Big Chemical Encyclopedia

Chemical substances, components, reactions, process design ...

Articles Figures Tables About

Silyl ketene acetals reaction with aldehydes, diastereoselectivity

Lewis-acid-promoted alkylations of silylenol ethers and silyl ketene acetals [195] with Co-complexed acetylenic acetals [196] and acetylenic aldehydes [197,198] (Scheme 4-56) also proceed with fair to excellent syn diastereoselectivity, in contrast to the low selectivity reactions of the free acetylenic derivatives [199, 200]. Reactions of the complexed aldehydes with lithium enolates are stereospecific, with (Z)-enolates giving syn selectivity and ( )-enolates anti selectivity [201]. The complementary stereoselectivity of the crossed aldol reactions of free and cobalt-complexed propynals with silyl ketene 0,S-acetals has been elaborated by Hanoaka exclusive syn selectivity is exhibited by the complexes and high anti selectivity is found with pro-... [Pg.125]

Besides their application in asymmetric alkylation, sultams can also be used as good chiral auxiliaries for asymmetric aldol reactions, and a / -product can be obtained with good selectivity. As can be seen in Scheme 3-14, reaction of the propionates derived from chiral auxiliary R -OH with LICA in THF affords the lithium enolates. Subsequent reaction with TBSC1 furnishes the 0-silyl ketene acetals 31, 33, and 35 with good yields.31 Upon reaction with TiCU complexes of an aldehyde, product /i-hydroxy carboxylates 32, 34, and 36 are obtained with high diastereoselectivity and good yield. Products from direct aldol reaction of the lithium enolate without conversion to the corresponding silyl ethers show no stereoselectivity.32... [Pg.148]

Denmark utilized chiral base promoted hypervalent silicon Lewis acids for several highly enantioselective carbon-carbon bond forming reactions [92-98]. In these reactions, a stoichiometric quantity of silicon tetrachloride as achiral weak Lewis acid component and only catalytic amount of chiral Lewis base were used. The chiral Lewis acid species desired for the transformations was generated in situ. The phosphoramide 35 catalyzed the cross aldolization of aromatic aldehydes as well as aliphatic aldehydes with a silyl ketene acetal (Scheme 26) [93] with good yield and high enantioselectivity and diastereoselectivity. [Pg.362]

Three years after the discovery of the asymmetric BINOL phosphate-catalyzed Mannich reactions of silyl ketene acetals or acetyl acetone, the Gong group extended these transformations to the use of simple ketones as nucleophiles (Scheme 25) [44], Aldehydes 40 reacted with aniline (66) and ketones 67 or 68 in the presence of chiral phosphoric acids (R)-3c, (/ )-14b, or (/ )-14c (0.5-5 mol%, R = Ph, 4-Cl-CgH ) to give P-amino carbonyl compounds 69 or 70 in good yields (42 to >99%), flnfi-diastereoselectivities (3 1-49 1), and enantioselectivities (72-98% ee). [Pg.416]

First, chemoselective (Chapter 24) conjugate addition of the silyl ketene acetal on the enone is preferred to direct aldol reaction with the aldehyde. Then an aldol reaction of the intermediate silyl enol ether on the benzaldehyde follows. The stereoselectivity results, firstly, from attack of benzalde-hyde on the less hindered face of the intermediate silyl enol ether, which sets the two side chains trans on the cyclohexanone, and, secondly, from the intrinsic diastereoselectivity of the aldol reaction (this is treated in some detail in Chapter 34). This is a summary mechanism. [Pg.757]

Aldol Reactions of Ester Derivatives. The Titanium(IV) C/tlor/dc-catalyzed addition of aldehydes to 0-silyl ketene acetals derived from acetate and propionate esters proceeds with high stereoselectivity. Formation of the silyl ketene acetal was found to be essential for high diastereoselectivity. Treatment of the silyl ketene acetal, derived from deprotonation of the acetate ester with LICA in THF and silyl trapping, with a corresponding aldehyde in the presence of TiCU (1.1 equiv) afforded the addition products in 93 7 diastereoselectivity and moderate yield (51-67%). Similarly, the propionate ester provides the anti-aldol product in high antilsyn selectivity (14 1) and facial selectivity (eq 4). [Pg.279]

Evans et al. recently reported the use of structurally well-defined Sn(II) Lewis acids for the enantioselective aldol addition reactions of a-heterosubstituted substrates [47]. These complexes are readily assembled from Sn(OTf)2 and C2-symmetric bis(oxazoline) ligands. The facile synthesis of these ligands commences with optically active 1,2-diamino alcohols, which are themselves readily available from the corresponding a-amino acids. The Sn(II)-bis(oxazoline) complexes were shown to function optimally as catalysts for enantioselective aldol addition reactions with aldehydes and ketone substrates that are suited to putatively chelate the Lewis acid. For example, use of 10 mol % Sn(II) catalyst, thioacetate, and thiopropionate derived silyl ketene acetals added at -78 °C in dichloromethane to glyoxaldehyde to give hydroxy diesters in superb yields, enantioselectivity, and diastereoselectivity (Eq. 27). The process represents an unusual example wherein 2,3-ant/-aldol adducts are obtained stereoselec-tively. [Pg.406]

If rhodium enolates are used in a catalytic cycle they can promote aldol reactions under reasonably mild conditions. For example, the aldol reactions of trimethylsilyl enol ethers and ketene silyl acetals (37) with aldehydes can be catalyzed by various rhodium(I) complexes, under essentially neutral conditions, to give p-trimethylsiloxy ketones and esters (38 equation 14 and Table 6). The study of Matsuda and coworkers suggests that use of the rhodium complex Rlu(CO)i2 (39 at 2 mol %) in benzene at 100 C gives best results for the formation of adduct (38 Table 6, entries 1-7). There is negligible diastereoselectivity in most cases. Various cationic ihodium complexes such as (40) also catalyze the reaction. Reetz and Vougioukas have found that this aldol reaction proceeds well with the more reactive ketene silyl acetals, (37) for R = OMe or OEt, in CH2CI2 at room temperature (Table 6, entries 8-13). The intermediacy of an ti -O-bound rhodium enolate, such as (41), in the catalytic cycle is like-... [Pg.310]

Six-membered chiral acetals, derived from aliphatic aldehydes, undergo aldol-type coupling reactions with a-silyl ketones, silyl enol ethers," and with silyl ketene acetals " in the presence of titanium tetrachloride with high diastereoselectivities (equation 41) significant results are reported in Table 20. This procedure, in combination with oxidative destructive elimination of the chiral auxiliary, has been applied... [Pg.650]

Other reports deal with a pyrrolidine-catalysed homo-aldol condensation of aliphatic aldehydes (further accelerated by benzoic acid), a diastereoselective aldol-type addition of chiral boron azaenolates to ketones,the use of TMS chloride as a catalyst for TiCU-mediated aldol and Claisen condensations, a boron-mediated double aldol reaction of carboxylic esters, gas-phase condensation of acetone and formaldehyde to give methyl vinyl ketone, and ab initio calculations on the borane-catalysed reaction between formaldehyde and silyl ketene acetal [H2C=C(OH)OSiH3]. ... [Pg.24]

Akiyama and coworkers, who had pioneered BINOL-derived phosphoric acids, noticed that aldimines 365 available from 2-aminophenol and aromatic, heteroaromatic, and cinnamyl aldehydes can be activated by the chiral acid catalyst 367, so that they are electrophilic enough to react with silyl ketene acetals 366 in diastereoselective and enantioselective Mannich reactions. Thus, P-amino esters 368 are formed with a high preference for the syn-diastereomers that are obtained in high enantiomeric excess (Scheme 5.96) [182]. [Pg.362]

Helmchen [67] and Oppolzer [68] investigated and documented the use of camphor-derived auxiliaries in Mukaiyama aldol reactions. Silyl ketene acetals 106 and 108 participate in diastereoselective additions to aldehydes in the presence of TiCl4, affording adducts with up to 99 1 diastereoselectivity (Equations 7 and 8). [Pg.112]

The reactions proceeded efficiently under mild conditions in short time. The silyl enol ethers reacted with the activated acetals or aldehydes at -78 °C to give predominant erythro- or threo-products [136, 137] respectively. In the same manner, the aldol reaction of thioacetals, catalyzed by an equimolar amount of catalyst, resulted in <-ketosulfides [139] with high diastereoselectivity. In the course of this investigation, the interaction of silyl enol ethers with a,]3-unsaturated ketones, promoted by the trityl perchlorate, was shown to proceed regioselec-tively through 1,2- [141] or 1,4-addition [138]. The application of the trityl salt as a Lewis acid catalyst was spread to the synthesis of ]3-aminoesters [142] from the ketene silyl acetals and imines resulting in high stereoselective outcome. [Pg.373]

The aldol reaction of ketene silyl acetals with several aldehydes (Mukaiyama aldol reaction) assisted by Li has been described briefly by Reetz et al. Wirth 5.0 m LPDE a clean reaction began within 1 h with the sole formation of the silylated aldol 112, whereas the use of a catalytic amount (3 mol %) of LiC104 in Et20 (3 mol % LPDE) required a reaction time of 5 days for 86 % conversion. As observed in the hetero-Diels-Alder reaction of a-alkoxyaldehyde, the higher rate of reaction of 79 compared with that of benzaldehyde can be attributable to chelation. Indeed, the use of 3 mol % LPDE required only 20 h at room temperature for complete uptake of 79 with a diastereoselectivity (syn-113lanti-113) of >96 % (Sch. 55). [Pg.45]

Aldol reaction of ketene silyl acetals with aldehydes (Mukaiyama aldol reaction) mediated by Li Lewis acid has been disclosed (Scheme 3.16) [42]. With 5.0 M LPDE the reaction proceeds smoothly (rt, 1 hour, >99%), although the reaction with 3 mol% of LPDE is less reactive (rt, 5 days, 86%). In the case of a substrate that induces chelation, the reaction is accelerated. Even in a condition with 3 mol% of LPDE, the reaction proceeded relatively smoothly (rt, 20 hours, 67%), and the high diastereoselectivity was gained (synjaniti = >9SI2) from the chelation. [Pg.118]

Examples of the Bronsted-acid catalysts and hydrogen-bond catalysts are shown in Figure 2.1. We have recently reported the Mannich-type reaction of ketene silyl acetals with aldimines derived from aromatic aldehyde catalyzed by chiral phosphoric acid 7 (Figure 2.2, Scheme 2.6) [12]. The corresponding [5-amino esters were obtained with high syn-diastereoselectivities and excellent enantioselectivities. [Pg.9]


See other pages where Silyl ketene acetals reaction with aldehydes, diastereoselectivity is mentioned: [Pg.245]    [Pg.412]    [Pg.118]    [Pg.882]    [Pg.94]    [Pg.531]    [Pg.360]    [Pg.109]    [Pg.822]    [Pg.143]    [Pg.71]    [Pg.72]    [Pg.505]    [Pg.311]    [Pg.378]    [Pg.531]    [Pg.313]    [Pg.547]    [Pg.213]    [Pg.664]    [Pg.21]    [Pg.410]    [Pg.313]    [Pg.6]   
See also in sourсe #XX -- [ Pg.2 , Pg.637 ]

See also in sourсe #XX -- [ Pg.2 , Pg.637 ]




SEARCH



Acetals reactions with

Acetals with aldehydes

Acetates reactions with

Acetic aldehyde

Aldehyde acetals

Aldehyde reaction with keten

Aldehyde-ketene

Aldehydes acetalization

Aldehydes acetate

Aldehydes diastereoselectivity

Aldehydes reaction with ketene acetals

Aldehydes reaction with ketenes

Aldehydes reaction with silyl ketene acetals

Diastereoselective reaction

Diastereoselective reactions Diastereoselectivity

Diastereoselectivity reaction

Keten acetal

Ketene acetal

Ketene reaction

Ketene reaction with aldehyde

Ketenes acetals

Ketenes reactions

Ketenes silyl acetals

Reaction with ketene

Reaction with ketene acetals

Reaction with ketenes

Reaction with silyl ketene acetals

Silyl acetate

Silyl ketene acetals

Silyl ketene acetals, reaction

Silyl ketenes

Silyl ketenes, reactions

Silyl reactions with

Silylation reactions

Silylative aldehyde

Silyls reactions with

© 2024 chempedia.info