首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
In this communication, we report the synthesis of a new chiral spiro‐bisoxazoline ligand, i.e., β‐naphthylmethyl‐substituted spiro‐BOX [(Ra,S,S)‐ L7 ] and have successfully applied it to the palladium‐catalyzed enantioselective cyclization reaction of simple allenes with o‐aminoiodobenzenes, affording highly optically active 3‐alkylideneindolines in good yields with excellent enantiomeric excesses.  相似文献   

2.
A β‐ketoacyl‐ACP reductase (FabG) gene from Bacillus sp. ECU0013 was heterologously overexpressed in Escherichia coli and the encoded protein was purified to homogeneity. The recombinant reductase could reduce a broad spectrum of prochiral ketones including aromatic ketones and keto esters and showed the highest activity in the asymmetric reduction of ethyl 2‐oxo‐4‐phenylbutyrate (OPBE). Using E. coli cells coexpressing both FabG and glucose dehydrogenase (GDH) genes, as much as 620 g⋅L−1 of OPBE was almost stoichiometrically converted to ethyl (S)‐2‐hydroxy‐4‐phenylbutyrate [(S)‐HPBE] with excellent (>99%) enantiomeric excess. More importantly, the process could be performed smoothly without external addition of an expensive cofactor as usually done and could be scaled up very easily. All these positive features demonstrate the applicability of this reductase for the large‐scale production of optically active α‐hydroxy acids/esters.  相似文献   

3.
Prochiral bicyclic diketones were transformed to a single diastereomer of 3‐substituted cyclohexylamine derivatives via three consecutive biocatalytic steps. The two chiral centres were set up by a C C hydrolase (6‐oxocamphor hydrolase) in the first step and by an ω‐transaminase in the last step. The esterification of the intermediate keto acid was catalysed by a lipase in the second step if possible. For two substrates the C C hydrolytic step as well as the esterification could be run simultaneously in a one‐pot cascade in an organic solvent. In one example, the reaction mixture of the first two steps could be directly subjected to bio‐amination in an organic solvent without the need to change the reaction medium. Depending on the choice of the ω‐transaminase employed and the substrate the cis‐ as well as the trans‐diastereomers could be obtained in optically pure forms.  相似文献   

4.
A practical asymmetric synthesis of enantiopure spiro[4,4]nonane‐1,6‐dione, a valuable precursor for chiral ligand development, is reported. This synthetic strategy includes a kinetic resolution of the readily synthesized ketone precursor with a chiral quaternary carbon center by bioreduction with baker’s yeast as the key step, followed by a hydroformylation, oxidation, esterification and Dieckmann cyclization reaction sequence to generate the spiro five‐membered ring. It was found that the masking of the β‐ketone carbonyl group of enantiopure ethyl 1‐allyl‐2‐oxocyclopentanecarboxylate via formation of a ketal with 1,3‐diol derivative is necessary during the process of Dieckmann condensation in order to prevent its racemization under basic conditions. This method allows the gram‐scale preparation of both enantiomers of spiro[4,4]nonane‐1,6‐dione ( 1 ) with excellent enantiopurities (up to >99% ee) in the overall yields of 54% [(R)‐ 1 ] and 42% [(S)‐ 1 ], respectively. The practicality of the present synthetic procedure has provided a fundamental platform for the development of spiro[4,4]nonane‐1,6‐dione‐based chiral chemistry.  相似文献   

5.
Both enantiomers of optically pure 4‐bromo‐3‐hydroxybutanoate, which is an important chiral building block in the syntheses of various biologically active compounds including statins, were synthesized from rac‐4‐bromomethyl‐β‐lactone through kinetic resolution. Candida antarctica lipase B (CAL‐B) enantioselectively catalyzes the ring opening of the β‐lactone with ethanol to yield ethyl (R)‐4‐bromo‐3‐hydroxybutanoate with high enantioselectivity (E>200). The unreacted (S)‐4‐bromomethyl‐β‐lactone was converted to ethyl (S)‐4‐bromo‐3‐hydroxybutanoate (>99% ee), which can be further transformed to ethyl (R)‐4‐cyano‐3‐hydroxybutanoate, through an acid‐catalyzed ring opening in ethanol. Molecular modeling revealed that the stereocenter of the fast‐reacting enantiomer, (R)‐bromomethyl‐β‐lactone, is ∼2 Å from the reacting carbonyl carbon. In addition, the slow‐reacting enantiomer, (S)‐4‐bromomethyl‐β‐lactone, encounters steric hindrance between the bromo substituent and the side chain of the Leu278 residue, while the fast‐reacting enantiomer does not have any steric clash.  相似文献   

6.
Various (R)‐ and (S)‐C‐allylglycine derivatives were synthesized by means of an auxiliary controlled diastereoselective aza‐Claisen rearrangement. Starting from (S)‐configured auxiliaries derived from optically active proline, an aza‐Claisen rearrangement enabled us to synthesize α(R)‐configured γ,δ‐unsaturated amides. Since (R)‐allylglycine derivatives could be directly generated by reacting N‐allylproline derivatives and various protected glycine fluorides, the corresponding (S)‐enantiomers were built‐up via an initial α‐chloroacetyl chloride rearrangement and a subsequent chloride azide substitution with complete inversion of the configuration. High diastereoselectivities were obtained (>15 : 1). The auxiliary could be efficiently removed by organolithium reactions of the amides furnishing α‐amino ketones. Another allyllithium addition allowed us to introduce a second allyl chain with high diastereoselectivity. Final ring closures by means of metatheses using Grubbs' (I) catalyst gave raise to the formation of enantiopure phenanthridines and cyclohexenes displaying defined substitution patterns ready for alkaloid total syntheses.  相似文献   

7.
The chemical synthesis of 3‐substituted tyrosine derivatives requires a minimum of four steps to access optically enriched material starting from commercial precursors. Attempting to short‐cut the cumbersome chemical synthesis of 3‐substituted tyrosine derivatives, a single step biocatalytic approach was identified employing the tyrosine phenol lyase from Citrobacter freundii. The enzyme catalyses the hydrolysis of tyrosine to phenol, pyruvate and ammonium as well as the reverse reaction, thus the formation of tyrosine from phenol, pyruvate and ammonium. Since the wild‐type enzyme possessed a very narrow substrate spectrum, structure‐guided, site‐directed mutagenesis was required to change the substrate specificity of this C C bond forming enzyme. The best variant M379V transformed, for example, o‐cresol, o‐methoxyphenol and o‐chlorophenol efficiently to the corresponding tyrosine derivatives without any detectable side‐product. In contrast, all three phenol compounds were non‐substrates for the wild‐type enzyme. Employing the mutant, various L ‐tyrosine derivatives (3‐Me, 3‐OMe, 3‐F, 3‐Cl) were obtained with complete conversion and excellent enantiomeric excess (>97%) in just a single ‘green’ step starting from pyruvate and commercially available phenol derivatives.  相似文献   

8.
The enantioselective acylation of racemic diisopropyl α‐ and β‐hydroxyphosphonates by hydrolases in t‐butyl methyl ether with isopropenyl acetate as acyl donor is limited by the narrow substrate specificity of the enzymes. High enantiomeric excesses (up to 99%) were obtained for the acetates of (S)‐diisopropyl 1‐hydroxy‐(2‐thienyl)methyl‐, 1‐hydroxyethyl‐ and 1‐hydroxyhexylphosphonate and (R)‐diisopropyl 2‐hydroxypropylphosphonate. The hydrolysis of a variety of β‐chloroacetoxyphosphonates by the lipase from Candida cylindracea and protease subtilisin in a biphasic system gives (S)‐β‐hydroxyphosphonates (ee 51–92%) enantioselectively. (S)‐2‐Phenyl‐2‐hydroxyethyl‐ and (S)‐3‐methyl‐2‐hydroxybutylphosphonates (ee 96% and 99%, respectively) were transformed into (R)‐2‐aminophosphonic acids of the same ee.  相似文献   

9.
A new strategy was developed for the synthesis of a valuable class of α‐aminomethylacrylates via the Baylis–Hillman reaction of different aldehydes with methyl acrylate followed by acetylation of the resulting allylic alcohols and SN2′‐type amination of the allylic acetates. Asymmetric hydrogenation of these diverse olefinic precursors using rhodium(Et‐Duphos) catalysts provided the corresponding β2‐amino acid derivatives with excellent enantioselectivities and exceedingly high reactivities (up to >99.5% ee and S/C=10,000). The first hydrogenation of (Z)‐configurated substrates was studied for the synthesis of β2‐amino acid derivatives. The high influence of the substrate geometry and steric hindrance on the reactivity and enantioselectivity was also disclosed for this reaction. This protocol provides a highly practical, facile and scalable method for the preparation of optically pure β2‐amino acids and their derivatives under mild reaction conditions.  相似文献   

10.
N,N′‐(Pyromellitoyl)‐bis‐(L ‐leucine) diacid was reacted with ethyl chloroformate in the presence of triethylamine followed by reaction with activated sodium azide and gave N,N′‐(pyromellitoyl)‐bis‐(L ‐leucine) diacylazide in high yield. This diacylazide was heated in dry benzene and gave the unstable N,N′‐(pyromellitoyl)‐bis‐(L ‐leucine) diisocyanate ( 5 ) in quantitative yield. Thus, diisocyanate 5 was generated in situ and polycondensation reaction of this monomer with several aromatic diols, such as 4,4′‐dihydroxybiphenyl, 1,4‐hydroquinone, bisphenol A, phenolphthalein and 1,4‐dihydroxyanthraquinone, was performed in dry toluene under refluxing in the presence of 1,4‐diazabicyclo[2.2.2]octane (triethylenediamine) as a catalyst. The polymerization reactions proceeded within 48 h, producing a series of optically active poly(imide–urethane)s with good yield and moderate inherent viscosity in the range 0.18–0.28 dl g?1. All of the above polymers were fully characterized by infrared spectra, elemental analyses and specific rotation. Some structural characterization and physical properties of these optically active poly(imide–urethane)s are reported Copyright © 2003 Society of Chemical Industry  相似文献   

11.
C2‐Symmetrical, enantiopure 2,6‐di[1‐(1‐aziridinyl)alkyl]pyridines (DIAZAPs) were prepared by a high‐yielding, three‐step sequence starting from 2,6‐pyridinedicarbaldehyde and (S)‐valinol or (S)‐phenylglycinol. The new compounds were tested as ligands in palladium‐catalyzed allylation of carbanions in different solvents. Almost quantitative yield and up to 99 % enantiomeric excess were obtained in the reactions of the enolates derived from malonate, phenyl‐ and benzylmalonate dimethyl esters with 1,3‐diphenyl‐2‐propenyl ethyl carbonate.  相似文献   

12.
The phenylalanine aminomutase (PAM) from Taxus chinensis catalyses the conversion of α‐phenylalanine to β‐phenylalanine, an important step in the biosynthesis of the N‐benzoyl phenylisoserinoyl side‐chain of the anticancer drug taxol. Mechanistic studies on PAM have suggested that (E)‐cinnamic acid is an intermediate in the mutase reaction and that it can be released from the enzyme's active site. Here we describe a novel synthetic strategy that is based on the finding that ring‐substituted (E)‐cinnamic acids can serve as a substrate in PAM‐catalysed ammonia addition reactions for the biocatalytic production of several important β‐amino acids. The enzyme has a broad substrate range and a high enantioselectivity with cinnamic acid derivatives; this allows the synthesis of several non‐natural aromatic α‐ and β‐amino acids in excellent enantiomeric excess (ee >99 %). The internal 5‐methylene‐3,5‐dihydroimidazol‐4‐one (MIO) cofactor is essential for the PAM‐catalysed amination reactions. The regioselectivity of amination reactions was influenced by the nature of the ring substituent.  相似文献   

13.
Tetralin‐1,4‐dione, the stable tautomer of dihydroxynaphthalene, was reduced with catecholborane in the presence of 3,3‐diphenyl‐1‐butyltetrahydro‐3H‐pyrrolo[1,2‐c][1,3,2]oxazaborole as catalyst to give enantiomerically highly enriched 4‐hydroxy‐1‐tetralone (99% ee) in an efficient one‐pot procedure. The R‐enantiomer provided a rapid access to sertraline while the S‐enantiomer was converted into 2‐epicatalponol and catalponol. A more selective enantioselective route to the antithermitic catalponol made use of the planar chiral tricarbonylchromium complex of hydroxytetralone. Its precursor chromium(tricarbonyl)[η6‐(1‐4,4a,8a)‐tetralin‐5,8‐dione] was obtained via direct complexation of 1,4‐dihydroxynaphthalene using chromium(tricarbonyl)‐ tris(ammonia) and boron trifluoride etherate as source of the chromium(tricarbonyl) fragment. Enolate prenylation was best carried out in the presence of a tetraamine ligand. Complete inversion of the stereogenic center bearing the prenyl group of the initially obtained tetralone complex was achieved via enolate formation followed by protonation.  相似文献   

14.
A highly efficient and enantioselective hydrogenation of unprotected β‐ketoenamines catalyzed with ruthenium(II) dichloro{(S)‐(−)‐2,2′‐bis[di(3,5‐xylyl)phosphino]‐1,1′‐binaphthyl}[(2S)‐(+)‐1,1‐bis(4‐methoxyphenyl)‐3‐methyl‐1,2‐butanediamine] {Ru[(S)‐xylbinap][(S)‐daipen]Cl2} has been successfully developed. This methodology provides a straightforward access to free γ‐secondary amino alcohols, which are key building blocks for a variety of pharmaceuticals and natural products, with high yields (>99%) and excellent enantioselectivities (up to 99% ee) in all cases.  相似文献   

15.
Starting from ethyl 4‐bromobutyrate, the chemoenzymatic synthesis of 6‐vinyl‐tetrahydro‐pyran‐2‐one has been accomplished. A screening of a number of available alcohol dehydrogenases and intense optimization of reaction parameters enabled us to establish an efficient synthesis of either enantiomer of the vinyllactone with excellent enantiomeric excess (>99%). The scope of possible applications of enantiopure vinyllactones has been verified by subjection to cross‐metathesis resulting in the total synthesis of the insect pheromone (S)‐5‐hexadecanolide and the cytotoxic styryllactone goniothalamine as well as derivatives thereof.  相似文献   

16.
An efficient multi‐enzyme cascade reaction for the synthesis of (R)‐ or (S)‐2‐hydroxybutyric acid [(R)‐ or (S)‐2‐HB] from l ‐threonine was developed by using recombinant Escherichia coli cells expressing separately or co‐expressing l ‐threonine deaminase from Escherichia coli K‐12 (ilvA), formate dehydrogenase (FDH) from Candida boidinii and l ‐lactate dehydrogenase (l ‐LDH) from Oryctolagus cuniculus or d ‐lactate dehydrogenase (d ‐LDH) from Staphylococcus epidermidis ATCC 12228. Up to 750 mM of l ‐threonine were completely transformed to (R)‐ or (S)‐2‐HB in optically pure form (>99% ee) with high isolated yields. This one‐pot multi‐enzyme transformation provides a new practical method for the synthesis of these important optically pure compounds.

  相似文献   


17.
The catalytic asymmetric inverse‐electron‐demand (IED) [4+2] cycloaddition of various salicylaldehyde‐derived N‐arylimines with electron‐rich alkenes in the presence of chiral BINOL‐derived phosphoric acid catalysts has been studied with the aim of obtaining optically active 4‐aminobenzopyran derivatives. Dienophiles such as 2,3‐dihydro‐2H‐furan, benzyl N‐vinylcarbamate and 2‐vinylindole have been employed.  相似文献   

18.
Cross‐dehydrocoupling reactions of (R)‐methyl(1‐naphthyl)phenylsilane (>99%ee) with (S)‐methyl(1‐naphthyl)phenylsilanol (>99% ee) proceeded with 82–99% retention of configuration of chiral silicon centres in the presence of various Rh‐catalysts. Cross‐dehydrocoupling polymerization of 1,3‐dimethyl‐1,3‐diphenyl‐1,3‐disiloxanediol with 1,3‐dihydro‐1,3‐dimethyl‐1,3‐diphenyl‐1,3‐disiloxane gave poly(methylphenylsiloxane) of moderate molecular weight in toluene at 60 °C in the presence of [RhCl(cod)]2 (5.0 mol%) and triethylamine (1.0 equivalent). Assignment of the triad signals of the resulting polymer was made by 1H NMR spectroscopy of the methyl proton (I = 0.04, H = 0.09 and S = 0.14 ppm) and 13C NMR spectroscopy of the ipso carbon of the phenyl group (S = 136.7, H = 136.9, and I = 137.1 ppm). Although the reaction of optically pure (S,S)‐1,3‐dimethyl‐1,3‐diphenyl‐1,3‐disiloxanediol with 1,3‐dihydro‐1,3‐dimethyl‐1,3‐diphenyl‐1,3‐disiloxane [(S,S):(S,R):(R,R)] = 84:16:0] gave a poly(methylphenylsiloxane) of rather low molecular weight, its triad tacticity was found to be rich in syndiotacticity (S:H:I = 60:32:8) by 13C NMR spectroscopy. © 2001 Society of Chemical Industry  相似文献   

19.
Optically pure (S,S)‐1,2‐bis[(o‐alkylphenyl)phenylphosphino]ethanes 1a–d were prepared in four steps from phenyldichlorophosphine via phosphine‐boranes as the intermediates. The rhodium complexes 5a–d of these diphosphines were used for the asymmetric hydrogenations of α‐(acylamino)acrylic derivatives including β‐disubstituted derivatives. Markedly high enantioselectivity (78–>99%) was observed for the reduction of β‐monosubstituted derivatives. β‐Disubstituted derivatives were also reduced in considerably high enantioselectivity (up to 90%). The single crystal X‐ray analysis of the rhodium complex 5c of (S,S)‐1,2‐bis[phenyl(5′,6′,7′,8′‐tetrahydronaphthyl)phosphino]ethane ( 1c) revealed its δ‐type structure with face orientation of the two tetrahydronaphthyl groups and edge orientation of the two phenyl groups. This conformation corresponds to that of the rhodium complex of 1,2‐bis[(o‐methoxyphenyl)phenylphosphino]ethane (DIPAMP); the rhodium complex of (R,R)‐DIPAMP, whose chirality at phosphorus is opposite that of 5c , exhibits a λ‐type structure with the face orientation of the two o‐methoxyphenyl groups and the edge orientation of the two phenyl groups. The conformational similarity of these rhodium complexes as well as the stereochemical outcome in the asymmetric hydrogenations means that the coordinative interaction of the methoxy group of DIPAMP with rhodium metal is not the main factor that affects asymmetric induction.  相似文献   

20.
Oxidoreductases from Streptomyces sp. GF3546 [3546‐IRED], Bacillus cereus BAG3X2 (BcIRED) and Nocardiopsis halophila (NhIRED) each reduce prochiral 2‐methylpyrroline (2MPN) to (S)‐2‐methylpyrrolidine with >95 % ee and also a number of other imine substrates with good selectivity. Structures of BcIRED and NhIRED have helped to identify conserved active site residues within this subgroup of imine reductases that have S selectivity towards 2MPN, including a tyrosine residue that has a possible role in catalysis and superimposes with an aspartate in related enzymes that display R selectivity towards the same substrate. Mutation of this tyrosine residue—Tyr169—in 3546‐IRED to Phe resulted in a mutant of negligible activity. The data together provide structural evidence for the location and significance of the Tyr residue in this group of imine reductases, and permit a comparison of the active sites of enzymes that reduce 2MPN with either R or S selectivity.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号