Big Chemical Encyclopedia

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

Articles Figures Tables About

Enantioselectivities evaluation

Recently, the direct aldol reaction between commercially available ethyl diazoacetate (EDA) and an aldehyde represents an attractive process in synthetic chemistry. In this context, Trost and co-workers (141) surveyed a series of metal and the magnesium/prophenol ligand 88-derived catalyst system was found to give the highest enantioselectivity. Evaluation of the reaction revealed that high levels of enantioselectivities (87-98% ee) could be achieved for a wide range of aldehydes (Scheme 28). [Pg.2222]

Enantioselective processes involving chiral catalysts or reagents can provide sufficient spatial bias and transition state organization to obviate the need for control by substrate stereochemistry. Since such reactions do not require substrate spatial control, the corresponding transforms are easier to apply antithetically. The stereochemical information in the retron is used to determine which of the enantiomeric catalysts or reagents are appropriate and the transform is finally evaluated for chemical feasibility. Of course, such transforms are powerful because of their predictability and effectiveness in removing stereocenters from a target. [Pg.51]

In recent years, together with enantioselective analysis, the determination of the natural abundance of stable isotopes by means of stable isotope ratio mass spectrometry (TRMS) can be very useful for the assignment of the origin of foods and food ingredients, and of authenticity evaluation (24). [Pg.223]

Authenticity evaluation has recently received increased attention in a number of industries. The complex mixtures involved often require very high resolution analyses and, in the case of determining the authenticity of natural products, very accurate determination of enantiomeric purity. Juchelka et al. have described a method for the authenticity determination of natural products which uses a combination of enantioselective multidimensional gas chromatography with isotope ratio mass spectrometry (28). In isotope ratio mass spectrometry, combustion analysis is combined with mass spectrometry, and the ratio of the analyte is measured versus a... [Pg.422]

The purpose of this study is only intended to illustrate and evaluate the decision tree approach for CSP prediction using as attributes the 166 molecular keys publicly available in ISIS. This assay was carried out a CHIRBASE file of 3000 molecular structures corresponding to a list of samples resolved with an a value superior to 1.8. For each solute, we have picked in CHIRBASE the traded CSP providing the highest enantioselectivity. This procedure leads to a total selection of 18 CSPs commercially available under the following names Chiralpak AD [28], Chiral-AGP [40], Chiralpak AS [28], Resolvosil BSA-7 [41], Chiral-CBH [40], CTA-I (microcrystalline cellulose triacetate) [42], Chirobiotic T [43], Crownpak CR(-i-) [28], Cyclobond I [43], DNB-Leucine covalent [29], DNB-Phenylglycine covalent [29], Chiralcel OB [28], Chiralcel OD [28], Chiralcel OJ [28], Chiralpak OT(-i-) [28], Ultron-ES-OVM [44], Whelk-0 1 [29], (/ ,/ )-(3-Gem 1 [29]. [Pg.120]

To evaluate the economics of this process, a cost model has been developed to estimate the separation costs for a specific racemate [68, 69]. For this purpose, the sensitivity of the separation costs for several key process parameters have been established as compared to a base-case separation in which a purity of 99 % is required at an enantioselectivity of 1.15. The maximum solubility of the drug is set... [Pg.143]

Schemes 16-19 present the details of the enantioselective synthesis of key intermediate 9. The retrosynthetic analysis outlined in Scheme 5 identified aldoxime 32 as a potential synthetic intermediate the construction of this compound would mark the achievement of the first synthetic objective, for it would permit an evaluation of the crucial 1,3-dipolar cycloaddition reaction. As it turns out, an enantioselective synthesis of aldoxime 32 can be achieved in a straightforward manner by a route employing commercially available tetronic acid (36) and the MEM ether of allyl alcohol (74) as starting materials (see Scheme 16). Schemes 16-19 present the details of the enantioselective synthesis of key intermediate 9. The retrosynthetic analysis outlined in Scheme 5 identified aldoxime 32 as a potential synthetic intermediate the construction of this compound would mark the achievement of the first synthetic objective, for it would permit an evaluation of the crucial 1,3-dipolar cycloaddition reaction. As it turns out, an enantioselective synthesis of aldoxime 32 can be achieved in a straightforward manner by a route employing commercially available tetronic acid (36) and the MEM ether of allyl alcohol (74) as starting materials (see Scheme 16).
In a first report [24], the enantioselectivities of various proteases were evaluated by comparing the biocatalyzedhydrolysis of2-chloroethyl esters of N-acetyl-i- and D-amino acids in water and their transesterification with w-propanol in butyl ether. By comparing the ratio of the kc t/Ku values for the l- and D-enantiomers in the two reactions, a remarkable relation of the proteases enantioselectivity was observed apparently, in this case, the organic solvents destroyed the selectivity of the tested enzymes. This finding... [Pg.9]

Several reports regarding the directed evolution of enantioselective epoxide hydrolases (EHs) have appeared [23,57-59]. These enzymes constitute important catalysts in synthetic organic chemistry [4,60]. The first two reported studies concern the Aspergillus niger epoxide hydrolase (ANEH) [57,58]. Initial attempts were made to enhance the enantioselectivity of the AN E H -catalyzed hydrolytic kinetic resolution of glycidyl phenyl ether (rac-19). The WT leads to an Evalue of only 4.6 in favor of (S)-20 (see Scheme 2.4) [58]. [Pg.41]

Interest in enantioselective epoxidation continues and 2,2-dimethylchromenes appear to be particularly suitable substrates for the evaluation of the catalytic system <96JCS(P1)1757, 96SL1079, 96TL3895>. [Pg.292]

Copper-complexes prepared with other type of N-chelating ligands have been also prepared and evaluated as catalysts for the Diels-Alder reaction. Eng-berts et al. [103] studied enantioselective Diels-Alder reaction of 3-phenyl-l-(2-pyridyl)-2-propen-l-one with cyclopentadiene in water (Scheme 39). By using coordinating chiral, commercially available a-amino-adds and their derivatives with copper salts as catalysts, they obtained the desired product with yields generally exceeding 90%. With L-abrine (72 in Scheme 39) as chiral moiety, an enantiomeric excess of 74% could be achieved. Moreover, the catalyst solution was reused with no loss of enantioselectivity. [Pg.124]

Subsequently, these catalysts were evaluated in the enantioselective desymmetri-sation of achiral trienes, and three distinct trends in catalyst selectivity were found. Firstly, catalysts 56a-b with two phenyl moieties on the backbone of the A -heterocycle exhibited higher enantioselectivity than those with a fused cyclohexyl group as the backbone 55a-b. Secondly, mono-ort/io-substituted aryl side chains induced greater enantioselectivity than symmetrical mesityl wing tips. Thirdly, changing the halide ligands from Cl to I" increased the enantioselectivity. As a result, catalyst 56b turned out to be the most effective. For example, 56b in the presence of Nal was able to promote the desymmetrisation of 57 to give chiral dihydrofuran 58 in up to 82% conversion and 90% ee (Scheme 3.3). [Pg.78]

In 2008, Rykowski et al. reported the synthesis of optically active 2,2 -bipyridine alkyl sulfoxides by asymmetric oxidation of their corresponding readily accessible 2,2 -bipyridine alkyl sulfides. These sulfoxides were further evaluated as ligands for the enantioselective addition of ZnEt2 to benzaldehyde, providing only low enantioselectivities of up to 14% ee (Scheme 3.34). [Pg.127]

The use of expensive and unstable ZnPli2 in the preparation of chiral di-arylmethanol derivatives, with electronically and sterically similar aryl rings, made this approach less attractive for the enantioselective synthesis. In order to avoid this inconvenience, other alternative preparations of arylzinc reagents were evaluated.As a first choice, Yus et al. proposed the use of arylboronic adds as a viable source of phenyl (Scheme 4.19). Thus, the reaction of various boronic acids with an excess of ZnEt2 at 70 °C gave the corresponding arylzinc intermediates (probably aryl(ethyl)zincs), which were trapped by reaction with dif-... [Pg.170]

As an extension of this methodology, the efficiency of these ligands was also evaluated by these authors for the Cu-catalysed cyclopropanation of styrene derivatives with EDA, providing the corresponding cyclopropanes with similar enantioselectivities of up to 97% ee (Scheme 6.4). ... [Pg.211]

In 2004, ruthenium-catalysed asymmetric cyclopropanations of styrene derivatives with diazoesters were also performed by Masson et al., using chiral 2,6-bis(thiazolines)pyridines. These ligands were prepared from dithioesters and commercially available enantiopure 2-aminoalcohols. When the cyclopropanation of styrene with diazoethylacetate was performed with these ligands in the presence of ruthenium, enantioselectivities of up to 85% ee were obtained (Scheme 6.6). The scope of this methodology was extended to various styrene derivatives and to isopropyl diazomethylphosphonate with good yields and enantioselectivities. The comparative evaluation of enantiocontrol for cyclopropanation of styrene with chiral ruthenium-bis(oxazolines), Ru-Pybox, and chiral ruthenium-bis(thiazolines), Ru-thia-Pybox, have shown many similarities with, in some cases, good enantiomeric excesses. The modification... [Pg.213]

In 2004, a series of other chiral thioether-phosphine ligands based on a cyclopropane backbone were evaluated in the rhodium-catalysed hydrogenation of a dehydroamino acid by Molander el al As shown in Scheme 8.2, even if these ligands were generally active, only moderate enantioselectivities of up to 47% ee were obtained. [Pg.244]

In 1998, Ruiz et al. reported the synthesis of new chiral dithioether ligands based on a pyrrolidine backbone from (+ )-L-tartaric acid. Their corresponding cationic iridium complexes were further evaluated as catalysts for the asymmetric hydrogenation of prochiral dehydroamino acid derivatives and itaconic acid, providing enantioselectivities of up to 68% ee, as shown in Scheme 8.18. [Pg.255]

In the same study, several ligands variously functional on both the nitrogen and the sulfur atoms have been developed, providing a new class of cyclo-hexylamino sulfide ligands derived from cyclohexene oxide. All the ligands depicted in Scheme 9.7 were evaluated for the Ir-catalysed hydride-transfer reduction of acetophenone in the presence of i-PrOH as the hydrogen donor, providing enantioselectivities of up to 70% ee. [Pg.273]

Novel chiral thiolated amino alcohols have been recently synthesised and then evaluated by Vilaivan el al. as a potential new class of ligands for Cu-catalysed nitroaldol reactions. Amino alcohol ligands bearing Ai-(2-alkyl-thio)benzyl substituents provided only modest enantioselectivities (22-46% ee) while those carrying Al-2-thienylmethyl substituents provided better enantioselectivities of up to 75% ee for the nitroaldol reaction between p-nitro-benzaldehyde and nitromethane. A range of aromatic aldehydes were acceptable substrates giving moderate to high enantioselectivities of up to 88% ee, as shown in Scheme 10.32. [Pg.319]


See other pages where Enantioselectivities evaluation is mentioned: [Pg.212]    [Pg.168]    [Pg.212]    [Pg.168]    [Pg.74]    [Pg.322]    [Pg.534]    [Pg.221]    [Pg.23]    [Pg.28]    [Pg.133]    [Pg.243]    [Pg.309]    [Pg.309]    [Pg.9]    [Pg.38]    [Pg.40]    [Pg.41]    [Pg.43]    [Pg.45]    [Pg.47]    [Pg.48]    [Pg.59]    [Pg.61]    [Pg.100]    [Pg.188]    [Pg.252]    [Pg.260]    [Pg.273]    [Pg.275]    [Pg.278]    [Pg.298]    [Pg.300]   
See also in sourсe #XX -- [ Pg.461 ]




SEARCH



Enantioselectivity evaluation

Enantioselectivity evaluation

© 2024 chempedia.info