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 the present study we investigated the enantioselectivity in the pharmacokinetics of metoprolol administered in a multiple‐dose regimen as the racemate. The study was conducted on 10 patients of both sexes with mild to severe essential hypertension, aged 28 to 76 years, with normal hepatic and renal function and phenotyped as extensive metabolizers of debrisoquine (urine debrisoquine to 4‐hydroxydebrisoquine ratios of 0.28 to 6.56). The patients were treated with racemic metoprolol (two 100 mg tablets every 24 h) for 7 days. Serial blood samples were collected at times zero, 0.5, 1, 1.5, 2, 3, 4, 6, 8, 10, 12, 16, 20, 22, and 24 h and urine at each 6 h period until 24 h after metoprolol administration. The plasma concentrations of the (−)‐(S)‐ and (+)‐(R)‐metoprolol enantiomers were determined by HPLC using a chiral stationary phase (Chiralpak AD, 4.6 × 250 mm) and fluorescence detection. The enantiomeric ratios differing from one were evaluated by the paired t test and the results are reported as means (95% CI). No differences were observed between metoprolol enantiomers in half‐lives and absorption, distribution and elimination rate constants. However, the following differences (p < 0.05) were observed between the (−)‐(S) and (+)‐(R) enantiomers: maximum plasma concentration, Cmax, 179.99 (123.33–236.64) versus 151.30 (95.04–207.57) ng/mL; area under the plasma concentration versus time curve, AUC, 929.85 (458.02–1401.70) versus 782.11 (329.80–1234.40) ng h/mL; apparent total clearance, ClT/f, 1.70 (0.79–2.61) versus 2.21 (1.06–3.36) L/h/kg, apparent distribution volume, Vd/f, 10.51 (6.35–14.68) versus 13.80 (6.93–20.68) L/kg, and renal clearance, ClR, 0.06 (0.05–0.08) versus 0.07 (0.05–0.09) L/kg. The enantiomeric ratios AUC(−)‐(S)/AUC(+)‐(R) ranged from 1.14 to 1.44, with a mean of 1.29. The data obtained demonstrate enantioselectivity in the kinetic disposition of metoprolol, with plasma accumulation of the pharmacologically more active (−)‐(S)‐metoprolol enantiomer in hypertensive patients phenotyped as extensive metabolizers of debrisoquine. Chirality 11:591–597, 1999. © 1999 Wiley‐Liss, Inc.  相似文献   

2.
Carvedilol is an antihypertensive drug available as a racemic mixture. (?)‐(S)‐carvedilol is responsible for the nonselective β‐blocker activity but both enantiomers present similar activity on α1‐adrenergic receptor. To our knowledge, this is the first study of carvedilol enantiomers in human plasma using a chiral stationary phase column and liquid chromatography with tandem mass spectrometry. The method involves plasma extraction with diisopropyl ether using metoprolol as internal standard and direct separation of the carvedilol enantiomers on a Chirobiotic T® (Teicoplanin) column. Protonated ions [M + H]+ and their respective ion products were monitored at transitions of 407 > 100 for the carvedilol enantiomers and 268 > 116 for the internal standard. The quantification limit was 0.2 ng ml?1 for both enantiomers in plasma. The method was applied to study enantioselectivity in the pharmacokinetics of carvedilol administered as a single dose of 25 mg to a hypertensive patient. The results showed a higher plasma concentration of (+)‐(R)‐carvedilol (AUC0–∞ 205.52 vs. 82.61 (ng h) ml?1), with an enantiomer ratio of 2.48. Chirality, 2012. © 2012 Wiley Periodicals, Inc.  相似文献   

3.
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   

4.
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   

5.
A high-performance liquid chromatographic (HPLC) technique is described for quantification of R(+)- and S(−)-propranolol from 100-μl rat blood samples. The procedure involves chiral derivatization with tert.-butoxycarbonyl- -leucine anhydride to form diastereomeric propranolol- -leucine derivatives which are separated on a reversed-phase HPLC column. The method as previously reported has been modified for assaying serial blood microsamples obtained from the rat for pharmacokinetic studies. An internal standard, cyclopentyldesisopropylpropranolol, has been incorporated into the assay and several derivatization parameters have been altered. Standard curves for both enantiomers were linear over a 60-fold concentration range in 100-μl samples of whole rat blood (12.5–750 ng/ml; r=0.9992 for each enantiomer). Inter- and intra-assay variability was less than 12% for each enantiomer at 25 ng/ml. No enantiomeric interference or racemization was observed as a result of the derivatization. No analytical interference was noted from endogenous components in rat blood samples. Preliminary data from two male Sprague-Dawley rats given a 2.0 mg/kg intravenous dose of racemic propranolol revealed differential disposition of the two enantiomers. R(+)-Propranolol achieved higher initial concentration but was eliminated more rapidly than S(−)-propranolol. Terminal half-lives of R(+)- and S(−)-propranolol were 19.23 and 51.95 min, respectively, in one rat, and 14.50 and 52.07 min, respectively, in the other.  相似文献   

6.
Metoprolol is available for clinical use as a racemic mixture. The S‐(?)‐metoprolol enantiomer is the one expressing higher activity in the blockade of the β1‐adrenergic receptor. The α‐hydroxymetoprolol metabolite also has activity in the blockade of the β1‐adrenergic receptor. The present study describes the development and validation of a stereoselective method for sequential analysis of metoprolol and of α‐hydroxymetoprolol in plasma using high‐performance liquid chromatography with tandem mass spectrometry (LC‐MS/MS). 1‐ml aliquots of plasma were extracted with dichloromethane : diisopropyl ether (1:1, v/v). Metoprolol enantiomers and α‐hydroxymetoprolol isomers were separated on a Chiralpak AD column (Daicel Chemical Industries, New York, NY, USA) and quantitated by LC‐MS/MS. The limit of quantitation obtained was 0.2 ng of each metoprolol enantiomer/ml plasma and 0.1 ng/ml of each α‐hydroxymetoprolol isomer/ml plasma. The method was applied to the study of kinetic disposition of metoprolol in plasma samples collected up to 24 h after the administration of a single oral dose of 100‐mg metoprolol tartrate to a hypertensive parturient with a gestational age of 42 weeks. The clinical study showed that the metoprolol pharmakokinetics is enantioselective, with the observation of higher area under the curve (AUC)0?∞ values for S‐(?)‐metoprolol (AUCS‐(?)/AUCR‐(+) = 1.81) and the favoring of the formation of the new chiral center 1′R of α‐hydroxymetoprolol (AUC0?∞1′R/1′S = 2.78). Chirality, 25:1–7, 2013. © 2012 Wiley Periodicals, Inc.  相似文献   

7.
The purpose of this study was to elucidate the pharmacokinetics of terazosin enantiomers in healthy Chinese male subjects. After a single oral dose of 2‐mg terazosin, the plasma concentrations of terazosin enantiomers were measured over the course of 48 h in 12 healthy subjects. The plasma concentrations of (+)‐(R)‐terazosin at all time points were higher than those of (?)‐(S)‐terazosin. The area under the plasma concentration–time curve (AUC0–∞) and maximum plasma concentration of (+)‐(R)‐terazosin were significantly greater than those of the (?)‐(S)‐terazosin (P < 0.01, respectively). The R/S ratio of AUC0–∞ of terazosin was 1.68. For the first time, it was proven that the pharmacokinetics of terazosin was stereoselective in healthy Chinese male subjects. Chirality 24:1047–1050, 2012. © 2012 Wiley Periodicals, Inc.  相似文献   

8.
Propylisopropyl acetamide (PID), an amide analogue of the major antiepileptic drug valproic acid (VPA), possesses favorable anticonvulsant and CNS properties. PID contains one chiral carbon atom and therefore exists in two enantiomeric forms. The purpose of this work was to synthesize the two PID enantiomers and evaluate their enantiospecific teratogenicity. Enantioselective synthesis of PID enantiomers was achieved by coupling valeroyl chloride with optically pure (4S)‐ and (4R)‐benzyl‐2‐oxazolidinone chiral auxiliaries. The two oxazolidinone enolates were alkylated with isopropyl triflate, hydrolyzed, and amidated to yield (2R)‐ and (2S)‐PID. These two PID enantiomers were obtained with excellent enantiomeric purity, exceeding 99.4%. Unlike VPA, both (2R)‐ and (2S)‐PID failed to exert teratogenic effects in NMRI mice following a single 3 mmol/kg subcutaneous injection. From this study we can conclude that individual PID enantiomers do not demonstrate stereoselective teratogenicity in NMRI mice. Due to its better anticonvulsant activity than VPA and lack of teratogenicity, PID (in a stereospecific or racemic form) has the potential to become a new antiepileptic and CNS drug. Chirality 11:645–650, 1999. © 1999 Wiley‐Liss, Inc.  相似文献   

9.
(1R)‐Normetanephrine is the natural stereoisomeric substrate for sulfotransferase 1A3 (SULT1A3)‐catalyzed sulfonation. Nothing appears known on the enantioselectivity of the reaction despite its potential significance in the metabolism of adrenergic amines and in clinical biochemistry. We confronted the kinetic parameters of the sulfoconjugation of synthetic (1R)‐normetanephrine and (1S)‐normetanephrine by recombinant human SULT1A3 to a docking model of each normetanephrine enantiomer with SULT1A3 and the 3′‐phosphoadenosine‐5′‐phosphosulfate cofactor on the basis of molecular modeling and molecular dynamics simulations of the stability of the complexes. The KM, Vmax, and kcat values for the sulfonation of (1R)‐normetanephrine, (1S)‐normetanephrine, and racemic normetanephrine were similar. In silico models were consistent with these findings as they showed that the binding modes of the two enantiomers were almost identical. In conclusion, SULT1A3 is not substrate‐enantioselective toward normetanephrine, an unexpected finding explainable by a mutual adaptability between the ligands and SULT1A3 through an “induced‐fit model” in the catalytic pocket. Chirality, 25:28‐34, 2012.© 2012 Wiley Periodicals, Inc.  相似文献   

10.
The occurrence of pharmaceuticals in the environment represents a challenge of emerging concern. Many pharmaceuticals are chiral compounds; however, few studies have examined the relative toxicity of pharmaceutical enantiomers to wildlife. Further, our understanding of stereospecific pharmacokinetics remains largely informed by research on humans and a few well‐studied laboratory test animals, and not by studies conducted with environmentally relevant species, including fish. The objective of this study was to investigate whether rainbow trout display stereospecific in vitro metabolism of three common chiral pharmaceuticals. Metabolism by trout liver S9 fractions was evaluated using a substrate depletion approach, which provides an estimate of intrinsic hepatic clearance (CLIN VITRO,INT). No biotransformation was observed for rac‐, R‐, or S‐fluoxetine. Ibuprofen, including both enantiomers and the racemic mixture, appeared to undergo slow metabolism, but the resulting substrate depletion curves did not differ significantly from those of inactive controls. Contrary to relative clearance rates in humans, S(?)‐propranolol was more rapidly cleared than the R(+)‐ enantiomer. This work demonstrates that relative clearance rates and the effects of racemic mixtures in trout could not have been predicted based on human data. Additional research describing species differences and exploring tools for species extrapolation in biomedical and environmental studies is needed. Chirality 25:763–767, 2013, © 2013 Wiley Periodicals, Inc.  相似文献   

11.
A biotransformation process was developed for the production of (S)-ketoprofen by enantioseletive hydrolysis of racemic ketoprofen ester using the mutant Trichosporon laibacchii strain CBS 5791. A satisfactory result was obtained, in which the E was 82.5, with an ee of 0.94 and a conversion of 0.47 under the optimum hydrolysis conditions [E is enantiomeric ratio, E=ln[1–X(1+ee)]/ln[1–X(1–ee)]; ee is enantiomeric excess, ee=(CSCR)/(CS+CR): temperature of hydrolysis was 23°C]. The medium used in biotransformation was a mixture of growth broth and biotransformation broth at a ratio of 1:9, the concentration of Tween 80 was 15 g/l, the time of hydrolysis, 72 h. These results are promising for further scale-up. Tween 80 significantly improved lipase enantioselectivity and activity at the optimum concentration.  相似文献   

12.
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   

13.
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   

14.
Racemic RS‐4‐(4‐hydroxyphenyl)‐2‐butanol (rhododendrol, RD) was used as a topical skin‐whitening agent until it was recently reported to induce leukoderma. We then showed that oxidation of RD with mushroom tyrosinase rapidly produces RD‐quinone, which is quickly converted to RD‐cyclic quinone and RD‐hydroxy‐p‐quinone. In this study, we examined whether either or both of the enantiomers of RD can be oxidized by human tyrosinase. Using a chiral HPLC column, racemic RD was resolved optically to R(?)‐RD and S(+)‐RD enantiomers. In the presence of a catalytic amount of l ‐dopa, human tyrosinase, which can oxidize l ‐tyrosine but not d ‐tyrosine, was found to oxidize both R(?)‐ and S(+)‐RD to give RD‐catechol and its oxidation products. S(+)‐RD was more effectively oxidized than l ‐tyrosine, while R(?)‐RD was less effective. These results support the notion that the melanocyte toxicity of RD depends on its tyrosinase‐catalyzed conversion to toxic quinones and the concomitant production of reactive oxygen species.  相似文献   

15.
Zhang Y  Shi K  Wen J  Fan G  Chai Y  Hong Z 《Chirality》2012,24(3):239-244
Tetrahydroberberine (THB), a racemic mixture of (+)‐ and (?)‐enantiomer, is a biologically active ingredient isolated from a traditional Chinese herb Rhizoma corydalis (yanhusuo). A chiral high performance liquid chromatography method has been developed for the determination of THB enantiomers in rat plasma. The enantioseparation was carried out on a Chiral®‐AD column using methanol:ethanol (80:20, v/v) as the mobile phase at the flow rate 0.4 ml/min. The ultraviolet detection was set at 230 nm. The calibration curves were linear over the range of 0.01–2.5 μg/ml for (+)‐THB and 0.01‐5.0 μg/ml for (?)‐THB, respectively. The lower limit of quantification was 0.01 μg/ml for both (+)‐THB and (?)‐THB. The stereoselective pharmacokinetics of THB enantiomers in rats was studied after oral and intravenous administration at a dose of 50 and 10 mg/kg racemic THB (rac‐THB). The mean plasma levels of (?)‐THB were higher at almost all time points than those of (+)‐THB. (?)‐THB also exhibited greater Cmax, and AUC0–∞, smaller CL and Vd, than its antipode. The (?)/(+)‐enantiomer ratio of AUC0–∞ after oral and intravenous administration were 2.17 and 1.43, respectively. These results indicated substantial stereoselectivity in the pharmacokinetics of THB enantiomers in rats. Chirality, 2012. © 2012 Wiley Periodicals, Inc.  相似文献   

16.
The purpose of this study was to investigate whether CYP2C19 activity can be estimated from plasma concentrations of lansoprazole enantiomers 4 h (C4h) after single administration by oral and enteral routes. Sixty‐nine subjects, 22 homozygous extensive metabolizers (homEMs), 32 heterozygous EMs (hetEMs), and 15 poor metabolizers (PMs), participated in the study. After a single oral or enteral dose of racemic lansoprazole (30 mg), plasma concentrations of lansoprazole enantiomers were measured 4 h postdose. The R/S ratio of lansoprazole at 4 h differed significantly among the three groups (P < 0.0001) regardless of the administration route. The R/S ratio of lansoprazole in CYP2C19 PMs ranged from 3.0 to 13.7, whereas in homEMs and hetEMs the ratio ranged from 8.6 to 90 and 2.1 to 122, respectively. The relationship between (S)‐lansoprazole concentration and R/S ratio of lansoprazole at C4h is given by the following formula: log10 [R/S ratio] = 2.2 – 0.64 × log10 [C4h of (S)‐lansoprazole] (r = 0.867, P < 0.0001). Thus, phenotyping CYP2C19 using the R/S enantiomer ratio of lansoprazole seems unlikely. However, to obtain a pharmacological effect similar to that in CYP2C19 PMs, we can presume that lansoprazole has a sufficient effect in the patient with an R/S enantiomer ratio at 4 h ≤ 13.70 and (S)‐lansoprazole concentration at 4 h ≥ 50 ng/ml. Chirality 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

17.
Paclobutrazol, with two stereogenic centers, but gives only (2R, 3R) and (2S, 3S)‐enantiomers because of steric‐hindrance effects, is an important plant growth regulator in agriculture and horticulture. Enantioselective degradation of paclobutrazol was investigated in rat liver microsomes in vitro. The degradation kinetics and the enantiomer fraction were determined using a Lux Cellulose‐1 chiral column on a reverse‐phase liquid chromatography–tandem mass spectrometry system. The t1/2 of (2R, 3R)‐paclobutrazol is 18.60 min, while the t1/2 of (2S, 3S)‐paclobutrazol is 10.93 min. Such consequences clearly indicated that the degradation of paclobutrazol in rat liver microsomes was stereoselective and the degradation rate of (2S, 3S)‐paclobutrazol was much faster than (2R, 3R)‐paclobutrazol. In addition, significant differences between the two enantiomers were also observed in enzyme kinetic parameters. The Vmax of (2S, 3S)‐paclobutrazol was more than 2‐fold of (2R, 3R)‐paclobutrazol and the Clint of (2S, 3S)‐paclobutrazol was higher than that of (2R, 3R)‐paclobutrazol after incubation in rat liver microsomes. These results may have potential implications for better environmental and ecological risk assessment for paclobutrazol. Chirality 27:344–348, 2015. © 2015 Wiley Periodicals, Inc.  相似文献   

18.
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   

19.
It was shown that racemic (±)‐ 2 [1′‐benzyl‐3‐(3‐fluoropropyl)‐3H‐spiro[[2]benzofuran‐1,4′‐piperidine], WMS‐1813 ] represents a promising positron emission tomography (PET) tracer for the investigation of centrally located σ1 receptors. To study the pharmacological activity of the enantiomers of 2 , a preparative HPLC separation of (R)‐2 and (S)‐2 was performed. The absolute configuration of the enantiomers was determined by CD‐spectroscopy together with theoretical calculations of the CD‐spectrum of a model compound. In receptor binding studies with the radioligand [3H]‐(+)‐pentazocine, (S)‐2 was thrice more potent than its (R)‐configured enantiomer (R)‐2 . The metabolic degradation of the more potent (S)‐enantiomer was considerably slower than the metabolism of (R)‐2 . The structures of the main metabolites of both enantiomers were elucidated by determination of the exact mass using an Orbitrap‐LC‐MS system. These experiments showed a stereoselective biotransformation of the enantiomers of 2 . Chirality, 2011. © 2010 Wiley‐Liss, Inc.  相似文献   

20.
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   

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