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.
Tang CG  Lin H  Zhang C  Liu ZQ  Yang T  Wu ZL 《Biotechnology letters》2011,33(7):1435-1440
Whole cells of Rhodotorula glutinis reduced N-methyl-3-oxo-3-(thiophen-2-yl) propanamide at 30 g/l to (S)-N-methyl-3-hydroxy-3-(2-thienyl) propionamide, an intermediate in the production of (S)-duloxetine, a blockbuster antidepressant drug, in 48 h. The reaction had excellent enantioselectivity (single enantiomer, >99.5% enantiomeric excess [ee]) with a >95% conversion.  相似文献   

2.
In order to define an enantioselective nuclear magnetic resonance (NMR) method for the antiasthmatic drug montelukast, a series of nine easily available products were evaluated as NMR chiral solvating agents (CSAs): D‐dibenzoyltartaric acid, D‐ditoluoyltartaric acid, (+)‐camphorsulfonic acid, (S)‐BINOL, (S)‐3,3’‐diphenyl‐2,2’‐binaphthyl‐1,1’‐diol, (R)‐3,3'′‐di‐9‐anthracenyl‐1,1'′‐bi‐2‐naphthol, (R)‐3,3'′‐di‐9‐phenanthrenyl‐1,1'′‐bi‐2‐naphthol, Pirkle's alcohol, and (?)‐cinchonidine. It was proved that most of the studied agents constitute diastereomeric complexes with both drug enantiomers in CD2Cl2 or CDCl3 solutions, thus permitting the direct 1H NMR detection of the unwanted S‐enantiomer, even at levels of 0.75%. (?)‐Cinchonidine was found to be the more convenient CSA in terms of NMR enantiodiscrimination power and ease of experimental requirements. The final method was validated and applied to the fast monitoring of the optical purity of montelukast “in‐process” samples, circumventing the need for tedious and slower analytical procedures like enantioselective chromatography or capillary electrophoresis. In addition, a method for the enantiopurity control of the commercial drug (montelukast sodium salt) was also established using (S)‐BINOL as NMR CSA. Chirality 25: 780–786, 2013. © 2013 Wiley Periodicals, Inc.  相似文献   

3.
A novel strategy for a more efficient synthesis of difficult sequence‐containing peptides, the S‐acyl isopeptide method, was developed and successfully applied. A model pentapeptide Ac–Val–Val–Cys–Val–Val–NH2 was synthesized via its water‐soluble S‐acyl isopeptide using an S‐acyl isodipeptide unit, Boc–Cys(Fmoc–Val)–OH. An S‐acyl isopeptide possessing excellent water solubility could be readily and quantitatively converted to the native peptide via an SN intramolecular acyl migration reaction at pH 7.4. Thus, the S‐acyl isopeptide method provides a useful tool in peptide chemistry. Copyright © 2008 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

4.
Stereoselective metabolism of propranolol side‐chain glucuronidation was studied for two recombinant human uridine diphosphate glucuronosyltransferases (UGTs), UGT1A9 and UGT2B7. The S‐ and R‐propranolol side‐chain glucuronides produced in the incubation mixtures were assayed simultaneously by RP‐HPLC with fluorescent detector. The excitation and emission wavelengths were set at 310 nm and 339 nm, respectively. UGT1A9 prefers catalyzing S‐enantiomer to R‐enantiomer and the intrinsic clearance (CLint) ratios of S‐enantiomer to R‐enantiomer are 3.8 times and 6.5times for racemic propranolol and individual enantiomers, respectively. UGT2B7, however, catalyzes slightly less S‐enantiomer than R‐enantiomer and the CLint ratio of S‐enantiomer to R‐enantiomer is 0.8 times. The high concentration of racemic propranolol (>0.57 mmol/l) and individual enantiomers (>0.69 mmol/l) exhibited substrate inhibition of glucuronidation for UGT2B7, but only the S‐enantiomer (>0.44 mmol/l) in racemic propranolol exhibited substrate inhibition for UGT1A9. The substrate inhibition constants (Ksi) were all similar (P > 0.05). Drug–drug interactions were also found between S‐ and R‐enantiomer glucuronidation metabolisms by UGT1A9 and UGT2B7. Chirality 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

5.
The acetylcholinesterase inhibition by enantiomers of exo‐ and endo‐2‐norbornyl‐Nn‐butylcarbamates shows high stereoselelectivity. For the acetylcholinesterase inhibitions by (R)‐(+)‐ and (S)‐(?)‐exo‐2‐norbornyl‐Nn‐butylcarbamates, the R‐enantiomer is more potent than the S‐enantiomer. But, for the acetylcholinesterase inhibitions by (R)‐(+)‐ and (S)‐(?)‐endo‐2‐norbornyl‐Nn‐butylcarbamates, the S‐enantiomer is more potent than the R‐enantiomer. Optically pure (R)‐(+)‐exo‐, (S)‐(?)‐exo‐, (R)‐(+)‐endo‐, and (S)‐(?)‐endo‐2‐norbornyl‐Nn‐butylcarbamates are synthesized from condensations of optically pure (R)‐(+)‐exo‐, (S)‐(?)‐exo‐, (R)‐(+)‐endo‐, and (S)‐(?)‐endo‐2‐norborneols with n‐butyl isocyanate, respectively. Optically pure norborneols are obtained from kinetic resolutions of their racemic esters by lipase catalysis in organic solvent. Chirality 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

6.
The addition reaction to N-methyl-(S)-alanine or N-methyl-(S)-phenylalanine N-car-boxyanhydride (NCA) of 3-methyl-5-substituted hydantoin (HDT) catalyzed by a tertiary amine was investigated as a model reaction for the propagation reaction of NCA according to the activated-NCA mechanism. Several activated HDTs having the (S)-configuration of the asymmetric carbon atom were found to react more rapidly than their activated enantiomers. This experimental result indicates that the enantiomer selection by terminal-unit control takes place in the propagation reaction according to the activated-NCA mechanism in which an activated NCA is added to a terminal acylated NCA ring of the growing chain. The enantiomer excess of the HDT recovered from the reaction mixture of N-methyl-(S)-phenylalanine NCA and racemic HDTs activated by a tertiary amine was determined. The extent of the enantiomer selection in the polymerization was found to be 3–10 times as large as that in the model reaction. From these results, it was concluded that the chirality of the penultimate unit, as well as that of the terminal NCA ring, plays an important role in determining the enantiomer selection in the NCA polymerization.  相似文献   

7.
The enantiomeric purity of escitalopram oxalate ESC and its “in‐process impurities,” namely, ESC‐N‐oxide, ESC‐citadiol, and R(?)‐enantiomer were studied in drug substance and products using high‐performance liquid chromatography (HPLC)‐UV (Method I), synchronous fluorescence spectroscopy (SFS) (Method IIA), and first derivative SFS (Method IIB). Method I describes as an isocratic HPLC‐UV for the direct resolution and determination of enantiomeric purity of ESC and its “in‐process impurities.” The proposed method involved the use of αl‐acid glycoprotein (AGP) chiral stationary phase. The regression plots revealed good linear relationships of concentration range of 0.25 to 100 and 0.25 to 10 μg mL?1 for ESC and its impurities. The limits of detection and quantifications for ESC were 0.075 and 0.235 μg mL?1, respectively. Method II involves the significant enhancement of the fluorescence intensities of ESC and its impurities through inclusion complexes formation with hydroxyl propyl‐β‐cyclodextrin as a chiral selector in Micliavain buffer. Method IIA describes SFS technique for assay of ESC at 225 nm in presence of its impurities: R(?)‐enantiomer, citadiol, and N‐oxide at ?λ of 100 nm. This method was extended to (Method IIB) to apply first derivative SFS for the simultaneous determination of ESC at 236 nm and its impurities: the R(?)‐enantiomer, citadiol, and N‐oxide at 308, 275, and 280 nm, respectively. Linearity ranges were found to be 0.01 to 1.0 μg mL?1 for ESC and its impurities with lower detection and quantification limits of 0.033/0.011 and 0.038/0.013 μg mL?1 for SFS and first derivative synchronous fluorescence spectra (FDSFS), respectively. The methods were used to investigate the enantiomeric purity of escitalopram.  相似文献   

8.
A profoundly time‐efficient chemoenzymatic method for the synthesis of (S)3‐(4‐chlorophenoxy)propan‐1,2‐diol and (S)1‐chloro‐3‐(2,5‐dichlorophenoxy)propan‐2‐ol, two important pharmaceutical intermediates, was successfully developed using Pseudomonas fluorescens lipase (PFL). Kinetic resolution was successfully achieved using vinyl acetate as acylating agent, toluene/hexane as solvent, and reaction temperature of 30°C giving high enantioselectivity and conversion. Under optimized condition, PFL demonstrated 50.2% conversion, enantiomeric excess of 95.0%, enantioselectivity (E = 153) in an optimum time of 1 hour and 50.3% conversion, enantiomeric excess of 95.2%, enantioselectivity (E = 161) in an optimum time of 3 hours, for the two racemic alcohols, respectively. Docking of the R‐ and S‐enantiomers of the intermediates demonstrated stronger H‐bond interaction between the hydroxyl group of the R‐enantiomer and the key binding residues of the catalytic site of the lipase, while the S‐enantiomer demonstrated lesser interaction. Thus, docking study complemented the experimental outcome that PFL preferentially acylated the R form of the intermediates. The present study demonstrates a cost‐effective and expeditious biocatalytic process that can be applied in the enantiopure synthesis of pharmaceutical intermediates and drugs.  相似文献   

9.
Chiral distinction in the proton pump inhibitor drugs omeprazole and in its chiral‐switch esomeprazole magnesium was studied employing the Density Functional Theory (DFT) method. At B3LYP/6‐311G(d,p), the 6‐methoxy???6‐methoxy and 5‐methoxy???5‐methoxy homochiral and heterochiral dimers were calculated. The chiral distinction free energies (ΔΔG298,(RS‐SS)) between the cyclic C2‐(S,S)‐ and Ci‐(R,S)‐dimers with two intermolecular hydrogen bonds are 3.8, 1.9 (with BSSE counterpoise correction), and –6.9 (with D3 dispersion and BSSE counterpoise corrections) kJ/mol. Adding water as an implicit solvent (polarized continuum model [PCM] model) resulted in a chiral distinction energy of –3.3 kJ/mol, indicating a reversal of the order of the relative stabilities of C2‐(S,S)‐ and Ci‐(R,S)‐dimers. The chiral distinction free energies between the corresponding (less stable) C1‐dimers with one intermolecular hydrogen bond are –9.3, –5.8 (with BSSE CC), 17.6 (D3 + BSSE CC), and –3.2 (H2O) kJ/mol. The results highlight the contention that omeprazole is not just a superposition of its enantiomer constituents. They are consistent with the pharmacological evidence of enantiomer–enantiomer interactions in omeprazole versus esomeprazole and the differences between the drugs omeprazole and esomeprazole magnesium and support the lodged application for regulatory supplementary protection certificate (SPC) exclusivity for the esomeprazole‐related combination drug Vimovo. Chirality 26:214–227, 2014. © 2014 Wiley Periodicals, Inc.  相似文献   

10.
A biocatalytic route for the synthesis of a potential β‐blocker, (S)‐moprolol is reported here. Enantiopure synthesis of moprolol is mainly dependent on the chiral intermediate, 3‐(2‐methoxyphenoxy)‐propane‐1,2‐diol. Various commercial lipases were screened for the enantioselective resolution of (RS)‐3‐(2‐methoxyphenoxy)propane‐1,2‐diol to produce the desired enantiomer. Among them, Aspergillus niger lipase (ANL) was selected on the basis of both stereo‐ and regioselectivity. The optimized values of various reaction parameters were determined such as enzyme (15 mg/mL), substrate concentration (10 mM), organic solvent (toluene), reaction temperature (30 °C), and time (18 h).The optimized conditions led to achieving >49% yield with high enantiomeric excess of (S)‐3‐(2‐methoxyphenoxy)propane‐1,2‐diol. The lipase‐mediated catalysis showed regioselective acylation with dual stereoselectivity. Further, the enantiopure intermediate was used for the synthesis of (S)‐moprolol, which afforded the desired β‐blocker. Chirality 28:313–318, 2016. © 2016 Wiley Periodicals, Inc.  相似文献   

11.
Flurbiprofen (F) is a nonsteroidal anti‐inflammatory drug (NSAID) used therapeutically as the racemate of (R)‐enantiomer and (S)‐enantiomer. The inversion of RF to SF and vice versa was investigated in C57Bl/6 and SJL mice and Dark Agouti and Lewis rats. The enzyme α‐methylacyl‐CoA racemase (AMACR) is involved in the chiral inversion pathway that converts members of the 2‐arylpropionic acid NSAIDs from the R‐enantiomer to the S‐enantiomer. We studied C57Bl/6 mice deficient in AMACR postulating that they should show reduced inversion of RF to SF. In line with the data of others in mice, (R)‐inversion to (S)‐inversion was relatively high in both the C57Bl/6 and SJL mice (fraction inverted, FI = 37.7% and 24.7%, respectively). In contrast, in AMACR deficient mice, there was no measurable peak for SF after administration of RF. The results in both rat strains (Dark Agouti and Lewis rats, FI = 1.4% and 4.1%, respectively) confirm the low chiral inversion of the enantiomers of flurbiprofen in the rat, as observed by other authors in the Sprague‐Dawley strain (<5%). From the present results, we conclude that for the study of flurbiprofen enantiomers, the rat is more suitable than the mouse as a model for the human in which (R)‐inversion to (S)‐inversion is negligible.  相似文献   

12.
A simple, sensitive, and robust normal‐phase isocratic HPLC‐UV method was developed and validated for the enantiomeric separation of rasagiline mesylate and its (S)‐enantiomer. The rasagiline and its (S)‐enantiomer were resolved on a Chiralcel‐OJ‐H (4‐methylbenzoate cellulose coated on silica) column using a mobile phase consisting of n‐hexane:isopropyl alcohol:ethanol:diethyl amine (96:2:2:0.01) at a flow rate of 1.0 ml/min. The column temperature was maintained at 27 °C and elution was monitored at 215 nm. The resolution (Rs) between the enantiomers was found to be more than 2.0. The limit of detection and the limit of quantification of the (S)‐enantiomer were found to be 0.35 and 1.05 µg/ml, respectively. The developed method was validated as per ICH guidelines with respect to linearity, limit of detection and quantification, accuracy, precision, and robustness—and satisfactory results were obtained. The sample solution and mobile phase were found to be stable up to 48 h. The method is useful for routine evaluation of the quality of rasagiline mesylate in bulk drug‐manufacturing units. Chirality 25:324–327, 2013. © 2013 Wiley Periodicals, Inc.  相似文献   

13.
Reactions of lithium halide (LiX, X = F, Cl, Br and I) and methyl halide (CH3X, X = F, Cl, Br and I) have been investigated at the B3LYP/6-31G(d) level of theory using the microhydration model. Beginning with hydrated lithium ion, four or two water molecules have been conveniently introduced to these aqueous-phase halogen-exchange SN2 reactions. These water molecules coordinated with the center metal lithium ion, and also interacted with entering and leaving halogen anion via hydrogen bond in complexes and transition state, which to some extent compensated hydration of halogen anion. At 298 K the reaction profiles all involve central barriers ΔE cent which are found to decrease in the order F > Cl > Br > I. The same trend is also found for the overall barriers (ΔE ovr ) of the title reaction. In the SN2 reaction of sodium iodide and methyl iodide, the activation energy agrees well with the aqueous conductometric investigation.  相似文献   

14.
Fluoxetine is used clinically as a racemic mixture of (+)‐(S) and (–)‐(R) enantiomers for the treatment of depression. CYP2D6 catalyzes the metabolism of both fluoxetine enantiomers. We aimed to evaluate whether exposure to gasoline results in CYP2D inhibition. Male Wistar rats exposed to filtered air (n = 36; control group) or to 600 ppm of gasoline (n = 36) in a nose‐only inhalation exposure chamber for 6 weeks (6 h/day, 5 days/week) received a single oral 10‐mg/kg dose of racemic fluoxetine. Fluoxetine enantiomers in plasma samples were analyzed by a validated analytical method using LC‐MS/MS. The separation of fluoxetine enantiomers was performed in a Chirobiotic V column using as the mobile phase a mixture of ethanol:ammonium acetate 15 mM. Higher plasma concentrations of the (+)‐(S)‐fluoxetine enantiomer were found in the control group (enantiomeric ratio AUC(+)‐(S)/(–)‐(R) = 1.68). In animals exposed to gasoline, we observed an increase in AUC0‐∞ for both enantiomers, with a sharper increase seen for the (–)‐(R)‐fluoxetine enantiomer (enantiomeric ratio AUC(+)‐(S)/(–)‐(R) = 1.07), resulting in a loss of enantioselectivity. Exposure to gasoline was found to result in the loss of enantioselectivity of fluoxetine, with the predominant reduction occurring in the clearance of the (–)‐(R)‐fluoxetine enantiomer (55% vs. 30%). Chirality 25:206–210, 2013. © 2013 Wiley Periodicals, Inc.  相似文献   

15.
C2‐symmetric N,N‐bis(phosphinomethyl)amines were prepared by the thermal reaction of aromatic aldehydes with ammonia and hypophosphorus acid as previously described. Both enantiomers of C2‐symmetric N,N‐bis(phosphinomethyl)amine were obtained in a high enantiomeric purity through the diastereomeric salt formation with (–)‐quinine, and subsequent fractional crystallization. X‐ray crystallographic analysis of one of the diastereomeric salts clearly revealed that (–)‐quinine could be an efficient resolving agent for obtaining the single enantiomer (R,R)‐N,N‐bis(phosphinomethyl)amine. Chirality 27:71–74, 2015. © 2014 Wiley Periodicals, Inc.  相似文献   

16.
The titration of the (S)‐enantiomer of omeprazole with the (R)‐enantiomer in chloroform‐d1 is monitored by continuous‐flow capillary microcoil 1H NMR spectroscopy employing a microcoil with a detection volume of 1.5 µl. The observed changes of the 1H NMR chemical shifts indicate the formation of a heterochiral (R,S) dimer of omeprazole via its sulfinyl group and the NH group of the benzimidazole ring. Chirality 24:1074–1076, 2012. © 2012 Wiley Periodicals, Inc.  相似文献   

17.
An efficient method of peptide thioester synthesis is described. The reaction is based on an N‐4,5‐dimethoxy‐2‐mercaptobenzyl (Dmmb) auxiliary‐assisted NS acyl shift reaction after assembling a peptide chain by Fmoc‐solid phase peptide synthesis. The Dmmb‐assisted NS acyl shift reaction proceeded efficiently under mildly acidic conditions, and the peptide thioester was obtained by treating the resulting S‐peptide with sodium 2‐mercaptoethanesulfonate. No detectable epimerization of the amino acid residue adjacent to the thioester moiety in the case of Leu was found. The reactions were also amenable to the on‐resin preparation of peptide thioesters. The utility was demonstrated by the synthesis of a 41‐mer peptide thioester, a phosphorylated peptide thioester and a 33‐mer peptide thioester containing a trimethylated lysine residue. Copyright © 2009 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

18.
The pharmacokinetics of (?)‐N‐(trans‐4‐isopropylcyclohexanecarbonyl)‐D ‐phenylalanine (nateglinide) and its enantiomer (L‐enantiomer) was studied in Goto‐Kakizaki (GK) rats after intravenous administration of nateglinide or L‐enantiomer at a dose of 40 μmol/kg body weight. Nateglinide, its L‐enantiomer and their metabolites in serum, bile and urine were determined. The total clearance (CLtot) and the volume of distribution (Vd) was slightly higher for nateglinide than those for L‐enantiomer in control rats, although the differences were not statistically significant. The cumulative excretions of L‐M1 (major metabolite of L‐enantiomer) and L‐M2 (major metabolite of L‐enantiomer) into bile were almost the same as that of M1 (major metabolite of nateglinide)and M2 (major metabolite of nateglinide). In GK rats, CLtot and Vd were higher for nateglinide than those for L‐enantiomer. The cumulative excretion of L‐M1 and L‐M2 were not different from those of M1 and M2, respectively, into bile or urine. CLtot and Vd for nateglinide or L‐enantiomer in GK rats were not different from those in control rats. The total excretion of M1, M2, L‐M1, and L‐M2 into bile or urine in GK rats was not substantially different from that of control rats. These results suggest that the L‐enantiomer of nateglinide shows higher CLtot and Vd compared with nateglinide, especially in the diabetic state. Chirality, 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

19.
The present study describes the development of two approaches for the determination of the enantiopurity of both enantiomers of indatraline. Initially, a method was developed using different chiral solvating agents (CSAs) for diastereomeric discrimination regarding signal separation in 1H nuclear magnetic resonance (NMR) spectroscopy, revealing MTPA as a promising choice for the differentiation of the indatraline enantiomers. This CSA was also tested for its ideal molar ratio, temperature, and solvent. Optimized conditions could be achieved that made determination of enantiopurity for (1R,3S)‐indatraline up to 98.9% enantiomeric excess (ee) possible. To quantify even higher enantiopurities, a high‐performance liquid chromatography (HPLC) method based on a modified β‐cyclodextrine phase was established. The influence of buffer type, concentration, pH value, percentage and kind of organic modifier, temperature, injection volume as well as sample solvent on chromatographic parameters was investigated. Afterwards, the reliability of the established HPLC method was demonstrated by validation according to the ICH guideline Q2(R1) regarding specificity, accuracy, precision, linearity, and quantitation limit. The developed method proved to be strictly linear within a concentration range of 1.25–1000 μM for the (1R,3S)‐enantiomer and 1.25‐750 μM for its mirror image that enables a reliable determination of enantiopurities up to 99.75% ee for the (1R,3S)‐enantiomer and up to 99.67% ee for the (1S,3R)‐enantiomer. Chirality 25:923–933, 2013. © 2013 Wiley Periodicals, Inc.  相似文献   

20.
Compounds based on the pyrroloquinoxaline system can interact with serotonin 5‐HT3, cannabinoid CB1, and μ‐opioid receptors. Herein, a chiral pool synthesis of diastereomerically and enantiomerically pure bromolactam (S,R,R,R)‐ 14A is presented. Introduction of the cyclohexenyl ring at the N‐atom of (S)‐proline derivatives 8 or methyl (S)‐pyroglutamate ( 12 ) led to the N‐cyclohexenyl substituted pyrrolidine derivatives 4 and 13 , respectively. All attempts to cyclize the (S)‐proline derivatives 4 with a basic pyrrolidine N‐atom via [3 + 2] cycloaddition, aziridination, or bromolactamization failed. Fast aromatization occurred during treatment of cyclohexenamines under halolactamization conditions. In contrast, reaction of a 1:1 mixture of diastereomeric pyroglutamates (S,R)‐ 13bA and (S,S)‐ 13bB with LiOtBu and NBS provided the tricyclic bromolactam (S,R,R,R)‐ 14A with high diastereoselectivity from (S,R)‐ 13bA , but did not transform the diastereomer (S,S)‐ 13bB . The different behavior of the diastereomeric pyroglutamates (S,R)‐ 13bA and (S,S)‐ 13bB is explained by different energetically favored conformations. Chirality 26:793–800, 2014. © 2014 Wiley Periodicals, Inc.  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