-substituted 2-oxindole has been widely

    employed as key intermediate for the synthesis of enantioenriched 3,3-disubstituted oxindoles.

    These oxindoles are frequently observed as the core structure in natural products and extensively

    investigated in drug discovery programs due to their potent antibacterial and anticancer

    activities.10

    Catalytic enantioselective -oxidation of 3-substituted-2-oxindole is a direct and

    attractive approach to access 3-substituted-3-hydroxy-2-oxindole core structures.11

    In 2008, Itoh

    reported the preparation of -hydroxyoxindoles using a Cinchonidinium catalyst under phase

    transfer conditions.12

    The product was obtained with moderate enantioselectivity and triethyl

    phosphite was required to reduce the peroxide intermediate. Encouraged by the efficiency and

    excellent stereocontrol of pentanidium, we investigated the -hydroxylation of 3-substituted-2-

    oxindole 7 with air as oxidant (Scheme 4a).13

    Pentanidium 2b, bearing bulkier side arms was

    developed for this reaction (Scheme 1). Initially, we were perplexed that no reductant such as

    triethyl phosphite was required and -hydroxyoxindoles were obtained directly with high

    enantioselectivities. Using isotope labelling experiments with 0.55 equivalent of 18

    O2, the role of

    molecular oxygen was verified (Scheme 4b). Mass spectrometry was employed to characterize

    the isolated hydroxyloxindole [18

    O]-8a and 84% level of 18

    O incorporation was observed.

  • 8

    Scheme 4. a) Pentanidium 2b-catalyzed -hydroxylation of 3-substituted-2-oxindoles, b) isotope

    labelling experiment and c) effect of the amount of oxygen on chemoselectivity.

    We found that the amount of molecular oxygen in the reaction affects the amount of

    hydroperoxide oxindole formed (Scheme 4c). When the reaction was conducted with excess

    oxidant by using an O2 balloon, hydroperoxide oxindole 9a was obtained in 85% yield with good

    enantioselectivity. Hydroperoxide oxindole 9a and hydroxyloxindole 8a were found to have the

    same absolute configuration (R). To gain insights into the reaction, racemic hydroperoxide

    oxindole 9a was prepared and used as oxidant in the -hydroxylation of oxindole 7a in the

    absence of air (Scheme 5a). With 5 mol % of pentanidium 2b and two equivalents of racemic 9a

    as oxidant, the -hydroxylated product (R)-8a was achieved with 76% ee, while the remained

    hydroperoxide oxindole was determined to be (S)-9a with 51% ee. On the basis of the

    experimental results (Scheme 4c and Scheme 5a), we proposed a two-step reaction that includes

    a kinetic resolution step (Scheme 5b). In the first step, in the presence of pentanidium 2b and

    base, the enolate generated from 3-substituted-2-oxindole 7a, adds to O2 in an enantioselective

    fashion to give hydroperoxide oxindole (R)-9a. This is followed by a kinetic resolution step, in

    NO

    PMB

    2b (5 mol%)air (0.5 equiv. O2)

    50% aq. KOH

    toluene, -60 oC12-96 h

    No reductant

    R2

    RN

    O

    PMB

    R2

    R

    HO

    NO

    PMB

    MeHO

    80% yield, 95% ee

    NO

    PMB

    MeHO

    75% yield, 86% ee

    NO

    PMB

    MeHO

    80% yield, 98% ee

    F MeO

    NO

    PMB

    n-BuHO

    72% yield, 85% ee

    NO

    PMB

    HO

    78% yield, 91% ee

    NO

    PMB

    HO

    78% yield, 94% ee

    NO

    PMB

    2b (5.0 mol%),

    0.55 equiv. O2

    conditionsMe N

    O

    PMB

    MeHO

    2b (1.0 mol%),

    excess O2

    conditionsN

    O

    PMB

    MeHOO

    7a

    8a80% yield, 95% ee

    a) b)

    c)

    9a85% yield, 80% ee

    NO

    PMB

    2b (5 mol%),

    0.55 equiv.18O250% aq. KOH

    toluene, -60 oC72 h, 87% yield

    95% ee

    Me

    NO

    PMB

    MeHO

    7a [18O]-8a

    84%18O incorporation

    7 major 8

    8a 8b 8c

    8d 8e 8f

  • 9

    which hydroperoxide oxindole (R)-9a is reduced by a second enolate of 3-substituted-2-oxindole

    7a, furnishing hydroxyloxindole product (R)-8a with improved enantiopurity.

    Scheme 5. a) Racemic hydroperoxide 9a as oxidant and b) proposed reaction pathway involving

    a kinetic resolution process.

    2.3 Pentanidium-catalyzed Michael addition of 3-alkylsubstituted oxindoles

    The direct Michael addition of 3-substituted oxindole to activate alkenes is an attractive

    approach to prepare 3,3-disubstitued oxindoles containing a quaternary carbon stereocenter.

    Recently, successful attempts using commercially available vinyl sulfones as Michael acceptors

    have been reported using organocatalysis.14

    However, only limited examples of highly

    enantioenriched 3,3-dialkyl oxindoles have been obtained.15

    We found that pentanidium 2 are

    efficient phase transfer catalysts for Michael addition of 3-alkyloxindoles 7 to phenyl vinyl

    sulfone 10 at low catalyst loading (Scheme 6).

    NO

    PMB

    NO

    PMB

    MeHOO

    7a1.0 equiv.

    rac-9a2.0 equiv.

    toluene, -60 oC72 h, no O2

    S = 5

    (S,S)-2b (5.0 mol%)50% aq. KOH

    NO

    PMB

    MeHO

    (R)-8a76% ee

    NO

    PMB

    MeHOO

    (S)-9a51% ee

    +

    Pentanidium 2b

    NO

    PMB

    Me2b

    Me

    7a

    base

    +

    NO

    PMB

    MeHOO

    (R)-9a

    NO

    PMB

    Me2b

    (S)-9a + (R)-8a

    kinetic resolution process

    O2

    a)

    b)

  • 10

    Scheme 6. Effect on enantioselectivities using halogenated pentanidiums.

    Scheme 7. Pentanidium 2f-catalyzed Michael addition of 3-alkylsubstituted oxindoles.

    Pentanidium 2c, with four 3,5-di-tert-butylbenzyl side-arms, was first tested and a moderate

    level of enantioselectivity was obtained (Scheme 6).16

    Further modification, by installing

    different halogens to the two benzyl groups, led to an enhancement in enantioselectivity and a

    high level of enantiocontrol was achieved with brominated-pentanidium 2f or iodinated-

    NO

    PMB

    Me

    7a

    SO2Ph+

    10

    2c-g (0.25 mol %)

    K3PO4 (10 equiv.)

    CPME, -60 oCN

    O

    PMB

    Me SO2Ph

    R2 =

    2c X = H 2d X = F2e X = Cl2f X = Br2g X = I

    X

    tBu

    tBu

    N

    N

    N

    R1

    Ph

    PhN

    N

    R2

    Ph

    Ph

    2Pentanidium

    R1 R2

    Cl

    R1 = 3,5-(tBu)2C6H3CH2

    11a

    X ee (%)

    7678808390

    substituent variation

    increasing enantioselectivity

    NO

    PMB

    Me SO2Ph

    NO

    Bn

    Et SO2Ph

    NO

    PMB

    SO2Ph

    NO

    PMB

    SO2PhPh

    99% yield, 95% ee

    NO

    PMB

    SO2Ph

    NO

    PMB

    SO2Ph

    77% yield, 94% ee

    85% yield, 92% ee 78% yield, 98% ee

    N

    Me

    O

    SO2PhEtO

    O

    MeO

    99% yield, 90% ee

    NO

    PG

    R1

    SO2Ph+

    2f (0.25 mol %)

    K3PO4(10 equiv.)

    CPME/m-xylene(1:2)

    -60 oC

    R2

    NO

    PG

    R1

    R2

    SO2Ph

    91% yield, 95% ee

    99% yield, 99% ee

    N

    Me

    MeON

    MeH

    CH3

    12an analogue of esermethole

    11a 11b 11c

    11d 11e 11f

    7 10 11

    11g

  • 11

    pentanidium 2g. With 0.25 mol% of pentanidium 2f, the reaction between a series of 3-

    alkylsubstituted oxindoles 7 and phenyl vinyl sulfone 10 proceeded in a highly enantioselective

    manner to furnish corresponding Michael adducts 11 with high functional group tolerance

    (Scheme 7). Gram quantity of oxindole 11g can be easily obtained from scale-up experiments.

    The adduct 11g then underwent simple transformations to provide pyrroloindoline derivative 12,

    an analogue of the natural product precursor esermethole.17

    2.4 Pentanidium-catalyzed alkylation of sulfenate

    Sulfenate anion18

    is an unstable reaction intermediate but it can be generated in situ; it has a

    nucleophilic sulfur center, which can be used for the construction of S-C bond. This intermediate

    can be used to prepare sulfoxide in a complementary way to the classical Andersen method and

    direct sulfoxidation method. There are a few methods to generate sulfenate anion in situ,

    including the use of base, fluoride or heat (Scheme 8).19

    Perrio and co-workers reported a phase

    transfer alkylation of sulfonate using Cinchonidinium catalyst and it provided chiral sulfoxides

    with moderate enantioselectivity.20

    Using pentanidium 2c as catalyst, we investigated the

    enantioselective benzylation of 2-thienyl sulfenate generated in situ from sulfinyl methyl esters

    13a (Scheme 9).21

    After optimization of reaction parameters, such as sulfenate anion precursors,

    solvent, inorganic base and temperature, the enantioselectivities remained moderate. Several

    pentanidiums with different steric or electronic patterns were prepared and it was found that

    levels of enantioselectivities significantly increases with the introduction of halogen to the

    pentanidiums (2h-2j). On the contrary, pentanidium 2k, containing OMe substituted side arms,

    results in significant deterioration of enantioselectivity.

  • 12

    Scheme 8. Methods for the in situ generation of sulfenate.

    Scheme 9. Development of halogenated pentanidiums for sulfenate alkylation.

    The generality of this reaction was systematically evaluated with a variety of sulfenate

    precursor 13 and alkyl halides using halogenated pentanidiums 2i or 2j. By trapping 2-thienyl

    sulfenate with various alkyl bromides, various sulfoxides 14a-i bearing benzyl, alkyl, allylic,

    propargylic, 2-naphthylmethyl groups were produced in good enantioselectivities. When thienyl

    or benzothienyl sulfenates were utilized, sulfoxides 14j-o were obtained in high yields and

    excellent enantioselectivities (Scheme 10). It is noteworthy that the reaction between sulfenate

    and m-xylene dibromide furnished bis-sulfoxide 14k in good yield (87%, dl/meso = 4.75:1) and

    excellent enantioselectivity. Sulfenates functionalized with furyl, benzofuryl and benzimidazole

    moieties were also smoothly converted to their corresponding heterocyclic sulfoxides 14p-s.

    RS

    O

    EWGEWG+

    a) Base-mediated retro-Michael elimination

    EWG = SO2Ph, CN, COR', CO2R'

    RS

    O

    TMS+

    b) Fluoride-mediated retro-Michael elimination

    RS

    O

    +

    c) Heat-mediated elimination

    RS

    Obase

    fluoride

    heat

    RS

    O

    RS

    O

    TMSF +

    R =

    2c X = H 2h X = Cl2i X = Br2j X = I2k X = OMe

    X

    tBu

    tBu

    N

    N

    N

    R

    Ph

    PhN

    N

    R

    Ph

    Ph

    2Pentanidium

    R R

    Cl

    X ee (%)

    6179889025substituent

    variation

    S

    O

    CO2Me67 wt% aq. CsOH

    CPME, -60 oC

    SS S

    2c, 2h-2k (1.0 mol %)BnBr (1.2 equiv.)

    13a 14a

    O

  • 13

    Remarkably, benzimidazole sulfoxide 14s, reminiscence to the drug esomeprazole,22

    was

    obtained in excellent enantioselectivity.

    Scheme 10. Halogenated pentanidium 2i or 2j-catalyzed alkylation of sulfenate anion.

    Preliminary mechanistic studies on the retro-Michael/alkylation reaction found that the base-

    promoted in situ generated sulfenate anion, could undergo sulfur alkylation, oxygen Michael

    addition or alkaline hydrolysis, depending on reaction conditions (Figure 4a). When a less active

    alkylating reagent, benzyl chloride, was used, only oxa-Michael adduct 15 was obtained even in

    the presence of pentanidium 2c. On the contrary, sulfoxide 14a was produced through sulfur

    alkylation using iodinated-pentanidium 2j. These experimental studies indicated that iodinated-

    RS

    O

    CO2Me

    2i or 2j (1.0 mol %), R'Br or R'I

    saturated aq. CsOH, CPME/Et2O (1:3)

    -70 oC or -40 oC

    RS

    R'

    O

    13 14

    S

    O

    S SMe

    O

    S S

    O

    S S

    O

    S

    S

    O

    S S

    O

    S S

    O

    S S

    O

    S

    S

    O

    S

    CF3

    S

    O

    S S

    O

    S

    Me

    S

    O

    S S

    O

    S

    S

    O

    O

    Me

    Cl

    S

    O

    O

    S

    O

    O

    Cl

    S

    O

    N

    N

    Me

    Me

    S

    O

    S S

    O

    SMe Me

    Me

    Me

    S

    O

    S

    Me

    14a87% yield92% ee

    14b65% yield77% ee

    14c66% yield81% ee

    14d69% yield81% ee

    OMe

    14e82% yield92% ee

    14f77% yield82% ee

    14g83% yield94% ee

    14h84% yield90% ee

    14l85% yield91% ee

    14m93% yield95% ee

    14n89% yield92% ee

    14o83% yield95% ee

    14p95% yield81% ee

    14q73% yield81% ee

    14s99% yield90% ee

    Me

    MeO

    14r83% yield79% ee

    14i84% yield92% ee

    14j85% yield91% ee

    14k87% yield99% ee

    Me

  • 14

    pentanidium 2j activated and stabilized both the electrophilic alkylating reagents and the

    sulfenate.

    Figure 4. a) Diverse reaction pathways of sulfenate anion, b) optimized most R-TS and c)

    proposed mechanistic model.

    Computational studies were performed to obtain the most stable transition state (TS) structure

    for the R- and S-sets of the TS at M06/BS1:UFF calculated with ONIOM method in

    Gaussian09A2. It was found that the R-TS, which leads to the experimental product 14a with R

    configuration, is more stable than the S-TS by 1.2 kcal mol-1

    in terms of the Gibbs free energy

    (Figure 4b). Theoretical study on stereoselectivity is consistent with experimental observations.

    Further analysis of the non-covalent interactions between iodinated-pentanidium 2j and the

    substrates revealed that Br-I halogen bonding23

    between the leaving Br of benzyl bromide and

    iodide on the iodinated-pentanidium 2j, plays a key role to stabilize the TS (Figure 4c).

    S

    O

    S

    sulfenate anion

    BnCl or BnBr

    2j

    Alkaline hydrolysis

    15

    OCO2MeSS

    base

    2c

    CO2MeBnCl,Oxa-Michael addition

    Sulfur alkylation to 14a

    SS

    O

    N

    IX

    d-d+

    X= Br, Cl

    Proposed Mechanistic Model

    Pentanidium 2j

    Halogen bond

    c)

    a) Diverse reaction pathways of sulfenate anion b) Optimized most stable (in terms of G) R-TS

  • 15

    2.5 Pentanidium-catalyzed alkylation of dihydrocoumarins

    Scheme 11. a) Silylamide BSA as a probase, b) effect of halogenated pentanidiums and c)

    identification of silyl ketene acetal 18 as an intermediate.

    We found that halogenated pentanidiums 2h-j exhibited excellent stereocontrol in the

    alkylation of dihydrocoumarins 16, using silylamide as the probase (Scheme 11).24

    In this

    strategy, a base with strong basicity but weak nucleophilicity would be generated in situ and

    consumed immediately for the formation of an enolate (Scheme 11a).25

    The potential

    background and side reactions are thus suppressed when employing this transient strong base.

    The approach will allow the utilization of base-sensitive substrates and reagents, consequently

    expanding the scope of asymmetric phase transfer reactions. For instance, dihydrocoumarins

    16a, which will undergo alkaline hydrolysis in the presence of alkali metal alkoxides or

    R = 2c X = H 2h X = Cl2i X = Br2j X = I

    X

    tBu

    tBu

    N

    N

    N

    R

    Ph

    PhN

    N

    R

    Ph

    Ph

    2Pentanidium

    R R

    Cl

    X ee (%)

    40919793substituent

    variation

    O O

    Ph

    2c, 2h-2j (10 mol %)BnBr (2.0 equiv.)

    CsF (4.0 equiv.)

    BSA (5.0 equiv.)

    THF, -40 oC, 24 hBn

    Ph

    OO

    16a 17a

    Si

    SiN

    O

    Bis(trimethylsilyl)acetamide (BSA)

    chiral cation catalyst

    SiN

    O

    (CH3)3SiF

    [chiral cation]

    probase

    strong basicity weak nucleophilicity

    a)

    b)

    c)O O

    Ph

    2i (5.0 mol %), BnBrCsF (2.0 equiv.)

    BSA (3.0 equiv.)

    THF, -40 oC

    83% yield, 95% eeBn

    Ph

    OO

    16a 17a

    O OTMS

    Ph

    LiHMDSTMSCl

    THF, 0 oC

    2i (5.0 mol %), BnBr

    CsF (3.0 equiv.)

    THF, -40 oC

    55% yield, 91% ee

    F

    18

  • 16

    hydroxides, due to the labile lactone moiety, was smoothly converted to the alkylated product

    17a in high yield and excellent enantioselectivity. Silyl ketene acetal 18 was identified as a key

    intermediate via NMR analysis of the crude reaction mixture. This was further verified by the

    direct benzylation of pre-prepared silyl ketene acetal 18 (Scheme 11c). This reaction is suitable

    for lactones including dihydrocoumarins, 3,4-dihydro-2H-benzo[h]chromen-2-ones and 1,2-

    dihydro-3H-benzo[f]chromen-3-ones (Scheme 12).

    Scheme 12. Halogenated pentanidium 2i-catalyzed enantioselective alkylation of lactones.

    3. BISGUANIDINIUM-CATALZYED ION PAIRING

    ENANTIOSELECTIVE REACTIONS

    In 2013, we found that bicyclic guanidinium salt 1·HI can operate under phase transfer

    conditions for the alkylation of 3-substituted 2-oxoindoles (Scheme 13).26

    Recently, we

    +

    O O

    Bn

    O O

    Bn

    O

    R1

    O O

    R1

    O

    R2 R2

    O O

    Ph

    16b-m 17b-m

    85% yield, 90% ee

    81% yield, 93% ee 78% yield, 94% ee

    O O

    Ph

    84% yield, 90% ee

    17c 17d

    17f 17g

    O O

    Bn

    85% yield, 87% ee

    17b

    O

    OBn

    80% yield, 93% ee

    17k

    Ph

    O O O O

    Bn Bn

    75% yield, 95% ee 95% yield, 90% ee

    17h 17i O O

    Bn

    85% yield, 96% ee

    17jMeO

    O OO

    OEtBn

    87% yield, 86% ee

    17e

    Br R3

    O

    O

    Ph

    Bn

    89% yield, 95% ee

    17lBr

    THF, -40 oC

    R3

    O

    O

    Ph

    Bn

    80% yield, 96% ee

    17m

    MeO

    2i (5.0 mol %)CsF (2.0 equiv.)BSA (3.0 equiv.)

    Ph Ph

  • 17

    developed dicationic bisguanidinium 3, which was initially designed with the intention to work

    with dianions or multiple anions. Bisguanidinium salts27

    (BG) 3 features two guanidinium

    moieties linked with various spacers (Scheme 14). For ease of synthesis, commercially available

    cyclic secondary diamines such as piperazine were examined as spacers between the two chiral

    imidazolinium salts. The dicationic bisguanidinium salts 3 were thus achieved in excellent yields

    and are amenable to modification. We hope that the dicationic moiety will lead us to discover

    new chemistries.

    Scheme 13. Guanidinium 1HI-catalyzed phase-transfer alkylation reaction.

    Scheme 14. General route to bisguanidinium salts 3 with piperazine as spacer.

    In conventional cationic phase transfer catalysis, most reactions are Brønsted base reactions,

    which the functional anion is usually either hydroxide or carbonate. Occasionally, other

    functional anions such as cyanide (CN) and hypochlorite (ClO

    ) are used in phase transfer

    Strecker or oxidation reactions respectively. An innovative example is the use of (hypo)iodite

    NO

    Me

    Ph

    7

    Br CO2Me+

    19

    1.HI (10 mol %)

    K2HPO4 (3.0 equiv.)

    ZnI2 (0.5 equiv.)

    mesitylene, 0 oC, 96 h

    NO

    Me

    Ph CO2MeNH

    N

    NHI

    93% yield94% ee

    20a

    Imidazolinium chloride

    +MeCN reflux

    Et3N

    N

    N

    Cl

    R2

    Ph

    Ph

    R1

    Cl

    NN

    N

    N N

    N

    R1

    R2

    R1

    R2

    Ph

    PhPh

    Ph

    3Bisguanidinium

    ClCl

    HN

    NH

    Spacer

    (S,S)-3a R1, R2 = CH2Ar

    (S,S)-3b R1, R2 = CH2Arp-F

    (S,S)-3c R1, R2 = CH2Aro-F(S,S)-3d R1 = CH2Arp-F R2 = CH2Aro-F

    tBu

    tBu

    tBu

    tBu

    tBu

    tBu

    F F

    Ar Arp-F Aro-F

  • 18

    (IO), generated in situ from iodide, by Ishihara for the phase transfer oxidative

    cycloetherification.28

    However, these reactions are but a subset of a more general strategy of

    chiral cationic ion pairing catalysis, which is defined as reactions using a catalyst compose of an

    organic chiral cation and an inorganic anionic salt (Figure 5).29

    In theory, all functional inorganic

    anions are possible counterparts to the chiral cations including organometallic anions. This mode

    of catalysis is complementary to the anion-directed catalysis using chiral phosphates.30

    Motivated by the desire to exploit more cation-directed anionic inorganic reagents, we began to

    systematically investigate metal-centered anions, beginning with metal oxides.

    Figure 5. Chiral cationic ion pairing catalysis.

    3.1 Bisguanidinium-catalyzed dihydroxylation and oxohydroxylation of -aryl acrylates

    Permanganate (MnO4) oxidation of alkenes under phase transfer condition was first reported

    using stoichiometric amount of a Cinchonidinium salt.31

    Moderate enantioselectivities and low

    yields were reported. The lack of catalytic activity of Cinchonidinium phase transfer catalyst is

    ascribed to its decomposition under the oxidation conditions. In the early stage of our research,

    ChiralCationicIonPairCatalysis

    Inorganicbasesi.e.,hydroxide,carbonate

    Otherinorganicsaltsi.e.,cyanide,iodide

    Metalanionicspeciesi.e.,permanganate,tungstate,molybdate

    RadicalGeneratingSalts

    i.e.,CAN(futurework)

    Organo-

    catalysis

    PhaseTransferCatalysis Organo-

    metallicCatalysis

  • 19

    we conducted simple experiments to determine the stability of dicationic bisguanidiniums in the

    presence of large excess KMnO4; we found that bisguanidiniums 3a-d (Scheme 14) were

    compatible with the oxidation conditions and were recoverable after the reactions. With 2.0

    mol% BG (S,S)-3d and 20 wt% aqueous potassium iodide (KI), the oxidation of the ,-

    unsaturated ester 21a proceeded smoothly to afford diol 22a with excellent enantioselectivity.32

    The yield is moderate as a side product 23a is formed due to C-C cleavage (Scheme 15).

    Generally, substrates bearing electron-rich aryl groups results in higher enantioselectivities than

    electron-deficient ones. We were surprised that thioether group is well tolerated in this

    methodology. We also demonstrated that this methodology is highly selective for ,-

    unsaturated esters over simple alkenes.

    Scheme 15. Bisguanidinium 3d-catalyzed dihydroxylation of -aryl acrylates.

    Trisubstituted enoates were also investigated for the permanganate-mediated alkene oxidation

    (Scheme 16). With 2.0 mol% of BG (S,S)-3c, a mixture of Z,E-trisubstituted enoates 24a

    provided two diastereoisomers, diols 25a and 25a’ in high enantioselectivities with a significant

    Ar COOtBu

    2.0 mol% (S,S)-3d

    KMnO4 (1.5 equiv.)

    20 wt% aq. KI

    TBME, -60 oC or -70 oC

    Ar COOtBu

    OH

    HO

    Ar COOtBu

    O

    21a-k 22a-k 23a-k

    +

    Ph

    Me

    MeO

    Me

    Me

    Et

    O

    O

    MeS

    22k60% yield92% ee

    22a65% yield92% ee

    22c72% yield85% ee

    22e67% yield89% ee

    22f64% yield90% ee

    22j71% yield86% ee

    22i63% yield86% ee

    22h71% yield86% ee

    MeO

    22d65% yield96% ee

    MeO

    OMe

    22j63% yield94% ee

    Me 22b62% yield90% ee

  • 20

    amount of the cleavage product observed. In an attempt to improve the yield of the desired diols,

    various additives were examined. When the pH of the reaction was lowered through the addition

    of acetic acid, 2-hydroxy-3-oxocarboxylic ester 26a was unexpectedly obtained exclusively in

    high yield and high enantioselectivity (Scheme 16). The observed high enantioselectivity

    indicates that both Z- and E-trisubstituted enoates 24a were transformed to the same enantiomer

    26a under these conditions. The reaction pathways to diols and C-C cleavage products were

    efficiently suppressed and we named this transformation, oxohydroxylation. Substrates bearing

    alkyl, thienyl, furyl, alkene moieties are well tolerated and a series of 2-hydroxy-3-oxocarboxylic

    esters 26a-n with high levels of enantioselectivities were achieved. Enantioenriched -

    hydroxymalonates 26o-p were prepared using methoxy-substituted enoates E-24o-p as starting

    materials.

    COOtBu

    O

    OMe

    OH

    COOtBu

    O

    OMe

    OH

    COOtBu

    O

    OMe

    OH

    COOtBu

    O

    OMe

    OH

    COOtBu

    nC5H11 O

    OMe

    OH

    COOtBu

    O

    OMe

    OH

    COOtBu

    O

    OMe

    OHi-Pr

    COOtBu

    nC11H23 O

    OMe

    OH

    COOtBu

    O

    OMe

    OH

    BnO

    COOtBu

    O

    OMe

    OH

    COOtBu

    MeO O

    OH

    COOtBu

    MeO O

    OMe

    OH

    COOtBu

    O

    OMe

    OHMeO

    COOtBu

    O

    OH

    S

    Ar COOtBu

    2.0 mol% (S,S)-3c

    KMnO4 (1.5 equiv.)

    AcOH (3.5 equiv.)

    TBME/H2O (20:1)

    -60 oC

    R

    Ar COOtBu

    OR

    OH

    2426

    R = alkyl (Z, E mixtures)R = OMe (E isomer)

    26a85% yield92% ee

    26b92% yield93% ee

    26c79% yield92% ee

    26d75% yield92% ee

    26e98% yield94% ee

    26f93% yield91% ee

    26g85% yield93% ee

    26h88% yield94% ee

    26i73% yield91% ee

    26j77% yield91% ee

    26k49% yield96% ee

    26l77% yield84% ee

    26o88% yield74% ee

    26p87% yield97% ee

    R = alkyl, OMe

    COOtBu

    O

    OHO

    O26n

    81% yield94% ee

    COOtBu

    O

    OH

    O 26m47% yield87% ee

    Ar COOtBu

    OHR

    OH

    25Not detected

  • 21

    Scheme 16. Bisguanidinium 3c-catalyzed oxohydroxylation of trisubstituted enoates.

    Scheme 17. a) Selectivity between TBA+MnO4

    − and KMnO4, b) cyclic manganate diester 27 as

    an intermediate and c) proposed ion pairing working model.

    A control reaction was attempted using 1.5 equivalent of tetrabutylammonium permanganate,

    an oxidant soluble in organic solvent (Scheme 17a). The reaction provided high level of chiral

    induction, which led us to propose that rate acceleration through ion pairing rather than the phase

    transfer step is crucial for asymmetric induction. Cyclic manganate diester 2733

    is widely

    recognized as the common intermediate in permanganate reaction and we proposed that it is

    Ph COOtBu

    (S,S)-3d (2.0 mol%) 1.5 equiv. oxidant

    TBME/20 wt% aq. KI

    -60 oC

    Ph COOtBu

    OH

    HO

    83% ee with TBA+MnO4-

    92% ee with KMnO4

    a)

    R CHOAr COOtBu

    OC-C Cleavage

    ArO

    MnO

    COOtBu

    O O

    H

    R

    COOtBu

    Ar

    OH

    HOCOOtBu

    Ar

    OH

    R

    OR

    Dihydroxylation

    H+

    Oxohydroxylation

    22 or 25 26

    27

    +

    b)

    21a

    c)

    MnO O

    O O

    Mn OO

    O O-

    COOtBu

    Ar

    R

    Mn OO

    O O-

    Ar

    COOtBu

    R

    MnO O

    O O

    Ar

    -O

    OtBuR COOtBu

    Ar

    R

    OMn

    O

    O O-

    BG2+

    BG2+

    accelerated pathway

    MnO O

    O O

    ArO-

    OtBu

    R

    BG2+

    low energy TS I

    high energy TS II

    BG2+

    COOtBu

    Ar

    R

    OMn

    O

    O O-BG2+

    KMnO4BG2+

    phase transfer

    Re

    COOtBu

    Ar

    R

    E, Z

    Si

    oxidationproducts

    Ar

    COOtBu

    R

    E, Z

    22a

    ´

  • 22

    through this intermediate that different products are obtained, via diverse pathways (Scheme

    17b). We found that the isolated diols 25 were not further oxidized to 26 under oxohydroxylation

    conditions, indicating that the diols are not intermediates in oxohydroxylation.

    The absolute configuration of the α-chiral center of oxidation products 26, derived from both

    Z- and E-trisubstituted enoates 24, is identical. On the basis of Lee’s prior work34

    and our

    experimental results, a working model of reaction acceleration through the formation of an

    intimate ion pair between bisguanidinium and enolate anion in the transition states was proposed

    (Scheme 17c). The results reveal that rate acceleration is mainly attributed to transition state

    stabilization through strong electrostatic interaction between dicationic BG and enolate anion.

    The disclosure of this catalytic asymmetric permanganate oxidation opens up new paradigms for

    chiral cation-directed catalysis. Exploration of other functional anions will undoubtedly lead to

    the discovery of interesting new transformations for asymmetric synthesis.

    3.2 Bisguanidinium diphosphatobisperoxotungstate-catalyzed sulfoxidation

    Peroxotungstate species have been reported to serve as catalyst for oxidation reactions such as

    epoxidation35

    and sulfoxidation.36

    Since peroxotungstate are anionic species, we hypothesized

    that their reactions can be accelerated and modulated using chiral cations. We found that with 2.0

    mol% of BG (S,S)-3a and 2.0 mol% of Ag2WO4, oxidation of heterocyclic sulfides 28a went

    smoothly to produce sulfoxides 29a with excellent enantioselectivities (Scheme 18).37

    Notably,

    the amount of additive, NaH2PO4 or NH4H2PO4 is crucial to achieve high yield and high

    stereoinduction. A series of enantioenriched sulfoxides 29a-q bearing benzimidazole,

    benzothiazole, pyridine, thiophene moieties were obtained, which are of potential interest for

  • 23

    drug development. The practical utility was further illustrated in the preparation of proton-pump

    inhibitor (S)-Lansoprazole 29l.

    Scheme 18. Bisguanidinium diphosphatobisperoxotungstate-catalyzed sulfoxidation.

    Experimentally, we found that a ratio of greater than 2 equivalent of dihydrogen phosphate

    (H2PO4−) with respect to tungstate is essential for achieving high enantioselectivity. Raman

    spectra of the active catalyst was obtained (Figure 6) and an experimentally peak observed at 711

    cm-1

    corresponds well to a peak at 713 cm-1

    in the computed Raman spectra of

    28 29

    (S,S)-3a (2.0 mol%)

    Ag2WO4 (2.0 mol% or 5.0 mol%)

    NaH2PO4 or NH4H2PO4 (10 mol%)

    35%wt H2O2 (1.05 equiv.)

    solvent, 0 oC

    N

    N S

    Me

    29a96% yield92% ee

    N

    N S

    MeCl

    29c81% yield97% ee

    N

    N S

    MeF

    29d78%yield 89%ee

    N

    N S

    MeNO2

    29e74%yield88%ee

    N

    N S

    Me

    29f 95%yield

    99%ee

    N

    N S

    Me

    29g 84%yield

    92%ee

    N

    N S

    Me

    29h72%yield 95%ee

    O

    O

    N

    N S

    Me

    29i71%yield82%ee

    N

    N S CO2Et

    Me

    29j 92%yield 90%ee

    N

    N S CO2tBu

    Me

    29k 86%yield

    80%ee

    CN

    ArS R

    ArS R

    N

    S S

    29m78%yield 93%ee

    N

    S S

    29o85%yield92%ee

    N

    S S

    29n70%yield 91%ee

    S

    29p76%yield 80%ee

    S S

    29q53%yield94%ee

    N

    O

    N

    HN S

    29l81% yield 90% ee

    N

    O

    CF3

    S

    29r79%yield90%ee

    N

    N S

    Me

    29b94% yield92% ee

    C6F5

    Me

    O

    OOO

    O O O

    OOO

    O O O

    OOO

    O O

  • 24

    [{PO2(OH)2}2{WO(O2)2}]2-

    . This peak is attributed to the twisting of P-O-H group in the

    phosphate ligand. A more thorough analysis of the [{PO2(OH)2}2{WO(O2)2}]2-

    revealed that an

    intramolecular hydrogen bond interaction between the two phosphate ligands is important for

    this peak; absence of such interaction will shift the vibrational frequency towards 800 cm-1

    . We

    thus proposed that the active anion is diphosphatobisperoxotungstate

    [{PO2(OH)2}2{WO(O2)2}]2-

    (Scheme 19a). Based on both computational and experimental

    results, a simple working model is proposed detailing the formation of

    diphosphatobisperoxotungstate and its transport from the aqueous to the organic phase (Scheme

    19b).

    Figure 6. Experimental Raman spectrum of [{PO2(OH)2}2{WO(O2)2}]2-

    .

  • 25

    Scheme 19. a) Diphosphatobisperoxotungstate 30 and b) proposed working model.

    3.3 Bisguanidinium dinuclear oxodiperoxomolybdosulfate-catalyzed sulfoxidation

    Similar to peroxotungstates, peroxomolybdates38

    are well-known catalysts for oxidation

    reactions.39

    However, they are typically complex mixtures and it is widely recognized as the

    challenging obstacle for elaborating them into highly enantioselective catalyst. In the attempt to

    prepare enantiopure sulfoxides of 2-sulfinyl esters through alkylation of sulfenate anions using

    -halogenated carboxylate, an unsatisfactory outcome was observed. Thus, direct sulfoxidation

    strategy was adopted to gain access to 2-sulfinyl ester sulfoxides 31a, which can be further

    elaborated to a commercial drug armodafinil. We found that with 2.5 mol % Na2MoO4·2H2O and

    potassium hydrogen sulfate (KHSO4), the oxidation proceeded efficiently with 1.0 mol% of BG

    (S,S)-3a and H2O2 as oxidant (Scheme 20a).40

    Its practical utility was successfully demonstrated

    in the gram-scale synthesis of (R)-modafinil (armodafinil) using a 0.25 mol% loading of (R,R)-

    3a (Scheme 20b).

    BG2+

    Ag2WO4

    organic phase

    water phase[{PO2(OH)2}2{WO(O2)2}]2-

    BG2+2Cl-

    W

    OP(O)(OH)2

    O

    OP(O)(OH)2O

    O

    O O

    2-Ar

    S

    R

    ArS R

    O

    BG2+[{PO2(OH)2}2{WO(O2)2}]2-

    W

    OP(O)(OH)2

    O

    OP(O)(OH)2O

    O

    O

    2-

    H2O2

    W

    O

    OO

    O

    O

    (HO)2(O)PO

    OP(O)(OH)2

    30Diphosphatobisperoxotungstate

    2-

    [{PO2(OH)2}2{WO(O2)2}]2-

    + H2O2 + NaH2PO4 (aq.)

    a) b)

  • 26

    Scheme 20. a) Catalytic molybdate-mediated sulfoxidation and b) preparative scale synthesis of

    armodafinil.

    The reactive catalytic species can be prepared by mimicking the reaction conditions in the

    absence of sulfide substrate (Scheme 21). The active catalyst was collected as a pale-yellow

    precipitate through filtration and the structure of the catalyst, (R,R)-3a-[Mo], was confirmed

    using X-ray analysis (Figure 7)41 and 95Mo NMR.42 The achiral anionic molybdenum species

    [(2-SO4){Mo2O2(2-O2)2(O2)2}]2-

    is revealed by X-ray crystallography to be embedded within

    the chiral cavity formed by two side arms of the chiral bisguanidinium dication.

    Scheme 21. Identification of the active catalyst.

    S CO2MePh

    Ph

    SPh

    Ph

    (R,R)-3a (0.25 mol%)Na2MoO4.2H2O (2.5 mol%)

    ent-31a91% yield91% ee

    35% aq. H2O2 (1.05 equiv.)

    KHSO4 (0.5 equiv.)

    nBu2O (0.05 M), rt, 8 h

    30a5 mmol

    MeOH (2 M)rt, 24 h

    NH3(10.0 equiv.)

    Armodafinil, 1.19 g95% yield, 91% ee

    32

    CO2Me

    O

    SPh

    Ph

    O

    NH2

    Ob)

    S CO2MePh

    Ph

    SPh

    Ph

    (S,S)-3a (0.25 mol%)Na2MoO4.2H2O (2.5 mol%)

    31a99% yield94% ee

    35% aq. H2O2 (1.05 equiv.)

    KHSO4 (0.5 equiv.)

    iPr2O (0.05 M), 0 oC, 1 h

    30a0.2 mmol

    CO2Me

    Oa)

    (R,R)-3a

    BG2+[2Cl]2-

    (1.0 mol%)

    0.04 mmol

    (R,R)-3a-[Mo]

    BG2+[(m2-SO4){Mo2O2(m2-O2)2(O2)2}]2-

    91% yield

    Na2MoO4·2H2O (2.5 mol%), 35% aq. H2O2 (1.0 equiv.)

    KHSO4 (0.5 equiv.) or H2SO4 (0.25 equiv.)Et2O (2 mL), rt, 2 h

  • 27

    Figure 7. X-Ray crystallographic structure of [BG]2+

    [(2-SO4){Mo2O2(2-O2)2(O2)2}]2-

    (R,R)-

    3a-[Mo](ellipsoids at 50% probability)

    The catalyst (R,R)-3a-[Mo] displayed excellent catalytic activity in sulfoxidation with H2O2 as

    oxidant (Scheme 22). (R,R)-3a-[Mo] was used directly as a stoichiometric oxidant and it

    provided sulfoxide ent-31a in 90% yield and 80% ee. This result confirm that (R,R)-3a-[Mo] is

    the actual oxidant and high level of enantiodiscrimination can be achieved. However, using just

    0.25 equivalent of (R,R)-3a-[Mo] led to the formation of ent-31a in 50% yield with 31% ee;

    demonstrating that two out of four peroxo moieties on (R,R)-3a-[Mo] are active oxygen donors

    as two equivalents of active oxygen from (R,R)-3a-[Mo] are transferred to the sulfides. On the

    basis of the experimental results, a possible mechanistic model was proposed (Scheme 23).43

    It is

    proposed that the second oxygen transfer is slower and less enantioselective than the first. In the

    presence of H2O2, the dimeric structure of the catalyst will be maintained and it is this structure

    that provides the highly effective enantiofacial discrimination.

    NN

    N

    N N

    N

    R1

    R2

    R1

    R2

    Ph

    PhPh

    Ph

    OS

    O

    O O

    Mo

    OO

    Mo

    O

    O

    O

    OO

    OO

    O

    [(m2-SO4){Mo2O2(m2-O2)2(O2)2}]2-

    (R,R)-3a-[Mo]

    2-

    [Mo]2-

    [Mo]2- =

    S CO2MePh

    Ph

    SPh

    Ph

    (R,R)-3a-[Mo] (x equiv.)

    35% aq. H2O2 (y equiv.)

    ent-31aiPr2O, rt, 8 h30a

    CO2Me

    O

    x = 0.01, y = 1.05 x = 1.0, y = 0 x = 0.25, y = 0

    yield 95%, ee 91%yield 90%, ee 80%yield 50%, ee 31%

  • 28

    Scheme 22. Mechanistic studies by using [BG]2+

    [(-SO4)Mo2O2(-O2)2(O2)2]2(R,R)-3a-[Mo].

    Scheme 23. Proposed catalytic cycle of bisguanidinium dinuclear oxodiperoxomolybdosulfate.

    3.4 Bisguanidinium-catalyzed alkylation of cyclic ketones

    Previously, we used the silylamide probase strategy with pentanidium as catalyst, to achieve

    the highly enantioselective alkylation of lactones24

    . For cyclic and linear ketones, we found that

    bisguanidiniums are more suitable. The probase can act as a silylation reagent to generate silyl

    enol ether, which is a key intermediate for the alkylation. With 10 mol% of BG (S,S)-3a and 2

    equivalent of probase, bis(tert-butyldimethylsilyl)acetamide (BTBSA), a variety of -benzyl-1-

    indanones and -benzyl-tetralones 32a-i can be smoothly converted to alkylation product 33a-i

    in high yields and enantioselectivities (Scheme 24a).24

    TBS enol ether 34 was prepared from -

    benzyl-1-indanone 32a and submitted to the alkylation condition (Scheme 24b). The

    enantioselectivity and yield obtained were similar to the conditions using probase directly,

    demonstrating that TBS enol ether is an intermediate in the reaction. For the alkylation of linear

    BG2+

    BG2+

    H2O2

    sulfide

    sulfoxide

    BG2+

    sulfidesulfoxide

    first oxygen transfer ishighly enantioselective

    second oxygen is less enantioselective

    [O]

    [O]

    (R,R)-3a-[Mo]

    formation of B was inhibited in the presence of

    excess H2O2

    OS

    O

    O O

    Mo

    OO

    Mo

    O

    O

    O

    OO

    OO

    O

    2-

    OS

    O

    O O

    Mo

    OO

    Mo

    O

    OO

    OO

    O

    2-

    O

    OS

    O

    O O

    Mo

    OO

    Mo

    O

    OO

    O

    2-

    O

    O

    BA

  • 29

    ketones, such as simple phenyl ethyl ketone, due to their low reactivity, the corresponding

    alkylation reaction can only be achieved via the preparation of their silyl enol ethers.

    Scheme 24. a) Bisguanidinium 3a-catalyzed enantioselective alkylation of cyclic ketones using

    probase strategy and b) identification of silyl enol ether as an intermediate.

    CsF (5.0 equiv.)

    Et2O, -40 oC

    O(S,S)-3a (10 mol%)BTBSA (2.0 equiv.)

    32a-n 33b-un

    n = 1,2

    O

    Ph

    98% yield, 95% ee

    33a

    O

    Ph

    99% yield, 89% ee

    33b

    S

    O

    Ph

    95% yield, 85% ee

    33c

    O

    O

    98% yield, 91% ee

    33d

    BrO

    95% yield, 95% ee

    33e

    S

    O

    99% yield, 90% ee

    33f

    CF3

    OMe

    O

    O

    Ph

    F

    99% yield, 92% ee

    33i

    O

    CF3

    75% yield, 90% ee

    33j

    O

    87% yield, 92% ee

    33k

    S

    O

    O

    CF3

    87% yield, 77% ee

    33lOMe

    Br R3

    O

    Ph

    90% yield, 98% ee

    33g

    F3C

    O

    Ph

    96% yield, 92% ee

    33hCl

    R2

    n = 1,2

    R1

    OTBS

    NTBS

    (BTBSA)

    O

    n

    R1R3

    b)

    a)

    O

    Ph

    PhBrO

    Ph

    Bn

    32a 33a

    (S,S)-3a (10 mol%)

    CsF(5.0 equiv.)

    Et2O, -40 oC

    (S,S)-3a (10 mol%)BTBSA (2.0 equiv.)

    CsF( 5.0 equiv.)

    Et2O, -40 oC yield 98%, ee 95%

    PhBrKHMDSTBSCl

    34 84% yield, 94% eeTHF, -78 oC

    OTBS

    Ph

    R2

  • 30

    Scheme 25. Proposed working model with BG hypervalent silicates ion pair.

    Based on the experimental results, a working model for enantioselective alkylation using

    silylamide as probase is putatively proposed (Scheme 25). When the probase is activated with

    fluoride, BG silylamide is first formed followed by BG enolate A. Silyl enol ether B is

    subsequently formed via silylation with another silylamide. We found that the steric features of

    the probase affects strongly the enantioselectivities and bis(tert-butyldimethylsilyl)acetamide

    (BTBSA) afforded better results than bis(trimethylsilyl)acetamide (BSA). A BG hypervalent

    silicates ion pair C is proposed to be the key intermediate, which is responsible for the

    enantiofacial differentiation of the incoming electrophiles.

    4. CONCLUSION AND PERSPECTIVES

    Over the past few years, we have developed N-sp2 hybridized guanidinium-type salts as

    efficient phase transfer catalysts. We have shown that these are general catalysts for a variety of

    transformations and reactions can be performed at a preparative scale with low loading of

    catalyst. In some cases, commercially important compounds are prepared as a demonstration of

    Br R3

    BG2+[2Cl]2-

    probase

    N

    O

    Si

    Si

    N

    O

    Si

    BG2+

    O

    Bn

    CsF

    O

    Bn

    -BG2+

    O

    Bn

    Si

    BG2+[2Cl]2-

    CsF

    C

    O

    Bn

    Si

    F

    L

    LL

    F

    BG2+O

    Bn

    N

    O

    Si

    BG2+

    chiral organic base

    chiral organic base

    A

    B

    R3

    32a

    33a

    2-

  • 31

    industrial applicability. We have also articulated a new mode of catalysis at the interface of

    phase transfer catalysis, organocatalysis and organometallic catalysis. We describe the possibility

    of activating and modulating organometallic catalysts through ionic interactions and not through

    the use of ligands. Chiral cationic ion pairing catalysis is in its infancy and we hope that more

    colleagues will find this approach refreshing and join us in exploring and defining its scope.

    AUTHOR INFORMATION

    Corresponding Author

    *E-mail: [email protected]

    Notes

    The authors declare no competing financial interest.

    Biographical Information

    Lili Zong was born in Henan (China) in 1985. She received her B.S. in engineering from Harbin

    Institute of Technology (China) in 2006. She studied organic chemistry under the supervision of

    Professor Yixiang Cheng during her M.Sc. in Nanjing University (China). In 2010, she started

    her graduate studies at National University of Singapore under the tutelage of Professor Choon-

    Hong Tan. She is currently a postdoctoral fellow in the laboratory of Professor Stefan Matile at

    University of Geneva.

    Choon-Hong Tan was born in Singapore in 1971. He received his BSc(Hons) First Class from

    the National University of Singapore in 1995. He obtained financial support from Trinity College

    (Cambridge), the Cambridge Commonwealth Trust and Tan Kah Kee Postgraduate Scholarship

    to pursue his PhD and he graduated from the University of Cambridge in 1999. Following that,

    he carried out postdoctoral training at Harvard University. He was a Research Associate at

    Harvard Medical School before joining the National University of Singapore in 2003 to start his

  • 32

    independent career. He joined Nanyang Technological University (Singapore) in 2012 and was

    promoted to Full Professor in 2016.

    ACKNOWLEDGMENT

    We are grateful to the technical and intellectual contributions of our co-workers whose names are

    listed in the relevant references. Financial support for the research was provided by National

    University of Singapore and Nanyang Technological University (Singapore).

    REFERENCES

    (1) Riordan, J. F. Arginyl residues and anion binding sites in proteins. Mol. Cell. Biochem.

    1979, 26, 71-92.

    (2) (a) Müller, G.; Riede, J.; Schmidtchen, F. P. Host-Guest Bonding of Oxoanions to

    Guanidinium Anchor Groups. Angew. Chem., Int. Ed. 1988, 27, 1516-1518. (b) Berger, M.;

    Schmidtchen, F. P. Zwitterionic Guanidinium Compounds Serve as Electroneutral Anion Hosts.

    J. Am. Chem. Soc. 1999, 121, 9986-9993.

    (3) Chinchilla, R.; Nájera, C.; Sánchez-Agulló, P. Enantiomerically pure guanidine-catalysed

    asymmetric nitroaldol reaction. Tetrahedron: Asymmetry 1994, 5, 1393-1402.

    (4) Iyer, M. S.; Gigstad, K. M.; Namdev, N. D.; Lipton, M. Asymmetric Catalysis of the

    Strecker Amino Acid Synthesis by a Cyclic Dipeptide. J. Am. Chem. Soc. 1996, 118, 4910-4911.

    (5) Corey, E. J.; Grogan, M. J. Enantioselective Synthesis of α-Amino Nitriles from N-

    Benzhydryl Imines and HCN with a Chiral Bicyclic Guanidine as Catalyst. Org. Lett. 1999, 1,

    157-160.

    (6) (a) Terada, M.; Nakano, M.; Ube, H. Axially Chiral Guanidine as Highly Active and

    Enantioselective Catalyst for Electrophilic Amination of Unsymmetrically Substituted 1,3-

    Dicarbonyl Compounds. J. Am. Chem. Soc. 2006, 128, 16044-16045. (b) Shen, J.; Nguyen, T. T.;

    Goh, Y. P.; Ye, W. P.; Fu, X.; Xu, J. Y.; Tan, C.-H. Chiral bicyclic guanidine-catalyzed

    enantioselective reactions of anthrones. J. Am. Chem. Soc. 2006, 128, 13692-13693. (c) Fu, X.;

    Jiang, Z.; Tan, C.-H. Bicyclic guanidine-catalyzed enantioselective phospha-Michael reaction:

    synthesis of chiral beta-aminophosphine oxides and beta-aminophosphines. Chem. Commun.

    2007, 5058-5060. (d) Ye, W. P.; Jiang, Z. Y.; Zhao, Y. J.; Goh, S. L. M.; Leow, D.; Soh, Y. T.;

    Tan, C.-H. Chiral bicyclic guanidine as a versatile Bronsted base catalyst for the enantioselective

    Michael reaction of dithomalonates and beta-keto thioesters. Adv. Synth. Catal. 2007, 349, 2454-

    2458. (e) Leow, D. S.; Lin, S. S.; Chittimalla, S. K.; Fu, X.; Tan, C.-H. Enantioselective

    protonation catalyzed by a chiral bicyclic guanidine derivative. Angew. Chem., Int. Ed. 2008, 47,

    5641-5645. (f) Jiang, Z. Y.; Pan, Y. H.; Zhao, Y. J.; Ma, T.; Lee, R.; Yang, Y. Y.; Huang, K. W.;

    Wong, M. W.; Tan, C.-H. Synthesis of a Chiral Quaternary Carbon Center Bearing a Fluorine

    Atom: Enantio- and Diastereoselective Guanidine-Catalyzed Addition of Fluorocarbon

    Nucleophiles. Angew. Chem., Int. Ed. 2009, 48, 3627-3631. (g) Liu, H. J.; Leow, D.; Huang, K.

    W.; Tan, C.-H. Enantioselective Synthesis of Chiral Allenoates by Guanidine-Catalyzed

    Isomerization of 3-Alkynoates. J. Am. Chem. Soc. 2009, 131, 7212-7213. (h) Wang, J. M.; Chen,

  • 33

    J.; Kee, C. W.; Tan, C.-H. Enantiodivergent and gamma-Selective Asymmetric Allylic

    Amination. Angew. Chem., Int. Ed. 2012, 51, 2382-2386.

    (7) (a) Leow, D.; Tan, C.-H. Chiral Guanidine Catalyzed Enantioselective Reactions. Chem.

    Asian J. 2009, 4, 488-507. (d) Leow, D.; Tan, C.-H. Catalytic Reactions of Chiral Guanidines

    and Guanidinium Salts. Synlett 2010, 1589-1605. (e) Wang, C.; Goh, C. M. T.; Xiao, S. H.; Ye,

    W. P.; Tan, C.-H. Enantioselective Protonation Catalyzed by Chiral Bronsted Bases. J. Synth.

    Org. Chem. Jpn. 2013, 71, 1145-1151.

    (8) (a) Shirakawa, S.; Maruoka, K. Recent Developments in Asymmetric Phase-Transfer

    Reactions. Angew. Chem., Int. Ed. 2013, 52, 4312-4348. (b) Kaneko, S.; Kumatabara, Y.;

    Shirakawa, S. A new generation of chiral phase-transfer catalysts. Org. Biomol. Chem. 2016, 14,

    5367-5376. (c) Albanese, D. C. M.; Foschi, F.; Penso, M. Sustainable Oxidations under Phase-

    Transfer Catalysis Conditions. Org. Process Res. Dev. 2016, 20, 129-139.

    (9) Ma, T.; Fu, X. A.; Kee, C. W.; Zong, L. L.; Pan, Y. H.; Huang, K. W.; Tan, C.-H.

    Pentanidium-Catalyzed Enantioselective Phase-Transfer Conjugate Addition Reactions. J. Am.

    Chem. Soc. 2011, 133, 2828-2831.

    (10) Fonseca, G. O.; Cook, J. M. Modern Methods for Total Synthesis of Important Oxindole

    Alkaloids. Org. Chem. Insights 2016, 6, 1-55.

    (11) Peddibhotla, S. 3-Substituted-3-hydroxy-2-oxindole, an Emerging New Scaffold for Drug

    Discovery with Potential Anti-Cancer and other Biological Activities. Curr. Bioact. Compd.

    2009, 5, 20-38.

    (12) Sano, D.; Nagata, K.; Itoh, T. Catalytic Asymmetric Hydroxylation of Oxindoles by

    Molecular Oxygen Using a Phase-Transfer Catalyst. Org. Lett. 2008, 10, 1593-1595.

    (13) Yang, Y.; Moinodeen, F.; Chin, W.; Ma, T.; Jiang, Z.; Tan, C.-H. Pentanidium–Catalyzed

    Enantioselective α-Hydroxylation of Oxindoles Using Molecular Oxygen. Org. Lett. 2012, 14,

    4762-4765.

    (14) Li, H.; Song, J.; Liu, X.; Deng, L. Catalytic Enantioselective C−C Bond Forming

    Conjugate Additions with Vinyl Sulfones. J. Am. Chem. Soc. 2005, 127, 8948-8949.

    (15) (a) Ohmatsu, K.; Kiyokawa, M.; Ooi, T. Chiral 1,2,3-Triazoliums as New Cationic

    Organic Catalysts with Anion-Recognition Ability: Application to Asymmetric Alkylation of

    Oxindoles. J. Am. Chem. Soc. 2011, 133, 1307-1309. (b) Zhu, Q.; Lu, Y. Stereocontrolled

    Creation of All-Carbon Quaternary Stereocenters by Organocatalytic Conjugate Addition of

    Oxindoles to Vinyl Sulfone. Angew. Chem., Int. Ed. 2010, 49, 7753-7756.

    (16) Zong, L.; Du, S.; Chin, K. F.; Wang, C.; Tan, C.-H. Enantioselective Synthesis of

    Quaternary Carbon Stereocenters: Addition of 3-Substituted Oxindoles to Vinyl Sulfone

    Catalyzed by Pentanidiums. Angew. Chem., Int. Ed. 2015, 54, 9390-9393.

    (17) Pallavicini, M.; Valoti, E.; Villa, L.; Resta, I. New asymmetric synthesis of (−)-

    esermethole. Tetrahedron: Asymmetry 1994, 5, 363-370.

    (18) (a) O'Donnell, J. S.; Schwan, A. L. Generation, structure and reactions of sulfenic acid

    anions. J. Sulfur Chem. 2004, 25, 183-211. (b) Schwan, A. L.; Söderman, S. C. Discoveries in

    Sulfenic Acid Anion Chemistry. Phosphorus, Sulfur Silicon Relat. Elem. 2013, 188, 275-286.

    (19) (a) Caupène, C.; Boudou, C.; Perrio, S.; Metzner, P. Remarkably Mild and Simple

    Preparation of Sulfenate Anions from β-Sulfinylesters: A New Route to Enantioenriched

    Sulfoxides. J. Org. Chem. 2005, 70, 2812-2815. (b) Maitro, G.; Vogel, S.; Sadaoui, M.; Prestat,

    G.; Madec, D.; Poli, G. Enantioselective Synthesis of Aryl Sulfoxides via Palladium-Catalyzed

    Arylation of Sulfenate Anions. Org. Lett. 2007, 9, 5493-5496. (c) Foucoin, F.; Caupène, C.;

  • 34

    Lohier, J.-F.; Sopkova de Oliveira Santos, J.; Perrio, S.; Metzner, P. 2-(Trimethylsilyl)ethyl

    Sulfoxides as a Convenient Source of Sulfenate Anions. Synthesis 2007, 2007, 1315-1324.

    (20) Gelat, F.; Jayashankaran, J.; Lohier, J.-F.; Gaumont, A.-C.; Perrio, S. Organocatalytic

    Asymmetric Synthesis of Sulfoxides from Sulfenic Acid Anions Mediated by a Cinchona-

    Derived Phase-Transfer Reagent. Org. Lett. 2011, 13, 3170-3173.

    (21) Zong, L.; Ban, X.; Kee, C. W.; Tan, C.-H. Catalytic Enantioselective Alkylation of

    Sulfenate Anions to Chiral Heterocyclic Sulfoxides Using Halogenated Pentanidium Salts.

    Angew. Chem., Int. Ed. 2014, 53, 11849-11853.

    (22) Cotton, H.; Elebring, T.; Larsson, M.; Li, L.; Sörensen, H.; von Unge, S. Asymmetric

    synthesis of esomeprazole. Tetrahedron: Asymmetry 2000, 11, 3819-3825.

    (23) Gilday, L. C.; Lang, T.; Caballero, A.; Costa, P. J.; Félix, V.; Beer, P. D. Angew. Chem.,

    Int. Ed. 2013, 52, 4356-4360.

    (24) Teng, B.; Chen, W.; Dong, S.; Kee, C. W.; Gandamana, D. A.; Zong, L.; Tan, C.-H.

    Pentanidium- and Bisguanidinium-Catalyzed Enantioselective Alkylations Using Silylamide as

    Brønsted Probase. J. Am. Chem. Soc. 2016, 138, 9935-9940.

    (25) Inamoto, K.; Okawa, H.; Taneda, H.; Sato, M.; Hirono, Y.; Yonemoto, M.; Kikkawa, S.;

    Kondo, Y. Organocatalytic deprotonative functionalization of C(sp2)-H and C(sp

    3)-H bonds

    using in situ generated onium amide bases. Chem. Commun. 2012, 48, 9771-9773.

    (26) Chen, W.; Yang, W.; Yan, L.; Tan, C.-H.; Jiang, Z. Bicyclic guanidinium-catalyzed

    enantioselective phase-transfer alkylation: direct access to pyrroloindolines and furoindolines.

    Chem. Commun. 2013, 49, 9854-9856.

    (27) Fu, X.; Loh, W. T.; Zhang, Y.; Chen, T.; Ma, T.; Liu, H. J.; Wang, J. M.; Tan, C.-H.

    Chiral Guanidinium Salt Catalyzed Enantioselective Phospha-Mannich Reactions. Angew.

    Chem., Int. Ed. 2009, 48, 7387-7390.

    (28) Uyanik, M.; Okamoto, H.; Yasui, T.; Ishihara, K. Quaternary Ammonium (Hypo)iodite

    Catalysis for Enantioselective Oxidative Cycloetherification. Science 2010, 328, 1376-1379.

    (29) Brak, K.; Jacobsen, E. N. Asymmetric Ion-Pairing Catalysis. Angew. Chem., Int. Ed.

    2013, 52, 534-561.

    (30) (a) Hamilton, G. L.; Kang, E. J.; Mba, M.; Toste, F. D. A Powerful Chiral Counterion

    Strategy for Asymmetric Transition Metal Catalysis. Science 2007, 317, 496-499. (b) Phipps, R.

    J.; Hamilton, G. L.; Toste, F. D. The progression of chiral anions from concepts to applications

    in asymmetric catalysis. Nat. Chem. 2012, 4, 603-614. (c) Mahlau, M.; List, B. Asymmetric

    Counteranion-Directed Catalysis: Concept, Definition, and Applications. Angew. Chem., Int. Ed.

    2013, 52, 518-533.

    (31) Bhunnoo, R. A.; Hu, Y.; Lainé, D. I.; Brown, R. C. D. An Asymmetric Phase-Transfer

    Dihydroxylation Reaction. Angew. Chem., Int. Ed. 2002, 41, 3479-3480.

    (32) Wang, C.; Zong, L.; Tan, C.-H. Enantioselective Oxidation of Alkenes with Potassium

    Permanganate Catalyzed by Chiral Dicationic Bisguanidinium. J. Am. Chem. Soc. 2015, 137,

    10677-10682.

    (33) (a) Fatiadi, A. J. The Classical Permanganate Ion: Still a Novel Oxidant in Organic

    Chemistry. Synthesis 1987, 1987, 85-127. (b) Dash, S.; Patel, S.; Mishra, B. K. Oxidation by

    permanganate: synthetic and mechanistic aspects. Tetrahedron 2009, 65, 707-739.

    (34) Lee, D. G.; Brown, K. C. Oxidation of hydrocarbons. 11. Kinetics and mechanism of the

    reaction between methyl (E)-cinnamate and quaternary ammonium permanganates. J. Am. Chem.

    Soc. 1982, 104, 5076-5081.

  • 35

    (35) Sato, K.; Aoki, M.; Ogawa, M.; Hashimoto, T.; Noyori, R. A Practical Method for

    Epoxidation of Terminal Olefins with 30% Hydrogen Peroxide under Halide-Free Conditions. J.

    Org. Chem. 1996, 61, 8310-8311.

    (36) Sato, K.; Hyodo, M.; Aoki, M.; Zheng, X.-Q.; Noyori, R. Oxidation of sulfides to

    sulfoxides and sulfones with 30% hydrogen peroxide under organic solvent- and halogen-free

    conditions. Tetrahedron 2001, 57, 2469-2476.

    (37) Ye, X.; Moeljadi, A. M. P.; Chin, K. F.; Hirao, H.; Zong, L.; Tan, C.-H. Enantioselective

    Sulfoxidation Catalyzed by a Bisguanidinium Diphosphatobisperoxotungstate Ion Pair. Angew.

    Chem., Int. Ed. 2016, 55, 7101-7105.

    (38) Dickman, M. H.; Pope, M. T. Peroxo and Superoxo Complexes of Chromium,

    Molybdenum, and Tungsten. Chem. Rev. 1994, 94, 569-584.

    (39) (a) Shi, X.; Wei, J. Preparation, characterization and catalytic oxidation properties of bis-

    quaternary ammonium peroxotungstates and peroxomolybdates complexes. Appl. Organomet.

    Chem. 2007, 21, 172-176. (b) Chakravarthy, R. D.; Ramkumar, V.; Chand, D. K. A molybdenum

    based metallomicellar catalyst for controlled and selective sulfoxidation reactions in aqueous

    medium. Green Chem. 2014, 16, 2190-2196.

    (40) Zong, L.; Wang, C.; Moeljadi, A. M. P.; Ye, X.; Hirao, H.; Tan, C.-H. Bisguanidinium

    dinuclear oxodiperoxomolybdosulfate ion pair-catalyzed enantioselective sulfoxidation. Nat.

    Commun. 2016, 7, 13455.

    (41) Salles, L.; Robert, F.; Semmer, V.; Jeannin, Y.; Bregeault, J. M. Novel di- and trinuclear

    oxoperoxosulfato species in molybdenum(VI) and tungsten(VI) chemistry: The key role of pairs

    of bridging peroxo groups. Bull. Soc. Chim. Fr. 1996, 133, 319-328.

    (42) Nardello, V.; Marko, J.; Vermeersch, G.; Aubry, J. M. 90

    Mo NMR and kinetic studies of

    peroxomolybdic intermediates involved in the catalytic disproportionation of hydrogen peroxide

    by molybdate ions. Inorg. Chem. 1995, 34, 4950-4957.

    (43) Thompson, D. J.; Cao, Z.; Judkins, E. C.; Fanwick, P. E.; Ren, T. Peroxo-dimolybdate

    catalyst for the oxygenation of organic sulfides by hydrogen peroxide. Inorg. Chim. Acta 2015,

    437, 103-109.