chapter 5 cadmium-proline catalyzed aldol...

52
165 CHAPTER 5 CADMIUM-PROLINE CATALYZED ALDOL REACTIONS 5.1 Introduction The aldol reaction 1,2 is one of the most venerable reactions in organic chemistry, firstly discovered by Wurtz in 1872, 1b although Kane previously described the prevently known Aldol condensation. 1a This useful transformation allows the formation of a C–C bond by reaction of an enolizable carbonyl compound acting as a source of nucleophile with itself or another carbonyl compound acting as an electrophile to give a β-hydroxy carbonyl compound known as aldol. Without any doubt, catalytic enantioselective methods are the most attractive alternative for providing chiral compounds with high selectivity and atom efficiency. 3 Biochemical methods based on the use of aldolase enzymes 4 and antibodies 5 have shown their usefulness to perform this task, with the scope of substrates being very narrow. In search of a wider substrate scope, different enantioselective chemical methods have been extensively developed in recent years, especially after the introduction of the Mukaiyama-aldol version of this reaction. 6 In this case, the generation of a silyl enol ether (or chemical equivalent) is compulsory, requiring the use of stoichiometric amounts of bases and silylating reagents and therefore having low atom efficiencies. In order to enhance the global efficiency of the process, an effort has been made to develop enantioselective direct aldol reactions, 7 where the use of performed enolates (or their equivalents) is avoided. In general, the asymmetric aldol reactions can be categorized into the following five types, (i) chiral auxiliary assisted aldol reaction based on the use of stoichiometric quantities of the chiral appendages; (ii) chiral Lewis acid and Lewis base catalysed (iii) heterobimetallic bifunctional Lewis acid/Bronsted

Upload: others

Post on 29-Jan-2021

2 views

Category:

Documents


0 download

TRANSCRIPT

  • 165

    CHAPTER 5

    CADMIUM-PROLINE CATALYZED ALDOL

    REACTIONS

    5.1 Introduction The aldol reaction1,2 is one of the most venerable reactions in organic

    chemistry, firstly discovered by Wurtz in 1872,1b although Kane previously described

    the prevently known Aldol condensation.1a This useful transformation allows the

    formation of a C–C bond by reaction of an enolizable carbonyl compound acting as a

    source of nucleophile with itself or another carbonyl compound acting as an

    electrophile to give a β-hydroxy carbonyl compound known as aldol. Without any

    doubt, catalytic enantioselective methods are the most attractive alternative for

    providing chiral compounds with high selectivity and atom efficiency.3 Biochemical

    methods based on the use of aldolase enzymes4and antibodies5 have shown their

    usefulness to perform this task, with the scope of substrates being very narrow. In

    search of a wider substrate scope, different enantioselective chemical methods have

    been extensively developed in recent years, especially after the introduction of the

    Mukaiyama-aldol version of this reaction.6 In this case, the generation of a silyl enol

    ether (or chemical equivalent) is compulsory, requiring the use of stoichiometric

    amounts of bases and silylating reagents and therefore having low atom efficiencies.

    In order to enhance the global efficiency of the process, an effort has been made to

    develop enantioselective direct aldol reactions,7 where the use of performed enolates

    (or their equivalents) is avoided. In general, the asymmetric aldol reactions can be

    categorized into the following five types, (i) chiral auxiliary assisted aldol reaction

    based on the use of stoichiometric quantities of the chiral appendages; (ii) chiral Lewis

    acid and Lewis base catalysed (iii) heterobimetallic bifunctional Lewis acid/Bronsted

  • 166 base catalysed aldol reeactions (iv) Enzyme or antibody catalysed reaction and (v)

    organocatalysis with L-proline or its structural analogs. Out of these methods the

    above three methods (i-iii) have successfully been used with the unmodified carbonyl

    compounds while the two methods (iv-v) require some sort of pre-activation of the

    carbonyl group of the ketones used for successful operations.

    A remarkable success has been achieved on the substrate scope and reaction

    selectivities by the use of so-called enantioselective organocatalytic direct aldol

    processes. The adjective organocatalytic is applied to processes in which reagents and

    catalysts are all small organic molecules containing only C, H, O, N, S, P, and halogen

    atoms.8Therefore, processes in which a metal (including boron and silicon atom) is

    involved should be excluded in strict sense. Furthermore, those processes involving

    high molecular weight organic compounds such as enzymes, antibodies, even

    polymers, and dendrimers, should be excluded.Although according to the above strict

    definition, reactions involving silicon atoms, polymers, and inorganic materials should

    be excluded. We considered that processes where the silyl group has only a steric role

    (or increase the solubility) should be included as well as cases in which the catalyst is

    immobilized in a polymer, dendrimer, or inorganic material, since in these cases the

    matrix does not play any relevant role in the reaction, and is only important in the

    recovery of catalysts.

    The enantioselective aldol reaction is a powerful method for constructing one

    or two successive chiral carbon centers.9 Progress in catalytic, enantioselective aldol

    reaction has been made by chiral Lewis acid-catalyzed reactions of trimethylsilyl enol

    ethers, which act

    as aldol donors and are prepared from parent carbonyl compounds.In the last decades,

    catalytic enantioselective direct aldol reactions, which do not require masked enol

    ethers to be prepared beforehand from ketones or esters (aldol donors), have attracted

    much attention due to their simple manipulation and high atom economy.10 Two types

    of direct aldol reactions have been reported: the reactions catalyzed by chiral metal

    alkoxide complexes, which were initially reported by Shibasaki,11,12 and the reactions

    involving enamine process, which were pioneered by List and Barbas.13,14 Both

    strategies have been extensively investigated, and are commonplace in the

    development of asymmetric reactions.15 Shunsuke et al.16 reported a novel type of

    enantioselective direct aldol reaction of cyclohexanone derivatives (1) and

    benzaldehyde derivatives (2) in high stereoselectivities with tetrachlorosilane and

  • 167 diisopropylethylamine in dichloromethane at rt using BINAPO (BINAP dioxide) as an

    organocatalyst (Scheme 1). Benzaldehyde was slowly added to the mixture containing

    all the other components at rt, 17 which afforded the aldol adduct (3, 4) with good anti-

    selectivity, but in low chemical and optical yields (rt, 4 h, 30% yield, syn/anti = 1/5,

    16% ee (anti)).

    O

    PhCHO

    P

    O

    P

    O

    PhPh

    BINAPO(10 mol%)

    SiCl4iPr2NEt

    syn- anti-

    +

    PhPh

    Ph Ph

    OH OHO O

    +

    1 2 3 4

    Scheme 1

    The direct asymmetric aldol reaction is one of the most important C–C bond-

    forming reactions in nature, and it is catalyzed by aldolase enzymes with excellent

    stereocontrol.18 The enzyme’s ability to control the enantioselectivity of the direct

    aldol reaction has inspired chemists and raised this transformation to prominence in

    the asymmetric assembly of complex natural products.19,20 In particular, the

    development of catalytic stereoselective methods for the asymmetric directed aldol

    reaction has recently been the subject of intense research.21 For example, the

    utilization of organometallic complexes and Lewis bases as catalysts has been highly

    successful for the asymmetric Mukaiyama-type aldol reaction between activated silyl

    enol ethers and aldehydes.22a-g Furthermore, the enantioselective aldol reaction

    between unmodified ketones and aldehydes is catalyzed by chiral organometallic

    complexes.22h-k Another approach to the catalysis of the direct asymmetric aldol

    reaction is the use of aldolase enzymes.18,23 Recently, organocatalysis has experienced

    a renaissance in asymmetric synthesis.24 In this context, proline and its derivatives

    have proved to be the best catalysts for the direct intermolecular asymmetric aldol

    reaction.25,26 Armando et al.27 described the linear amino acid-catalyzed direct

    asymmetric intermolecular aldol reaction; simple amino acids such as alanine, valine,

    isoleucine, aspartate, alanine tetrazole 3 and serine catalyzed the direct catalytic

    asymmetric intermolecular aldol reactions between unmodified ketones and aldehydes

    with excellent stereocontrol and furnished the corresponding aldol products in up to

  • 168 98% yield and with up to 99% ee. They reported the L-alanine catalyzed asymmetric

    aldol reaction between different ketones (5) and acceptor aldehydes (6) affording aldol

    product (7) (Scheme 2).

    L-alanine(30 mol%)

    H2O(10 equiv)

    DMSO,rt3-4 days

    O

    + H R R

    O OH

    R2 R2R1

    O

    R1 5 6 7

    Scheme2

    Maruoka and coworkers28 reported a novel amino acid derived from optically

    pure binaphthol. Various amines, most of which are synthesized from optically pure

    proline, have also been employed in the direct aldol reactions.29a,29b,30 Recently,

    Cheng31 developed a simple primary-tertiary diamine-Brønsted acid catalyst that has

    been successfully applied to direct aldol reactions, and this catalytic system was highly

    efficient for both linear and cyclic aliphatic ketones.Prolinamides have also received

    great success for their easy preparation and high efficiency.29c,29d-29f,32 Zhao et al.33

    reported enantioselective direct aldol reactions catalyzed by prolinamide derived from

    cinchonine and optically pure proline (i-iii). Both cyclic and acyclic ketones (5,8) were

    reacted with various aldehydes (6) furnishing the desired aldol products (9) in up to

    90% yield with excellent enantioselectivities (up to 95%) and moderate

    diastereoselectivities (up to 3.6/1) in the case of cyclic ketones (Scheme 3).

    N

    N

    HH2N

    N

    N

    HHN

    ONH

    N

    N

    HHN

    ONH

    CHO10 mol%cat.10 mol% TFA

    - 18oC

    Cat.= i. ii. iii.

    +R

    OOOH

    6 5,89

    (i-iii)

    Scheme 3

  • 169

    After the discovery of a proline-catalyzed enantioselective intramolecular aldol

    reaction in the 1970s, the corresponding intermolecular enantioselective direct aldol

    reactions between ketones and aldehydes were reported by Barbas et al. in 2000.34a,34b

    Although proline is a rather good catalyst, it is not without some potential drawbacks,

    such as (1) a low solubility that limits its reactivity in typical organic solvents, (2)

    potential side reactions and established parasitic equilibria with substrates,35 and (3)

    low selectivities with planar aromatic aldehydes in direct aldol reactions. Therefore,

    considerable effort has been directed at the development of proline analogs in order to

    improve their reactivity, selectivity and scope. Considering the practical synthetic

    issues, the carboxylic acid moiety of proline has been targeted as a site for

    modification. Its reactivity and selectivity is enhanced in custom-made catalysts, even

    though the identification of a good catalyst in turn requires the synthesis of various

    analogs of a proposed design in order to identify the optimum one. Moreover, the

    improved catalyst is usually obtained through the modification of proline with chiral

    molecules that have additional functionality, and are much more precious than the

    proline itself.36

    ‘‘To make a good asymmetric catalyst perfect’’, the role of suitable additives,

    or co-catalysts, can be crucial in enhancing the reactivity and stereoselectivity of the

    catalytic system. For example, it has been shown that the addition of a small amount

    of water often accelerates the reaction rate and increases the enantioselectivity of

    proline-catalyzed aldol reactions.37b-37i Recently, Shan has shown that using chiral

    diols as additives can improve the enantioselectivity of proline catalyzed aldol

    reactions, probably through their involvement

    in the transition state, via the formation of a hydrogen bonding network.37a Miller et

    al. showed the cooperative effect of a co-catalyst in a proline-catalyzed Baylis–

    Hillmann reaction, where the proline and co-catalyst were proposed to interact in a

    transition state assembly, in turn forming a catalytic system that was ‘‘greater than the

    sum of its parts’’.38

    Prolinecatalyzed direct aldol reactions have been shown experimentally and

    computationally to proceed through enamine intermediates, in which the transition

    state is highly stabilized by hydrogen bond donation from the carboxylic acid moiety

    to the electrophile, with concurrent development of partial iminium and carboxylate

    ions on the proline .39 Moreover, proline is known to exist as a zwitterion, forming a

  • 170 highly insoluble network of hydrogen-bonded units. Omer Reis et.al40 reported a

    proline–thiourea host–guest complex catalyze direct asymmetric aldol reactions of

    various aromatic aldehydes (6) and cyclohexanone (10) in non-polar solvents giving

    product (11) with high diastereo- and enantioselectivities (up to 94 : 6 dr and 499% ee)

    (Scheme 4). O

    proline:thiourea10:10

    Hexane,rt+

    O

    Ar

    OH

    Ar H

    O

    6 10 11

    Scheme 4

    The development of enantioselective reactions in water was long thought to be

    mainly confined to the realm of enzymes. The aldol condensation is a key carbon–

    carbon bond forming reaction, which creates the β-hydroxy carbonyl structural unit

    found in many natural products and drugs.42 In nature, type I and II aldolases catalyze

    this reaction in water with perfect enantiocontrol through an enamine mechanism and

    by using a zinc cofactor, respectively.43 A Zn-proline catalyzed aldol reaction occurs

    in aqueous media with moderate enantiomeric excess.44 However, proline can catalyze

    direct aldol reactions in polar organic solvents with high enantioselectivity,41,45 but it

    affords the racemate in water.46 Although several chiral organocatalysts have been

    developed for the aldol reaction,41,47 and some of them provide aldols

    enantioselectively in aqueous organic solvents,48 they still require the use of an

    organic solvent.49 It was found that only enzymes and antibodies of very high

    molecularweight have been able to catalyze the direct aldol reaction in water with high

    enantioselectivity.42,43 Yujiro et.al50 reported that 4-tert-butyldimethylsiloxyproline,

    which is easily prepared from commercially available trans-4-hydroxyproline, is a

    highly active proline surrogate and used as a catalyst to the asymmetric aldol reaction

    of cyclohexanone (10) and benzaldehyde (2) in the presence of water and the anti-

    aldol product (12) was obtained with excellent diastereoselectivity in a nearly optically

    pure form (Scheme 5). The effectiveness of the siloxyproline catalysts compared to

    proline and hydroxyproline can be attributed to the solubility of the catalysts.

    Although proline and hydroxyproline dissolve in water, siloxyproline is only partially

    soluble in water and forms an organic phase with the aldehyde andketone in which the

    aldol reaction proceeds efficiently.

  • 171

    ONH

    10 mol%

    water,RT 18 h

    synisomer

    anti isomer

    +H

    OO OH

    CHO2H

    THBDPSO

    +

    10 2 12

    Scheme 5

    Evans,51 Heathcock,52 Masamune53 and Mukaiyama54 have established aldol

    reaction as the principal chemical method for the stereoselective construction of

    complex polyol architecture. Recently, studies by Barbas,55 Evans,56 List,57 Shair,58

    Shibasaki,59 and Trost60 have outlined the first examples of enantioselective direct

    aldol reactions, an important class of metal or proline catalyzed transformation that

    does not require the pregeneration of enolates or enolate equivalents.With these

    remarkable advances in place, a fundamental goal for asymmetric aldol technology has

    become the development of catalytic methods that would allow the direct coupling of

    aldehyde substrates.61 R. Ian Storer et.al 62 reported an asymmetric proline catalyzed

    aldol reaction with a-thioacetal aldehydes (13). Thioacetal bearing aldehydes readily

    participate as electrophilic cross-aldol partners with a broad range of aldehyde and

    ketone (6, 5) donors. High levels of reaction efficiency as well as diastereo- and

    enantiocontrol are observed in the production of anti-aldol adduct (14) (Scheme 6).

    L-proline, DMF

    slow addition of donorH

    X

    O

    H

    SR

    SR

    O

    H

    X SR

    SR

    O OH

    +

    6, 5 13 14

    Scheme 6

    Chemically, aldol reaction is dominated by approaches that utilize preformed enolate

    equivalents in combination with a chiral catalyst.63Typically, a metal is involved in the

    reaction mechanism.63d Most enzymes, however, use a fundamentally different

    strategy and catalyze the direct aldolization of two unmodified carbonyl compounds.

    Class I aldolases utilize an enamine based mechanism,64 while Class II aldolases

  • 172 mediate this process by using a zinc cofactor.65 The development of aldolase

    antibodies that use an enamine mechanism and accept hydrophobic organic substrates

    has demonstrated the potential inherent in amine-catalyzed asymmetric aldol

    reactions.66 Recently, the first small-molecule asymmetric class II aldolase mimics

    have been described in the form of zinc, lanthanum, and barium complexes.67,68

    Benjamin et al. 69 reported that the amino acid proline is an effective asymmetric

    catalyst for the direct aldol reaction between unmodified acetone (15) and a variety of

    aldehydes (6) affording aldol adduct (16) with 68% yield and 76% ee (Scheme 7).

    20 vol%

    DMSO

    30 mol%

    68 % (76% ee)

    +

    NH

    COOH

    O

    NO2

    O OH

    NO2

    O

    15 6 16

    Scheme 7

    The Mukaiyama aldol reaction of silyl enol ethers with aldehydes catalyzed by

    Lewis acid transition metals and main group elements chiral complexes is regarded as

    the first efficient approach in the field of asymmetric aldol reaction.70 This strategy

    involves a preliminary transformation of the ketone into a more active silyl enol ether,

    whereas biocatalytic pathways based on aldol reactions between a ketone and an

    aldehyde do not require this chemical activation. Considering a possible interaction

    between Lewis acids, L-proline and the aldehyde generating a more activated catalytic

    system has recently been described by Darbre and co-workers with a Zn (prolinate)2

    complex prepared under basic conditions71a And by Mlynarski and co-workers with

    bis(prolinamides)/zinc(II) complexes.71b Aldol reactions catalyzed by Zn(prolinate)2

    were found to be moderately stereoselective (ee up to 56% and de up to 54%). Bis

    (prolinamides)/zinc (II) complexes proved to be one of the more stereoselective

    catalytic systems for direct aldol reactions. In both cases, the use of water as a

    cosolvent is required in order to obtain a complete solubilization of the complex. The

    enamine pathway in L-proline catalysis is analogous to the one observed in class I

    aldolases and Lewis acid activation is similar to class II aldolases mechanism

    involving a zinc (II) cofactor.

  • 173

    Lewis acid catalysis in the presence of water is not trivial, since most of them

    are decomposed under aqueous conditions. Nevertheless, a few examples of water-

    tolerant Lewis acids have previously been reported by Kobayashi et al.72 Furthermore,

    Lewis acids in aqueous media are known to coordinate to a molecule of water to

    generate metallo-hydroxonium species leading to a nearly neutral pKa value.73 In the

    presence of ligands, those metallohydroxonium species are able to dissociate and

    equilibrate with different ligand–metal complexes. Such dynamic behaviours are

    observed in metallo-enzymes. Moreover, L-proline could activate the ketone via an

    enamine intermediate and at the same time interact with various metal salts to activate

    the aldehyde partner.Furthermore; the nature of the metal could have important effects

    on both stereo- and chemical selectivities. Mael et.al 74 described chloride salts from

    group 75 elements (ZnCl2, CdCl2, HgCl2) based on combinations of various water

    compatible Lewis acids and L-proline co-catalysts has been evaluated for the direct

    asymmetric aldol reaction of cyclohexanone (10) with various aromatic aldehydes (6)

    with optimized catalytic conditions (catalytic system: L-proline: 20%/ZnCl2: 10%;

    solvent mixture: DMSO/H2O, 8:2) gave anti aldol products (17) with improved

    enantioselectivity (>99% ee) along with syn aldol products (18) compared to a

    moderately stereoselective procedure based on proline activation only (Scheme 8).

    O

    NH

    COOH

    (20 %)

    Lewis acid,DMSO/H2O

    (8:2)R.T., 24 h

    anti(1'R,2S) and (1'S,2R)

    syn(1'R,2R) and (1'S,2S)

    +O

    NO2

    OH

    H

    O

    O2N

    O

    NO2

    OH

    +10

    6

    17

    18 Scheme 8

    The enantioselective aldol reaction with small organic molecules in an

    aqueous medium had limited success until recently when Barbas76 and Hayashi77

    independently reported efficient proline-derived chiral catalysts which catalyzed the

    aldol reaction with high enantiocontrol in the presence of a large excess of water.78

  • 174 Most of the studies have been done with 10 mol % catalyst loading except in one

    example where Hayashi has shown that the catalyst loading can be reduced to 1 mol

    %, but at the cost of a longer reaction time (2 days). Therefore, there is a great need

    for efficient chiral organocatalysts, which can work at a lower loading without

    affecting the enantioselectivity and the reaction time. The catalysts should also have a

    wide substrate scope, with respect to both ketones and aldehydes. Vishnu et.al 79

    reported that L-proline-derived organocatalysts iv and v (Scheme 9) are very effective

    organocatalysts which catalyzed the direct aldol reaction of both acyclic and cyclic

    ketones (5, 8) with different aldehydes (6) in an excess of water/brine affording aldol

    product (19) with excellent enantioselectivities up to >99% and diastereoselectivities

    up to 99% with very good yields were obtained by using much lower catalyst loadings

    (0.5 mol %).

    1 or 2 (0.5 mol%),brine,-5oC

    iv: R= i-Buv: R= Ph

    >99% ee

    +

    NH

    NH

    Ph

    OHPh

    O R

    O

    R2

    O OH

    R2 H

    O

    5, 8 6 19

    Scheme 9

    Direct catalytic and enantioselective aldol reactions of unmodified ketones or

    aldehydes were reported by the research groups of Shibasaki,80 Trost,81 Jorgensen,82

    MacMillan,83 List,84 Barbas III85 and Cordova86 using organometallic or purely

    organic catalysts.Recent work has been attempted by using a recyclable ionic liquid as

    the solvent,87 buffered aqueous media,88 Zn-proline complexes in aqueous media or

    aqueous micelles.89 It is always economical if the catalytic reaction is performed in an

    ecofriendly solvent, which allows both solvent and catalyst to recycle. S.

    Chandrasekhar et.al90 described an efficient synthesis of chiral β-hydroxy ketones

    from various aldehydes (6) and acetone (15) in poly (ethylene glycol)-400 catalysed

    by L-proline giving aldol adduct (20) (Scheme 10). They studied the asymmetric aldol

    reaction by using 4-nitrobezaldehyde, acetone and L-proline (10 mol %) in PEG-400.

    The reaction was completed in 30 min and yielded 94% of product with 67% ee.

    Enantiomeric excess was determined using chiralcel OB-H column. After workup

  • 175 (extraction with ether) mother liquor (PEG+proline) was kept aside for further runs.

    The transformation in conventional solvent (DMSO) took 4 h for completion of the

    reaction. Several groups studied the mechanism of L-proline catalysed direct

    asymmetric aldol reaction and proposed an enamine mechanism based on the Hajos–

    Parrish–Eder–Sauer–Wiechert reaction mechanism.91

    L-proline(10 mol%)acetone (4 eq.)

    PEG,r.t.30 minO2N

    H

    O2N

    OO OH

    15

    6 20 Scheme 10

    Catalytic asymmetric aldol reactions of aldehydes with silyl enol ethers (the

    Mukaiyama aldol reaction92) mediated by chiral Lewis acids have been elaborated into

    the most powerful and efficient asymmetric aldol methodology. Tomoaki et al.93

    reported catalytic asymmetric aldol reactions in aqueous media using Pr (OTf)3 and

    chiral bis-pyridino-18-crown-6vi. The binding ability of the crown ether with the RE

    cation and the catalytic activity of the complex are important for attaining high

    selectivity in the asymmetric aldol reaction. Various aromatic and α, β-unsaturated

    aldehydes (6) and silyl enol ethers (21) derived from ketones and a thioester can be

    employed in the catalytic asymmetric aldol reactions using Pr (OTf)3 and vi, to

    provide the aldol adducts (22) in good to high yields and stereoselectivities. In the

    case using the silyl enol ether derived from the thioester, 2, 6-di-tert-butylpyridine

    significantly improved the yields of the aldol adducts (Scheme 11).

  • 176

    vi.(24 mol%)Pr(OTf)320 mol%

    H2O/EtOH=1/90oC,18 h

    90% yieldsyn/anti=90/10

    79% ee

    +

    NO

    O

    O

    ON

    PhPh

    OOH

    HPh

    O OSiMe3

    Ph

    Scheme 11

    Cordova et al., Amedjkouh andTeo et al. achieved excellent stereoselectivities

    with other aminoacids too when the enantioselective aldol reactions were performed in

    ionic liquids or DMSO in the presence of small amounts of water or in

    water,respectively.97-101 In intramolecular aldol reaction phenyl- alanine was even

    more efficient than proline.102, 103 Protonated arginine and lysine also performed well

    in aldol reactions in IL.104 On the other hand,manyderivatives of proline were

    developed and succcessfully applied in aldol reactions.94-96,105 Sadaf et al 106 reported

    on IL-tagging of (S)-proline by 1, 2, 3- triazolium salts and the successful application

    of these IL-tagged organo catalysts in direct aldol reactions of aromatic aldehydes (6)

    with cyclic as well as open chain ketones (5, 8) giving aldol adduct (23) with high

    diastereo- and enantioselectivities. The 1, 2, 3-triazolium tag substituents were limited

    to unbranched alkyl groups (Scheme 12, viia, viib (R=alkyl)).

  • 177

    (20 mol% catalyst) viia, viib, viic

    argon,rt

    viia (R =H)viib (R =Me)

    viic.

    vii=viia,viib,viic

    +

    O

    R1R3

    R2

    OH

    R3

    OO

    R1 R2

    NN

    N

    O

    NH

    OH

    O

    RO

    O

    BF4

    NN

    N

    O

    NH

    OH

    O

    BF4N2H

    O OH

    Scheme 12

    Direct asymmetric catalytic aldol reactions have been successfully performed

    using aldehydes and unmodified ketones together with chiral cyclic secondary amines

    as catalysts.107 L-proline and 5,5-dimethylthiazolidinium-4-carboxylate (DMTC) were

    found to be the most powerful amino acid catalysts for the reaction of both acyclic and

    cyclic ketones as aldol donors with aromatic and aliphatic aldehydes to afford the

    corresponding aldol products with high regio-, diastereo-, and enantioselectivities.

    Reactions employing hydroxyacetone as an aldol donor provide anti-1,2-diols as the

    major product with ee values up to >99%. The observed stereochemistry of the

    products was explained by a metal-free Zimmerman-Traxler-type transition state and

    involves an enamine intermediate.The reactions tolerate a small amount of water (

  • 178 metabolite capable of being used as a catalyst (Scheme 13).109 Various aldehydes and

    ketones including carbohydrate derivatives can be chosen as the substrates. It is

    predicted that the synthetic method will get further attention as a prebiotic route to

    sugar.110 Proline catalyzed Aldol reaction of nitrobenzaldehydes with various ketones

    was investigated in aqueous anionic micelles. Satisfactory reaction yields were

    obtained with SDS (sodium dodecylsulfonate), SDBS (sodium

    dodecylbenzenesulfonate), and SLS (sodium laurylsulfate), all anionic surfactants.

    Other surfactants such as Triton-100, CTAB (cetyltrimethylammonium bromide), and

    OTAC (octadecyltrimethylammonium chloride), which are either neutral or cationic

    surfactants, did not promote the reaction in aqueous media even with proline as

    catalyst.11

    (30 mol%)

    phosphate buffer+

    NO2

    OHO

    NO2

    O

    O

    NH

    5 6 24

    Scheme 13

    Tang et al.112 reported L-Prolinamides (viiia-viiid), prepared from L-proline

    and simple aliphatic and aromatic amines which is a class of organic catalysts, (S)-

    pyrrolidine-2-carboxamides (L-prolinamides) have been found to be active catalysts

    for the direct aldol reaction of 4-nitrobenzaldehyde (6) with neat acetone (15) at room

    temperature. They give aldol adduct (25) with moderate enantioselectivities of up to

    46% enantiomeric excess (ee). The enantioselectivity increases as the amide NOH

    becomes a better hydrogen bond donor (Scheme 14).

  • 179

    20 mol% viiia,viib, viic, viid

    25oC

    NH

    O

    NH2

    NH

    O

    NHCH3

    CH3CH3

    NH

    O

    NH CH2

    NH

    O

    NCH2CH3

    CH2CH3

    viii= a

    b

    c

    d

    +

    OOH

    OH

    O

    H2N6

    15 25

    Scheme 14

    Wolfgang et al.113 reported enamine-based direct aldol reaction of α, α-

    disubstituted aldehydes (26) with aryldehydes (6) by using diamine and trifluroacetic

    acid as additives giving aldol products (27) in excellent yield and very high

    enantiomeric excess (Scheme 15).

    (5 mol%)

    TFA(5 mol%)DMSO,rt

    R= Me,Et,Pr X=NO2,CN,Br

    OHO

    RX

    NH

    N

    OHC

    X

    H

    O

    R

    +

    26 6 27

    Scheme 15

    The direct aldol reaction avoids the pre- formation of silyl enol ethers.

    Recently proline and chiral zinc-proline complexes have been described as an

    enantioselective catalyst in organic synthesis especially in direct aldol addition of

    acetone and a variety of aldehydes, 114 with formation of an iminium ion which is

    converted to the corresponding enamine nucleophile,mimicking the class I

    aldolases.The methods utilizing Lewis acids rely on the catalysis of metal complexes

    bearing chiral ligands ,such as the heterobimetallic LaLi3tris(binaphthoxide) and the

  • 180 Zn-BINOL homobimetallic catalysts developed by Shibasaki 115 as well as Trost’s Zn

    III-semi crown ether.116 The reaction described above have been carried out under

    anhydrous conditions in organic solvents and the metal complexes were reported to be

    water sensitive. Tamis et al.117 tried to develop water soluble Lewis acids having

    unprotected amino acids as chiral lihands and to investigate their activities as catalysts

    for the direct aldol reaction. They reported the aldol reaction of acetone (15) and p-

    nitrobenzaldehyde (6) catalyzed by a Zn-proline complex in the presence of water

    giving aldol product (28) quantitative yields and enantiomeric excesses up to 56%

    with 5 mol% of the catalyst at room temperature. The catalytic ability of Zn-proline

    complexes bearing other amino acids such as lysine and arginine is also reported

    (Scheme 16).

    [(L)-Proline]2Zn5 mol%

    H2O66 vol%

    100% yield,56% ee33 vol%

    +

    NO2

    OHO

    NO2

    H

    O

    O

    15 6 28

    Scheme 16

    In this last chapter the aldol reactions of unmodified ketones with

    aldehydes catalyzed by Cd-proline complex in the presence of water are reported

    at room temperature.

    5.2 RESULTS AND DISSCUSSION The aldol reactions of unmodified ketones (5) with aldehydes (6) catalyzed by

    Cd-proline complex in the presence of water are reported (Scheme 17). At first,

    several substituted benzaldehyde derivatives (6) with acetone (5b) were investigated

    and aldol reaction products (29) have been isolated after simple extraction of the

    reaction mixture.

  • 181

    +

    R2

    R

    OHO

    R

    OCd(Pro)2

    H2O R1

    R1

    CHOR2

    5 6 29

    a: R = CH3, R1 = CH3 a: R2 = NO2 a: R = CH3, R1 = H, R2 = NO2 b: R = CH3, R1 = H b: R2 = Cl b: R = CH3, R1 = H, R2 = Cl

    c: R = R1 = (CH2)4 c: R2 = OCH3 c: R = CH3, R1 =CH3, R2 = NO2 d: R = CH3, R1 = CH3CO d: R2 = Br d: R = CH3, R1 = CH3, R2 = Cl e: R = CH3, R1 = PhCO e: R2 = H e: R = R1 = (CH2)4, R2 = NO2

    f: R = CH3, R1 = CH3CO, R2 = NO2

    g: R = CH3, R1 = CH3CO, R2 = Cl

    h: R = CH3, R1 = PhCO, R2 = NO2

    i: R = CH3, R1 = R2 = H

    j: R = CH3, R1 = H, R2 = OCH3

    k: R = CH3, R1 = H, R2 = Br

    l: R = CH3, R1 =H, R2= 2-Cl (ortho Cl)

    Scheme 17

    Preliminary studies revealed that this aldol reaction was indeed possible to

    provide the aldol adduct in 99% ee. Gratifyingly, the aldol union can be

    comprehensively occurred involving both the enamine and enolate mechanism via the

    slow addition of 4-Nitrobenzaldehyde (6a) to an excess of the acetone, (5b) ketone.

    The results for some of the screening investigations for the enantioselective direct

    aldol condensation of the unmodified ketones 5 and various aldehydes 6 catalyzed by

    Cd-proline complex in the presence of water are presented in Table 1. The Cd-proline

    complex was prepared by adding triethylamine to a mixture of proline (5 mmol) in

    methanol (10 mL); cadmium acetate (2.5 mmol) was then added to the reaction

    mixture after 10 min. After stirring for half an hour, a white precipitate was collected

    by filtration. Thus various aldol products can be synthesized by using Cd-proline as

    catalyst in water medium.

  • 182 Table1. Aldol reaction of unmodified ketones and various aldehydes catalyzed by

    Cd-proline complex

    1 100

    entry Product Reaction timea (h) Yieldb eec

    24

    NO2

    OHO

    2Cl

    OHO

    24 98

    NO2

    OHO

    Me3 26 100

    4

    Cl

    OHO

    Me120 97

    5

    O

    NO2

    OH

    240 75

    6

    Cl

    OH

    COMe

    O

    NO2

    OH

    COMe

    O

    NO2

    OH

    COPh

    O

    Br

    OHOOMe

    OHO

    OHO

    19285

    7 168 82

    8 120 92

    9

    10

    11

    > 99

    83

    88 65

    8324

    32

    96 7263

    65

    > 99

    56

    58

    72

    48

    12OHO Cl

    24 95 85

    > 99

  • 183 The aldol addition product, 4-Hydroxy-4-nitrophenyl)-butan-2-one, (29a) was

    prepared by stirring acetone (5 ml) with 4-nitrobenzaldehyde (1 m mol,0.1512 g) in

    the presence of cadmium-proline complex (50 µ mol,0.016 g) in water (10 ml) at room

    temperature for 24 hours with quantitative yield and 78% ee.The structure of (29a)

    was assigned assigned by the presence of hydroxyl group band at 3292.60 cm-1, the

    presence of aromatic C-H stretching band at 3100-3000 cm-1, the presence of carbony

    stretching band at 1716 cm-1 , the presence of aromatic asymmetric and symmetric N-

    O stretching band in the range of 1524-1360 cm-1 in its IR spectrum. It was further

    confirmed by the presence of broad singlet at δ 3.70 due to O-H protons in its 1 H

    NMR spectrum, presence of singlet at δ 2.22 due to the presence of 3H protons due to

    methyl protons and also the presence of mutiplet signals at δ 2.90-2.81 due methylene

    protons in its 1H NMR spectrum. It was further confirmed by its 13C NMR spectrum

    due to the presence of carbon atoms at δ 208.55, 150.04, 147.28, 126.43, 123.76,

    68.89, 51.51, 30.72. The enantiomeric excess was determined by using chiral HPLC

    using chiral column-ChiraDex and found to be 80 % ee.

    Fig.1. HPLC chromatogram of compound (29a)

  • 184

    When 4-Chlorobenzaldehyde (0.5 m mol, 0.07g) (6b) was made to stir with

    acetone (5b) (5 ml) in the presence of Cadmium-proline catalyst (50 µmol,0.016 g) in

    water medium (10 ml) for 24 hours at room temperature giving the aldol addition

    product , 4-Hydroxy-4-(4- Chlorophenyl)-butan-2-one, (29b). The structure of

    compound 29b was determined by the presence of hydroxyl band at 3338cm-1,

    aromatic C-H band at 3011, 2901cm-1, also the band at 1716 due to the presence of

    carbonyl and also at 1417-835 cm-1 due to aromatic asymmetric and symmetric N-O

    stretching in its IR spectrum. Its structure was further assigned by its 1H NMR

    spectrum due to the presence of multiplet signals due to methine protons at δ 5.15-5.12

    and also by the presence of broad single signal at δ 3.40 due to –OH protons. The

    compound, (29b) was further confirmed by its 13C NMR and Mass spectral data. Its

    purity was also confirmed by HPLC using chiral column –Lichro Cart 250-4, Chira

    Dex and optical rotation by measuring on an Autopol II, serial number 30415.

    Fig.2. 1H NMR spectrum of compound (29b)

  • 185

    Fig.3. 13C NMR spectrum of compound (29b)

    When 4-Nitrobenzaldehyde (0.0756g, 0.5 mmol) (6a) was made to stir with 2-

    butanone (10 ml) (5a) in the presence of cadmium-proline complex (0.016g, 50µmol)

    in water medium for 26 hours giving 4-Hydroxy-3-methyl-(4’Nitrophenyl)-butan-2-

    one, (29c). Its structure was assigned by the presence of hydroxyl band at 3398cm-1,

    the presence of aromatic C-H stretching band at 3112, 2923 cm-1, band at 1701cm-1

    due to carbonyl stretching and also the band in the range of 1419-869 cm-1 due to the

    presence of aromatic asymmetric and symmetric stretching due to N-H band in its IR

    spectrum. The compound was further confirmed by its 1H NMR spectral data by

    showing signals at δ 8.21, 7.51, 3.41, 2.22, 0.99 due to aromatic protons, -OH protons

    and two methy protons. Its structure was further assingned by its 13C NMR, Mass

    spectral data, optical rotation and also the purity of the compound was confirmed by

    HPLC using chiral column –Lichro Cart 250-4, Chira Dex.

  • 186

    Fig.4. IR spectrum of compound (29c)

    Fig.5. 1H NMR spectrum of compound (29c)

  • 187

    Fig.6. 13 C NMR spectrum of compound (29c)

    The aldol addition product,4-Hydroxy-3-methyl-(4-Chlorophenyl)-butan-2-one

    ,(29d) was prepared by stirring 2-Butanone (5a) ( 5 ml) with 4-Chlorobenzaldehyde

    (6b) (0.5 mmol, 0.07 g) in 10 ml of water medium for 120 hours .Its structure was

    confirmed by its IR spectral data by showing bands at 3497 cm-1 due to –OH

    stretching , 2974 cm-1 due to aromatic C-H stretching , the presence of carbonyl group

    is shown by showing band at 1711 cm-1. The structure of compoung (29d) was further

    confirmed by the presence of broad singlet at δ 3.24 due to –OH protons and two

    singlets at δ 2.17 and δ1.07 due to the presence of two methyl protons in its 1H NMR

    spectrum.Its structure was further confirmed by its 13C NMR and mass spectral data

    and also the purity of the compound (29d) by HPLC by using chiral column –Lichro

    Cart 250-4, Chira Dex.

  • 188

    Fig.7. IR spectrum of compound (29d)

    Fig.8. 1H NMR spectrum of compound (29d)

  • 189

    Fig.9.NMR spectrum of compound (29d)

    Similary , the aldol addition product , (29e), 2-((R-Hydroxy(4-

    Nitrophenyl)methyl) cyclohexanone (5c) was prepared by reacting 4-

    Nitrobenzaldehyde (6a) (0.5 mmol, 0.076g) with cyclohexanone (3 ml) in the presence

    of cadmium-proline complex (50 µmol, 0.016g) in 6 ml of water for 240 hours at room

    temperature . The compound (29e) was obtained as white crystalline solid and its

    structure was assigned by the presence of hydroxyl group at 3506 cm-1, the presence of

    aromatic C-H stretching at 2949 cm-1 and also the presence of carbonyl stretching at

    1693 cm-1, the presence of aromatic asymmetric and symmetric C-N stretching band in

    the range of 1344-800 cm-1 in its IR spectrum. Its structure was further confirmed by

    its 1H NMR, 13C NMR and mass spectral data. The compound (29e) was further

    purified by calculating the diastereomeric anti/syn ratio from its 1H NMR analysis of

    the crude sample: δ 5.48(d, J=1.8 Hz, 1H, syn, minor), 4.89 (d, J=8.8Hz, 1H, anti,

    major). Enantiomeric excess was determined by HPLC with a Chiralcel column (n-

    hexane/i-PrOH 90/10, 1.0 mL/min, λ = 254 nm, 25oC); tR= 26.3 min (minor) and 34.9

    min (major).

    When 4-Nitrobenzaldehyde (6a) (0.5 mmol, 0.0756g) was reacted with

    acethylacetone (5d) (5mL) in the presence of Cadmium-proline complex ( 50µmol,

    0.016g) in 10 mL of water afforded the aldol addition product, 4-Hydroxy-3-Acetyl-

  • 190 (4’Nitrophenyl)-butan-2-one, (29f). The compound (29f) was assigned by the presence

    of hydroxyl group at 3500 cm-1, the presence of aromatic C-H stretching band at 2974,

    2910 cm-1, the presence of carbonyl stretching band at 1695 cm-1 and also the presence

    of aromatic asymmetric and symmetric N-O stretching band in the range of 1348 -798

    cm-1 in its IR spectrum. Its structure was further assigned by the presence of a singlet

    at δ 3.31 due to the presence of O-H protons and also presence of two singlets at δ1.76

    and δ 1.22 due to the presence of two methyl protons in its 1H NMR spectrum. Its

    structure was further assigned by its 13C NMR and mass spectral data. Its purity was

    also confirmed by HPLC using chiral column –Lichro Cart 250-4, Chira Dex and

    optical rotation by measuring on an Autopol II, serial number 30415.

    Similarly, all other aldol addition compounds from (29f-29l) were prepared

    and their structures were confirmed by their IR, 1H NMR, 13C NMR, Mass spectral

    data and HPLC by using chiral column.The 4-chlorobenzaldehyde (entries 2, 4 and 7)

    aldol reaction products with different ketones were found to obtain in 82-98% yields

    but lower enantioselectivities (50-72% ee) compared to that of 4-nitrobenzaldehyde

    (except entry 8). The benzaldehyde (entry 9) aldol reaction product with acetone was

    obtained in 83% yield and 78% ee which was found to be of better reaction as

    compared with the literature.118 We are also able to improve the yield and

    enantioselectivity by using the Cd-proline complex for the asymmetric aldol reaction

    of p-anisaldehyde with acetone (entry 9) compared to results obtained by using ionic

    liquid.119 However, we could not improve the enantioselectivity of the aldol products

    (entries 7 and 8), although the yields are relatively high. Thus a variety of substituted

    benzaldehydes (6) and unmodified ketones (5) were employed and aldol products (29)

    were in good yields (72-100%) and reasonable enantioselectivities (48-83% ee).

  • 191

    Fig.10. IR spectrum of compound (29l)

    Fig11. 1H NMR spectra of compound (29l)

  • 192

    Fig.12. 13C NMR spectrum of compound (29l)

    The mechanistic paths for the Cd-proline catalysis may be assumed to occur

    through both enamine and enolate type mechanisms ((Scheme 18).120

    N

    OO

    Cd

    H

    2+

    OO

    NH

    N

    OO

    Cd

    H

    2+

    OO

    NH

    ON

    OO

    Cd 2+

    OO

    NH

    O

    O

    Cd 2+

    OO

    NH

    N

    O

    OH

    O

    Cd 2+

    OO

    NH

    N

    O

    OH2

    RCHO

    si- Facial attack

    O

    Cd 2+

    OO

    NH

    N

    O

    OH

    R

    OH

    N

    OO

    Cd

    H

    2+

    OO

    NHR

    OHO

    +

    O

    Cd 2+

    OO

    NH

    OHN

    O

    HO

    RH

    Scheme 18

  • 193

    5.3 Experimental 5.3.1 General

    Organic solutions were concentrated under reduced pressure on a Buchi rotary

    evaporator. Chromatographic purification of products was accomplished using column

    chromatography using silica gel 60-120 mesh size. Thin-layer chromatography (TLC)

    was performed on glass Plates using silica gel –G. Visualization of the developed

    chromatograms was performed by ultraviolet irradiation (254 nm) or stained using

    iodine vapours, alkaline potassium permanganate solution or 2, 4-

    dinitrophenylhydrazine solution. The IR spectra were recorded as KBr pellets on FT–

    IR Shimazdu IR-408 spectrometer. Absorption maxima were recorded in wave

    numbers (cm-1). 1H NMR spectra were recorded on Bruker AC-400 spectrometers.

    Residual non-deuterated solvent was used as an internal reference and all chemical

    shifts (δ H and δ C) are quoted in parts per million (ppm) downfield from

    tetramethylsilane (TMS). Mass spectra were recorded on a Kratas concept-IS mass

    spectrometer couples to a Mach 3 data system, or on a Jeol-D 300 mass

    spectrometer.Chemical shifts (δ) are given in parts per million (ppm), and coupling

    constants (J) are given in Hertz (Hz). The proton spectra are reported as follows:

    chemical shift (d/ppm) multiplicity (s=singlet, d=doublet, t=triplet, q=quartet,

    p=pentet, sept=septet, m=multiplet), coupling constant (J/Hz), number of protons,

    assignment).13C NMR spectra were recorded in CDCl3 at ambient temperature on

    Bruker AC-100MHz spectrometer at 75 MHz, with the central peak of CHCl3 as 13

    the internal reference (dC=77.3 ppm). Data for 13C NMR are reported in terms of

    chemical shift. Where a compound has been characterized as an inseparable mixture of

    diastereoisomers, the NMR data for the major isomer has been reported. Optical

    rotations were measured on an Autopol II, serial number 30415, manufactured by

    Rudolph Research Analytical, Automatic Polarimeter equipped with a Sodium lamp

    (589 nm) and a 10 cm microcell. The purity of compounds were performed by reverse

    phase HPLC (Merck Hitachi) using C-18 column and a UV Detector L-2400, pump L-

    2130 (Merck). Compounds prepared and used subsequently without further

    purification were judged to be of suitable purity by NMR analysis. The mobile phase

    used for HPLC was a mixture of Methanol: Water in the ratio of 40:60 at oven

    temperature 20oC.The working flow rate was 0.8 ml/min.High performance liquid

    chromatography (HPLC) was also performed on Hewlett-Packard 1100 Series

    chromatographs using a Chiralcel AD column (25 cm) and AD guard (5 cm), a

  • 194 Chiralcel OJ column (25 cm) and OJ guard (5 cm), a Chiralcel AS column (25 cm)

    and AS guard (5 cm), or a Chiralcel ODH column (25 cm) and ODH guard (5 cm) as

    noted. Syringe pump additions were made using a 10 syringe parallel pump (in all

    cases the syringe needle tip was submerged below the surface of the liquid in the

    receiver vessel to ensure continuous mixing).

    All the commercial chemicals were distilled before used.

    5.3.2 Synthesis of Cd-proline complex The Cd-proline complex was prepared by adding triethylamine (0.7 mL) to a

    mixture of L-proline (0.58g, 5 mmol) in methanol (10mL), followed after 10 min by

    cadmium acetate (0.67g,2.5 mmol).After stirring for 45 minutes a white precipitate

    was collected by filtration.The complex was thus obtained as a white amorphous

    compound. The presence of –OH stretching band is shown at 3265.59 cm-1 and also

    bands at 3198.08 cm-1,2960.83 cm-1 due to C-H stretching and the presence of

    carboxylate ion is shown by the presebce of strong bant at 1566.25 cm-1 and weak

    band at 1431.23 cm-1 in its IR specrtrum. The presence of N-H proton is shown by the

    presence of singlet at δ 2.758 and also the preence of CH2 protons at δ 3.72-3.69, δ

    3.04-3.02, 2.14-2.09, 1.71-1.59, 1.64-4.46 in its 1H NMR spectrum. It was further

    confirmed by the presence of peaks at δ 25.69, 30.12, 47.82, 60.66 due to carbon

    atoms in its 13C NMR spectrum.

    5.3.3 Synthesis of 4-Hydroxy-4- nitrophenyl)-butan-2-one, 29a Yellow liquid.Yield = 100%. IR (KBr, cm-1): 3580, 3292, 3100-3000, 1716,

    1523-858. 1H NMR (300 MHz, CDCl3): δ 8.20 (d, J=7.0 Hz, 2H, Ar-H), 7.52 (d,

    J=7.0 Hz, 2H, Ar-H), 5.28-5.22 (m, 1H, -CH-OH), 3.70 (br, s, 1H, -OH), 2.90-2.81

    (m, 2H, -CH2CO), 2.22 (s, 3H,-COCH3): 13 C NMR: (75MHz, CDCl3): δ 208.55,

    150.04, 147.28, 126.43, 123.75, 68.89, 51.51, 30.72; Mass (EI): m/z 209 (M+), 43 ;

    Optical rotation = -0.010 which was measured on an Autopol II,serial number

    30415,manufactured by Rudolph Research Analytical , Hackettstown , NJ , USA.

    Enantiomeric excess : > 99% , which was determined by HPLC analysis using Lichro

    Cart 250-4 Chira dex column ( methanol/water 40/60) UV-VIS 254 nm , flow rate 0.8

    mL/min: major isomer, tR 9.46 min and minor isomer ,tR 11.25 min.

    5.3.4 Synthesis of 4-Hydroxy-4-(4- Chlorophenyl)-butan-2-one, 29b Colourless liquid.Yield = 98%. IR (KBr, cm-1): 3338, 3011, 2901, 1716, 1417-

    835; 1H NMR ( 300 MHz, CDCl3) : δ 7.333 (dd, 1H, Ar-H), 7.278 (dd, 1H, Ar-H),

    5.14-5.12 ( m, 1H, -CHOH ), 3.403 (br, s, 1H, -OH), 2.83-2.78 ( m, 2H,-CH2CO ),

  • 195 2.20 (s, 3H,-COCH3) ; 13 C NMR (75 MHz, CDCl3) : δ 209.18, 141.33,133.53, 128.87

    , 127.21 , 69.35 , 51.97 , 30.95; Mass (EI) : m/z 198 (M+), 43: Optical rotation

    measured at 589 nm = -0.010 which was which was measured on an Autopol II , serial

    number 30415 , manufactured by Rudolph Research.

    5.3.5 Synthesis of 4-Hydroxy-3-methyl-(4-Nitrophenyl)-butan-2-one,

    29c White crystalline solid compound.Yield = 100 % ; IR (KBr, cm-1): 3398,

    3113, 2924, 1701and 1342-869; 1H NMR ( 300 MHz , CDCl3 ) : δ 8.21 ( d, J=8.3 Hz,

    2H) , 7.51 ( d, J= 8.3 Hz , 2H ), 4.87-4.84 ( m, anti-0.72 H), 3.41 (d, J=4.8 Hz, 1H),

    2.93-2.86 ( m, 1H), 2.05 (s, 3H), 0.98 (d, J=7.4 Hz, 3H): 13C NMR ( 75 MHz, CDCl3) :

    δ 213.09, 127.65, 123.82, 149.47, 147.67, 53.43, 30.28, 14.22; ee = 83%. (anti), the ee

    value of compound 29c was determined by HPLC analysis using a DIACEL

    CHIRALPAK AS Column ( hexane / i-PrOH , 90:10 , λmax 280 nm , flow rate=2.0

    mL/ min), tR=16.2 min (minor) and 22.4 min (major). Due to the small amount of the

    syn , the enantiomeric excess could not be determined.

    5.3.6 Synthesis of 4-Hydroxy-3-methyl-(4-Chlorophenyl)-butan-2-one,

    29d Colourless liquid.Yield = 97%. IR (KBr, cm-1): 3496, 2974, 2933, 1710, 1355-

    825. 1H NMR ( 300 MHz, CDCl3) : δ 7.83- 7.21 ( m , 5 H , Ar-H ) , 5.14- 5.09 ( m ,1H

    ,-CH-OH ), 3.24 ( br , s, -OH ) , 2.82-2.75 ( m, 1H , -CH-CO ) , 2.12 ( s , 3H , CH3CO

    ) , 1.07 ( m , 3H , CH3CH ) ; 13C NMR ( 75 MHz , CDCl3) :δ 213.60 , 190.95 , 141.35

    , 1330.01-127.03 , 72.17 , 69.30 , 52.89 , 36.85 , 29.35 , 9.89 ; ee = > 99 % . The

    enantiomeric excess was determined by HPLC using chiral column –Lichro Cart 250-

    4, Chira Dex. ( Methanol / Water , 40 : 60 , λmax 254 nm , flow rate = 0.8 ml/min ) , tR

    = 3.99 min ( major ) and minor very small peak negligible.

    5.3.7 Synthesis of 2-(R-Hydroxy (4-Nitrophenyl) methyl)

    cyclohexanone, 29e White amorphous solid compound.Yield = 75 %. ; mp = 129-130 oC; IR (KBr,

    cm -1): 3506, 2948, 1693, 1344-856. 1H NMR (300 MHz, CDCl3): δ 8.20 (d, J=8.6 Hz,

    2H), 7.50(d, J=8.7 Hz, 2H), 4.89 (dd, J=8.4 Hz, 1H), 4.04 (s, 1H), 2.63-2.33 (m, 2H),

    2.12-1.36 (m, 6H). 13C NMR ( 75 MHz, CDCl3): δ 214.8, 148.4, 147.6, 127.9, 123.5,

    74.0, 57.2, 42.7, 30.8, 27.7, 24.7 ppm The diastereomeric anti/syn ratio was

    determined by 1H NMR analysis of the crude product : δ 5.48 ( d, J=1.8 Hz, 1H, syn,

  • 196 minor), 4.89 ( d, J=8.8 Hz, 1H, anti, major ). Enantiomeric excess was determined by

    HPLC with a Chiracel AD column (n-hexane/ i-PrOH 90/10, 1.0 mL/min, λ = 254 nm,

    25oC): tR =26.3 min (minor) and 34.9 min (major). Optical rotation= - 0.012 which

    was measured at λ= 589 nm on an Autopol II, serial number 30415, manufactured by

    Rudolph Research Analytical, Hackettstown.

    5.3.8 Synthesis of 4-Hydroxy-3-Acetyl-(4-Nitrophenyl)-butan-2-one,

    29f White amorphous solid compound.Yield = 85 %. IR (KBr, cm-1):3500, 2974,

    2910, 1695.31, 1348– 798; 1H NMR (300 MHz, CDCl3):δ 8.20 ( d , J= 7.0 Hz, 2H, Ar-

    H), 7.48(d, J= 7.0 Hz, 2H, Ar-H ), 4.302 -4.135 ( m , 1H , CH-Ph), 3.340- 3.256 ( m,

    1H, -OH), 2.85-2.77( d, 1H, -CH-CO), 1.75( s, 3H, CH3), 1.254(s, 3H, CH3); 13C

    NMR(75 MHz , CDCl3): δ 214.43, 209.38, 151.99, 128.99-124.42, 108.93, 69.24,

    63.23, 44.31, 34.95, 27.05; ee = 58% which was determined by chiral HPLC using

    Chira dex column.

    5.3.9 Synthesis of 4-Hydroxy-3-Acetyl-(4-Chlorophenyl)-butan-2-one,

    29g White amorphous solid compound.Yield = 82 %. IR ( KBr, cm-1 ): 3412, 2972,

    1718, 1695; 1H NMR(300 MHz, CDCl3): δ 7.32 - 7.24( m, 3H, Ar-H); 7.18(dd, 1H,

    Ar-H); δ 7.09(dd, 1 H, Ar-H), 4.09-3.86 (m, 1H, -CH-Ph ), 3.45(s, 1H, -OH), 2.8 -

    2.43(m, 1H, -CHCO), 1.67(m,3H,CH3), 1.39 (m, 3H, CH3): 13C NMR(75 MHz,

    CDCl3): δ 215.53, 203.56, 179.27, 142.59, 136.91 -128.99, 73.88, 69.20, 63.35, 53.61,

    44.14, 34.58, 30.65, 27.84. ee=72 %; The enantiomeric excess was determined by

    HPLC using chiral column –Lichro Cart 250-4, Chira Dex. (Methanol/Water, 40:60,

    λmax 254 nm, flow rate = 0.8 ml/min), tR = 2.23 min (major) and tR= 4.01 min (minor).

    5.3.10 Synthesis of 4-Hydroxy-3-Benzoyl-(4-Nitrophenyl)-butan-2-

    one, 29h White amorphous solid compoundYield = 92 %. IR (KBr, cm-1): 3439, 3049,

    1678, 1659, 1234-914; 1H NMR (300 MHz, CDCl3): δ 8.10(d, J= 6.7 Hz, 2H, Ar-H),

    7.90(dd, J=7.0 Hz, 2H, Ar-H), δ 7.59 -7.45 (m, 5H, Ar-H), δ 2.41(s, 3H, CH3), δ

    1.61(s, 3H, CH3); 13C NMR (75 MHz, CDCl3): δ 196.98, 195.08, 148.21, 142.67,

    139.20, 137.71, 135.44, 134.77, 130.65, 129.24, 129.24, 129.17, 123.95, 27.65; ee

    =>99 %. The enantiomeric excess was determined by HPLC using chiral column –

  • 197 Lichro Cart 250-4, Chira Dex. (Methanol/Water, 40:60, λ = 254 nm, flow rate=0.8

    ml/min), tR= 3.98 min (major) associated with a very small negligible peak for minor.

    5.3.11 Synthesis of (4R)-Hydroxy-4-phenyl-butan-2-one, 29i Colourless oil.Yield = 83%. 1H NMR (300 MHz, CDCl3): d 7.33–7.17(m, 5H,

    Ar-H), 5.15–5.04 (m, 1H, –CHOH), 3.17 (br s, 1 H,–OH), 2.80–2.75 (m, 2H, –

    CH2CO), 2.17 (s, 3H, –COCH3); Mass(EI): m/z 164(MC), 43; IR(neat): 3413, 2932,

    1718, 1450, 890cm K1; [a]D25 C60.0 (c 1, CHCl3) for 83% ee [lit.value].

    5.3.12 Synthesis of (4R)-Hydroxy-4-(4-methoxyphenyl)-butan-2-one,

    29j Colourless oil.Yield = 88%; 1H NMR (300 MHz, CDCl3): d 7.27 (d, J=8.8 Hz,

    2H, Ar-H), 6.88 (d, J=8.8 Hz, 2H, Ar-H), 5.10 (dd, J=9.0, 3.3 Hz, 1H, –CHOH), 3.80

    (s, 3H,–OCH3), 3.22 (br s, 1H, –OH), 2.86–2.78 (m, 2H,–CH2CO), 2.19 (s, 3H, –

    COCH3); Mass (EI): m/z 194 (MC), 43; IR (neat): 3424, 2917, 1730, 1450, 1100, 1070

    cmK1; ee = 65%. The enantiomeric excess was determined by HPLC using chiral

    column –Lichro Cart 250-4, Chira Dex. (Methanol / Water, 50:50, λ = 254 nm, flow

    rate = 0.8 ml/min).

    5.3.13 Synthesis of (4R)-Hydroxy-4-(4-bromophenyl)-butan-2-one,

    29k Colourless oil.Yield = 72 %. 1H NMR (300 MHz, CDCl3): d 7.47(d, J=8.4 Hz,

    2H, Ar-H), 7.22 (d, J=8.4 Hz, 2H, Ar-H),5.08 (dd, J=5.6, 7.8 Hz, 1H, –CHOH), 3.38

    (br s, 1H,–OH), 2.80–2.70 (m, 2H, –CH2CO), 2.20 (s, 3H, –COCH3); Mass (EI): m/z

    243 (MC), 43; IR (neat): 3418, 2934, 1713,1489, 1369, 1077, 538 cmK1; [α] 25D C53.3

    (c 1, CHCl3) for 90% ee [lit. value]. Enantiomeric excess: 63%. The enantiomeric

    excess was determined by HPLC using chiral column –Lichro Cart 250-4, Chira Dex.

    (Methanol / Water, 50:50, λ = 254 nm, flow rate = 0.8 ml/min), (compared with the

    literature value).

    5.3.14 Synthesis of (4R)-Hydroxy-4-(2-Chlorophenyl)-butan-2-one,

    29l Colourless oil.Yield = 95%. IR (KBr, cm-1): 3492, 3064, 2916, and 1708,

    1429-948. 1H NMR ( 300 MHz, CDCl3): δ 7.61 (d, J=8.6 Hz, 1H, Ar-H), 7.32–7.19

    (m, 3H, Ar-H), 5.49 (d, J=10.4 Hz, 1H, -CH-OH ), 3.64 (s,1H, -OH ), 2.71-2.65 (m,

    1H, -CH2OH ), 2.21 (s, 3H, -COCH3); 13C NMR (75 MHz, CDCl3): δ 209.31, 140.08,

    131.08,129.38-127.05, 66.56, 50.01, 50.61.[α] 25D = + 97.0 (c 1,CHCl3) for 85% ee [lit

  • 198 value].ee., which was determined by optical rotation ( compared with the literature

    value ).

    5.4 Conclusion In summary, we have concluded an enantioselective direct aldol reaction which

    enables flexible design of asymmetric chiral organometallic catalyst- Cadmium-

    proline complex based on proline architectures, prepared by reacting L-proline ( 0.58

    g, 5 mmol ) with cadmium-acetate ( 0.67g, 2.5 mmol) in the presence of triethylamine

    ( 0.7 mL ) in methanol ( 10 mL ) for 45 minutes at room temperature. This catalyst

    was found to promote the aldol reactions of unmodified ketones with various

    aldehydes in water medium, giving aldol products in excellent yield and ee. Since this

    kind of reaction is easily carried out in water, it is notable that the reaction is

    developed on the basis of Green Chemistry protocol. Green Chemistry that possesses

    the spirit of sustainable development was booming in the 1990s, 1and has attracted

    more and more interest in the 21st century. Large amounts of organic solvents are used

    in chemical processes, many of which are volatile, flammable, and toxic. The use of

    non-hazardous and renewable materials is one of the most important goals of Green

    Chemistry. With the increasing concerns about the environmental protection;

    development of direct aldol reaction in aqueous media through a Green Chemistry

    procedure is desirable.

    The Cadmium-proline-catalyzed asymmetric aldol reaction was proposed to

    proceed in the enamine mechanism, where imine functionality converts the aldol

    donor to enamine, whereas the carboxylic acid group provides a hydrogen bond to the

    acceptor.121 The enamine attacks the carbonyl group of aldehyde to generate the

    transition state, whose stereochemistry is stabilized by the chiral carboxyl and leads to

    an enantioselectivity. In the reaction of acetone with 4-nitrobenzaldehyde in aqueous

    media, the reactions with Cadmium-proline typically lead to S-aldol, whereas the

    reaction with L-proline provides R-aldol product. Hence, direct aldolization process

    using water soluble chiral asymmetric catalyst, Cadmium-proline complex were found

    to be atom economic, and thus they serve as attractive approaches for the synthesis of

    versatile β-hydroxy carbonyl compounds.

    5.5 References:

  • 199 1. Kane, R. Ann. Phys. Chem., Ser. 2 1838, 44, 475; (b) Wurtz, A. Bull. Soc. Chim.

    Fr. 1872, 17, 436–442.

    2. Modern Aldol Reactions; Mahrwald, R., Ed.; Wiley-VCH: Weinheim, 2004;

    Vols. 1–2.

    3. Trost, B. M. Science 1991, 254, 1471–1477; (b) Sheldon,R. A. Pure Appl.

    Chem. 2000, 72, 1233–1246; (c) Trost, B.M. Acc. Chem. Res. 2002, 35, 695–

    705; (d) Guillena, G.;Ramo n, D. J.; Yus, M. Angew. Chem., Int. Ed. 2007, 46,

    2358–2364.

    4. For recent reviews, see: (a) Silvestri, M. G.; Desantis, G.; Mitchell, M.; Wong,

    C.-H. Top. Stereochem. 2003, 23, 267–342; (b) Sukumaran, J.; Hanefeld, U.

    Chem. Soc. Rev. 2005,34, 530–542; (c) Samland, A. K.; Sprenger, G. A.

    Appl.Microbiol. Biotechnol. 2006, 71, 253–264.

    5. Reymond, J.-L. J. Mol. Catal. B: Enzym.1998, 5, 331–337; (b) Hertweck, C. J.

    Prakt. Chem. 2000, 342, 832–835;(c) Zhu, X.; Tanaka, F.; Hu, Y.; Heine, A.;

    Fuller, R.;Zhong, G.; Olson, A. J.; Lerner, R. A.; Barbas, C. F., III;Wilson, I. A.

    J. Mol. Biol. 2004, 343, 1269–1280.

    6. See for example: Carreira, E. M. In Comprehensive Asymmetric Catalysis;

    Jacobsen, E. N., Pfaltz, A., Yamamoto, H., Eds.; Springer: Berlin, 1999; Vol. 3,

    pp 997–1065.

    7. Alcaide, B.; Almendros, P. Eur. J. Org. Chem. 2002, 1595–1601; (b) Alcaide,

    B.; Almendros, P. Angew. Chem., Int. Ed. 2003, 42, 858–860.

    8. For reviews covering different aspects of enantioselective organocatalytic

    reactions, see: (a) Dalko, P. I.; Moisan, L.Angew. Chem., Int. Ed. 2001, 40,

    3726–3748; (b) Jarvo, E.R.; Miller, S. J. Tetrahedron 2002, 58, 2481–2495; (c)

    Berkessel, A. Curr. Opin. Chem. Biol. 2003, 7, 409–419; (d) Gro ger, H.;

    Wilken, J.; Berkessel, A. In Organic Synthesis Highlights V; Wiley-VCH:

    Weinheim, 2003; pp 178–186; (e)Schreiner, P. R. Chem. Soc. Rev. 2003, 32,

    289–296; (f)Oestreich, M. Nachr. Chem. 2004, 52, 35–38; (g) Pihko, P.M.

    Angew. Chem., Int. Ed. 2004, 43, 2062–2064; (h) Miller,S. J. Acc. Chem. Res.

    2004, 37, 601–610; (i) Bolm, C.;Rantanen, T.; Schiffers, I.; Zani, L. Angew.

    Chem., Int. Ed.2005, 44, 1758–1763; (j) Berkessel, A.; Groger, H. Asymmetric

    Organocatalysis—From Biomimetic Concepts to Applications in Asymmetric

    Synthesis; Wiley-VCH: Weinheim, 2005; (k) Palomo, C.; Mielgo, A. Angew.

    Chem., Int.Ed. 2006, 45, 7876–7880; (l) Taylor, M. S.; Jacobsen, E. N.Angew.

  • 200

    Chem., Int. Ed. 2006, 45, 1520–1543; (m) Lelais, G.; MacMillan, D. W. C.

    Aldrichim. Acta 2006, 39, 79–87; (n)Marcelli, T.; van Maarseveen, J. H.;

    Hiemstra, H. Angew.Chem., Int. Ed. 2006, 45, 7496–7504; (o) Marigo,

    M.;Jorgensen, K. A. Chem. Commun. 2006, 2001–2011; (p)Guillena, G.;

    Ramon, D. J. Tetrahedron: Asymmetry 2006,17, 1465–1492; (q) Gaunt, M. J.;

    Johansson, C. C. C.;McNally, A.; Vo, N. T. Drug Discovery Today 2007, 12, 8–

    27; (r) Rueping, M. Nachr. Chem. 2007, 55, 35–37; (s) Enders, D.; Grondal, C.;

    Hu ttl, R. M. Angew. Chem., Int.Ed. 2007, 46, 1570–1581; (t) Enantioselective

    Organocatalysis; Dalko, P. I., Ed.; Wiley-VCH: Weinheim, 2007; (u)

    Guillena,G.; Ramo´ n, D. J.; Yus, M. Tetrahedron: Asymmetry 2007, 18, 693–

    700; (v) Pellissier, H. Tetrahedron 2007, 63,9267–9331.

    9. For recent reviews on enantioselective aldol reaction, see: (a) Nelson, S.

    G.Tetrahedron: Asymmetry 1998, 9, 357–389; (b) Groger, H.; Vogl, E. M.;

    Shibasaki,M. Chem. Eur. J. 1998, 4, 1137–1141; (c) Mahrwald, R. Chem. Rev.

    1999, 99,1095–1120; (d) Machajewski, T. D.; Wong, C.-H. Angew. Chem., Int.

    Ed. 2000, 39, 1352–1374; (e) Palomo, C.; Oiarbide, M.; Garcia, J. M. Chem.

    Soc. Rev. 2004, 33,65–75.

    10. For reviews on direct aldol reaction, see: (a) Alcaide, B.; Almendros, P. Eur.

    J.Org. Chem. 2002, 1595–1601; (b) Saito, S.; Yamamoto, H. Acc. Chem. Res.

    2004,37, 570–579; (c) Guillena, G.; Najera, C.; Ramon, D. J. Tetrahedron:

    Asymmetry 2007, 18, 2249–2293.

    11. Yamada, Y. M. A.; Yoshikawa, N.; Sasai, H.; Shibasaki, M. Angew. Chem.,

    Int.Ed. 1997, 36, 1871–1873; (b) Shibasaki, M.; Yoshikawa, N. Chem. Rev.

    2002, 102, 2187–2210; (c) Shibasaki, M.; Matsunaga, S. Chem. Soc. Rev. 2006,

    35, 269–279.

    12. For related catalyses, see: (a) Trost, B. M.; Ito, H. J. Am. Chem. Soc. 2000,

    122,12003–12004; (b) Suzuki, T.; Yamagiwa, N.; Matsuo, Y.; Sakamoto,

    S.;Yamaguchi, K.; Shibasaki, M.; Noyori, R. Tetrahedron Lett. 2001, 42, 4669–

    4671.

    13. List, B.; Lerner, R. A.; Barbas, C. F. J. Am. Chem. Soc. 2000, 122, 2395–2396;

    (b)Notz, W.; Tanaka, F.; Barbas, C. F. Acc. Chem. Res. 2004, 37, 580–591;

    (c)Mukherjee, S.; Yang, J. W.; Hoffmann, S.; List, B. Chem. Rev. 2007, 107,

    5471–5569.

  • 201 14. For related catalyses, see: (a) Torii, H.; Nakadai, M.; Ishihara, K.; Saito, S.;

    Yamamoto, H. Angew. Chem., Int. Ed. 2004, 43, 1983–1986; (b) Cobb, A. J. A.;

    Longbottom, D. A.; Shaw, D. M.; Ley, S. V. Chem. Commun. 2004, 1808–1809;

    (c) Kano, T.; Takai, J.; Tokuda, O.; Maruoka, K. Angew. Chem., Int. Ed. 2005,

    44, 3055–3057.

    15. Although the substrates are limited to glycine Schiff bases, the aldol reactions

    promoted by phase transfer catalysts have been reported, see: Ooi, T.;

    Taniguchi, M.; Kameda, M.; Maruoka, K. Angew. Chem., Int. Ed. 2002, 41,

    4542–4544.

    16. Shunsuke Kotani, Yasushi Shimoda, Masaharu Sugiura, Makoto Nakajima,

    Tetrahedron Lett. 50,2009,4602-4605.

    17. Although adding tetrachlorosilane to the mixture of the other components gave

    the same results, adding a ketone to the mixture decreased the stereoselectivities.

    18. W.-D. Fessner, in Stereoselective Biocatalysis; R. N. Patel, Ed.;Marcel Dekker,

    New York, 2000, p. 239; (b) T. D. Machajewski and C.-H. Wong, Angew.

    Chem., Int. Ed., 2000, 39, 1352.

    19. Comprehensive Organic Synthesis, Vol. 2, B. M. Trost, I. Fleming, C.-

    H.Heathcock, Eds.; Pergamon, Oxford, 1991.

    20. For examples of application in total synthesis see: (a) T.Mukaiyama, Angew.

    Chem., Int. Ed., 2004, 43, 5590; (b) K. C. Nicolaou,D. Vourloumis, N.

    Winssinger and P. S. Baran, Angew. Chem., Int.Ed., 2000, 39, 44.

    21. Modern Aldol Reactions, Vol. 1 & 2, R. Mahrwald, Ed.; Wiley-VCH,Weinheim,

    2004; E. M. Carreira, in Comprehensive Asymmetric Catalysis; E. N. Jacobsen,

    A. Pfaltz, H. Yamamoto, Eds.; Springer,Heidelberg, 1999; J. S. Johnson and D.

    A. Evans, Acc. Chem. Res.,2000, 33, 325.

    22. K. Mikami and S. Matsukawa, J. Am. Chem. Soc., 1994, 116, 4077; (b) G. E.

    Keck, X.-Y. Li and D. Krishnamurthy, J. Org. Chem., 1995,60, 5998; (c) D. A.

    Evans, D. M. Fitch, T. E. Smith and V. J. Cee,J. Am. Chem. Soc., 2000, 122,

    10033; (d) K. Juhl, N. Gathergood and K. A. Jørgensen, Chem. Commun., 2000,

    2211; (e) E. M. Carreira,R. A. Singer and W. S. Lee, J. Am. Chem. Soc., 1994,

    116, 8837; (f)H. Ishita, Y. Yamashita, H. Shimizu and S. Kobayashi, J. Am.

    Chem.Soc., 2000, 122, 5403; (g) S. E. Denmark and R. A. Stavanger, J.

    Am.Chem. Soc., 2000, 122, 8837; (h) N. Kumagai, S. Matsunaga,T. Kinoshita, S.

    Harada, S. Okada, S. Sakamoto, K. Yamaguchi and M. Shibasaki, J. Am. Chem.

  • 202

    Soc., 2003, 125, 2169; (i) B.M. Trost, H. Ito and E. R. Silcoff, J. Am. Chem.

    Soc., 2001, 123, 3367; (j) Y. M. A.Yamada, N. Yoshikawa, H. Sasai and M.

    Shibasaki, Angew. Chem.,Int. Ed., 1997, 36, 1871; (k) D. A. Evans, C. W.

    Downey and J. L. Hubbs, J. Am. Chem. Soc., 2003, 125, 8706.

    23. R. Schoevaart, F. Van Rantwijk and R. A. Sheldon, J. Org. Chem.,2000, 65,

    6940; H. J. M. Gijsen, L. Qiao, W. Fitz and C.-H Wong,Chem. Rev., 1996, 96,

    443.

    24. Reviews see: P. I. Dalko and L. Moisan, Angew. Chem., Int. Ed., 2004, 43, 5138;

    B. List, Tetrahedron, 2002, 58, 5573.

    25. For the proline-catalyzed intermolecular aldol reaction see: (a) Z. G. Hajos and

    D. R. Parrish, J. Org. Chem., 1974, 39, 1615; (b) U. Eder, R. Sauer and R.

    Wiechert, Angew. Chem., Int. Ed., 1971, 10,496; (c)C. Pidathala, L. Hoang, N.

    Vignola and B. List, Angew. Chem., Int. Ed., 2003, 42, 2785. For the

    stoichiometric use of phenylalanine and tyrosine to mediate asymmetric

    intermolecular aldol condensations see:(d) S. Danishefsky and P. Cain, J. Am.

    Chem. Soc., 1975, 97, 5282; (e) I. Shimizu, Y. Naito and J. Tsuji, Tetrahedron

    Lett., 1980, 21, 4975; (f)H. Hagiwara and H. Uda, J. Org. Chem., 1988, 53,

    2308; (g) E. J. Corey and S. C. Virgil, J. Am. Chem. Soc., 1990, 112, 6429.

    26. B. List, R. A. Lerner and C. F. Barbas, III, J. Am. Chem. Soc., 2000,122, 2395;

    (b)W. Notz and B. List, J. Am. Chem. Soc., 2000, 122, 7386;(c) K. S.

    Sakthivel,W. Notz, T. Bui and C. F. Barbas, III, J. Am. Chem.Soc., 2001, 123,

    5260; (d) B. List, P. Porjarliev and C. Castello, Org.Lett., 2001, 3, 573; (e) A.

    Cordova, W. Notz and C. F. Barbas, III,Chem. Commun., 2002, 3024; (f) A.

    Cordova, W.Notz and C. F. Barbas,III, J. Org. Chem., 2002,67, 301; (g) A. B.

    Northrup and D. W. C.MacMillan, J. Am. Chem. Soc., 2002, 124, 6798; (h) A.

    Bøgevig,N. Kumaragurubaran and K. A. Jørgensen, Chem. Commun., 2002,620;

    S. Saito, M. Nakadai and H. Yamamoto, Tetrahedron, 2002, 58,8167; (i) Z.

    Tang, F. Jiang, L.-T. Yu, X. Cui, L.-Z. Gong, A.-Q. Mi, Y.-Z. Jiang and Y.-D.

    Wu, J. Am. Chem. Soc., 2003, 125, 5262; (j)H. Torii, M. Nakadai, K. Ishihara, S.

    Saito and H. Yamamoto, Angew.Chem., Int. Ed., 2004, 43, 1983; (k) A. Hartikaa

    and P. I. Arvidsson,Tetrahedron: Asymmetry, 2004, 15, 1831; (l) A. Berkessel,

    B. Koch and J. Lex, Adv. Synth. Catal., 2004, 346, 1141; (m) A. J. A.

    Cobb,D.M. Shaw, D. A.Longbottom, J.B.Gold and S. V.Ley, Org.

    Biomol.Chem., 2005, 3, 84.

  • 203 27. Armando Cordova, Weibiao Zou, Ismail Ibrahem, Efraim Reyes, Magnus

    Engqvist and Wei-Wei Liao, Chem. Commun.,2005,3586-3588.

    28. Kano, T.; Takai, J.; Tokuda, O.; Maruoka, K. Angew.Chem., Int. Ed. 2005, 44,

    3055.Kano, T.; Tokuda, O.; Maruoka, K. Tetrahedron Lett.2006, 47, 7423.

    29. Mase, N.; Nakai, Y.; Ohara, N.; Yoda, H.; Takabe, K.;Tanaka, F.; Barbas, C. F.,

    III. J. Am. Chem. Soc. 2006, 128,734.(b) Hayashi, Y.; Sumiya, T.; Takahashi, J.;

    Gotoh, H.;Urushima, T.; Shoji, M. Angew. Chem., Int. Ed. 2006, 45, (c) Chimni,

    S. S.; Mahajan, D.; Babu, V. V. S. TetrahedronLett. 2005, 46, 5617 (d) Chen, J.-

    R.; Lu, H.-H.; Li, X.-Y.; Cheng, L.; Wan, J.;Xiao, W.-J. Org. Lett. 2005, 7,

    4543. (e) Samanta, S.; Liu, J.; Dodda, R.; Zhao, C.-G. Org. Lett.2005, 7, 5321.

    (f) Gryko, D.; Lipinski, R. Adv. Synth.Catal. 2005, 347, 1948.

    30. (a)Hayashi, Y.; Sekizawa, H.; Yamagushi, J.; Gotoh, H. J.Org. Chem., 2007, 72,

    6493.Davies, S. G.; Sheppard, R. L.; Smith, A. D.; Thoson, J.E. Chem. Commun.

    2005, 3802.Tang, Z.; Cun, L.-F.; Mi, A.-Q.; Jiang, Y.-Z.; Gong, L.-Z.Org. Lett.

    2006, 8, 1263.

    31. Luo, S.-Z.; Xu, H.; Li, J.-Y.; Zhang, L.; Cheng, J.-P. J.Am. Chem. Soc. 2007,

    129, 3074.(b) Luo, S.; Li, Y.; Xu, H.; Cheng, J.-P. Chem. Eur. J. 2008, 14, 1273.

    32. Tang, Z.; Jiang, F.; Yu, L.-T.; Mi, A.-Q.; Jiang, Y.-Z.; Wu, Y.-D. J. Am. Chem.

    Soc. 2003, 125, 5262.(b) Tang, Z.; Jiang, H.; Cui, X.; Gong, L.-Z.; Mi, A.-Q.;

    Jiang,Y.-Z.; Wu, Y.-D. Proc. Natl. Acad. Sci. U. S. A. 2004,101, 5755.(c) Guo,

    H.-M.; Cun, L.-F.; Gong, L.-Z.; Mi, A.-Q.; Jiang,Y.-Z. Chem. Commun. 2005,

    1450.(d) Tang, Z.; Yang, Z.-H.; Chen, X.-H.; Cun, L.-F.; Mi,A.-Q.; Jiang, Y.-Z.;

    Gong, L.-Z. J. Am. Chem. Soc. 2005,127, 9285.He, L.; Tang, Z.; Cun, L.-F.; Mi,

    A.-Q.; Jiang, Y.-Z.;Gong, L.-Z. Tetrahedron 2006, 62, 346.(f) Jiang, M.; Zhu,

    S.-F.; Yang, Y.; Gong, L.-Z.; Zhou, X.-G.;Zhou, Q.-L. Tetrahedron: Asymmetry

    2006, 17, 384.(g) Tang, Z.; Marx, A. Angew. Chem., Int. Ed. 2007, 46, 7297.

    33. Zhao, Jing, Chen, Aijun, Liu and Quanzhong, Chinese Journal of Chemistry,

    2009, 27, 930-936.

    34. B. List, A. R. Lerner and C. F. Barbas, J. Am. Chem. Soc., 2000,122, 2395; (b)

    K. Sakthivel, W. Notz, T. Bui and C. F. Barbas,J. Am. Chem. Soc., 2001, 123,

    5260

    35. C. Isart, J. Bures and J. Vilarrasa, Tetrahedron Lett., 2008, 49,5414; (b) B. List,

    L. Hoang and H. J. Martin, Proc. Natl. Acad. Sci.U. S. A., 2004, 101, 5839; (c)

  • 204

    F. Orsini, F. Pelizzoni, M. Forte,R. Destro and P. Gariboldi, Tetrahedron, 1988,

    44, 519.

    36. Z. Tang, F. Jiang, X. Cui, L.-Z. Gong, A.-Q. Mi, Y.-Z. Jiang and Y. Dong, Proc.

    Natl. Acad. Sci. U. S. A., 2004, 101, 5755;(b) A. Hartikka and P. I. Arvidsson,

    Eur. J. Org. Chem., 2005, 11,4287; (c) A. Berkessel, B. Koch and J. Lex, Adv.

    Synth. Catal., 2004,346, 1141; (d) M. R. Vishnumaya, S. K. Ginotra and V. K.

    Singh,Org. Lett., 2006, 8, 4097; (e) F. Jiang, L.-T. Yu, X. Cui, L.-Z. Gong, A.-

    Q. Mi, Y.-Z. Jiang and Y.-D. Wu, J. Am. Chem. Soc., 2005, 127,9285; (f) D. E.

    Ward, V. Jheengut and O. T. Akinnusi, Org. Lett.,2005, 7, 1181; (g) H. Zhang,

    S. Mitsumori, N. Utsumi, M. Imai,N. Garcia-Delgado, M. Mifsud, M.

    Albertshofer, P. H.-Y. Cheong,K. N. Houk, F. Tanaka and C. F. Barbas, J. Am.

    Chem. Soc., 2008,130, 875; (h) S. Luo, H. Xu, L. Zhang, J. Li and J.-P. Cheng,

    Org.Lett., 2008, 10, 653; (i) M. Pouliquen, J. Blanchet, M.-C. Lasne and J.

    Rouden, Org. Lett., 2008, 10, 1029; (j) J. Wang, H. X. Xie, H. Li,L. S. Zu and

    W. Wang, Angew. Chem., Int. Ed., 2008, 47, 4177. For excellent reviews of

    proline-catalyzed reactions, see: (k) B. List, Tetrahedron, 2002, 58, 5573; (l) S.

    Mukherjee, J.-W. Yang,S. Hoffmann and B. List, Chem. Rev., 2007, 107, 5471.

    37. Y. Zhou and Z. Shan, J. Org. Chem., 2006, 71, 9510;(b) A. I. Nyberg, A. Usano

    and P. M. Pihko, Synlett, 2004, 1891;(c) H. Torii, M. Nakadai, K. Ishihara, S.

    Saito and H. Yamamoto,Angew. Chem., Int. Ed., 2004, 43, 1983; (d) M.

    Amedjkouh, Tetrahedron: Asymmetry, 2005, 16, 1411; (e) Z. Tang, Z.-H.

    Yang,L.-F. Cun, L.-Z. Gong, A.-Q. Mi and Y.-Z. Jiang, Org. Lett., 2004,6, 2285;

    (f) D. E.Ward and V. Jheengut, Tetrahedron Lett., 2004, 45,8347; (g) P. M.

    Pihko, K. M. Laurikainen, A. Usano, A. I. Nyberg and J. A. Kaavi, Tetrahedron,

    2006, 62, 317; (h) P. Dziedzic,W. Zou, J. Ha fren and A. Co rdova, Org. Biomol.

    Chem., 2006, 4,38. For a review on additive-improved asymmetric catalysis, see:

    E.M. Vogl, H. Gro ger and M. Shibasaki, Angew. Chem., Int. Ed., 1999, 38,

    1570.

    38. M. M. Vasbinder, J. E. Imbriglio and S. J. Miller, Tetrahedron,2006, 62, 11450;

    (b) J. E. Imbriglio, M. M. Vasbinder and S. J. Miller, Org. Lett., 2003, 5, 3741.

    39. L. Hoang, S. Bahmanyar, K. N. Houk and B. List, J. Am. Chem.Soc., 2003, 125,

    16; (b) S. Bahmanyar, K. N. Houk, H. J. Martin and B. List, J. Am. Chem. Soc.,

    2003, 125, 2475; (c) F. R. Clemente and K. N. Houk, Angew. Chem., Int. Ed.,

  • 205

    2004, 37, 5766; (d) C. Allemann, R. Gordillo, F. R. Clemente, P. H.-Y. Cheong

    and K. N. Houk, Acc. Chem. Res., 2004, 37, 558.

    40. Omer Reis, Serkan Eymur, Barbaros Reis and Ayhan S. Demir, Chem. Commun.

    2009, 1088-1090.

    41. A. Berkssel, H. Groger, Asymmetric Organocatalysis, Wiley-VCH, Weinheim,

    2005; b) P. I. Dalko, L. Moisan, Angew.Chem.2004, 116, 5248; Angew.Chem.

    Int. Ed. 2004, 43, 5138; c) Y.Hayashi, J. Syn. Org. Chem. Jpn. 2005, 63, 464.

    42. Modern Aldol Reactions, Vols. 1, 2 (Ed.: R. Mahrwald), Wiley-VCH,

    Weinheim, 2004.[6] T. D. Machajewski, C.-H. Wong, Angew. Chem. 2000, 112,

    1406; Angew. Chem. Int. Ed. 2000, 39, 1352.

    43. T. Hamada, K. Manabe, S. Ishikawa, S. Nagayama, M. Shiro, S.Kobayashi, J.

    Am. Chem. Soc. 2003, 125, 2989.

    44. T. Darbre, M. Machuqueiro, Chem. Commun. 2003, 1090.

    45. For representative reports, see; a) Z. G. Hajos, D. R. Parrish, J.Org. Chem. 1974,

    39, 1615; b) U. Eder, G. Sauer, R. Wiechert, Angew. Chem. 1971, 83, 492;

    Angew. Chem. Int. Ed. Engl. 1971,10, 496; c) B. List, R. A. Lerner, C. F. Barbas

    III, J. Am. Chem.Soc. 2000, 122, 2395; d) A. B. Northup, D. W. C. MacMillan,

    J.Am. Chem. Soc. 2002, 124, 6798.

    46. A.Cordova, W. Notz,C. F. Barbas III, Chem. Commun. 2002, 3024.

    47. Z. Tang, Z.-H. Yang, X.-H. Chen, L.-F. Cun, A.-Q. Mi, Y.-Z.Jiang, L.-Z. Gong,

    J. Am. Chem. Soc. 2005, 127, 9285, and references therein.

    48. H. Torii, M. Nakadai, K. Ishihara, S. Saito, H. Yamamoto, Angew. Chem. 2004,

    116, 2017; Angew. Chem. Int. Ed. 2004, 43,1983; b) A. I. Nyberg, A. Usano, P.

    M. Pihko, Synlett 2004, 1891;c) Z. Tang, Z.-H. Yang, L.-F. Cun, L.-Z. Gong,

    A.-Q. Mi, Y.-Z.Jiang, Org. Lett. 2004, 6, 2285; d) J. Casas, H. Sunden,

    A.Cordova, Tetrahedron Lett. 2004, 45, 6117; e) D. E. Ward, V.Jheengut,

    Tetrahedron Lett. 2004, 45, 8347; f) I. Ibrahem, A.Cordova, Tetrahedron Lett.

    2005, 46, 3363; g) M. Amedjkouh, Tetrahedron: Asymmetry 2005, 16, 1411; h)

    A. Cordova, W. Zou,I. Ibrahem, E. Reyes,M. Engqvist,W.-W. Liao, Chem.

    Commun.2005, 3586.

    49. Nornicotine was reported to promote the aldol reaction in water,but low

    asymmetric induction (ca. 20%ee) was observed: a) T. J.Dickerson, K. D. Janda,

    J. Am. Chem. Soc. 2002, 124, 3220;b) C. J. Rogers, T. J. Dickerson, A. P.

    Brogan, K. D. Janda, J. Org.Chem. 2005, 70, 3705.

  • 206 50. Yujiro Hayashi, Tatsunobu Sumiya, Junichi Takahashi, Hiroaki Gotoh, Tatsuya

    Urushima, and Mitsuru Shoji, Angew. Chem. Int.Ed.2006, 45, 958-961.

    51. Evans, D. A.; Bartroli, J.; Shih, T. L. J. Am. Chem. Soc.1981, 103, 2127. (b)

    Evans, D. A.; Nelson, J. V.; Vogel, E.;Taber, T. R. J. Am. Chem. Soc. 1981, 103,

    3099. (c) Evans,D. A.; Vogel, E.; Nelson, J. V. J. Am. Chem. Soc.

    1979,101,6120.

    52. Danda, H.; Hansen, M. M.; Heathcock, C. H. J. Org. Chem.1990, 55, 173. (b)

    Heathcock, C. H. Science 1981, 214, 395.(c) Heathcock, C. H. Asymmetric

    Synthesis; Morrion, J. D.,Ed.; Academic: New York, 1984; Vol. 3, p 111. (d)

    Heathcock, C. H.; White, C. T. J. Am. Chem. Soc. 1979, 101. (e) Kleschick,W.

    A.; Buse, C. T.; Heathcock, C. H. J. Am. Chem. Soc. 1977,99.

    53. Kim, B. M.; Williams, S. F.; Masamune, S. Comprehensive Organic Synthesis;

    Trost, B. M., Fleming, I., Heathcock, C. H., Eds.; Pergamon: Oxford, 1991; Vol.

    2, p 239, Chapter 1.7. (b) Masamune, S.; Ali, S.; Snitman, D. L.; Garvey, D.

    S.Angew. Chem. 1980, 92, 573. (c) Masamune, S.; Choy, W.; Kerdesky, F. A. J.;

    Imperiali, B. J. Am. Chem. Soc. 1981, 103, 1566. (d) Masamune, S.; Sato, T.;

    Kim, B. M.; Wollmann,T. A. J. Am. Chem. Soc. 1986, 108, 8279–8281.

    54. Kobayashi, S.; Uchiro, H.; Shiina, I.; Mukaiyama, T.Tetrahedron 1993, 49,

    1761. (b) Mukaiyama, T. The Directed Aldol Reaction. Organic Reactions;

    Wiley: New York, 1982; Vol. 28, p 203. (c) Mukaiyama, T.; Banno, K.;

    Narasaka, K.R. Ian Storer, D. W. C. MacMillan / Tetrahedron ,60 ,(2004) 7705–

    7714 7713 J. Am. Chem. Soc. 1974, 96, 7503. (d) Mukaiyama, T.; Narasaka, K.;

    Banno, K. Chem. Lett. 1973, 9, 1011.

    55. (a) Chowdari, N. S.; Ramachary, D. B.; Cordova, A.; Barbas,C. F. Tetrahedron

    Lett. 2002, 43, 9591. (b) Cordova, A.; Notz, W.; Barbas, C. F. Chem. Commun.

    2002, 3024. (c) List, B.; Lerner, R. A.; Barbas, C. F. J. Am. Chem. Soc. 2000,

    122, 2395. (d) Sakthivel, K.; Notz, W.; Bui, T.; Barbas, C. F. J. Am.Chem. Soc.

    2001, 123, 5260.

    56. Evans, D. A.; Tedrow, J. S.; Shaw, J. T.; Downey, C. W. J. Am.Chem. Soc.

    2002, 124, 392.

    57. (a)List, B. Tetrahedron 2002, 58, 5573. (b) List, B.; Pojarliev, P.; Castello, C.

    Org. Lett. 2001, 3, 573. (c) Notz, W.; List, B.J. Am. Chem. Soc. 2000, 122, 7386.

    (d) Pidathala, C.; Hoang, L.; Vignola, N.; List, B. Angew. Chem., Int. Ed. 2003,

    42, 2785.

  • 207 58. Lalic, G.; Aloise, A. D.; Shair, M. D. J. Am. Chem. Soc. 2003,125, 2852.

    59. (a) Kumagai, N.; Matsunaga, S.; Yoshikawa, N.; Ohshima, T.; Shibasaki, M.

    Org. Lett. 2001, 3, 1539. (b) Yamada, Y. M. A.;Yoshikawa, N.; Sasai, H.;

    Shibasaki, M. Angew.Chem., Int.Ed.Engl.1997,36,1871.(c) Yoshikawa, N.;

    Kumagai, N.;Matsunaga, S.; Moll, G.; Ohshima, T.; Suzuki, T.; Shibasaki,M. J.

    Am. Chem. Soc. 2001, 123, 2466. (d) Yoshikawa, N.; Shibasaki, M. Tetrahedron

    2001, 57, 2569. (e) Yoshikawa, N.;Yamada, Y. M. A.; Das, J.; Sasai, H.;

    Shibasaki, M. J. Am.Chem. Soc. 1999, 121, 4168.

    60. (a)Trost, B. M.; Fettes, A.; Shireman, B. T. J. Am. Chem. Soc.2004, 126, 2660.

    (b) Trost, B. M.; Ito, H. J. Am. Chem. Soc.2003, 122, 1. (c) Trost, B. M.; Ito, H.;

    Silcoff, E. R. J. Am.Chem. Soc. 2001, 123, 3367. (d) Trost, B. M.; Silcoff, E. R.;

    Ito, H. Org. Lett. 2001, 3, 2497.

    61. Denmark, S. E.; Ghosh, S. K. Angew. Chem., Int. Ed. 2001, 40, 4759.

    62. R. Ian Storer and David W. C. MacMillan, Tetrahedron, 2004 60, 7705-7714.

    63. (1) Reviews: (a) Nelson, S. G. Tetrahedron: Asymmetry 1998, 9, 357-389. (b)

    Gro¨ger, H.; Vogl, E. M.; Shibasaki, M. Chem. Eur. J. 1998, 4, 1137.(c) Bach,

    T. Angew. Chem., Int. Ed. Engl. 1994, 33, 417. (d) A notable exception is

    Denmark’s chiral Lewis base-catalyzed Mukayama-type aldol reaction:

    Denmark, S. E.; Stavenger, R. A.; Wong, K.-T. J. Org. Chem. 1998, 63, 918-

    919.

    64. March, J. J.; Lebherz, H. G. TIBS 1992, 17, 110. (b) Rutter, W. J.Fed. Proc. Am.

    Soc. Exp. Biol. 1964, 23, 1248. (c) Lai, C. Y.; Nakai, N.; Chang, D. Science

    1974, 183, 1204. (d) Morris, A. J.; Tolan, D. R.; Biochemistry 1994, 33, 12291.

    65. Fessner, W.-D.; Schneider, A.; Held, H.; Sinerius, G.; Walter, C.; Hixon,M.;

    Schloss, J. D. Angew. Chem., Int. Ed. Engl. 1996, 35, 2219-

    2221.Phosphoenolpyruvate aldolases use a preformed enolate,

    phosphoenolpyruvate, to accomplish aldol addition reactions. For studies of this

    and other aldolase enzymes in organic synthesis see: Gijsen, H. J. M.; Qiao, L.;

    Fitz, W.; Wong, C.-H. Chem. ReV. 1996, 96, 443-473.

    66. (a) Wagner, J.; Lerner, R. A.; Barbas, C. F., III Science 1995, 270, 1797. (b)

    Barbas, C. F., III; Heine, A.; Zhong, G.; Hoffmann, T.; Gramatikova,S.;

    Bjo¨rnestedt, R.; List, B.; Anderson, J.; Stura, E. A.; Wilson, E. A.; Lerner,R. A.

    Science 1997, 278, 2085-2092. (c) Hoffmann, T.; Zhong, G.; List, B.;Shabat, D.;

    Anderson, J.; Gramatikova, S.; Lerner, R. A.; Barbas, C. F., III J.Am. Chem.

  • 208

    Soc. 1998, 120, 2768-2779. (d) List, B.; Shabat, D.; Barbas, C.F., III; Lerner, R.

    A. Chem. Eur. J. 1998, 881-885. (e) Zhong, G.; Shabat,D.; List, B.; Anderson,

    J.; Sinha, S. C.; Lerner. R. A.; Barbas, C. F., III Angew. Chem., Int. Ed. 1998,

    37, 2481-2484. (f) Zhong, G.; Lerner. R. A.; Barbas, C. F., III Angew. Chem.,

    Int. Ed. 1999, 38, 3738-3741. (g) Sinha, S. C.;Sun, J.; Miller, G.; Barbas, C. F.,

    III; Lerner, R. A. Org. Lett. 1999, 1, 1623-1626. (h) Turner, J. M.; Bui, T.;

    Lerner, R. A.; Barbas, C. F., III; List, B.Manuscript submitted for publication.

    67. (a) Nakagawa, M.; Nakao, H.; Watanabe, K.-I. Chem. Lett. 1985, 391-394. (b)

    Yamada, Y.; Watanabe, K.-I.; Yasuda, H. Utsunomiya Daigaku Kyoikugakubu

    Kiyo, Dai-2-bu 1989, 39, 25-31.

    68. Yamada, Y. M. A.; Yoshikawa, N.; Sasai, H.; Shibasaki, M. Angew.Chem., Int.

    Ed. Engl. 1997, 36, 1871. (b) Yamada, Y. M. A.; Shibasaki, M.Tetrahedron Lett.

    1998, 39, 5561. (c) Yoshikawa, N.; Yamada, Y. M. A.; Das, J.; Sasai, H.;

    Shibasaki, M. J. Am. Chem. Soc. 1999, 121, 4168-4178.

    69. Benjamin List, Richard A. Lerner, and Carlos F. Barbas III,

    J.Am.Chem.Soc.2000, 122, 2395-2396.

    70. Mukaiyama, T.; Banno, K.; Narasaka, K. J. Am. Chem. Soc. 1974, 96, 7503–

    7506; (b) Kobayashi, S.; Fujishita, Y.; Mukaiyama, T. Chem. Lett. 1990, 1455–

    1458.

    71. Darbre, T.; Machuqueiro, M. Chem. Commun. 2003, 1090–1091; (b) Kofoed, J.;

    Reymond, J.-L.; Darbre, T. Org.Biomol. Chem. 2005, 3, 1850–1855.

    72. Kobayashi, S.; Manabe, K. Pure App. Chem. 2000, 1373–1380; (b) Kobayashi,S.

    Eur. J. Org. Chem. 1999, 15227–15232; (c) Kobayashi, S.; Nagayama,

    S.;Busujima, T. J. Am. Chem. Soc. 1998, 120, 8287–8288; (d) Otto, S.;

    Bertoncin, F.;Engberts, J. B. F. N. J. Am. Chem. Soc. 1996, 118, 7702–7707; (e)

    Engberts, J. B. F.N.; Feringa, B. L.; Keller, E.; Otto, S. Rec. Trav. Chim. Pays-

    Bas 1996, 115, 457–464; (f) Kobayashi, S.; Ogawa, C. Chem. Eur. J. 2006, 12,

    5954–5960.

    73. For a review on the roles of zinc in zinc enzymes: Kimura, E.; Kikuta, E. J.

    Biol.Inorg. Chem. 2000, 5, 139–155.

    74. Mael Penhoat, Didier Barbry and Christian Rolando, Tetrahedron Lett. 52, 2011,

    159-162.

    75. Lutz, M.; Bakker, R. Acta Crystallogr., Sect. C 2003, 59, 18–20.

  • 209 76. Mase, N.; Nakai, Y.; Ohara, N.; Yoda, H.; Takabe, K.; Tanaka, F.;Barbas, C. F.,

    III. J. Am. Chem. Soc. 2006, 128, 734.

    77. (a) Hayashi, Y.; Sumiya, T.; Takahashi, J.; Gotoh, H.; Urushima,T.; Shoji, M.

    Angew. Chem., Int. Ed. 2006, 45, 958. (b) Hayashi, Y.; Aratake,S.; Okano, T.;

    Takahashi, J.; Sumiya, T.; Shoji, M. Angew. Chem., Int. Ed.2006, 45, 5527.

    78. Vishnu Maya, Monika Raj, and Vinod K. Singh, Organic Lett. 2007, Vol. 9, No.

    13, 2593-2595.

    79. For some recent selected references on aldol reaction in water via asymmetric

    organocatalysis, see: (a) Guizzetti, S.; Benaglia, M.; Raimondi, L.; Celentano, G.

    Org. Lett. 2007, 9, 1247. (b) Wu, Y.; Zhang, Y.; Yu, M.; Zhao, G.; Wang, S.

    Org. Lett. 2006, 8, 4417. (c) Jiang, Z.; Liang, Z.; Wu, X.; Lu, Y. Chem.

    Commun. 2006, 2801. (d) Dziedzic, P.; Zou, W.; Hafren,J.; Cardova, A. Org.

    Biomol. Chem. 2006, 4, 38. (e) Chimni, S. S.; Mahajan, D. Tetrahedron:

    Asymmetry 2006, 17, 2108. (f) Guillena, G.; Hita, M. D.C.; Najera, C.

    Tetrahedron: Asymmetry 2006, 17, 1493. (g) Fu, Y.-Q.; Li,Z.-C.; Ding, L.-N.;

    Tao, J.-C.; Zhang, S.-H.; Tang, M.-S. Tetrahedron: Asymmetry 2006, 17, 3351.

    (h) Pihko, P. M.; Laurikainen, K. M.; Usano,A.; Nyberg, A. I.; Kaavi, J. A.

    Tetrahedron 2006, 62, 317.

    80. (a) Yoshikawa, N.; Kumagai, N.; Matsunaga, S.; Moll, G.; Ohshima, T.; Suzuki,

    T.; Shibasaki, M. J. Am. Chem. Soc.2001, 123, 2466. (b) Yamada, M. A. Y.;

    Shibasaki, M.Tetrahedron Lett. 1998, 39, 5561.

    81. Trost, B. M.; Ito, H. J. Am. Chem. Soc. 2000, 122, 12003.

    82. Bogevig, A.; Kumaragurubaran, N.; Jorgensen, K. A. Chem.Commun. 2002,

    620.

    83. Northrup, A. B.; MacMillan, D. W. C. J. Am. Chem. Soc. 2002,124, 6798.

    84. List, B.; Lerner, R. A.; Barbas, C. F., III. J. Am. Chem. Soc.2000, 122, 2395.

    85. Cordova, A.; Notz, W.; Barbas, C. F., III. J. Org. Chem. 2002, 67, 301.

    86. Cordova, A. Tetrahedron Lett. 2004, 45, 3949.

    87. Kotrusz, P.; Kmentova, I.; Gotov, B.; Toma, S.; Solcaniova, E.Chem. Commun.

    2002, 2510.

    88. (a) Cordova, A.; Notz, W.; Barbas, C. F., III. Chem. Commun.2002, 3024. (b)

    Nyberg, A. I.; Usano, A.; Pihko, P. M. Synlett 2004, 1891. (c) Hartikka, A.;

    Arvidsson, P. I. Tetrahedron: Asymmetry 2004, 15, 1831. (d) Peng, Y.-Y.; Ding,

    Q. P.; Li, Z.;Wang, P. G.; Cheng, J.-P.Tetrahedron Lett. 2003, 44, 3871.(e)