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.
DL-Glutamic acid has been resolved into optically active forms as the diastereoisomeric salt of optically active neutral amino acid amides, and the salt is easily converted to the sodium salt of the active forms. By resolution with L-tyrosinamide and L-leucinamide, sodium L-glutamate was obtained in 65 and 80 per cent yield respectively. Attempts to extend this method to resolution of DL-glutamic acid using L-phenylalaninamide as resolving agent resulted in poor yield of less pure D-glutamic acid.

By the infrared spectroscopy and X-ray diffraction analysis it has been confirmed that the diastereoisomeric salts of L-leucinamide or L-tyrosinamide with L- or D-glutamic acid compose a combined salt structure in solid state, whereas L-phenylalaninamide with L- or D-glutamic acid does not compose a characteristic diastereoisomeric salt but rather the mechanical mixture of L-phenylalaninamide and L- or D-glutamate anion in solid state.

In the previous study1), it was reported that L-leucinamide forms the characteristic diastereoisomeric salts with racemic N-acyl mono amino acids, most of which are fairly resolved into their antipodes.  相似文献   

2.
Enzymatic resolution of 2-phenyl-1-propanol, enantiomeric ratios of which were recently improved up to 31 and 41 with lipase PS and PPL, respectively, by transesterification using vinyl 3-phenylpropanoate, was more excellently attained by PPL-catalyzed hydrolysis of its ester of 3-phenylpropanoic acid in 107 of E value.  相似文献   

3.
In this study an enantioseparation method for rac‐bambuterol (5‐(2‐(tert‐butylamino)‐1‐hydroxyethyl)‐1,3‐phenylene bis(dimethylcarbamate)) via diastereoisomeric salt formation with o‐chloromandelic acid was developed. The enantiomeric excess (ee) values and chemical purities of the desired products were confirmed by high‐performance liquid chromatography (HPLC) using chiral stationary phase and reverse‐phase HPLC analyses, respectively. The ee values and the chemical purities both exceeded 99%. Animal experiments showed that (R)‐bambuterol was a potent inhibitor for histamine‐induced asthma reactions. (S)‐bambuterol was ineffective in relaxing the airways. Both enantiomers increased heart rates in beagles. Therefore, replacing rac‐bambuterol with (R)‐bambuterol could be beneficial for asthma patients. Chirality 28:306–312, 2016. © 2016 Wiley Periodicals, Inc.  相似文献   

4.
As an example of acyclic P‐chiral phosphine oxides, the resolution of ethyl‐(2‐methylphenyl)‐phenylphosphine oxide was elaborated with TADDOL derivatives, or with calcium salts of the tartaric acid derivatives. Besides the study on the resolving agents, several purification methods were developed in order to prepare enantiopure ethyl‐(2‐methylphenyl)‐phenylphosphine oxide. It was found that the title phosphine oxide is a racemic crystal‐forming compound, and the recrystallization of the enantiomeric mixtures could be used for the preparation of pure enantiomers. According to our best method, the (R)‐ethyl‐(2‐methylphenyl)‐phenylphosphine oxide could be obtained with an enantiomeric excess of 99% and in a yield of 47%. Complete racemization of the enantiomerically enriched phosphine oxide could be accomplished via the formation of a chlorophosphonium salt. Characterization of the crystal structures of the enantiopure phosphine oxide was complemented with that of the diastereomeric intermediate. X‐ray analysis revealed the main nonbonding interactions responsible for enantiomeric recognition.  相似文献   

5.
Different strategies for stereoselective microbial preparation of various chiral aromatic compounds are described. Optically pure 2-methyl-3-phenyl-1-propanol, ethyl 2-methyl-3-phenylpropanoate, 2-methyl-3-phenylpropanal, 2-methyl-3-phenylpropionic acid and 2-methyl-3-phenylpropyl acetate have been prepared using different microbial biotransformations starting from different prochiral and/or racemic substrates. (S)-2-Methyl-3-phenyl-1-propanol and (S)-2-methyl-3-phenylpropanal were prepared by biotransformation of 2-methyl cinnamaldehyde using the recombinant strain Saccharomyces cerevisiae BY4741ΔOye2Ks carrying a heterologous OYE gene from Kazachstania spencerorum. (R)-2-Methyl-3-phenylpropionic acid was obtained by oxidation of racemic 2-methyl-3-phenyl-1-propanol with acetic acid bacteria. Kinetic resolution of racemic 2-methyl-3-phenylpropionic acid was carried out by direct esterification with ethanol using dry mycelia of Rhizopus oryzae CBS 112.07 in organic solvent, giving (R)-ethyl 2-methyl-3-phenylpropanoate as major enantiomer. Finally, (R,S)-2-methyl-3-phenylpropyl acetate was enantioselectively hydrolysed employing different bacteria and yeasts having cell-bound carboxylesterases with prevalent formation of (R)- or (S)-2-methyl-3-phenyl-1-propanol, depending on the strain employed.  相似文献   

6.
Summary Pig liver esterase (EC 3.1.1.1) catalyzed hydrolysis of the dimetrhy ester of meso-cis-1,2-cyclohexanedicarboxylic acid yielded the optically pure (1S,2R)-monoester. The corresponding diethyl ester yielded racemic monoester.The diethyl ester of racemic trans-1,2-cyclohexanedicarboxylic acid was kinetically resolved by partial hydrolysis with subtilisin (EC 3.4.21.14) or pig liver esterase. The (1R,2R)-monoester had an enantiomeric excess of 45% and was obtained in an enantiomerically pure form through recrystallisation. The remaining (1S,2S)-diester exhibited an enantiomeric excess of 83%. The nature of the ester function (methyl, ethyl, and propyl esters) had a great influence on the enantiomeric excess obtained and on the kinetic parameters.  相似文献   

7.
Mey B  Paulus H  Lamparter E  Blaschke G 《Chirality》1999,11(10):772-780
The enantiomers of the anorectic drug amfepramone [rac-diethylpropion, rac-2-(diethylamino)-1-phenyl-1-propanone; rac-DEP] were separated in the preparative scale by crystallization. With enantiopure di-O-benzoyltartaric acid as salt-forming chiral selector, diastereoisomeric salts of DEP enantiomers with a final purity of more than 97.5% were obtained. Analytical liquid chromatographic and electrophoretic methods for the control of the enantiomeric purity and the stoichiometry of the salts were developed. The enantioseparation of rac-DEP by capillary electrophoresis (CE) using hydroxypropyl-beta-cyclodextrin (HP-beta-CD) as chiral discriminator and phosphate buffer (pH 3.3) as run buffer led to good separations. HPLC methods were developed using polysaccharide chiral stationary phases (CSP). The separation of the two enantiomers and the two main degradation products (1-phenyl-1,2-propanedione and propiophenone), known from solid and liquid pharmaceutical preparations, was attained in one run on the silica-based CSP cellulose tris(3,5-dimethylphenylcarbamate) (Chiralcel OD). The conditions which might affect the enantioselectivity and the quality of the enantiomeric separation were investigated for Chiralcel OD and the related CSP amylose tris(3,5-dimethylphenylcarbamate) (Chiralpak AD). Both CSPs showed very similar chromatographic properties. The separation factors could be influenced significantly by varying the polar organic modifier added to the mobile phase.  相似文献   

8.
(1R,2R)-1-Phenyl-1-alkyl/arylamino-2-(N-alkylamino)propane hydrochloride salts have been synthesized with high degree of enantiomeric purity from (1S,2R)-(+)-1-phenyl-2-(N-alkylamino)-1-propanol through the corresponding chloro derivatives. This reaction sequence involves three inversions with overall inversion of configuration at C-1.  相似文献   

9.
Acetic acid bacteria (five strains of Acetobacter and five strains of Gluconobacter) were used for the biotransformation of different primary alcohols (2-chloropropanol and 2-phenylpropanol) and diols (1,3-butandiol, 1,4-nonandiol and 2,3-butandiol). Most of the tested strains efficiently oxidized the substrates. 2-Chloropropanol and 1,3-butandiol were oxidized with good rates and low enantioselectivity (enantiomeric excess=18–46% of the S-acid), while microbial oxidation of 2-phenylpropanol furnished (S)-2-phenyl-1-propionic acid with enantiomeric excess (e.e.) >90% with 10 strains. The dehydrogenation of 2,3-butandiol was strongly dependent on the stereochemistry of the substrate; the meso form gave S-acetoin with all the tested strains, the only exception being a Gluconobacter strain. The formation of diacetyl was observed only by using R,R-2,3-butandiol with Acetobacter strains. Oxidation of 1,4-nonandiol gave γ-nonanoic lactone in one step, although with moderate enantioselectivity.  相似文献   

10.
Summary A total of 1229 cultures, including 230 actinomycetes, 508 other bacteria, 12 fungi and 479 yeasts were screened for their ability to oxidize the isopropyl side chain of 2-phenyl propane (cumene). Four strains of actinomycetes and six strains of bacteria but no yeasts were found positive in converting 2-phenyl propane to its oxygenated products. Eight strains oxidized cumene through the alkyl side chain producing 2-phenyl-1-propanol. TwoBacillus strains oxidized cumene to an oxygenated product.Pseudomonas oleovorans NRRL B-3429 exhibited the highest alkyl side chain oxidation activity. The optimum reaction conditions for strain B-3429 are: 25 °C, pH 6.5 and 48 h of reaction. Octane-grown cells of strain B-3429 produced higher product yields (about 7.2-fold) than the glucose-grown cells. Prolonged incubation resulted in an increase in 2-phenyl-1-propionic acid production at the expense of 2-phenyl-1-propanol. The yield of 2-phenyl-1-propanol plus 2-phenyl-1-propionic acid was 5.1%. Reaction in the presence of methanol favored the accumulation of 2-phenyl-1-propionic acid and also increased the total yield. (The yield of 2-phenyl-1-propanol plus 2-phenyl-1-propionic acid was 14.9%.) Structures of the reaction products were confirmed by GC/MS and GC/IR analyses. Products contained 92% R(–) isomer.  相似文献   

11.
Racemic N-methylamphetamine (rac-MA) was resolved with 2R,3R-tartaric acid (TA) and its derivatives (O,O'-dibenzoyl-(2R,3R)-tartaric acid monohydrate (DBTA) and O,O'-di-p-toluoyl-(2R,3R)-tartaric acid (DPTTA)), individually and using them in different combinations. After partial diastereomeric salt formation, the free enantiomers were extracted by supercritical fluid extraction using carbon dioxide as solvent. DBTA and DPTTA are efficient resolving agents for rac-MA, the best chiral separation being obtained at a molar ratio of 0.25 resolving agent to racemic compound for both resolving agents (ee(E) = 82.5% and ee(E) = 57.9%, respectively). Compared with the two other acids, TA is practically unsuitable for enantiomer separation (ee(E) < 5%). Applying a mixture of one individually active and one ineffective acid in half the equivalent molar ratio, when the acids are in 1:1 ratio in the mixture, the resolution efficiency values obtained exceeded those obtained by using the components individually. Decreasing the molar ratio of resolving agent mixture to 0.25, at which the individual resolving agents give the best chiral separation, the obtained resolution efficiency values did not differ significantly from those expected. The outcome of the resolution process depended only on the amount of the individually active resolving agents in the mixture.  相似文献   

12.
The fluorides, chlorides and bromides of 3-halo-1-phenoxy-2-propanol, 3-halo-1-phenylmethoxy-2-propanol and 3-halo-1-(2-phenylethoxy)-2-propanol have been resolved by transesterification with various butanoates as acyl donors in hexane and lipase B from Candida antarctica (Novozyme 435) as catalyst. The enantiomeric ratio E depended on the hydroxy protecting groups in 1-position and the halogens in 3-position. For some substrates, the enantiomeric ratio was dependent on the acylating agent.  相似文献   

13.
Crown ethers are capable of complexing with primary amines and have been utilized in chromatography to separate amino acid racemates. This application has been extended to resolve (1‐amino‐1‐phenylmethyl)phosphonic acid and (1‐aminoethyl)phosphonic acid racemates, along with their aminocarboxylic acid analogs (2‐phenylglycine and alanine, respectively), via a ChiroSil RCA crown ether based chiral stationary phase. Effects of the organic modifier, temperature, and acid type and concentration on retention and selectivity were also investigated. Trends in retention and selectivity varied between aminophosponic acids and their aminocarboxylic analogs. Computer modeling and 1H NMR analyses were performed to potentially gain a better understanding of interactions of the aforementioned molecules with the ChiroSil RCA chiral stationary phase. Theoretical predictions of the most stable conformations for (R)‐ and (S)‐enantiomers were compared to elution order; it was found that the elution order agreed with molecular modeling such that the longest retention correlated with the predicted most stable complex between the enantiomer and crown ether. 1H NMR demonstrated interactions of aminophosphonic and aminocarboxylic racemates with (+)‐(18‐crown‐6)‐2,3,11,12‐tetracarboxylic acid in solution and was utilized to determine enantiomeric excess of (1‐amino‐1‐phenylmethyl)phosphonic acid after its enantioenrichment via crystallization through diastereomeric salt formation with the crown ether followed by filtration. Chirality 25:369–378, 2013. © 2013 Wiley Periodicals, Inc.  相似文献   

14.
The asymmetric resolution of various 1-aryl-3-buten-1-ols via microbial hydrolysis of the corresponding acetates has been investigated using different Rhizopus species. The chosen species, R. arrhizus (wild type), efficiently hydrolyzed 1-phenyl- and 1-para-substituted phenyl-3-buten-1-ol acetates, producing the enantiomerically pure (R)-alcohols with 53–65% yields. Although the antipode acetates were obtained with 9–52% enantiomeric excess, the (S)-alcohols were amenable in > 99% enantiomeric excess via a R. arrhizus mediated asymmetric reduction of the corresponding ketones.  相似文献   

15.
Maekawa H  Itoh K  Goda S  Nishiguchi I 《Chirality》2003,15(1):95-100
Anodic oxidation of 1-acetoxy-3,4-dihydronaphthalene (1) and alpha-acetoxy-beta-alkylstyrenes (3) at -78 degrees C in a mixed solvent of acetonitrile (CH(3)CN), tetrahydrofuran (THF), and acetic acid (AcOH) containing (S)-tetraethylammonium camphorsulfonate as a chiral supporting electrolyte brought about enantioselective formation of the corresponding 2-acetoxy-1-tetralones (2) and (R)-2-acetoxy-1-phenyl-1-alkanone (4) with maximum enantiomeric excess (ee) of 44% and 21%, respectively. Introduction of a 7-methoxy group into 1 and increase in bulkiness of a beta-alkyl group in 3 resulted in improvement of enantioselectivity of the reactions.  相似文献   

16.
Resolution of (R,S)-ibuprofen (2-(4-isobutylphenyl)propionic acid) enantiomers by esterification reaction with 1-propanol in different organic solvents was studied using native Aspergillus niger lipase. The main variables controlling the process (enzyme concentration and 1-propanol:ibuprofen molar ratio) have been optimized using response surface methodology based on a five-level, two-variable central composite rotatable design, in which the selected objective function was enantioselectivity. This enzyme preparation showed preferentially catalyzes the esterification of R(−)-ibuprofen, and under optimum conditions (7% w/v of enzyme and molar ratio of 2.41:1) the enantiomeric excess of active S(+)-ibuprofen and total conversion values were 79.1 and 48.0%, respectively, and the E-value was 32, after 168 h of reaction in isooctane.  相似文献   

17.
We studied the reaction between vinyl butyrate and 2-phenyl-1-propanol in acetonitrile catalyzed by Fusarium solani pisi cutinase immobilized on zeolites NaA and NaY and on Accurel PA-6. The choice of 2-phenyl-1-propanol was based on modeling studies that suggested moderate cutinase enantioselectivity towards this substrate. With all the supports, initial rates of transesterification were higher at a water activity (a(w)) of 0.2 than at a(w) = 0.7, and the reverse was true for initial rates of hydrolysis. By providing acid-base control in the medium through the use of solid-state buffers that control the parameter pH-pNa, which we monitored using an organo-soluble chromoionophoric indicator, we were able, in some cases, to completely eliminate dissolved butyric acid. However, none of the buffers used were able to improve the rates of transesterification relative to the blanks (no added buffer) when the enzyme was immobilized at an optimum pH of 8.5. When the enzyme was immobilized at pH 5 and exhibited only marginal activity, however, even a relatively acidic buffer with a pK(a) of 4.3 was able to restore catalytic activity to about 20% of that displayed for a pH of immobilization of 8.5, at otherwise identical conditions. As a(w) was increased from 0.2 to 0.7, rates of transesterification first increased slightly and then decreased. Rates of hydrolysis showed a steady increase in that a(w) range, and so did total initial reaction rates. The presence or absence of the buffers did not impact on the competition between transesterification and hydrolysis, regardless of whether the butyric acid formed remained as such in the reaction medium or was eliminated from the microenvironment of the enzyme through conversion into an insoluble salt. Cutinase enantioselectivity towards 2-phenyl-1-propanol was indeed low and was not affected by differences in immobilization support, enzyme protonation state, or a(w).  相似文献   

18.
Kozma D  Fogassy E 《Chirality》2001,13(8):428-430
Solvent-free optical resolution of N-methylamphetamine was developed by distillation after partial diastereoisomeric salt formation. From the 18 chiral acids tested by this method, five provide by this method resolution: O,O'-dibenzoyltartaric acid, O,O'-di-p-toluoyltartaric acid, 6-methoxy-alpha-methyl-2-naphthaleneacetic acid (Naproxen), the cis-permetrinic acid, and the 2-phenoxypropionic acid. Among them the O,O'-dibenzoyltartaric acid in water-free form provided the more effective resolution. The efficiency of this resolution S = 0.74 is in the range of the industrial-scale resolutions and not worse than the efficiency achieved by optical resolution via fractional crystallization.  相似文献   

19.
3-Chloro-1-phenyl-1-propanol is used as a chiral intermediate in the synthesis of antidepressant drugs. Various microbial reductases were expressed in Escherichia coli, and their activities toward 3-chloro-1-phenyl-1-propanone were evaluated. The yeast reductase YOL151W (GenBank locus tag) exhibited the highest level of activity and exclusively generated the (S)-alcohol. Recombinant YOL151W was purified by Ni-nitrilotriacetic acid (Ni-NTA) and desalting column chromatography. It displayed an optimal temperature and pH of 40°C and 7.5–8.0, respectively. The glucose dehydrogenase coupling reaction was introduced as an NADPH regeneration system. NaOH solution was occasionally added to maintain the reaction solution pH within the range of 7.0–7.5. By using this reaction system, the substrate (30 mM) could be completely converted to the (S)-alcohol product with an enantiomeric excess value of 100%. A homology model of YOL151W was constructed based on the structure of Sporobolomyces salmonicolor carbonyl reductase (Protein Data Bank ID: 1Y1P). A docking model of YOL151W with NADPH and 3-chloro-1-phenyl-1-propanone was then constructed, which showed that the cofactor and substrate bound tightly to the active site of the enzyme in the lowest free energy state and explained how the (S)-alcohol was produced exclusively in the reduction process.  相似文献   

20.
Previously, we reported that eight glucose-grown microbial cultures out of 1229 screened oxidize the alkyl side-chain of 2-phenylpropane (cumene) stereospecifically. Now, we have adapted these cultures to grow on n-octane and found that their cumene oxidation activities increased more than 30 times. We also found an additional 11 cultures (ten bacteria, one actinomycete) that oxidized cumene when grown on octane but not on glucose. In general, octane-grown cells were more active in cumene oxidation than glucose-grown cells. Rhodococcus rhodochrous NRRL B-2153 showed the best conversion yield (2-phenyl-1-propanol plus 2-phenyl-1-propionic acid was 5.5%) at 25°C, pH 8.0, 250 rpm, and 12 h of reaction. Structures of the reaction products were confirmed by gas chromatography (GC)/mass spectrometry and GC/infrared analyses. Products contained 84% ee (enantiomeric excess) of the R(–) isomer, as analyzed with a GC cyclodextrin chiral column. Strain B-2153 oxidized alkylbenzenes in the following order of reaction rate: ethylbenzene >amylbenzene > butylbenzene > cumene > propylbenzene > sec-butylbenzene. tert-Butylbenzene was not oxidized.  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