Acessibilidade / Reportar erro

Metal-catalyzed asymmetric aldol reactions

Abstracts

The aldol reaction is one of the most powerful and versatile methods for the construction of C-C bonds. Traditionally, this reaction was developed in a stoichiometric version; however, great efforts in the development of chiral catalysts for aldol reactions were performed in recent years. Thus, in this review article, the development of metal-mediated chiral catalysts in Mukaiyama-type aldol reaction, reductive aldol reaction and direct aldol reaction are discussed. Moreover, the application of these catalysts in the total synthesis of complex molecules is discussed.

aldol reactions; asymmetric induction; chiral ligands; total synthesis


A reação aldólica é uma das ferramentas mais poderosas e versáteis para a construção de ligações C-C. Tradicionalmente, esta reação foi desenvolvida em sua versão estequiométrica, no entanto, grandes esforços no desenvolvimento de catalisadores quirais para reações aldólicas foram realizados nos últimos anos. Desta forma, neste artigo de revisão, é discutido o desenvolvimento de catalisadores metálicos em reação aldólica do tipo Mukaiyama, reação aldólica redutiva e reação aldólica direta. Além disto, a aplicação destes catalisadores na síntese total de moléculas complexas será abordada.


REVIEW

Metal-catalyzed asymmetric aldol reactions

Luiz C. Dias* * e-mail: ldias@iqm.unicamp.br FAPESP has sponsored the publication of this article. ; Emílio C. de Lucca Jr.; Marco A. B. Ferreira; Ellen C. Polo

Institute of Chemistry, University of Campinas (UNICAMP), CP 6154, 13083-970 Campinas-SP, Brazil

ABSTRACT

The aldol reaction is one of the most powerful and versatile methods for the construction of C-C bonds. Traditionally, this reaction was developed in a stoichiometric version; however, great efforts in the development of chiral catalysts for aldol reactions were performed in recent years. Thus, in this review article, the development of metal-mediated chiral catalysts in Mukaiyama-type aldol reaction, reductive aldol reaction and direct aldol reaction are discussed. Moreover, the application of these catalysts in the total synthesis of complex molecules is discussed.

Keywords: aldol reactions, asymmetric induction, chiral ligands, total synthesis

RESUMO

A reação aldólica é uma das ferramentas mais poderosas e versáteis para a construção de ligações C-C. Tradicionalmente, esta reação foi desenvolvida em sua versão estequiométrica, no entanto, grandes esforços no desenvolvimento de catalisadores quirais para reações aldólicas foram realizados nos últimos anos. Desta forma, neste artigo de revisão, é discutido o desenvolvimento de catalisadores metálicos em reação aldólica do tipo Mukaiyama, reação aldólica redutiva e reação aldólica direta. Além disto, a aplicação destes catalisadores na síntese total de moléculas complexas será abordada.

1. Introduction

The aldol reaction is one of the most powerful and versatile methods in the chemistry of carbonyl compounds for the construction of C–C bonds in a regio-, stereo- and enantioselective manner.1 It is well known that the relative configuration of the aldol adduct (in those reactions that proceed by a cyclic six member transition state) is controlled by the geometry of the propionate-type enolate, in which Z-enolates lead to preferential formation of the 1,2-syn products and E-enolates to 1,2-anti products. These observations can be rationalized from the Zimmerman-Traxler model. In this proposal, the aldol reaction undergoes a chair-type six-membered cyclic transition state, being diastereoselectivity dependent on the steric demand of the enolate and the aldehyde substituents (Scheme 1).2


According to this model, the R3 aldehyde substituent occupies, preferably, the pseudo-equatorial position, eliminating unfavorable 1,3-diaxial interactions between the R3 group and the R1 and L substituents, thus providing a transition state of lower energy. In the case of the E-enolates, 1,2-anti aldol adducts are preferably formed, since in the transition state TS1 the 1,3-diaxial interactions are minimized with respect to transition state TS2, which leads to the formation of 1,2-syn aldol adduct.

In the case of the Z-enolates, the formation of 1,2-anti aldol adduct is disfavored due to 1,3-diaxial interactions present in the transition state TS3. Thus, the 1,2-syn aldol adduct is formed preferentially, since in the transition state TS4 these repulsions are minimized. Thus, when preformed chiral enolates are employed together with aldehydes, it is possible to obtain aldol adducts with excellent levels of asymmetric induction.3

The aim of this review article is to discuss representative studies involving stereoselective aldol reactions using metal-mediated chiral catalysis with special attention to selectivity, substrate scope, current limitations and application in the total synthesis of natural products.4,5 The literature is covered up to early 2012.

2. Crown Ethers and Related Ligands in Lanthanide-Catalyzed Mukaiyama Aldol Reactions

Reactions in aqueous media have several advantages over conventional dry synthetic procedures. In this context, rare earth metal triflates (RE(OTf)3), which are water-tolerant Lewis acids, have been used in aldol reactions and can be effortlessly recovered and reused.6

Kobayashi and co-workers7 showed that in the Mukaiyama aldol reactions involving chiral crown ether ligands that bind strongly with the larger rare earth metal cations (such as La, Ce, Pr and Nd), the ligands do not decrease the Lewis acid ability for the enantioselective transformations. The Mukaiyama aldol reactions in aqueous media between enolsilanes 1 and 2 and α,β-unsaturated aldehydes, mediated by Pr(OTf)3 and bis-pyridino-18-crown-6 ether 15, gave good to high yields and high diastereo- and enantioselectivities in favor of the corresponding syn aldol adducts (Table 1, entries 1-13). However, aliphatic aldehyde 12 was not a suitable substrate for this transformation (entry 14).

In 2010, Allen and co-workers8 designed a new class of multidentate ligand 16 for aqueous europium-catalyzed Mukaiyama aldol reactions involving enolsilanes 2. The complex Eu3+·16 yielded the corresponding aldol adducts in unprecedented enantioselectivities (90 to 96% ee) and diastereoselectivities (21:1 to 32:1) for aliphatic, aryl and α,β-unsaturated aldehydes (Table 1, entries 15-19).

For ligand 15, the authors proposed transition state TS5, taking into account the coordination of the aldehyde to the Pr3+ cation and the shielding of the Si face of the aldehyde by the axially oriented methyl substituent of the Pr3+·15 complex, thus directing the attack of the enolate to the Re face of the aldehyde (Scheme 2).7 For ligand 16, in which the aldehyde is complexed with the Eu3+ cation, the authors proposed that the Si face of the aldehyde is blocked by the ester substituent, and the Re face of the aldehyde becomes proper for the enol attack.8


In 2012, Allen and co-workers9 applied crown ether 16 as a ligand for neodymium-catalyzed Mukaiyama aldol reactions (Table 2).

As becomes evident from the results presented in Table 2, the Mukaiyama aldol reactions between enolsilane 1 and aromatic and α,β-unsaturated aldehydes, mediated by Nd(OTf)3, gave good to high yields and high diastereo- and enantioselectivities in favor of the corresponding syn aldol adducts (Table 2, entries 1-6). The aldol reactions using aliphatic aldehyde 14 (Table 2, entry 7) led to the formation of aldol adduct with good diastereo- and enantioselectivities in favor of the syn isomer, but in low yield.

3. Diastereo- and Enantioselective Catalytic Reductive Aldol Reactions

Catalytic reductive aldol reactions employing chiral ligands consist in a very exciting method to obtain aldol adducts in high diastereo- and enantioselective fashion.10 The aldol coupling between α,β-unsaturated ester or ketone and aldehydes are promoted using catalytic amounts of a transition metal complex under hydrogenation conditions. The tremendous advantage of this method is that the regioselective reductive formation of a transition metal enolate, required to the aldol reaction, is generated in situ by conjugated addition of a metal hydride to an unsaturated carbonyl compound (Scheme 3).11 The most common reductive agents are molecular hydrogen, silanes and stannanes in stoichiometric amounts. Among the reports involving Co-, Pt-, Pd-, Ni-, Cu-, Ir- and Rh-catalyzed reductive aldol reactions, so far Rh, Ir and Cu transition metals and chiral ligand partners are the most prominent for this transformation. Efforts in order to get better diastereo- and enantioselective aldol adducts under milder conditions reaching high turnover and avoiding carbonyl reduction of the substrates are the driving forces of this field.


The first example of a catalytic asymmetric reductive aldol reaction was reported by Morken and co-workers12 in 2000 employing acrylates and several aldehydes (Table 3). Using Rh catalyst and (R)-BINAP (25), syn aldol products 23 were obtained in good yields and good diastereo- and enantioselectivities (Table 3, entries 1-5). In the Morken's subsequent paper, the scope of the substrates showed synthetic flexibility, producing miscellaneous α,β-substituted β-hydroxyesters (Table 3, entries 6-13).13 Subsequently, a new iridium catalyst was developed using [(cod)IrCl]2 and In-pybox (26) as a chiral ligand partner, to selectively give syn aldol products 24 with benzaldehyde and alkoxy aldehydes (Table 3, entries 14-18).14 The aldol adducts 37 and 38 were the precursors to establish the stereogenic centers at C3-C4 and C10-C11 respectively, for the enantioselective synthesis of borrelidin (39) (Scheme 4).15


In 2008, Krische and co-workers16 introduced a new class of effective monodentate TADDOL-like phosphonite ligands with the ability to promote molecular hydrogen-mediated reductive aldol coupling of vinyl ketones with aldehydes (Scheme 5). High levels of syn-diastereoselectivity in the formation of aldol adducts 40a-e were observed with miscellaneous aldehydes using the preformed complex [Rh(cod)L2]OTf (41) as precatalyst. A few examples are showed in Scheme 5.


Nishiyama and co-workers17 showed the possibility to achieve aldol adducts in high anti-selectivity (Table 4).

The reductive aldol reactions between acrylate 42 and a number of aldehydes were promoted by the chiral Rh-Phebox (44) catalyst and alkoxyhydrosilane providing anti-43 in good to high levels of diastereo- and enantioselectivities.

The supposed stereochemical pathway for this transformation, supported by theoretical calculations, involves the Rh-(E)-enolate (identified by 1H NMR) in a Zimmerman-Traxler-type transition state TS7, with attack to the complexed aldehyde from the less sterically hindered face of the enolate (Scheme 6).18


4. Chiral Lewis Base Catalysis in Enantio­selective Aldol Reactions

The concept of Lewis base catalysis in aldol reactions involving trichlorosilyl enolates with aldehydes has been extensively explored by Denmark et al.19,20 since the first report in 1996 (Scheme 7).21 In the context of stereoselective aldol transformations, the electron-pair of a chiral Lewis base catalyst interacts with an acceptor silicon atom of the enolate making it more reactive. In addition, the new chiral complex should interact with the carbonyl oxygen of aldehyde producing aldol adducts in a stereoselective manner.


The scope and generality of this transformation were extensively studied by Denmark et al.19,20 using phosphoramide organocatalysts. Mukaiyama aldol reactions involving Lewis base catalysis were also reported involving trimethoxysilyl enol ethers activated by binaphtholate organocatalysts.22

An alternative strategy involving Lewis base catalysis in Mukaiyama aldol reactions involves the fluorine-catalyzed bifunctional approach reaction with chiral Lewis acid catalysts, examined by Yamamoto and co-workers23 using the catalyst system BINAP/AgOTf/KF/[18]crown-6 (Scheme 8). In this strategy, the fluoride ions act as an achiral Lewis base forming an anionic hypervalent silicate activating the trimethoxysilyl enol ether, as represented by the cyclic transition state TS8, explaining the diastereoselectivities of aldol adducts syn-50c or anti-50a-b from Z- or E-enolates, respectively.23


Another successful example involving catalytic enantioselective aldol reactions combines a weak achiral Lewis acid (SiCl4) with a chiral phosphoramide (R,R)-53 Lewis base catalyst generating a strong and activated chiral Lewis acid (Scheme 9). Denmark and Chung24 developed Mukaiyama aldol reactions version using this concept. The sense of diastereoselectivity is modulated by the size of the protecting group of silyl ketene acetals and high levels of diastereo- and enantioselectivities were obtained (Scheme 9). Theoretical calculations support the cationic opening transition states TS9 and TS10 for aldol reactions involving enolates 51 and 54, respectively.25


In 2012, Nakajima, Kotani and co-workers26 presented an efficient method for the enantioselective reductive aldol reaction of α,β-unsaturated ketones with aldehydes (Table 5). The conjugated reduction was performed using a tertiary amine and trichlorosilyl triflate. Then, the aldol reaction was made in the presence of BINAP dioxide (BINAPO).

As evident from the results presented in Table 5, the reductive aldol reactions with the chalcone derivatives proceeded smoothly to give the corresponding products in good yields with high diastereo- and enantioselectivities (entries 2-7). Isopropyl ketone 63 provided the aldol product in high yield and selectivity (entry 8), whereas the stereoselectivity of the reaction with cyclopropyl ketone 64 was found to decrease (entry 9).

In addition, reductive aldol reactions between various aldehydes and chalcone (56) were conducted (entries 10-15). The aromatic aldehydes tended to give the aldol adducts with good yields and high stereoselectivities (entries 10-12). The conjugate aldehyde 9 furnished the product in high yield with high selectivity (entry 13). Strikingly, the aliphatic aldehydes 12 and 28, which were generally less reactive in Lewis base-catalyzed reactions, gave the corresponding aldol adducts in good yields with high diastereoselectivities (entries 14 and 15). The rationalization for the observed selectivity is shown in Scheme 10.


The conjugate reduction of ketone gave the (Z)-trichlorosilyl enol ether (Scheme 10). The aldol reaction of (Z)-trichlorosilyl enol ether with aldehyde proceeded via a six-membered transition state TS11 involving hypervalent silicon species to afford the corresponding aldol adduct with high syn-diastereo- and enantioselectivities.

5. Metal-Catalyzed Direct Aldol Reactions

The so-called direct aldol reaction comprises an extraordinary category of aldol transformations developed aiming to atom economy through clean and economic reaction conditions.27 The exciting challenges involving this transformation are to find new catalytic systems that allow the C-C bond coupling by the reaction of enolizable carbonyl compounds with itself or with another carbonyl compound, without the preactivation of the enolate nucleophile, in high chemo-, regio- and stereoselective fashion. Despite the remarkable success of organocatalytic direct aldol processes, heterobimetallic-catalyzed direct aldol reactions shows milder conditions than enamine-based organocatalysts employing nucleophilic amines. However, the requirement of long times under low temperature conditions for these reactions remains as the biggest drawback of this methodology. New organocatalytic approaches involving chiral phosphoric Brønsted acid in direct aldol processes have also emerged as a very attractive alternative.28

Numerous heterobimetallic complexes with chiral BINOL-based ligands have been emerged as very suitable catalysts for direct aldol reactions. A tremendous contribution was given by Shibasaki and co-workers29 in this field. They reported the utilization of the heterobimetallic lanthanum-lithium-BINOL (LLB) complex catalyzing the aldol reaction between aromatic and aliphatic aldehydes with several equivalents of ketones in long reaction times. The heteropolymetallic catalyst LLB-KOH (70) was used in order to shorten the reaction times by enhancing the catalytic activity of LLB complex, giving aldol adducts 69a-d in modest to good enantioselectivities (Scheme 11).30


Shibasaki and co-workers31 reported the use of (S)-LLB (74) catalyst in the formal total synthesis of fostriecin (75) and 8-epi-fostriecin (8-epi-75). The best reaction condition to the system of interest for fostriecin (75) afforded the aldol adduct 73 in good yield using ketone 71 and aldehyde 72 and the two-center Lewis acid-Brønsted base catalyst (S)-LLB (74) (Scheme 12). On the other hand, a new study was performed in order to improve the selectivity for the desired aldol adduct 8-epi-73 used in the synthesis of 8-epi-fostriecin (8-epi-75), and the addictive LiOTf showed the best performance.


In early 2000, Trost and co-workers32 reported a new chiral dinuclear zinc catalyst, prepared from Et2Zn and chiral ligand 78 (Scheme 13).


In these works, the authors obtained aldol adducts with excellent levels of enantioselectivities using the version of direct aldol reaction between various ketones and aldehydes, mediated by a chiral dinuclear zinc catalyst. These results can be consistently explained by the proposed catalytic cycle (Scheme 14). The catalyst 83, prepared in situ by treatment of ligand 78 with 2 equivalents of diethylzinc, involves initiation by liberation of 3 equivalents of ethane followed by a fourth by reaction with the active methylene partner (acetophenone in this case). The chiral space derives from the conformational preferences of the diphenylcarbinol moieties. Thus, the role of the two proximal zinc species is to provide both a zinc to form the requisite enolate (zinc functioning as a Brønsted base) and a second zinc to function as a Lewis acid to coordinate the aldehyde.


In 2005, Trost and co-workers33 described a formal synthesis of fostriecin (75) (Scheme 15). One of the steps of the synthesis consisted in the direct aldol reaction between ketone 85 and aldehyde 84, mediated by chiral binuclear zinc catalyst 83, which led to the formation of aldol adduct 86 with excellent level of enantioselectivity. Compound 86 corresponds to the C8-C13 fragment of fostriecin (75).


Recently, Kumagai, Shibasaki and co-workers34 developed a direct catalytic asymmetric aldol reaction between thioamide 87 and aldehydes employing a soft Lewis acid/hard Brønsted base cooperative catalysis (Table 6 ).

As can be seen from Table 6 , independently of the steric nature of the aldehyde, the aldol adducts were obtained with yields ranging from moderate to good and excellent level of enantiomeric excess.

In 2012, Kumagai, Shibasaki and co-workers35 reported the total synthesis of duloxetine (92), a dual serotonin and norepinephrine reuptake inhibitor in presynaptic cells. The key step of the synthesis was a direct catalytic aldol reaction between thioamide 90 and aldehyde 7, mediated by chiral catalyst ent-88, which provided the aldol adduct 91 with high enantioselectivity (ee = 92%) (Scheme 16).


6. BINOL and Related Ligands in Catalytic Stereoselective Mukaiyama Aldol Reactions

The Lewis acid mediated aldol reactions involving silyl enol ethers with aldehydes are one of the most convenient methods to control the asymmetry in stereoselective catalytic aldol process. Catalytic asymmetric aldol reaction involving ligands possessing symmetry elements of pure rotation such as BINOL and derivatives have been extensively studied. Reetz et al.36 firstly reported enantioselective Mukaiyama aldol reactions involving BINOL-Ti(IV) complex as Lewis acid.

The research groups of Mikami37 and Keck38 gave important contributions to the development of the catalytic thioacetate Mukaiyama aldol reactions, involving, for example, BINOL-Ti(IV) complexes (R)-98 and (S)-95, respectively (Scheme 17). In general, these reactions showed high enantioselectivities with several aldehydes.


In 1994, Carreira and co-workers39 reported the design of a chiral tridentate Schiff base BINOL-derivative ligand 102 utilized for the preparation of the chiral complex 104 (Table 7). This complex presented a superior performance for Mukaiyama aldol reactions between silyl ketene acetals derived from O-alkyl acetates 105 and 106 and several aldehydes. As can be seen in Table 7, aromatic, unsaturated, and saturated aldehydes provided aldol adducts in high enatioselectivities in an in situ preparation of complex 104.40

The Carreira's catalyst 104 was successfully applied in the total synthesis of the antitumor dipsipeptide romidepsin (113)41 and the polyene macrolide roflamycoin (116)42 (Scheme 18). In both synthetic studies, the asymmetric aldol reaction furnished the aldol adducts in high yields, with high levels of enantioselectivities, which were utilized as precursors in the synthesis of the natural products.


In 2000, Kobayashi and co-workers43 developed a Mukaiyama aldol reaction involving zirconium-Lewis acid complex 119 based on the chiral 3,3'-I2-BINOL ligand. The Mukaiyama aldol reactions between several aldehydes and either silyl enol ether derived from O- or S-alkyl acetates preceded in high levels of diastereo- and enantioselectivities in good yields (Scheme 19). Notably, the best reaction condition involves the preparation of catalyst 119 with a small amount of water and in the presence of a primary alcohol.44 The Z- and E-silylketene acetals (Z-120 and E-120) react in a stereoconvergent manner providing anti-121 aldol adduct in high diastereo- and enantioselectivities. Further studies showed the development of an air-stable and storable Zr-BINOL catalyst, remaining unaltered the yield and stereoselectivities of aldol adducts.45


Chiral catalyst 119 was utilized by Inoue and co-workers46 in the total synthesis of the potent toxin antillatoxin (125) (Scheme 20). The aldol reaction between aldehyde 123 and silyl enol ether 122 afforded the intermediate 124 in high diastereo- and enantioselectivity to set up the C4 and C5 stereocenters of antillatoxin (125).


7. Asymmetric Induction in Mukaiyama Aldol Reactions with Bis(oxazolinyl) (BOX) and Bis(oxazolinyl)pyridine (PYBOX) as Chiral Ligands

Early studies involving the C2-symmetric chiral Lewis acid complexes in aldol reactions, such as bis(oxazolinyl) (BOX) and bis(oxazolinyl)pyridine (PYBOX) ligands have been developed by Evans and co-workers.47 In these works, the authors achieved excellent levels of regio-, diastereo- and enantioselectivities using electrophiles capable of chelation, for example, (benzyloxy)acetaldehyde (33) (Table 8).47

As can be seen, the reactions were found to be quite general with respect to the silylketene acetal structure. In all cases, excellent yields were obtained with enantiomeric excesses above 95%.

The requirement for a chelating substituent at the aldehyde partner is critical to catalyst selectivity, as (tert-butyldimethylsiloxy)-acetaldehyde gave low enantioselectivity (ee = 56%). Curiously, β-(benzyloxy)propionaldehyde provided racemic products, indicating a strict requirement for a five-membered catalyst-aldehyde chelate. The observed results can be rationalized based on a pyramidal square transition state TS12 with a penta-coordination geometry (Scheme 21).


As can be seen from the proposed transition state TS12, the aldehyde is preferentially attacked from the Si face, justifying the absolute configuration of the observed aldol adducts.

Evans and co-workers47 have obtained excellent results from the Mukaiyama type aldol reaction between silyl enol ethers and methyl pyruvate (128) in the presence of box complex 130 (Table 9).

Structural variations of the silyl enol ether are possible without loss in enantioselectivity. Both silylketene acetals and ketone-derived enolsilanes afford highly enantioselective additions (Table 9, entries 1-3, ee > 93%). The catalyzed aldol addition of substituted silylketene acetals to pyruvate esters mediated by box complex 130 provides succinate derivatives with high syn diastereoselectivities. The Z- and E-isomers of the illustrated silylketene acetals (entries 4-7) react in a stereoconvergent manner, providing the syn aldol adducts in high diastereo- and enantioselectivity (dr > 94:06 1,2-syn:1,2-anti, ee > 93%).

The observed results can be rationalized by invoking a square planar transition state TS13 (Scheme 22).


As can be observed from the proposed transition state TS13, the Si face is exposed to suffer an attack of the nucleophile, which is consistent with the results.

In addition, Evans and co-workers47 showed that the use of tin C2-symmetric complexes as chiral Lewis acid led to the formation of aldol adducts with 1,2-anti relationship in excellent levels of enantiomeric excess (Table 10).

The catalyzed addition to pyruvates is general with respect to the silyl enol ether. Both E- and Z-isomers of the silyl enol ether (Table 10, entries 1 and 2) react in a stereoconvergent manner providing the substituted succinate derivative with excellent diastereo- and enantioselectivity (dr = 99:01 1,2-anti:1,2-syn, ee > 96%). Variation in the size of the alkyl substituent of the enolsilane is possible without significant loss in stereoselectivity (entries 2-4, dr > 98:2 1,2-anti:1,2-syn, ee > 96%).

In 2002, Evans et al.48 reported the total synthesis of pectenotoxin-4 (139) and pectenotoxin-8 (140) (Scheme 23). The enantioselective Sn2+-catalyzed aldol reaction between silyl enol ether 133 and glyoxylate 134 led to the formation of aldol adduct 135, which corresponds to the C8-C11 fragment of pectenotoxins in excellent yields and enantiomeric excess (Scheme 23). Similarly, the aldol reaction between silyl enol ether 137 and glyoxylate 134, mediated by chiral Lewis acid 132, led to the formation of compound 138 (Scheme 23). This aldol adduct correspond to the C36-C39 fragment of pectenotoxins in excellent yields, diastereo- and enantiomeric excess.


In 2006, Jørgensen and co-workers49 concluded the total synthesis of nonnatural indolizine alkaloid 145 (Scheme 24). For this purpose, the authors performed a Mukaiyama-type aldol reaction between silyl enol ether 142 and aldehyde 141, mediated by a chiral (S,S)-t-Bu-BOX 144, leading to the formation of the aldol adduct 143 in excellent levels of diastereo- and enantiomeric excess (Scheme 24).


In 2006, Movassaghi and co-workers50 concluded the total synthesis of acylfulvene (148) and irofulven (149) (Scheme 25). One of the steps consisted of the Mukaiyama-type aldol reaction between enolsilane 146 and methyl pyruvate 147 mediated by (R,R)-t-Bu-BOX ent-130 leading to the formation of the aldol adduct 147 with excellent levels of enantioselectivities (Scheme 25). This aldol adduct corresponds to the C1-C4 fragment of acylfulvene (148) and irofulven (149).


8. Asymmetric Induction in Mukaiyama Aldol Reactions with Boron-Derivatives as Chiral Ligands

In the early 1990, Masamune and co-workers51 and Corey et al.52 developed, independently, an asymmetric aldol reaction catalyzed by amino acids derived oxazaborolidines (Scheme 26). In these works, the authors achieved good levels of enantioselectivities.


In 2010, Micoine and Fürstner53 concluded the total synthesis of the potent cell migration inhibitor lactimidomycin (159) (Scheme 27). In this paper, the authors performed the late-stage Mukaiyama-type aldol reaction between the silyl enol ether prepared from ketone 156 and aldehyde 157 mediated by oxazaborolidine 158, after work-up with HF-pyridine leading to the formation of lactimidomycin (159) in 60% yield and excellent selectivity.


9. Conclusions

In this review article, our group demonstrated representative examples of metal-mediated catalytic asymmetric aldol reactions. This type of aldol reactions using chiral catalysis are one of the most powerful methods to control the stereochemistry of aldol adducts. One of the major motivations for the development of new enantioselective catalysis is to reach higher catalytic efficiency under very mild reaction conditions. Thus, the design of new chiral catalyst systems continues to be an attractive field in organic synthesis.

Although, the applicability of the most asymmetric catalysis is limited in terms of substrate generality, we can predict a promising future for this field in the light of the search for synthetic ideality54 and the green chemistry.55

Acknowledgements

We are grateful to FAEP-UNICAMP, FAPESP, CNPq, CAPES and INCT-INOFAR (Proc. CNPq 573.564/2008-6) for financial support.

References

1. Cowden, C. J.; Paterson, I.; Org. React. 1997, 51, 1; Franklin, A. S.; Paterson, I.; Contemp. Org. Synth. 1994, 1, 317; Heathcock, C. H. In Comprehensive Organic Synthesis, vol. 2; Heathcock, C. H., ed.; Pergamon Press: New York, USA, 1991; Kim, B. M.; Williams, S. F.; Masamune, S. In Comprehensive Organic Synthesis, vol. 2; Heathcock, C. H., ed.; Pergamon Press: New York, USA, 1991; Evans, D. A.; Nelson, J. V.; Taber, T. R.; Top. Stereochem. 1982, 13, 1; Dias, L. C.; Aguilar, A. M.; Quim. Nova 2007, 30, 2007; Dias, L. C.; Aguilar, A. M.; Chem. Soc. Rev. 2008, 37, 451; Ariza, X.; Garcia, J.; Romea, P.; Urpí, F.; Synthesis 2011, 2175; Mahrwald, R.; Aldol Reaction, vol. 1; Springer-Science: New York, USA, 2009; Mahrwald, R.; Chem. Rev. 1999, 99, 1095; Mengel, A.; Reiser, O.; Chem. Rev. 1999, 99, 1191.

2. Zimmerman, H. E.; Traxler, M. D.; J. Am. Chem. Soc. 1957, 79, 1920.

3. The IUPAC defines asymmetric induction as "The traditional term describing the preferential formation in a chemical reaction of one enantiomer or diastereoisomer over the other as a result of the influence of a chiral feature present in the substrate, reagent, catalyst or environment". IUPAC; Compendium of Chemical Terminology, 2nd ed. (the "Gold Book"); Compiled by McNaught, A. D.; Wilkinson, A.; Blackwell Scientific Publications: Oxford, 1997. XML on-line corrected version http://goldbook.iupac.org (2006-) created by Nic, M.; Jirat, J.; Kosata, B.; updates compiled by Jenkins, A; ISBN 0-9678550-9-8 DOI 10.1351/goldbook.

4. For recent reviews on catalytic aldol reactions, see: Alcaide, B.; Almendros, P.; Eur. J. Org. Chem. 2002, 1595; Shibasaki, M.; Matsunaga, S.; Kumagai, N. In Modern Aldol Reactions, vol. 2; Mahrward, R., ed.; Wiley-VCH Verlag: Weinheim, Germany, 2004; Denmark, S. E.; Fujimori, S. In Modern Aldol Reactions, vol. 2; Mahrward, R., ed.; Wiley-VCH Verlag: Weinheim, Germany, 2004.

5. Organocatalytic processes are beyond the scope of this review. For recent reviews in this area, see: Notz, W.; Tanaka, F.; Barbas III, C. F.; Acc. Chem. Res. 2004, 37, 580; Mukherjee, S.; Yang, W. J.; Hoffmann, S.; List, B.; Chem. Rev. 2007, 107, 5471; Amarante, G. W.; Coelho, F.; Quim. Nova 2009, 32, 469; Bhanushali, M.; Zhao, C. G.; Synthesis, 2011, 1815; Pansare, S. V.; Paul, E. K.; Chem. Eur. J. 2011, 17, 8770; Taylor, J. E.; Bull, S. D.; Williams, J. M. J.; Chem. Soc. Rev. 2012, 41, 2109; Melchiorre, P.; Angew. Chem., Int. Ed. 2012, 51, 9748; Bisai, V.; Synthesis, 2012, 44, 1453; Ávila, E. P.; Amarante, G. W.; ChemCatChem 2012, 4, 1713; Mahlau, M.; List, B.; Angew. Chem., Int. Ed. 2013, 52, 518.

6. Kobayashi, S.; Sugiura, M.; Kitagawa, H.; Lam, W. W.-L.; Chem. Rev. 2002, 102, 2227.

7. Kobayashi, S.; Hamada, T.; Nagayama, S.; Manabe, K.; Org. Lett. 2001, 3, 165; Hamada, T.; Manabe, K.; Ishikawa, S.; Nagayama, S.; Shira, M.; Kobayashi, S.; J. Am. Chem. Soc. 2003, 125, 2989.

8. Mei, Y.; Dissanayake, P.; Allen, M. J.; J. Am. Chem. Soc. 2010, 132, 12871.

9. Mei, Y.; Averill, D. J.; Allen, M. J.; J. Org. Chem. 2012, 77, 5624.

10. Ngai, M.-Y.; Kong, J.-R.; Krische, M. J.; J. Org. Chem. 2007, 72, 1063; Adachi, S.; Harada, T.; Eur. J. Org. Chem. 2009, 3661.

11. Nishiyama, H.; Shiomi, T.; Top. Curr. Chem. 2007, 279, 105.

12. Taylor, S. J.; Duffey, M. O.; Morken, J. P.; J. Am. Chem. Soc. 2000, 122, 4528.

13. Russell, A. E.; Fuller, N.; Taylor, S. J.; Aurriset, P.; Morken, J. P.; Org. Lett. 2004, 6, 2309.

14. Zhao, C.-X.; Duffey, M. O.; Taylor, S. J.; Morken, J. P.; Org. Lett. 2001, 3, 1829.

15. Duffey, M. O.; LeTiran, A.; Morken, J. P.; J. Am. Chem. Soc. 2003, 125, 1458; Duffey, M. O.; LeTiran, A.; Morken, J. P.; J. Am. Chem. Soc. 2003, 125, 12666.

16. Bee, C.; Han, S. B.; Hassan, A.; Lida, H.; Krische, M. J.; J. Am. Chem. Soc. 2008, 130, 2746; Han, S. B.; Hassan, A.; Krische, M. J.; Synthesis 2008, 17, 2669.

17. Nishiyama, H.; Shiomi, T.; Tsuchiya, Y.; Matsuda, I.; J. Am. Chem. Soc. 2005, 127, 6972.

18. Yang, Y.-F.; Shi, T.; Zhang, X.-H.; Tang, Z.-X.; Wen, Z.-Y.; Quan, J.-M.; Wu, Y.-D.; Org. Biomol. Chem. 2011, 9, 5845.

19. Denmark, S. E.; Fujimori, S. In Modern Aldol Reactions, vol. 2; Mahrward, R., ed.; Wiley-VCH Verlag: Weinheim, Germany, 2004.

20. Denmark, S. E.; Beutner, G. L.; Angew. Chem., Int. Ed. 2008, 47, 1560.

21. Denmark, S. E.; Winter, S. B. D.; Su, X.; Wong, K. T.; J. Am. Chem. Soc. 1996, 118, 7404.

22. Nakajima, M.; Orito, Y.; Ishizuka, T.; Hashimoto, S.; Org. Lett. 2004, 6, 3763; Ichibakase, T.; Orito, Y.; Nakajima, M.; Tetrahedron Lett. 2008, 49, 4427.

23. Wadamoto, M.; Ozasa, N.; Yanagisawa, A.; Yamamoto, H.; J. Org. Chem. 2003, 68, 5593.

24. Denmark, S. E.; Chung, Chung, W.-j; Angew. Chem., Int. Ed. 2008, 47, 1890.

25. Denmark, S. E.; Chung, W.-j; J. Org. Chem. 2008, 73, 4582.

26. Osakama, K.; Sugiura, M.; Nakajima, M.; Kotani, S.; Tetrahedron Lett. 2012, 53, 4199.

27. Trost, B. M.; Brindle, C. S.; Chem. Soc. Rev. 2010, 39, 1600.

28. Pousse, G.; Cavelier, F. L.; Humphreys, L.; Rouden, J.; Blanchet, J.; Org. Lett. 2010, 12, 3582.

29. Yamada, Y. M. A.; Yoshikawa, N.; Sasai, H.; Shibasaki, M.; Angew. Chem., Int. Ed. 1997, 36, 1871.

30. Yoshikawa, N.; Yamada, Y. M. A.; Das, J.; Sasai, H.; Shibasaki, M.; J. Am. Chem. Soc. 1999, 121, 4168.

31. Maki, K.; Motoki, R.; Fujii, K.; Kanai, M.; Kobayashi, T.; Tamura, S.; Shibasaki, M.; J. Am. Chem. Soc. 2005, 127, 17111; Fujii, K.; Maki, K.; Kanai, M.; Shibasaki, M.; Org. Lett. 2003, 5, 733.

32. Trost, B. M.; Ito, H.; J. Am. Chem. Soc. 2000, 122, 12003; Trost, B. M.; Ito, H.; Silcoff, E. R.; J. Am. Chem. Soc. 2001, 123, 3367; Trost, B. M.; Silcoff, E. R.; Ito, H.; Org. Lett. 2001, 3, 2497; Trost, B. M.; Fettes, A.; Shireman, B. T.; J. Am. Chem. Soc. 2004, 126, 2660; Trost, B. M.; Malhotra, S.; Fried, B. A.; J. Am. Chem. Soc. 2009, 131, 1674.

33. Trost, B. M.; Frederiksen, M. U.; Papillon, J. P. N.; Harrington, P. E.; Shin, S.; Shireman, B. T.; J. Am. Chem. Soc. 2005, 127, 3666.

34. Iwata, M.; Yazaki, R.; Suzuki, Y.; Kumagai, N.; Shibasaki, M.; J. Am. Chem. Soc. 2009, 131, 18244.

35. Suzuki, Y.; Iwata, M.; Yazaki, R.; Kumagai, N.; Shibasaki, M.; J. Org. Chem. 2012, 77, 4496.

36. Reetz, M. T.; Kyong, S. H.; Bolm, C.; Zierke, T.; Chem. Ind. 1986, 824.

37. Mikami, K.; Matsukawa, S.; J. Am. Chem. Soc. 1994, 116, 4077.

38. Keck, G. E.; Krishnamurthy, D.; J. Am. Chem. Soc. 1995, 117, 2363.

39. Carreira, E. M.; Singer, R. A.; Lee, W.; J. Am. Chem. Soc. 1994, 116, 8837; Singer, R. A.; Shepard, M. S.; Carreira, E. M.;Tetrahedron 1998, 54, 7025.

40. Singer, R. A.; Carreira, E. M.; Tetrahedron Lett. 1997, 38, 927.

41. Li, K. W.; Wu, J.; Xing, W.; Simon, J. A.; J. Am. Chem. Soc. 1996, 118, 7237.

42. Rychnovsky, S. D.; Khire, U. R.; Yang, G.; J. Am. Chem. Soc. 1997, 119, 2058.

43. Ishitani, H.; Yamashita, Y.; Shimizu, H.; Kobayashi, S.; J. Am. Chem. Soc. 2000, 122, 5403.

44. Yamashita, Y.; Ishitani, H.; Shimizu, H.; Kobayashi, S.; J. Am. Chem. Soc. 2002, 124, 3292.

45. Kobayashi, S.; Saito, S.; Ueno, M.; Yamashita, Y.; Chem. Commun. 2003, 2016; Seki, K.; Ueno, M.; Kobayashi, S.; Org. Biomol. Chem. 2007, 5, 1347.

46. Okura, K.; Matsuoka, S.; Goto, R.; Inoue, M.; Angew. Chem., Int. Ed. 2010, 49, 329.

47. Evans, D. A.; Murry, J. A.; Kozlowski, M. C.; J. Am. Chem. Soc. 1996, 118, 5814; Evans, D. A.; Kozlowski, M. C.; Burgey, C. S.; MacMillan, D. W. C.; J. Am. Chem. Soc. 1997, 119, 7893; Evans, D. A.; MacMillan, D. W. C.; Campos, K. R.; J. Am. Chem. Soc. 1997, 119, 10859; Evans, D. A.; Kozlowski, M. C.; Murry, J. A.; Burgey, C. S.; Campos, K. R.; Connell, B. T.; Staples, R. J.; J. Am. Chem. Soc. 1999, 121, 669; Evans, D. A.; Burgey, C. S.; Kozlowski, M. C.; Tregay, S. W.; J. Am. Chem. Soc. 1999, 121, 686; Evans, D. A.; Masse, C. E.; Wu, J.; Org. Lett. 2002, 4, 3375.

48. Evans, D. A.; Rajapakse, H. A.; Stenkamp, D.; Angew. Chem., Int. Ed. 2002, 41, 4569; Evans, D. A.; Rajapakse, H. A.; Chiu, A.; Stenkamp, D.; Angew. Chem., Int. Ed. 2002, 41, 4573.

49. Landa, A.; Minkkilä, A.; Blay, G.; Jørgensen, K. A.; Chem. Eur. J. 2006, 12, 3472.

50. Movassaghi, M.; Piizzi, G.; Siegel, D. S.; Piersanti, G.; Angew. Chem., Int. Ed. 2006, 45, 5859; Siegel, D. S.; Piizzi, G.; Piersanti, G.; Movassaghi, M.; Piizzi, G.; J. Org. Chem. 2009, 74, 9292.

51. Parmee, E. R.; Tempkin, O.; Masamune, S.; Abiko, A.; J. Am. Chem. Soc. 1991, 113, 9365.

52. Corey, E. J.; Cywin, C. L.; Roper, T. D.; Tetrahedron Lett. 1992, 33, 6907.

53. Micoine, K.; Fürstner, A. J. Am. Chem. Soc.; 2010, 132, 14064.

54. Hendrickson, J. B.; J. Am. Chem. Soc. 1975, 97, 5784; Trost, B. M.; Science 1991, 254, 1471; Trost, B. M.; Angew. Chem., Int. Ed. Engl. 1995, 34, 259; Wender, P. A.; Croatt, M. P.; Witulski, B.; Tetrahedron 2006, 62, 7505; Wender, P. A.; Verma, V. A.; Paxton, T. J.; Pillow, T. H.; Acc. Chem. Res. 2008, 41, 40; Wender, P. A.; Miller, B. L.; Nature 2009, 460, 197; Burns, N. Z.; Baran, P. S.; Hoffmann, R. W.; Angew. Chem., Int. Ed. 2009, 48, 2854; Young, I. S.; Baran, P. S.; Nature Chem. 2009, 1, 193; Shenvi, R. A.; O'Malley, D. P.; Baran, P. S.; Acc. Chem. Res. 2009, 42, 530; Gaich, T.; Baran, P. S.; J. Org. Chem. 2010, 75, 4657.

55. Anastas, P. T.; Warner, J. C.; Green Chemistry: Theory and Practice, vol. 1; Oxford University Press: New York, USA, 1998.

Submitted: December 9, 2012

Published online: January 22, 2013

Luiz C. Dias was born in Balneário Camboriú, Santa Catarina State, Brazil. He has a degree in Chemistry (Federal University of Santa Catarina (UFSC, Brazil, 1988), a PhD in Chemistry at the University of Campinas (UNICAMP, Brazil, 1993) and was a Post-doctoral Fellow (Harvard University, USA, 1994-1995). He is a Full Professor at the Institute of Chemistry, UNICAMP, and CNPq (Brazilian National Council for Technological and Scientific Development) Researcher 1A. He has experience in the area of total synthesis of compounds with pharmacological activity, control of relative stereochemistry in acyclic systems and theoretical Chemistry (NBO). Professor Dias has more than 85 publications and two patents, and delivered more than 120 lectures. He concluded the guidance of 21 Master degree students, 11 PhD students and supervised ten Post-doctoral fellows. He was the General Secretary of the Brazilian Chemical Society (SBQ) for four years, Editor of the Electronic Bulletin of SBQ for 6 years. He was a member of the Advisory Board of Chemistry for CNPq and an Assistant Coordinator for the Chemistry Area of the Brazilian Federal Agency for Support and Evaluation of Graduate Education (CAPES). He is currently an Editor of the Journal of the Brazilian Chemical Society (JBCS) and the coordinator of the CAPES Chemistry Area. In 1999, he received the JBCS medal and the Academic Recognition Award Zeferino Vaz (UNICAMP) in 2009. He is a member of the Academy of Sciences of the São Paulo State (ACIESP, 2009), a Commander of the National Order of Scientific Merit (Presidency of the Republic, 2010) and Member of the Brazilian Academy of Sciences (ABC, 2011).

Emílio C. de Lucca Jr. was born in São Roque, São Paulo State, Brazil, in 1986. He earned a BSc in Chemistry at the University of São Paulo (FFCLRP, USP, Brazil, 2008) and MSc degree at the University of Campinas (UNICAMP, Brazil, 2011) under the supervision of Prof. Dias. Emílio is currently working at his PhD in the Prof. Dias lab at the UNICAMP.

Marco A. B. Ferreira was born in Ribeirão Preto, São Paulo State, Brazil, in 1983. He received his BSc degree in Chemistry from the University of São Paulo (FFCLRP, USP, Brazil, 2006). From 2006 to 2008, Marco held a Master degree at the University of Campinas (IQ, UNICAMP, Brazil), under Prof. Dias guidance. Also in the same research group, Marco held his PhD, completed in 2012. After that, he began his post-doctoral studies in the laboratory of Professors Roberto R. Neto and Claudio F. Tormena (UNICAMP), where he is currently working on the development of new methods to determine the relative stereochemistry of organic systems by applying methods of quantum chemistry. His research interests are focused on the development of new synthetic methodologies and their application in the total synthesis of natural products with potential biological activity.

Ellen C. Polo was born in São Bernardo do Campo, São Paulo State, Brazil, in 1985. She has a degree in Chemistry from the University of São Paulo (FFCLRP, USP, Brazil, 2007) and earned her MSc degree at the University of Campinas (IQ, UNICAMP, Brazil) in 2011 under the supervision of Prof. Dias. She is currently pursuing her doctorate in the same lab at the UNICAMP.

  • 1. Cowden, C. J.; Paterson, I.; Org. React. 1997, 51, 1;
  • Franklin, A. S.; Paterson, I.; Contemp. Org. Synth. 1994, 1, 317;
  • Heathcock, C. H. In Comprehensive Organic Synthesis, vol. 2;
  • Heathcock, C. H., ed.; Pergamon Press: New York, USA, 1991;
  • Kim, B. M.; Williams, S. F.; Masamune, S. In Comprehensive Organic Synthesis, vol. 2;
  • Heathcock, C. H., ed.; Pergamon Press: New York, USA, 1991;
  • Evans, D. A.; Nelson, J. V.; Taber, T. R.; Top. Stereochem. 1982, 13, 1;
  • Dias, L. C.; Aguilar, A. M.; Quim. Nova 2007, 30, 2007;
  • Dias, L. C.; Aguilar, A. M.; Chem. Soc. Rev. 2008, 37, 451;
  • Ariza, X.; Garcia, J.; Romea, P.; Urpí, F.; Synthesis 2011, 2175;
  • Mahrwald, R.; Aldol Reaction, vol. 1; Springer-Science: New York, USA, 2009;
  • Mahrwald, R.; Chem. Rev. 1999, 99, 1095;
  • Mengel, A.; Reiser, O.; Chem. Rev. 1999, 99, 1191.
  • 2. Zimmerman, H. E.; Traxler, M. D.; J. Am. Chem. Soc. 1957, 79, 1920.
  • 3. The IUPAC defines asymmetric induction as "The traditional term describing the preferential formation in a chemical reaction of one enantiomer or diastereoisomer over the other as a result of the influence of a chiral feature present in the substrate, reagent, catalyst or environment". IUPAC; Compendium of Chemical Terminology, 2nd ed. (the "Gold Book"); Compiled by McNaught, A. D.; Wilkinson, A.; Blackwell Scientific Publications: Oxford, 1997.
  • XML on-line corrected version http://goldbook.iupac.org (2006-) created by Nic, M.; Jirat, J.; Kosata, B.; updates compiled by Jenkins, A; ISBN 0-9678550-9-8 DOI 10.1351/goldbook.
    » link
  • 4. For recent reviews on catalytic aldol reactions, see: Alcaide, B.; Almendros, P.; Eur. J. Org. Chem. 2002, 1595;
  • Shibasaki, M.; Matsunaga, S.; Kumagai, N. In Modern Aldol Reactions, vol. 2;
  • Mahrward, R., ed.; Wiley-VCH Verlag: Weinheim, Germany, 2004;
  • Denmark, S. E.; Fujimori, S. In Modern Aldol Reactions, vol. 2;
  • Mahrward, R., ed.; Wiley-VCH Verlag: Weinheim, Germany, 2004.
  • 5. Organocatalytic processes are beyond the scope of this review. For recent reviews in this area, see: Notz, W.; Tanaka, F.; Barbas III, C. F.; Acc. Chem. Res. 2004, 37, 580;
  • Mukherjee, S.; Yang, W. J.; Hoffmann, S.; List, B.; Chem. Rev. 2007, 107, 5471;
  • Amarante, G. W.; Coelho, F.; Quim. Nova 2009, 32, 469;
  • Bhanushali, M.; Zhao, C. G.; Synthesis, 2011, 1815;
  • Pansare, S. V.; Paul, E. K.; Chem. Eur. J. 2011, 17, 8770;
  • Taylor, J. E.; Bull, S. D.; Williams, J. M. J.; Chem. Soc. Rev. 2012, 41, 2109;
  • Melchiorre, P.; Angew. Chem., Int. Ed. 2012, 51, 9748;
  • Bisai, V.; Synthesis, 2012, 44, 1453;
  • Ávila, E. P.; Amarante, G. W.; ChemCatChem 2012, 4, 1713;
  • Mahlau, M.; List, B.; Angew. Chem., Int. Ed. 2013, 52, 518.
  • 6. Kobayashi, S.; Sugiura, M.; Kitagawa, H.; Lam, W. W.-L.; Chem. Rev. 2002, 102, 2227.
  • 7. Kobayashi, S.; Hamada, T.; Nagayama, S.; Manabe, K.; Org. Lett. 2001, 3, 165;
  • Hamada, T.; Manabe, K.; Ishikawa, S.; Nagayama, S.; Shira, M.; Kobayashi, S.; J. Am. Chem. Soc. 2003, 125, 2989.
  • 8. Mei, Y.; Dissanayake, P.; Allen, M. J.; J. Am. Chem. Soc. 2010, 132, 12871.
  • 9. Mei, Y.; Averill, D. J.; Allen, M. J.; J. Org. Chem. 2012, 77, 5624.
  • 10. Ngai, M.-Y.; Kong, J.-R.; Krische, M. J.; J. Org. Chem. 2007, 72, 1063;
  • Adachi, S.; Harada, T.; Eur. J. Org. Chem. 2009, 3661.
  • 11. Nishiyama, H.; Shiomi, T.; Top. Curr. Chem. 2007, 279, 105.
  • 12. Taylor, S. J.; Duffey, M. O.; Morken, J. P.; J. Am. Chem. Soc. 2000, 122, 4528.
  • 13. Russell, A. E.; Fuller, N.; Taylor, S. J.; Aurriset, P.; Morken, J. P.; Org. Lett. 2004, 6, 2309.
  • 14. Zhao, C.-X.; Duffey, M. O.; Taylor, S. J.; Morken, J. P.; Org. Lett. 2001, 3, 1829.
  • 15. Duffey, M. O.; LeTiran, A.; Morken, J. P.; J. Am. Chem. Soc. 2003, 125, 1458;
  • Duffey, M. O.; LeTiran, A.; Morken, J. P.; J. Am. Chem. Soc. 2003, 125, 12666.
  • 16. Bee, C.; Han, S. B.; Hassan, A.; Lida, H.; Krische, M. J.; J. Am. Chem. Soc. 2008, 130, 2746;
  • Han, S. B.; Hassan, A.; Krische, M. J.; Synthesis 2008, 17, 2669.
  • 17. Nishiyama, H.; Shiomi, T.; Tsuchiya, Y.; Matsuda, I.; J. Am. Chem. Soc. 2005, 127, 6972.
  • 18. Yang, Y.-F.; Shi, T.; Zhang, X.-H.; Tang, Z.-X.; Wen, Z.-Y.; Quan, J.-M.; Wu, Y.-D.; Org. Biomol. Chem. 2011, 9, 5845.
  • 19. Denmark, S. E.; Fujimori, S. In Modern Aldol Reactions, vol. 2; Mahrward, R., ed.; Wiley-VCH Verlag: Weinheim, Germany, 2004.
  • 20. Denmark, S. E.; Beutner, G. L.; Angew. Chem., Int. Ed. 2008, 47, 1560.
  • 21. Denmark, S. E.; Winter, S. B. D.; Su, X.; Wong, K. T.; J. Am. Chem. Soc. 1996, 118, 7404.
  • 22. Nakajima, M.; Orito, Y.; Ishizuka, T.; Hashimoto, S.; Org. Lett. 2004, 6, 3763;
  • Ichibakase, T.; Orito, Y.; Nakajima, M.; Tetrahedron Lett. 2008, 49, 4427.
  • 23. Wadamoto, M.; Ozasa, N.; Yanagisawa, A.; Yamamoto, H.; J. Org. Chem. 2003, 68, 5593.
  • 24. Denmark, S. E.; Chung, Chung, W.-j; Angew. Chem., Int. Ed. 2008, 47, 1890.
  • 25. Denmark, S. E.; Chung, W.-j; J. Org. Chem. 2008, 73, 4582.
  • 26. Osakama, K.; Sugiura, M.; Nakajima, M.; Kotani, S.; Tetrahedron Lett. 2012, 53, 4199.
  • 27. Trost, B. M.; Brindle, C. S.; Chem. Soc. Rev. 2010, 39, 1600.
  • 28. Pousse, G.; Cavelier, F. L.; Humphreys, L.; Rouden, J.; Blanchet, J.; Org. Lett. 2010, 12, 3582.
  • 29. Yamada, Y. M. A.; Yoshikawa, N.; Sasai, H.; Shibasaki, M.; Angew. Chem., Int. Ed. 1997, 36, 1871.
  • 30. Yoshikawa, N.; Yamada, Y. M. A.; Das, J.; Sasai, H.; Shibasaki, M.; J. Am. Chem. Soc. 1999, 121, 4168.
  • 31. Maki, K.; Motoki, R.; Fujii, K.; Kanai, M.; Kobayashi, T.; Tamura, S.; Shibasaki, M.; J. Am. Chem. Soc. 2005, 127, 17111;
  • Fujii, K.; Maki, K.; Kanai, M.; Shibasaki, M.; Org. Lett. 2003, 5, 733.
  • 32. Trost, B. M.; Ito, H.; J. Am. Chem. Soc. 2000, 122, 12003;
  • Trost, B. M.; Ito, H.; Silcoff, E. R.; J. Am. Chem. Soc. 2001, 123, 3367;
  • Trost, B. M.; Silcoff, E. R.; Ito, H.; Org. Lett. 2001, 3, 2497;
  • Trost, B. M.; Fettes, A.; Shireman, B. T.; J. Am. Chem. Soc. 2004, 126, 2660;
  • Trost, B. M.; Malhotra, S.; Fried, B. A.; J. Am. Chem. Soc. 2009, 131, 1674.
  • 33. Trost, B. M.; Frederiksen, M. U.; Papillon, J. P. N.; Harrington, P. E.; Shin, S.; Shireman, B. T.; J. Am. Chem. Soc. 2005, 127, 3666.
  • 34. Iwata, M.; Yazaki, R.; Suzuki, Y.; Kumagai, N.; Shibasaki, M.; J. Am. Chem. Soc. 2009, 131, 18244.
  • 35. Suzuki, Y.; Iwata, M.; Yazaki, R.; Kumagai, N.; Shibasaki, M.; J. Org. Chem. 2012, 77, 4496.
  • 36. Reetz, M. T.; Kyong, S. H.; Bolm, C.; Zierke, T.; Chem. Ind. 1986, 824.
  • 37. Mikami, K.; Matsukawa, S.; J. Am. Chem. Soc. 1994, 116, 4077.
  • 38. Keck, G. E.; Krishnamurthy, D.; J. Am. Chem. Soc. 1995, 117, 2363.
  • 39. Carreira, E. M.; Singer, R. A.; Lee, W.; J. Am. Chem. Soc. 1994, 116, 8837;
  • Singer, R. A.; Shepard, M. S.; Carreira, E. M.;Tetrahedron 1998, 54, 7025.
  • 40. Singer, R. A.; Carreira, E. M.; Tetrahedron Lett. 1997, 38, 927.
  • 41. Li, K. W.; Wu, J.; Xing, W.; Simon, J. A.; J. Am. Chem. Soc. 1996, 118, 7237.
  • 42. Rychnovsky, S. D.; Khire, U. R.; Yang, G.; J. Am. Chem. Soc. 1997, 119, 2058.
  • 43. Ishitani, H.; Yamashita, Y.; Shimizu, H.; Kobayashi, S.; J. Am. Chem. Soc. 2000, 122, 5403.
  • 44. Yamashita, Y.; Ishitani, H.; Shimizu, H.; Kobayashi, S.; J. Am. Chem. Soc. 2002, 124, 3292.
  • 45. Kobayashi, S.; Saito, S.; Ueno, M.; Yamashita, Y.; Chem. Commun. 2003, 2016;
  • Seki, K.; Ueno, M.; Kobayashi, S.; Org. Biomol. Chem 2007, 5, 1347.
  • 46. Okura, K.; Matsuoka, S.; Goto, R.; Inoue, M.; Angew. Chem., Int. Ed. 2010, 49, 329.
  • 47. Evans, D. A.; Murry, J. A.; Kozlowski, M. C.; J. Am. Chem. Soc. 1996, 118, 5814;
  • Evans, D. A.; Kozlowski, M. C.; Burgey, C. S.; MacMillan, D. W. C.; J. Am. Chem. Soc. 1997, 119, 7893;
  • Evans, D. A.; MacMillan, D. W. C.; Campos, K. R.; J. Am. Chem. Soc. 1997, 119, 10859;
  • Evans, D. A.; Kozlowski, M. C.; Murry, J. A.; Burgey, C. S.; Campos, K. R.; Connell, B. T.; Staples, R. J.; J. Am. Chem. Soc. 1999, 121, 669;
  • Evans, D. A.; Burgey, C. S.; Kozlowski, M. C.; Tregay, S. W.; J. Am. Chem. Soc. 1999, 121, 686;
  • Evans, D. A.; Masse, C. E.; Wu, J.; Org. Lett. 2002, 4, 3375.
  • 48. Evans, D. A.; Rajapakse, H. A.; Stenkamp, D.; Angew. Chem., Int. Ed. 2002, 41, 4569;
  • Evans, D. A.; Rajapakse, H. A.; Chiu, A.; Stenkamp, D.; Angew. Chem., Int. Ed. 2002, 41, 4573.
  • 49. Landa, A.; Minkkilä, A.; Blay, G.; Jørgensen, K. A.; Chem. Eur. J. 2006, 12, 3472.
  • 50. Movassaghi, M.; Piizzi, G.; Siegel, D. S.; Piersanti, G.; Angew. Chem., Int. Ed. 2006, 45, 5859;
  • Siegel, D. S.; Piizzi, G.; Piersanti, G.; Movassaghi, M.; Piizzi, G.; J. Org. Chem. 2009, 74, 9292.
  • 51. Parmee, E. R.; Tempkin, O.; Masamune, S.; Abiko, A.; J. Am. Chem. Soc. 1991, 113, 9365.
  • 52. Corey, E. J.; Cywin, C. L.; Roper, T. D.; Tetrahedron Lett. 1992, 33, 6907.
  • 53. Micoine, K.; Fürstner, A. J. Am. Chem. Soc.; 2010, 132, 14064.
  • 54. Hendrickson, J. B.; J. Am. Chem. Soc. 1975, 97, 5784;
  • Trost, B. M.; Science 1991, 254, 1471;
  • Trost, B. M.; Angew. Chem., Int. Ed. Engl. 1995, 34, 259;
  • Wender, P. A.; Croatt, M. P.; Witulski, B.; Tetrahedron 2006, 62, 7505;
  • Wender, P. A.; Verma, V. A.; Paxton, T. J.; Pillow, T. H.; Acc. Chem. Res. 2008, 41, 40;
  • Wender, P. A.; Miller, B. L.; Nature 2009, 460, 197;
  • Burns, N. Z.; Baran, P. S.; Hoffmann, R. W.; Angew. Chem., Int. Ed. 2009, 48, 2854;
  • Young, I. S.; Baran, P. S.; Nature Chem. 2009, 1, 193;
  • Shenvi, R. A.; O'Malley, D. P.; Baran, P. S.; Acc. Chem. Res. 2009, 42, 530;
  • Gaich, T.; Baran, P. S.; J. Org. Chem. 2010, 75, 4657.
  • 55. Anastas, P. T.; Warner, J. C.; Green Chemistry: Theory and Practice, vol. 1; Oxford University Press: New York, USA, 1998.
  • *
    e-mail:
    FAPESP has sponsored the publication of this article.
  • Publication Dates

    • Publication in this collection
      08 Feb 2013
    • Date of issue
      Dec 2012

    History

    • Received
      09 Dec 2012
    • Accepted
      22 Jan 2013
    Sociedade Brasileira de Química Instituto de Química - UNICAMP, Caixa Postal 6154, 13083-970 Campinas SP - Brazil, Tel./FAX.: +55 19 3521-3151 - São Paulo - SP - Brazil
    E-mail: office@jbcs.sbq.org.br