Big Chemical Encyclopedia

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

Articles Figures Tables About

Asymmetric Reactions using Chiral Auxiliaries

The use of chiral auxiliaries proved to be particularly efficient for the preparation of optically pure materials via radical reactions [29]. Yamamoto et al. [30] examined the free-radical reaction of an a-bromoglycine derivative 35 having a chiral auxiliary with allyltributyltin 30 to give allylated product 36 (Sch. 14). Although the use of 0.1 equiv. [Pg.69]

ZnCl2-OEt2 accelerated the reaction sufficiently, high diastereofacial selection needed 2 equiv. Lewis acid. It was noted that ethereal zinc halides acted both as radical initiators and as chelating agents, whereas BF3 OEt2 and SnC did not show such activity. [Pg.70]

The diphenylalaninol-derived oxazolidinone skeleton was thus effective as a chiral auxiliary, and was then applied to the /3-radical addition of the a,/3-unsaturated compound 39 (Sch. 16) [32]. The high diastereofacial selectivity of the /3-radical addition can be explained by a chelation model XVII similar to XVI for the preceding allylation reaction. [Pg.71]

Lewis acid (2 equiv) EtsB (10 equiv) BuaSnH (5 equiv) O2 [Pg.71]


A special case of asymmetric induction using chiral auxiliaries has been reported for the alkylation of /1-keto esters94,106. In this approach the reaction proceeds in a diastereoselective manner via a base-catalyzed opening of the corresponding chiral 1.2-cyclohexanedioxy or 1,2-cyclohep-tanedioxy acetal, e g., acetal 50. [Pg.716]

Although asymmetric versions of aza Diels-Alder reactions using chiral auxiliaries have been reported, only one example uses a stoichiometric amount of a chiral Lewis acid [44]. The first reported example of a catalytic enantioselective aza Diels-Alder reaction employed a chiral lanthanide catalyst [45]. A chiral ytterbium or scandium catalyst, prepared from Yb(OTf)3 or Sc(OTf)3, (i )-BINOL, and DBU, is effective in the enantioselective aza Diels-Alder reactions. The reaction of A-alkylidene- or N-arylidene-2-hydroxyaniline with cyclopentadiene proceeded in the presence of the chiral catalyst and 2,6-di-rerf-butyl-4-methylpyridine (DTBMP) to afford the corresponding 8-hydroxyquinoline derivatives in good to high yields with good to excellent diastereo- and enantioselectivity (Eq. 15). [Pg.894]

Our initial improvement in the synthesis of pyrrolidine acid 3 relied on a racemic 1,3 dipolar cycloaddition followed by resolution. Attempts to devise asymmetric protocols of this reaction using chiral auxiliaries were not productive. The results from our laboratories were consistent with literature findings, with a moderate diastereoselectivity of 3 to 4 1 at best obtained even when double chiral auxiliaries were used. Several other approaches, such as Aza-Cope/Mannich reaction, intramolecular C-H insertion, and asymmetric aryl 1,4 addition, did not bear fruit. [Pg.79]

Asymmetric Buchner reactions using chiral auxiliary have also been undertaken. The diazoketo substrate 126 for the chiral tethered Buchner reaction is prepared from optically pure (2/ ,4/f)-2,4-pentanediol in three steps the Mitsunobu reaction with 3,5-dimethylphenol, esterification with diketene, and diazo formation/deacetylation. Treatment of 126 with rhodium(II) acetate results in a quantitative yield of 127 with more than 99% ee. This compound is reduced with lithium aluminium hydride, and the resulting diol 128 undergoes epoxidation and concurrent acetal formation to give 129 as a single diastereomer. Hydrogenation of 129 with Raney nickel proceeds stereoselectively to yield saturated diol 130, which is converted to aldehyde 132 via acid hydrolysis followed by oxidation. Compound 132 is a versatile intermediate for natural product synthesis. [Pg.442]

Asymmetric reactions using nonnatural chiral auxiliaries with participation and formation of heterocycles 98YGK386. [Pg.205]

Mg11 complexes are also effective for controlling asymmetric radical reactions.33,34 Moreover, enantioselective radical reactions using chiral Mg11 complexes have been studied, and high enantioselectivities have been realized in the presence of stoichiometric or catalytic amounts of chiral auxiliaries such as bis-oxazolines (Scheme 8).35-39 In most cases, substrates having bidentate chelating moieties are required. [Pg.402]

Asymmetric Diels-Alder reactions. Unlike methyl crotonate, which is a weak dienophile, chiral (E)-crotonyl oxazolidinones when activated by a dialkylaluminum chloride (1 equiv.) are highly reactive and diastereoselective dienophiles. For this purpose, the unsaturated imides formed from oxazolidinones (Xp) derived from (S)-phenylalanol show consistently higher diastereoselectivity than those derived from (S)-valinol or (IS, 2R)-norephedrine. The effect of the phenyl group is attributed in part at least to an electronic interaction of the aromatic ring. The reactions of the unsaturated imide 1 shown in equation (I) are typical of reactions of unsaturated N-acyloxazolidinones with cyclic and acyclic dienes. All the Diels-Alder reactions show almost complete endo-selectivity and high diastereoselectivity. Oxazolidinones are useful chiral auxiliaries for intramolecular Diels-Alder... [Pg.244]

The chiral auxiliaries anchored to the substrate, which is subjected to diastereoselective catalysis, is another factor that can control these reactions. These chiral auxiliaries should be easily removed after reduction without damaging the hydrogenated substrate. A representative example in this sense is given by Gallezot and coworkers [268], They used (-)mentoxyacetic acid and various (S)-proline derivates as chiral auxiliaries for the reduction of o-cresol and o-toluic acid on Rh/C. A successful use of proline derivates in asymmetric catalysis has also been reported by Harada and coworkers [269,270], The nature of the solvent only has a slight influence on the d.e. [271],... [Pg.521]

The asymmetric synthesis of a-amino acids is an important topic due to their extensive use in pharmaceuticals and agrochemicals and as chiral ligands. The Strecker reaction is historically one of the most versatile ways to produce a-amino acids, but this method has a maximum yield of only 50% for a single enantiomer. Higher yields can be achieved by using chiral auxiliaries, but auxiliaries have other drawbacks, such as high cost, low availability, the need for purification, and high loss rates. A possible solution to these problems would be to use a chiral auxiliary in a crystallization-induced asymmetric transformation. [Pg.343]

The asymmetric [3 + 4] cycloaddition is readily achieved using chiral auxiliaries or catalysts [16]. The efficiency of the chiral auxiliary approach is illustrated in the [3-1-4] cycloaddition with cyclopentadiene. The vinyldiazoacetate 6, with (T)-pantolactone as the chiral auxiliary, generated the bicyclo[3.2.1]octadiene 75 in 87% yield and 76% dia-stereomeric excess (Eq. 10) [82]. Alternatively, the chiral rhodium prolinate Rh2(S-DOSP)4-catalyzed reaction of 4 generated the bicyclo[3.2.1]octadiene 76 in 77% yield and with 93% enantiomeric excess (Eq. 11) [83]. [Pg.316]

Asymmetric Mannich reactions provide useful routes for the synthesis of optically active p-amino ketones or esters, which are versatile chiral building blocks for the preparation of many nitrogen-containing biologically important compounds [1-6]. While several diastereoselective Mannich reactions with chiral auxiliaries have been reported, very little is known about enantioselective versions. In 1991, Corey et al. reported the first example of the enantioselective synthesis of p-amino acid esters using chiral boron enolates [7]. Yamamoto et al. disclosed enantioselective reactions of imines with ketene silyl acetals using a Bronsted acid-assisted chiral Lewis acid [8]. In all cases, however, stoichiometric amounts of chiral sources were needed. Asymmetric Mannich reactions using small amounts of chiral sources were not reported before 1997. This chapter presents an overview of catalytic asymmetric Mannich reactions. [Pg.143]

Asymmetric catalysis with chiral ligands [82] is commonly considered to be advantageous instead of using chiral auxiliaries. Catalytic asymmetric Michael reactions are known [83], but not with iron as the catalytically active metal. Only two reports on iron catalyzed catalytic asymmetric Michael reaction with dipeptides [84] or diamino thioethers [85] exist, but the enantioselectivities were disappointing (18% ee and 10% ee, respectively). [Pg.233]

Using a chiral auxiliary. The achiral substrate is combined with a pure enantiomer known as a chiral auxiliary to form a chiral intermediate. Treatment of this intermediate with a suitable reagent produces the new asymmetric centre. The chiral auxiliary causes, by steric or other means (see section 10.2.2), the reaction to favour the production of one of the possible stereoisomers in preference to the others. Completion of the reaction is followed by removal of the chiral auxiliary, which may be recovered and recycled, thereby cutting down development costs (Figure 10.10). An advantage of this approach is that where the reaction used to produce the new asymmetric centre has a poor stereoselectivity the two products of the reaction will be diastereoisomers, as they contain two different asymmetric centres. These diastereoisomers may be separated by crystallization or chromatography (see section 10.2.1) and the unwanted isomer discarded. [Pg.212]

By definition, a chiral auxiliary differs from a chiral template in that the auxiliary is capable of being recycled after the desired asymmetric reaction. Hence, chiral templates will not be included in this chapter. The uses of a chiral moiety as a ligand for a reagent have also been excluded. [Pg.443]

A stereocontrolled synthesis of the biologically active neolignan (+)-dehydrodiconiferyl alcohol, which was isolated from several Taxus species, was achieved via Evans asymmetric aldol condensation [58] using ferulic acid amide derived from D-phenylalanine. The reaction steps are shown in Fig. 9. This stereocontrolled reaction is also useful for preparing the enantiomer of (+)-dehydroconiferyl alcohol using chiral auxiliary oxazolidinone prepared from L-phenylalanine. This reaction also enables the syntheses of other natural products that possess the same phenylcoumaran framework. [Pg.121]


See other pages where Asymmetric Reactions using Chiral Auxiliaries is mentioned: [Pg.69]    [Pg.486]    [Pg.69]    [Pg.486]    [Pg.312]    [Pg.291]    [Pg.362]    [Pg.296]    [Pg.123]    [Pg.186]    [Pg.187]    [Pg.185]    [Pg.127]    [Pg.284]    [Pg.127]    [Pg.441]    [Pg.11]    [Pg.127]    [Pg.310]    [Pg.239]    [Pg.243]    [Pg.270]    [Pg.502]    [Pg.166]    [Pg.81]    [Pg.631]    [Pg.312]    [Pg.408]    [Pg.81]    [Pg.268]    [Pg.429]    [Pg.127]    [Pg.143]    [Pg.242]    [Pg.184]   


SEARCH



Asymmetric chirality

Asymmetric reactions chiral auxiliaries

Asymmetric using chiral auxiliaries

Chiral auxiliaries reaction

Chirality auxiliaries

Reaction auxiliaries

Reactions chiral

Using chiral auxiliaries

© 2024 chempedia.info