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.
The aim of this study was to examine the effect of clarithromycin, a CYP3A4 inhibitor, on the enantioselective disposition of lansoprazole among three different CYP2C19 genotype groups in healthy Japanese subjects. These subjects included 6 each of homozygous extensive metabolizers (homEMs), heterozygous extensive metabolizers (hetEMs), and poor metabolizers (PMs). In the EMs of CYP2C19, clarithromycin markedly increased Cmax and the AUC0-infinity of (S)-lansoprazole and (S)-hydroxylansoprazole compared with those of the corresponding (R)-enantiomers. Clarithromycin significantly increased Cmax and the AUC0-infinity of (S)-lansoprazole in the homEMs by 110% and 115%, respectively, and in the hetEMs by 105% and 103%, respectively, compared with placebo. Furthermore, clarithromycin slightly prolonged the elimination half-life of (R)-lansoprazole in the homEMs and hetEMs but did not alter that of (S)-lansoprazole. In the of PMs CYP2C19, clarithromycin significantly increased Cmax and the AUC0-infinity and significantly prolonged the elimination half-lives of (R)- and (S)-lansoprazole by 51% and 49%, respectively. The present study suggests that there are significant drug interactions between (R)- or (S)-lansoprazole and clarithromycin in EMs by inhibiting the CYP3A4-catalyzed sulfoxidation primarily during the first pass, whereas in PMs, the overall metabolism of lansoprazole is inhibited.  相似文献   

2.
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   

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.
Debrisoquine (D), an antihypertensive drug metabolized to 4-hydroxydebrisoquine (4-OHD) by CYP2D6, is commonly used as an in vivo probe of CYP2D6 activity and can be used to phenotype individuals as either extensive (EMs) or poor metabolizers (PMs) of such drugs as β-adrenergic blockers, tricyclic antidepressants, and class 1C antiarrhythmics. This report describes reversed-phase HPLC systems by which D and 4-OHD or S-(+) and R-(−)-4-OHD in urine are more selectively quantified without the need for derivatization techniques. We also studied the urinary excretion of R-(−)- and S-(+)-4-hydroxydebrisoquine in EM hypertensive patients in order to determine weather 4-OHD formation exhibits enantioselectivity. Twelve patients with mild to severe essential hypertension were admitted to the study. They received a single tablet of Declinax containing 10 mg debrisoquine sulfate. All the urine excreted during the following 8 h was collected. The debrisoquine metabolic ratio (DMR) was calculated as % of dose excreted as D/% of dose excreted as 4-OHD and the debrisoquine recovery ratio (DRR) was calculated as % of dose excreted as 4-OHD/% of dose excreted as D+4-OHD. Debrisoquine and its metabolite were determined in urine by HPLC using a reversed-phase Select B LiChrospher column, a mobile phase of 0.25 N acetate buffer, pH 5–acetonitrile (9:1, v/v) and a fluorescence detector. The limit of quantitation was determined to be 25.0 ng/ml for D and 18.75 ng/ml for 4-OHD. Intra- and inter-day relative standard deviations (RSDs) were less than 10%. All hypertensive patients studied showed a DMR of less than 12.6 or a DRR higher than 0.12 and were classified as EMs. Direct enantioselective separation on chiral stationary phase involved resolution of S-(+)-4-OHD and R-(−)-4-OHD on a Chiralcel OD-R column with a mobile phase of 0.125 N sodium perchlorate, pH 5–acetonitrile–methanol (85:12:3, v/v/v). The quantitation limit of each enantiomer was 3.75 ng/ml of urine. Intra- and inter-day RSDs were less than 10% for each enantiomer. A high degree of enantioselectivity in the 4-hydroxylation of D favouring the S-(+) enantiomer was observed, resulting in R-(−)-4-OHD not detected in the urine of the EM hypertensive patients studied.  相似文献   

5.
The sex pheromone of the cloaked pug moth, Eupithecia abietaria Götze, an important cone‐feeding pest in spruce seed orchards in Europe, was investigated. Chemical and electrophysiological analyses of pheromone gland extracts of female moths and analogous analyses of synthetic hydrocarbons and epoxides of chain length C19 and C21 revealed (3Z,6Z,9Z)‐3,6,9‐nonadecatriene (3Z,6Z,9Z‐19:H) and 3Z,6Zcis‐9,10‐epoxynonadecadiene (3Z,6Zcis‐9,10‐epoxy‐19:H) as candidate pheromone components, which were found in a gland extract in a ratio of 95 : 5. In field trapping experiments, conspecific males were only attracted to a combination of 3Z,6Z,9Z‐19:H and the (9S,10R)‐enantiomer of 3Z,6Zcis‐9,10‐epoxy‐19:H. The (9R,10S)‐enantiomer was not attractive, which is in agreement with studies on other Eupithecia species, for which males have only been attracted by the (9S,10R)‐enantiomer of epoxides. Subsequent experiments showed that E. abietaria males were attracted to a wide range of ratios of the two active compounds and that trap catches increased with increasing dose of the binary blend. A two‐component bait containing 300 μg 3Z,6Z,9Z‐19:H and 33 μg of the (9S,10R)‐enantiomer of 3Z,6Zcis‐9,10‐epoxy‐19:H was efficient for monitoring E. abietaria in spruce seed orchards in southern Sweden, where this species has probably been overlooked as an important pest in the past. With sex pheromones recently identified for two other moths that are major pests on spruce cones, the spruce seed moth, Cydia strobilella L., and the spruce coneworm, Dioryctria abietella Denis & Schiffermüller, pheromone‐based monitoring can now be achieved for the whole guild of cone‐feeding moths in European spruce seed orchards.  相似文献   

6.
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   

7.
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   

8.

Background

There are inconsistent conclusions about whether CYP2C19 variants could affect H. pylori eradication rate in patients treated with the proton pump inhibitor (PPI)-based therapy. We therefore performed a meta-analysis of randomized clinical trials (RCTs) to re-evaluate the impact of CYP2C19 variants on PPI-based triple therapy for the above indication.

Methods

All relevant RCTs in the PubMed, Cochrane Library, EMBASE, Web of Science and two Chinese databases (up to February 2013) were systematically searched, and a pooled analysis was performed with the odds ratio (OR) and 95% confidence interval (CI) by the STATA software.

Results

Sixteen RCT datasets derived from 3680 patients were included. There was no significant heterogeneity across the data available in this meta-analysis. There were significant differences in that rate between homozygous (HomEMs) and heterozygous (HetEMs) extensive metabolizers (OR 0.724; 95% CI 0.594–0.881), between HomEMs and poor metabolizers (PM) (OR 0.507; 95%CI 0.379–0.679), or between HetEMs and PMs (OR 0.688; 95%CI 0.515–0.920), regardless of the PPI being taken. Furthermore, sub-analysis of individual PPIs was carried out to explore the difference across all the PPIs used. A significantly low rate was seen in HomEMs vs. HetEMs taking either omeprazole (OR 0.329; 95%CI 0.195–0.553) or lansoprazole (OR 0.692; 95%CI 0.485–0.988), and also in HomEMs vs. PMs for omeprazole (OR 0.232; 95%CI 0.105–0.515) or lansoprazole (OR 0.441; 95%CI 0.252–0.771). However, there was no significant difference between HetEMs and PMs taking either one. No significant differences were observed for rabeprazole or esomeprazole across the CYP2C19 genotypes of interest.

Conclusions

Carriage of CYP2C19 loss-of-function variants is associated with increased H. pylori eradication rate in patients taking PPI-based triple therapies when omeprazole or lansoprazole is chosen. However, there is no a class effect after use of rabeprazole or esomeprazole.  相似文献   

9.

Background  

Omeprazole is metabolized by the hepatic cytochrome P450 (CYP) 2C19 enzyme to 5-hydroxyomeprazole. CYP2C19 exhibits genetic polymorphisms responsible for the presence of poor metabolizers (PMs), intermediate metabolizers (IMs) and extensive metabolizers (EMs). The defective mutations of the enzyme and their frequencies change between different ethnic groups; however, the polymorphism of the CYP2C19 gene has not been studied in Colombian mestizos. The aim of this study was to evaluate the genotype and phenotype status of CYP2C19 in Colombian mestizos, in order to contribute to the use of appropriate strategies of drug therapy for this population.  相似文献   

10.
Interactions between naringenin and the cytochrome P450 (CYP) system have been of interest since the first demonstration that grapefruit juice reduced CYP3A activity. The effects of naringenin on other CYP isoforms have been less investigated. In addition, it is well known that interactions with enzymes are often stereospecific, but due to the lack of readily available pure naringenin enantiomers, the enantioselectivity of its effects has not been characterized. We isolated pure naringenin enantiomers by chiral high‐performance liquid chromatography and tested the ability of (R)‐,(S)‐ and rac‐naringenin to inhibit several important drug‐metabolizing CYP isoforms using recombinant enzymes and pooled human liver microsomes. Naringenin was able to inhibit CYP19, CYP2C9, and CYP2C19 with IC50 values below 5 μM. No appreciable inhibition of CYP2B6 or CYP2D6 was observed at concentrations up to 10 μM. Whereas (S)‐naringenin was 2‐fold more potent as an inhibitor of CYP19 and CYP2C19 than (R)‐naringenin, (R)‐naringenin was 2‐fold more potent for CYP2C9 and CYP3A. Chiral flavanones like naringenin are difficult to separate into their enantiomeric forms, but enantioselective effects may be observed that ultimately impact clinical effects. Inhibition of specific drug metabolizing enzymes by naringenin observed in vitro may be exploited to understand pharmacokinetic changes seen in vivo. Chirality, 2011. © 2011 Wiley‐Liss, Inc.  相似文献   

11.
The application of several immobilized lipases has been explored in the enantioselective esterification of (R,S)‐2‐methylbutyric acid, an insect pheromone precursor. With the use of Candida antarctica B, using hexane as solvent, (R)‐pentyl 2‐methylbutyrate was prepared in 2 h with c 40%, eep 90%, and E = 35, while Thermomyces lanuginosus leads to c 18%, eep 91%, and E = 26. The (S)‐enantiomer was obtained by the use of Candida rugosa or Rhizopus oryzae (2‐h reaction, c 34% and 35%, eep 75 and 49%, and E = 10 and 4, respectively). Under optimal conditions, the effect of the solvent, the molar ratio, and the nucleophile were evaluated.  相似文献   

12.
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   

13.
Golo Storch  Oliver Trapp 《Chirality》2018,30(10):1150-1160
We present rhodium catalysts that contain stereodynamic axially chiral biphenol‐derived phosphinite ligands modified with non‐stereoselective amides for non‐covalent interactions. A chirality transfer was achieved with (R)‐ or (S)‐acetylphenylalanine methyl amide, and the interaction mechanism was investigated by NMR measurements. These interactions at the non‐stereoselective interaction sites and the formation of supramolecular complexes result in an enrichment of either the (Rax)‐ or (Sax) enantiomer of the tropos catalysts, which in turn provide the (R)‐ or (S)‐acetylphenylalanine methyl ester in the hydrogenation of (Z)‐methyl‐α‐acetamidocinnamate.  相似文献   

14.
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   

15.
Prolonged swimming performances of two as yet unnamed species of three‐spined stickleback, Gasterosteus spp., were compared. The two fishes (not yet formally described, referred to here as benthic and limnetic) inhabit different niches within Paxton Lake, Texada Island, British Columbia, Canada, and are recent, morphologically distinct species. Limnetics had longer endurance during prolonged swimming than did benthics. The mean regression of the log10 of fatigue time (Ft, s) on swimming speed (U, standard length, LS s?1) for limnetics (log10Ft = 7·03 ? 0·46U) had a similar slope, but a significantly higher intercept than that for benthics (log10Ft = 5·55 ? 0·43U). Adult benthics were larger, heavier and deeper‐bodied fish than limnetics. Limnetics, however, had a significantly greater pectoral fin edge:base ratio (mean ± s .e .: limnetics, 4·58 ± 0·43; benthics, 3·63 ± 0·27). In addition, limnetics had significantly lower drag coefficients (CD) than benthics (limnetics, log10CD = ?0·49log10Re + 0·66; benthics, log10CD = ?0·26log10Re ? 0·30) where Re is the Reynolds number [(LSU?1), where U and ν are swimming velocity and the kinematic viscosity of the water, respectively]. Compared to their ancestral form, the anadromous three‐spined stickleback Gasterosteus aculeatus L., limnetics and benthics had significantly longer and shorter endurance times, respectively. In addition, both these fishes had significantly higher fast‐start velocities than their ancestral form. Selection due to differential resource use may have lead to divergence of body form, and, therefore, of steady swimming performance. Therefore predation may be the selective force for the similar high escape performance in these two fishes.  相似文献   

16.
The pyrrolidine side chain makes proline play a unique role in protein structure and function. The Cγ ring pucker preference and the cis trans peptidyl bond ratio can be mediated via stereoelectronic effects. Here we used a compact triple‐stranded antiparallel β‐sheet protein, the human Pin1 WW domain, to study the consequences of implanting a preorganized Cγ ring pucker on protein structure and function. The conserved Pro37 is a key residue involved in one hydrophobic core, plays an important role in the WW domain, and adopts a Cγendo ring pucker in the native structure. Pro37 was replaced with Cγexo biased pucker derivatives: (2S,4R)‐4‐hydroxyproline (4R‐Hyp), (2S,4R)‐4‐fluoroproline (4R‐Flp), (2S,4R)‐4‐methoxyproline (4R‐Mop), and Cγendo biased pucker derivatives: (2S,4S)‐4‐hydroxyproline (4S‐hyp), (2S,4S)‐4‐fluoroproline (4S‐flp), (2S,4S)‐4‐methoxyproline (4S‐mop) to examine how a preorganized pucker affects the folding stability and ligand‐binding affinity. Circular dichroism measurements indicate that among the variants, only the one with 4S‐flp substitution (P37flp) is more stable than the wild type, suggesting that the stabilization effects originated from preorganization of the backbone conformation and the hydrophobicity of C? F group. Analysis of ligand‐binding affinity using isothermal titration calorimetry revealed that only P37flp has a stronger ligand affinity than the wild type, showing that 4S‐flp can stabilize the WW domain and increase its ligand affinity. Together we have used 4‐substituted proline derivatives and the WW domain to demonstrate that proline ring puckering can be a key factor in determining the folding stability of a protein but the choice of the derivative groups is also critical. Proteins 2014; 82:67–76. © 2013 Wiley Periodicals, Inc.  相似文献   

17.
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   

18.
Two new chiral mononuclear Mn(III) complexes, [Mn L ( R )Cl (C2H5OH)]?C2H5OH ( 1 ) and [Mn L ( S ) (CH3OH)2]Cl?CH3OH ( 2 ), {H2 L = (R,R)‐or (S,S)‐N,N’‐bis‐(2‐hydroxy‐1‐naphthalidehydene)‐cyclohexanediamine} were synthesized and characterized by various physicochemical techniques. Bond valence sum (BVS) calculations and the Jahn‐Teller effect indicate that the Mn centers are in a +3 oxidation state. The statuses of the two complexes in the solution were confirmed as a pair of enantiomers by electrospray ionization, mass spectrometry (ESI‐MS) spectrum. The binding ability of the complexes with calf thymus CT‐DNA was investigated by spectroscopic and viscosity measurements. Both of the complexes could interact with CT‐DNA via an intercalative mode with the order of 1 ( R ‐enantiomer) > 2 ( S ‐enantiomer). Under the physiological conditions, the two compounds exhibit efficient DNA cleavage activities without any external agent, which also follows the order of R ‐enantiomer > S ‐enantiomer. Interestingly, the concentration‐dependent DNA cleavage experiments indicate an optimal concentration of 17.5 μM. In addition, the interaction of the compounds with bovine serum albumin (BSA) was also investigated, which indicated that the complexes could quench the intrinsic fluorescence of BSA by a static quenching mechanism. Chirality 27:142‐150, 2015. © 2014 Wiley Periodicals, Inc.  相似文献   

19.
Introduction: This study determined the pharmacokinetics and pharmacodynamics of (R)‐ and (S)‐ketamine and (R)‐ and (S)‐norketamine following a 5‐day moderate dose, as a continuous (R,S)‐ketamine infusion in complex regional pain syndrome (CRPS) patients. Materials and methods: Ketamine was titrated to 10–40 mg/h and maintained for 5 days. (R)‐ and (S)‐Ketamine and (R)‐ and (S)‐norketamine pharmacokinetic and pharmacodynamic studies were performed. Blood samples were obtained on Day 1 preinfusion, and at 60–90, 120–150, 180–210, and 240–300 min after the start of the infusion, on Days 2, 3, 4, 5, and on Day 5 at 60 min after the end of infusion. The plasma concentrations of (R)‐ and (S)‐ketamine and (R)‐ and (S)‐norketamine were determined using enantioselective liquid chromatography–mass spectrometry. Results: Ketamine and norketamine levels stabilized 5 h after the start of the infusion. (R)‐Ketamine clearance was significantly lower resulting in higher steady‐state plasma concentrations than (S)‐ketamine. The first‐order elimination for (S)‐norketamine was significantly greater than that of (R)‐enantiomer. When comparing the pharmacokinetic parameters of the patients who responded to ketamine treatment with those who did not, no differences were observed in ketamine clearance and the first‐order elimination of norketamine. Conclusion: The results indicate that (R)‐ and (S)‐ketamine and (R)‐ and (S)‐norketamine plasma concentrations do not explain the antinociceptive activity of the drug in patients suffering from CRPS. Chirality, 2011. © 2010 Wiley‐Liss, Inc.  相似文献   

20.
After cultivation on (R,S)‐2‐(2,4‐dichlorophenoxy)propionate, two α‐ketoglutarate‐dependent dioxygenases were isolated and purified from Delftia acidovorans MC1, catalysing the cleavage of the ether bond of various phenoxyalkanoate herbicides. One of these enzymes showed high specificity for the cleavage of the R‐enantiomer of substituted phenoxypropionate derivatives: the Km values were 55 μM and 30 μM, the kcat values 55 min–1 and 34 min–1 with (R)‐2‐(2,4‐dichlorophenoxy)propionate [(R)‐2,4‐DP] and (R)‐2‐(4‐chloro‐2‐methylphenoxy)propionate, respectively. The other enzyme predominantly utilised the S‐enantiomers with Km values of 49 μM and 22 μM, and kcat values of 50 min–1 and 46 min–1 with (S)‐2‐(2,4‐dichlorophenoxy)propionate [(S)‐2,4‐DP] and (S)‐2‐(4‐chloro‐2‐methylphenoxy)propionate, respectively. In addition, it cleaved phenoxyacetate herbicides (i.e. 2,4‐dichlorophenoxyacetate: Km = 123 μM, kcat = 36 min–1) with significant activity. As the second substrate, only α‐ketoglutarate served as an oxygen acceptor for both enzymes. The enzymes were characterised by excess substrate inhibition kinetics with apparent Ki values of 3 mM with (R)‐2,4‐DP and 1.5 mM with (S)‐2,4‐DP. The reaction was strictly dependent on the presence of Fe2+ and ascorbate; other divalent cations showed inhibitory effects to different extents. Activity was completely extinguished within 2 min in the presence of 100 μM diethylpyrocarbonate (DEPC).  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