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 31 毫秒
1.
β-methylaspartate ammonia-lyase, EC 4.3.1.2, (β-methylaspartase) from Clostridium tetanomorphum was used to produce a 40/60 molar ratio of (2S,3R) and (2S,3S)-3-methylaspartic acids, 2a and 2b , respectively, from mesaconic acid 1 as substrate, on a large scale. To prepare (3R,4R)-3-methyl-4-(benzyloxycarbonyl)-2-oxetanone (benzyl 3-methylmalolactonate) 6, 2a and 2b were transformed, in the first step, into 2-bromo-3-methylsuccinic acids 3a and 3b and separated. After three further steps, (2S,3S)- 3a yielded the α,β-substituted β-lactone (3R,4R) 6 with a very high diastereoisomeric excess (>95% by chiral gas chromatography). The corresponding crystalline polymer, poly[benzyl β-(2R,3S)-3-methylmalate] 8 , prepared by an anionic ring opening polymerization, was highly isotactic as determined by 13C NMR. Catalytic hydrogenolysis of lactone 6 yielded (3R,4R)-3-methyl-4-carboxy-2-oxetanone (3-methylmalolactonic acid) 7 , to which reactive, chiral, or bioactive molecules can be attached through ester bonds leading to polymers with possible therapeutic applications. Because of the ability of β-methylaspartase to catalyse both syn- and anti-elimination of ammonia from (2S,3RS)-3-methylaspartic acid 2ab at different rates, the (2S,3R)-stereoisomer 2a was retained and isolated for further reactions. These results permit the use of the chemoenzymatic route for the preparation of both optically active and racemic polymers of 3-methylmalic acid with well-defined enantiomeric and diastereoisomeric compositions. Chirality 10:727–733, 1998. © 1998 Wiley-Liss, Inc.  相似文献   

2.
(2RS,3SR)-2-Amino-3-chlorobutanoic acid hydrochloride [(2RS,3SR)-ACB · HCl] was found to exist as a conglomerate based on the melting point, infrared spectrum, and solubility. Optical resolution by preferential crystallization of (2RS,3SR)-ACB · HCl was achieved to yield both (2R,3S)- and (2S,3R)-ACB · HCl of 80–100% optical purities. The obtained (2R,3S)- and (2S,3R)-ACB · HCl were recrystallized, taking into account the solubility of (2RS,3SR)-ACB · HCl, to give efficiently optically pure (2R,3S)- and (2S,3R)-ACB · HCl. Treatment of the purified (2R,3S)- and (2S,3R)-ACB · HCl with triethylamine gave optically pure (2R,3S)- and (2S,3R)-2-amino-3-chlorobutanoic acid, respectively. Chirality 9:656–660, 1997. © 1997 Wiley-Liss, Inc.  相似文献   

3.
Racemic 5-methoxy-2-methyl-2-dipropylaminotetralin ( 3 ) has been prepared by a short synthetic route, in which the N,N-dipropyliminium perchlorate of 5-methoxy-2-tetralone ( 4 ) is a key intermediate. Racemic 3 was resolved by crystallization of the corresponding diastereomeric di-p-toluoyltartrates. The enantiomeric excess (%ee) of the phenolic derivatives of (+)-(R)- and (?)-(S)-3 [(+)-(R)- and (?)-(S)-2] was determined by 1HNMR spectroscopic analysis of the corresponding diastereomeric (?)-(R)-1,1′-binaphthyl-2,2′-diylphosphoric acid salts utilizing 13C satellites. X-ray crystallography established the absolute configuration of (?)-(S)-2 · HCl. The enantiomers of 2 were tested for hippocampal output of 5-hydroxytryptamine, 5-hydroxyindoleacetic acid, and dihydroxyphenylacetic acid in rats by use of in vivo microdialysis. The (?)-(S)-enantiomer appeared to affect 5-HT-turnover, whereas (+)-(R)- 2 was inactive. Results obtained provide support for the previously reported hypothesis that the inactivity of (?)-(S)- 2 at central DA receptors is caused by the steric bulk of the C(2)-methyl group. This makes it possible to define a “DA D2 receptor essential volume.” © 1993 Wiley-Liss, Inc.  相似文献   

4.
Copper(II) complexes of (S)-phenylalaninamide have been successfully used for the direct enantiomeric separation of unmodified (R,S)-α-hydroxy acids in reversed phase high-performance liquid chromatography (RP-HPLC). The effect of various parameters (pH, eluent polarity, selector concentration) on enantioselectivity is discussed. Evidence is provided that a mechanism of ligand exchange is actually occurring during the chromatographic separation. The method is very convenient and easy to use, and the chiral selector is commercially available and can be recovered at the end of the analysis. A conventional achiral RP-ODS-2 column is used and no pretreatment of the samples is required. This method allows the accurate determination of the enantiomeric excess of α-hydroxy acids in synthetic and biological samples. © 1995 Wiley-Liss, Inc.  相似文献   

5.
Both enantiomers of (3S)-(-)- and (3R)-(+)-Neodictyoprolenol [(3S,5Z,8Z)-(-)-1,5,8-undecatrien-3-ol] were successfully converted to the algal sex pheromone, (1S,2R)-(-)-dictyopterene B and (1R,2S)-(+)-dictyopterene B in high enantiomeric purities (e. e. > 99%), respectively, by the biomimetic reaction involving phosphorylation and elimination under a mild condition.  相似文献   

6.
A comparative study of four peptidomimetics of the sequence Phe-Met-Arg-Phe-amide (FMRFa) was performed to compare the conformational bias caused by trans-2,3-methanomethionine and α-methylmethionine stereoisomers. The specific compounds studied were F[(2S,3S)-cyclo-M] RFa, F[(2R,3R)-cyclo-M]RFa, F[(S)-α-MeM]RFa, and F[(R)-α-MeM]RFa. Molecular simulations based on CHARMm 22 indicate that γ-turn, inverse γ-turn, and α-helical conformations about the cyclo-M residue are accessible to the two F[cyclo-M]RFa stereoisomers. Similar calculations for F[(S)-α-MeM]RFa, and F[(R)-α-MeM]RFa indicate that the α-methylamino acids tend to favor α-helical conformations. The nmr data is presented for the four peptidomimetics. Most informative were the rotating frame nuclear Overhauser effect cross peaks between the NH protons proximal to the methionine surrogates, and the Cβ hydrogens. Overall, these nmr data indicate F[(2S,3S)-cyclo-M]RFa and F[(2R,3R)-cyclo-M]RFa preferentially adopt inverse γ-turn and γ-turn conformations, respectively, whereas F[(S)-α-MeM]RFa and F[(R)-α-MeM]RFa tend to form partial left- and right-handed helical structures (although energy differences between the two turn structures, and between the two helical structures are likely to be small). It is suggested that the wider NH-Cα-CO angle of cyclopropane amino acids and their more severe steric requirements around the Cβ carbons force the peptidomimetic N- and C-termini into the same region of conformational space. This favors C7 turns in the cyclopropane amino acid series relative to the less constrained α-methyl derivatives. © 1997 John Wiley & Sons, Inc. Biopoly 42: 439–453, 1997  相似文献   

7.
Abstract

The asymmetric acylation of (R, S)-3-n-butylphthalide could be efficiently catalyzed by Novozyme 435. The effect of various reaction parameters such as water activity, temperature, molar ratio of acetic anhydride to (R, S)-3-n-butylphthalide, and reaction time on the asymmetric acylation were studied. The optimums of the reaction parameters were water activity 0.62, temperature 30°C, molar ratio of acetic anhydride to (R, S)-3-n-butylphthalide 8:1, and reaction time 48 h, respectively. Under the optimum conditions, enantiopure 3-n-butylphthalide with an optical purity of 95.7% enantiomeric excess and 49.1% yield could be obtained. Furthermore, the enantiomeric excess of product was over 98%.  相似文献   

8.
(2S,3S)-3-methyl- and 3-isopropylaspartic acids were synthesized by bioconversion of the corresponding alkylfumarates (mesaconate and 3-isopropylfumarate) using β-methylaspartase from cell-free extracts of Clostridium tetanomorphum. Optically pure (2S,3S)-3-alkylaspartic acids were transformed in several steps to benzyl (3S,4R)-3-alkylmalolactonates without any racemization of the two chiral centers. These optically active α,β-substituted-β-lactones were polymerized by anionic ring opening polymerization yielding optically active semi-crystalline polyesters. 13C NMR analysis of poly[benzyl β-3-isopropylmalate] in CDCl3 has shown that only the iso-type stereosequence is present in the polymer, indicating that the macromolecular chain is constituted by the only units of benzyl β-(2S,3S)-3-isopropylmalate monomer. The polymerization reaction was done without any racemization of the two stereogenic centers as in the case of benzyl (3S,4R)-3-methylmalolactonate. © 1996 Wiley-Liss, Inc.  相似文献   

9.
The determination of enantiomeric purity of (R)- and (S)-2-hydroxy-4-phenylbutyric acid by chiral HPLC is described. Good resolution has been obtained on covalently bonded L-hydroxyproline saturated with Cu(II) ions. The method makes possible the determination of enantiomeric purity in media containing growing cells. © 1994 Wiley-Liss, Inc.  相似文献   

10.
A purified and partially characterized novel NADP+-dependent oxidoreductase from Clostridium tyrobutyricum DSM 1460 was applied for the preparative reduction of several 3-oxo acids to (S)-3-hydroxy acids. (R)-3-Hydroxybutyrate was prepared by the same enzyme selectively dehydrogenating the S enantiomer of (R,S)-3-hydroxybutyrate. The enantiomeric purity of the (S)- and (R)-3-hydroxy acids was at least 98% enantiomeric excess (e.e). NADPH for reductions and NADP+ for dehydrogenations were regenerated by applying artificial mediator accepting pyridine nucleotide oxidoreductases in the form of a crude extract of C. thermoaceticum cells. For NADP+ regeneration also the system 2-oxoglutarate/glutamate dehydrogenase was used for comparison. Instead of the purified (S)-3-hydroxycarboxylate oxidoreductase, resting cells of C. tyrobutyricum were also applied for reductions and dehydrogenations with substrate concentrations of 200–400 mM leading to products with e.e. values above 96%.Dedicated to Prof. H.G. Floss on the occasion of his 60th birthday  相似文献   

11.
Summary Stereoselective oxidation of (S)-isomer of rac-1-phenylethanol (1-PA) by the yeast Pachysolen tannophilus IFO 1007 immobilized into calcium-alginate gels was investigated to produce (R)-isomer. Continuous production of (R)-isomer was accomplished for more than 80 h with an enantiomeric excess of > 90% using a bioreactor of a fluidized-bed type.  相似文献   

12.
A direct, isocratic, and simple chromatographic method is described for the resolution of racemic albuterol using the α1-acid glycoprotein chiral stationary phase (AGP-CSP) under reverse phase conditions. The effect of various organic modifiers, temperature, and phosphate buffer ionic strength on the separation factor (α) and stereochemical resolution factor (Rs) has been studied. The enantiomeric separation of albuterol was also achieved using a urea-type CSP of (S)-indoline-2-carboxylic acid and (R)-1-(α-naphthyl)ethylamine, known as Chirex 3022, running in the normal phase mode. The effect of different organic acids added to the mobile phase was examined and the chiral recognition mechanism(s) is discussed. Solid phase extraction with C18 Sep-Pak cartridges was applied as a clean-up step to determine the enantiomeric ratio between (?)-R and (+)-S-albuterol in pharmaceutical formulations and in human plasma. © 1995 Wiley-Liss, Inc.  相似文献   

13.
Novozyme 435 could be a highly efficient catalyst in the asymmetric acylation of (R,S)-3-n-butylphthalide in tetrahydrofuran–hexane solvents. The effect of various reaction parameters such as agitation velocity, water content, mixed media, temperature, concentration of Novozyme 435, molar ratio of acetic anhydride to (R,S)-3-n-butylphthalide, reaction time, enantiomeric excess of substrate (eeS), enantiomeric excess of product (eeP), and enantioselective ratio (E) were studied. Tetrahydrofuran markedly improved (R,S)-3-n-butylphthalide conversion, enantiomeric excess of remaining 3-n-butylphthalide, and enantiomeric ratio. The optimum media were 50% (v/v) tetrahydrofuran and 50% (v/v) hexane. Other ideal reaction conditions were an agitation velocity of 150 rpm, 0.4% (v/v) water content, temperature of 30°C, 8 mg/mL dosage of Novozyme 435, 8:1 (0.4 mmol: 0.05 mmol) molar ratio of acetic anhydride to (R,S)-3-n-butylphthalide, and a reaction time of 48 hr. Under the optimum conditions, 96.4% eeS and 49.3% conversion of (R,S)-3-n-butylphthalide were achieved. In addition, enantiomeric excess of the product was above 98.0%.  相似文献   

14.
Synthesis, absolute configuration and antiviral activity of enantiomeric antiviral agents (R)-(-)- and (S)-(+)-synadenol (2 and 3a) are described.  相似文献   

15.
Both hitherto unknown (+)-(R)- and (?)-(S)-thioglycidyl esters, (R)-( 2 ) and (S)-( 2 ), have been synthesized with different high enantiomeric excesses (ee) by two routes from the corresponding rac-glycidyl esters rac-( 1 ). The first includes a porcine pancreatic lipase (PPL)-mediated kinetic resolution of these esters followed by sulfuration with practically complete inversion to the (+)-(R)-enantiomer (+)-(R)-( 2 ) (36–86% ee). (?)-(S)-Thioglycidyl esters (?)-(S)-( 2 ) are obtained by the reverse reaction sequence (43–80% ee). In the latter case the hydrolysis rate is lower than that of analogous glycidyl esters. Moreover, the dependence of enantiomeric excess on the size of the acyl-group is of the opposite tendency. Therefore, in both cases suitable selection of the acid residue gives rise to maximum enantioselectivity. The irreversible lipase-catalyzed acylation of rac-glycidol and rac-thioglycidol, however, was found to be a less suitable alternative. The enantiomeric excess of recovered homochiral esters was determined by chiral chromatography using modified cellulose stationary phases (OB, OD). © 1993 Wiley-Liss, Inc.  相似文献   

16.
During the investigation of chemical properties of the dicyclic system of insect juvenile hormone analogues, biotransformation of 2-(4-methoxybenzyl)-1-cyclohexanone (1) by plant cell cultures was studied. Among other components, the cis-(1S, 2S)- and cis-(1R, 2R)-2-(4-methoxybenzyl)-1-cyclohexanol enantiomers 2a and 2b were found in the reaction mixture. Higher stereoselectivity of the biotransformation was observed for trans-(1S, 2R)-enantiomer 3a formation, which occurred in at least 60% of calculated enantiomeric excess.  相似文献   

17.
cis- and trans-5-Ethoxy-1,3-oxathiolane-2-carboxylic acids were obtained in pure form. The cis isomer was resolved into its enantiomers through diastereoisomeric salt formation with enantiomerically pure α-methylbenzylamine. Reduction of the salt followed by benzoylation led to 2-benzoyloxymethyl-5-ethoxy-2(R)-5(S)-1,3-oxathiolane and 2-benzoyloxymethyl-5-ethoxy-2(S)-5(R)-1,3-oxathiolane, useful intermediates in nucleoside chemistry. © 1993 Wiley-Liss, Inc.  相似文献   

18.
The title compounds, 1a and 1b, have been synthesized in a three-step sequence starting from (-)-(S) and (+)-(R)-propylene oxide, respectively, in acceptable overall yields. The enantiomeric excess values for 1a and 1b were 96% and 93% respectively, as assessed by HPLC analysis on a chiral stationary phase of the corresponding N-acetyl derivatives. The synthetic route herein presented may represent a facile entry to highly enriched mexiletine enantiomers, alternative to those previously reported in the literature.  相似文献   

19.
从土壤中分离的1株产碱杆菌Alcaligenes sp.ECU0401具有扁桃酸脱氢酶活性,可以以扁桃酸、苯甲酰甲酸或苯甲酸为唯一C源生长,并且具有较高的脱氢酶活力。以外消旋扁桃酸为C源,采用分批补料策略培养(或反应)99h,扁桃酸累计投入量为30.4g/L,(S)-(+)-扁桃酸被完全降解,(R)-(-)-扁桃酸回收产率为32.8%,对映体过量值(e.e.)〉99.9%。利用静息细胞作为催化剂不对称降解外消旋扁桃酸的氯代衍生物,制备获得光学活性的(R)-(-)-邻氯扁桃酸、(S)-(+)-间氯扁桃酸和(S)-(+)-对氯扁桃酸,光学纯度均超过99.9%e.e.。  相似文献   

20.
Three different biochemical approaches were used for the synthesis of ethyl 5-(S)-hydroxyhexanoate 1 and 5-(S)-hydroxyhexanenitrile 2. In the first approach, ethyl 5-oxo-hexanoate 3 and 5-oxo-hexanenitrile 4 were reduced by Pichia methanolica (SC 16116) to the corresponding (S)-alcohols, ethyl (S)-5-hydroxyhexanoate 1 and 5-(S)-hydroxyhexanenitrile 2, with an 80-90% yield and >95% enantiomeric excess (e.e). In the second approach, racemic 5-hydroxyhexanenitrile 5 was resolved by enzymatic succinylation, leading to the formation of (R)-5-hydroxyhexanenitrile hemisuccinate and leaving the desired alcohol 5-(S)-hydroxyhexanenitrile 2 with a yield of 34% (50% maximum yield) and >99% e.e. In the third approach, enzymatic hydrolysis of racemic 5-acetoxy hexanenitrile 6 resulted in the hydrolysis of the R-isomer to provide 5-(R)-hydroxyhexanenitrile, leaving 5-(S)-acetoxyhexanenitrile 7 with a 42% yield (50% maximum yield) and >99% e.e.  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