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 609 毫秒
1.
D J Porter  E Abushanab 《Biochemistry》1992,31(35):8216-8220
The enantiomers of erythro-9-(2-hydroxy-3-nonyl)adenine [(+)- and (-)-EHNA) bound to adenosine deaminase (ADA) at pH 7 with concomitant changes in the optical properties of the enzyme. The association rate constant for (+)-EHNA was 2.9 x 10(6) M-1 s-1 and that for (-)-EHNA was 6.4 x 10(6) M-1 s-1. The dissociation of (-)-EHNA.ADA or (+)-EHNA.ADA in the presence of excess coformycin was monitored by the quenching of enzyme fluorescence as coformycin.ADA was formed. The dissociation rate constants of (+)- and (-)-EHNA.ADA were 0.0054 s-1 and 2.7 s-1, respectively. A similar value for the dissociation rate constant (0.005 s-1) for (+)-EHNA.ADA was calculated from the time course for the appearance of catalytic activity after dilution of (+)-EHNA.ADA into 100 microM adenosine. The Ki values of ADA for (+)- and (-)-EHNA were similar to the dissociation constants calculated from the ratio of the respective dissociation and association rate constants. The biphasic time-dependent inhibition of the catalytic activity of ADA by (+/- )-EHNA [Frieden, C., Kurz, L. C., & Gilbert, H. R. (1980) Biochemistry 19, 5303-5309] was confirmed. However, the catalytic activity of ADA was inhibited monophasically by (+)-EHNA. Thus, the biphasic nature of the time course for inhibition of ADA by (+/- )-EHNA was the result of the presence of both enantiomers of the inhibitor in this assay. These kinetic data were interpreted in terms of single-step mechanisms for binding of (+)- and (-)-EHNA.  相似文献   

2.
The (+)- and (?)-enantiomers of 3-isopropyl 5-(4-methylphenethyl) 1,4-dihydro-2,6-dimethyl-4-(2-pyridyl)-3,5-pyridinedicarboxylate were synthesized using an efficient highly enantioselective (ee ≥ 96%) variant of the Hantzsch dihydropyridine synthesis. The key step in this procedure involved the asymmetric Michael addition of a metalated chiral aminocrotonate, derived from D -valine or L -valine, respectively, to the Knoevenagel acceptor (Z)-2-isopropoxycarbonyl-1-(2-pyridyl)-but-1-en-3-one. Both enantiomers exhibited a dual cardioselective partial calcium channel agonist (positive inotropic)/smooth muscle selective calcium channel antagonist effect. The relative in vitro smooth muscle calcium channel antagonist activities of the (?):(+) enantiomers was 26:1. In contrast, the (+)-enantiomer exhibited a greater in vitro positive inotropic effect on guinea pig left atrium where the contractile force was maximally increased by 14.8% at a concentration of 1.63 × 10?8 M. © 1994 Wiley-Liss, Inc.  相似文献   

3.
The enantiomers of 1-methyl-3-(10H-phenothiazine-10-ylmethyl)-1-azoniabicyclo[2,2,2]octane iodide ( 1 ) were prepared by chiral chromatographic resolution of the precursor mequitazine ( 2 ). The (+)-(S)-enantiomer 1b is 10-fold more potent than (?)-(R)-enantiomer 1a as a histamine antagonist, while the two enantiomers show the same antimuscarinic activity in vitro. The absolute configuration of the more active dextrorotatory isomer has been determined by X-ray analysis. Conformational analysis and molecular modeling suggest that the (+)-(S)-enantiomer can adopt a conformation similar to that attributed to the receptor binding conformers of classical antihistamines. © 1994 Wiley-Liss, Inc.  相似文献   

4.
The purposes of this work were (1) to develop a high performance liquid chromatographic (HPLC) assay for the enantiomers of thalidomide in blood, (2) to study their inversion and degradation in human blood, and (3) to study the pharmacokinetics of (+)-(R)- and (?)-(S)-thalidomide after oral administration of the separate enantiomers or of the racemate to healthy male volunteers. The enantiomers of thalidomide were determined by direct resolution on a tribenzoyl cellulose column. Mean rate constants of chiral inversion of (+)-(R)-thalidomide and (?)-(S)-thalidomide in blood at 37°C were 0.30 and 0.31 h?1, respectively. Rate constants of degradation were 0.17 and 0.18 h?1. There was rapid interconversion in vivo in humans, the (+)-(R)-enantiomer predominating at equilibrium. The pharmacokinetics of (+)-(R)- and (?)-(S)-thalidomide could be characterized by means of two one-compartment models connected by rate constants for chiral inversion. Mean rate constants for in vivo inversion were 0.17 h?1 (R to S) and 0.12 h?1 (S to R) and for elimination 0.079 h?1 (R) and 0.24 h?1 (S), i.e., a considerably faster rate of elimination of the (?)-(S)-enantiomer. Putative differences in therapeutic or adverse effects between (+)-(R)- and (?)-(S)-thalidomide would to a large extent be abolished by rapid interconversion in vivo. © 1995 Wiley-Liss, Inc.  相似文献   

5.
For the first time, a method for enantiomer resolution of the anticonvulsant Galodif (1‐((3‐chlorophenyl)(phenyl)methyl) urea) by chiral HPLC was developed, whereas the enantiomeric composition of 1‐((3‐chlorophenyl)(phenyl)methyl) amine—precursor in Galodif synthesis—cannot be resolved by this method. However, starting 1‐((3‐chlorophenyl)(phenyl)methyl) amine quantitatively forms diastereomeric N‐((3‐chlorophenyl)(phenyl)methyl)‐1‐camphorsulfonamides in reaction with chiral (1R)‐(+)‐ or (1S)‐(?)‐camphor‐10‐sulfonyl chlorides. The diastereomeric ratio of obtained camphorsulfonamides can be easily determined by NMR 1H and 13C spectroscopy. The DFT calculations of specific rotation of Galodif enantiomers showed good agreement with experimental data. The absolute configuration of enantiomers was proposed for the first time.  相似文献   

6.
A retention model for the chiral separation of an uncharged solute, felodipine, on CHIRAL-AGP, using a micellar mobile phase is proposed. The model assumes the presence of two stereoselective sites and each enantiomer was found to interact with different sites. Addition of a chiral aliphatic alcohol, (+)-(S)-2-octanol, preferentially interacted with the binding site for (?)-(S)-felodipine. The monomeric form of the micellar agent (Tween® 20) competed with the enantiomers for the adsorption sites, and the formation of a 1:1 complex between the enantiomers and the micelles was assumed. The retention of the solutes was effectively controlled by adding small quantities (<1.63 × 10?3 M) of the nonionic detergent Tween 20 to the mobile phase. Baseline separation was achieved by addition of 1.0 mM n-octylamine to the mobile phase; 8.14 × 10?4 M Tween 20 in phosphate buffer pH 7.0. The separation factor (α = 1.74) was unaffected by the detergent concentration in the presence of 1.0 mM n-octylamine. © 1995 Wiley-Liss, Inc.  相似文献   

7.
1. The metabolism in vitro and microsomal interactions of (+)-amphetamine, (-)-amphetamine, (+)-benzphetamine and (-)-benzphetamine were studied with hepatic microsomes from phenobarbitone-pretreated male rabbits. 2. (+)-Benzphetamine was N-demethylated 30-35% faster than (-)-benzphetamine, but the apparent Michaelis constants for the two enantiomers were similar. 3. (-)-Amphetamine was deaminated about 200% faster than (+)-amphetamine. 4. The benzphetamine enantiomers gave qualitatively and quantitatively identical type I microsomal difference spectra (peak, 390nm; trough, 425nm) indicating identical apparent binding affinities for microsomes and identical spectral changes at maxima (DeltaE(max.) values). 5. The amphetamine enantiomers gave qualitatively identical type II microsomal difference spectra (peak, 433nm; trough, 395nm). However, the type II spectral data indicated that (+)-amphetamine had a markedly higher apparent binding affinity than (-)-amphetamine for microsomes. The amphetamine enantiomers gave identical DeltaE(max.) values. 6. The benzphetamine enantiomers (0.5mm) enhanced the rate of microsomal cytochrome P-450 reduction by NADPH by 400-500%, (+)-benzphetamine enhancing the rate 20-25% more than (-)-benzphetamine. 7. The amphetamine enantiomers decreased the rate of microsomal cytochrome P-450 reduction by NADPH. At a concentration of 2mm, (+)-amphetamine decreased the rate more than (-)-amphetamine. 7. All four enantiomers enhanced microsomal NADPH oxidation.  相似文献   

8.
A set of polymers was imprinted with (-)-ephedrine using UV initiation, under the influence of a constant external magnetic field with intensities ranging from 0 to 1.55 T. Synthesised materials were characterised by X-ray crystallography, infrared spectroscopy, swelling and surface area. Recognition properties were assessed by the ability to discriminate between (+) and (-)-ephedrine and by Scatchard analyses on chromatographic mode. It was shown that polymer morphology and recognition properties are affected by the magnetic field. This resulted in considerable improvements in the chromatographic resolution of ephedrine enantiomers by materials synthesised under the influence of magnetic field. Apparently the magnetic field improved the ordering of the polymer structure and facilitated the formation of more uniform imprinting sites.  相似文献   

9.
Stereoselective drug-channel interactions may help to elucidate the molecular basis of voltage-gated potassium channel block by local anesthetic drugs. We studied the effects of the enantiomers of bupivacaine on a cloned human cardiac potassium channel (hKv1.5). This channel was stably expressed in a mouse Ltk- cell line and studied using the whole-cell configuration of the patch-clamp technique. Both enantiomers modified the time course of this delayed rectifier current. Exposure to 20 microM of either S(-)-bupivacaine or R(+)-bupivacaine did not modify the activation time constant of the current, but reduced the peak outward current and induced a subsequent exponential decline of current with time constants of 18.7 +/- 1.1 and 10.0 +/- 0.9 ms, respectively. Steady-state levels of block (assessed with 250-ms depolarizing pulses to +60 mV) averaged 30.8 +/- 2.5% (n = 6) and 79.5 +/- 3.2% (n = 6) (p < 0.001), for S(-)- and R(+)-bupivacaine, respectively. The concentration dependence of hKv1.5 inhibition revealed apparent KD values of 27.3 +/- 2.8 and 4.1 +/- 0.7 microM for S(-)-bupivacaine and R(+)-bupivacaine, respectively, with Hill coefficients close to unity, suggesting that binding of one enantiomer molecule per channel was sufficient to block potassium permeation. Analysis of the rate constants of association (k) and dissociation (l) yielded similar values for l (24.9 s-1 vs. 23.6 s-1 for S(-)- and R(+)-bupivacaine, respectively) but different association rate constants (1.0 x 10(6) vs. 4.7 x 10(6) M-1 s-1 for S(-)- and R(+)-bupivacaine, respectively). Block induced by either enantiomer displayed a shallow voltage dependence in the voltage range positive to 0 mV, i.e., where the channel is fully open, consistent with an equivalent electrical distance delta of 0.16 +/- 0.01. This suggested that at the binding site, both enantiomers of bupivacaine experienced 16% of the applied transmembrane electrical field, referenced to the inner surface. Both bupivacaine enantiomers reduced the tail current amplitude recorded on return to -40 mV and slowed their time course relative to control, resulting in a "crossover" phenomenon.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

10.
Y Xu  P Tang  L Firestone    T T Zhang 《Biophysical journal》1996,70(1):532-538
Whether proteins or lipids are the primary target sites for general anesthetic action has engendered considerable debate. Recent in vivo studies have shown that the S(+) and R(-) enantiomers of isoflurane are not equipotent, implying involvement of proteins. Bovine serum albumin (BSA), a soluble protein devoid of lipid, contains specific binding sites for isoflurane and other anesthetics. We therefore conducted 19F nuclear magnetic resonance measurements to determine whether binding of isoflurane to BSA was stereoselective. Isoflurane chemical shifts were measured as a function of BSA concentration to determine the chemical shift differences between the free and bound isoflurane. KD was determined by measuring the 19F transverse relaxation times (T2) as a function of isoflurane concentration. The binding duration was determined by assessing increases in 1/T2 as a result of isoflurane exchanging between the free and bound states. The S(+) and R(-) enantiomers exhibited no stereoselectivity in chemical shifts and KD values (KD = 1.3 +/- 0.2 mM, mean +/- SE, for S(+), R(-), and the racemic mixture). Nonetheless, stereoselectivity was observed in dynamic binding parameters; the S(+) enantiomer bound with slower association and dissociation rates than the R(-).  相似文献   

11.
(3SR,4RS)-3,4-Epoxypiperidine-4-carboxylic acid (isoguvacine oxide) is a potent and specific GABAA receptor agonist. Isoguvacine oxide, originally designed as a potentially alkylating agonist, turned out to interact with the GABAA receptor in a fully reversible manner. The protected form of isoguvacine oxide, benzyl (3SR,4RS)-1-(benzyloxycarbonyl)-3,4-epoxypiperidine-4-carboxylate ( 1 ) (Scheme 1), has now been resolved by chiral chromatography using cellulose triacetate as a chiral stationary phase. The enantiomers of 1 (ee ≥ 98.8%) were subsequently deprotected by hydrogenolysis. Whereas both enantiomers of isoguvacine oxide were inactive as inhibitors of the binding of [3H]GABA to GABAB receptor sites (IC50 > 100 μM), (+)-isoguvacine oxide (IC50 = 0.20 ± 0.03 μM) and (?)-isoguvacine oxide (IC50 = 0.32 ± 0.05 μM) showed comparable potencies as inhibitors of the binding of [3H]GABA to GABAA receptor sites. Furthermore, (+)-isoguvacine oxide (EC50 = 6 μM; 33% relative efficacy) and (?)-isoguvacine oxide (EC50 = 5 μM; 38% efficacy relative to 10 μM muscimol) were approximately equipotent and equiefficacious as stimulators of the binding of [3H]diazepam to the GABAA receptor-associated benzodiazepine site. This latter effect is an in vitro estimate of GABAA agonist efficacy. These pharmacological data for isoguvacine oxide and its enantiomers do not seem to support our earlier conception of the topography of the GABAA recognition site(s), derived from extensive structure—activity studies on GABAA agonists. Thus, the model of the GABAA recognition site(s) comprising a narrow cleft or pocket, in which the anionic moiety of the zwitterionic GABAA agonists is assumed to be embedded during receptor activation, may have to be revised. © 1995 Wiley-Liss, Inc.  相似文献   

12.
The pharmacological activity of rac-ketoprofen and its enantiomers was investigated in vitro using different cellular models. The effect of these compounds on arachidonic acid metabolism was assessed by measuring the inhibition of prostanoid generation under the action of several agonists. Thus, we have evaluated the inhibition of (1) thromboxane B2 synthesis in rabbit platelets and human polymorphonuclear leukocytes (PMNs), (2) prostaglandin E2 synthesis in three cultured cells, namely human umbilical vein endothelial cells (HUVEC), human keratinocytes, and mouse macrophage-like P388D1 cells. The IC50 values found for (+)-(S)-ketoprofen were in the range between 0.1 nM and 0.8 μM, being slightly lower in all models than those found for rac-ketoprofen (0.4 nM–3 μM). On the other hand, (?)-(R)-ketoprofen showed inhibition of cyclooxygenase only at concentrations two or three orders of magnitude higher than those required for the (+)-(S) enantiomer. These results, obtained with cell types of relevance for inflammatory processes and with compounds of high optical purity, demonstrate that the prostanoid biosynthesis inhibition caused by the drug rac-ketoprofen is exclusively due to its dextrorotatory enantiomer. © 1993 Wiley-Liss, Inc.  相似文献   

13.
The in vitro aromatase activity in microsomal fractions from rat ovary and its inhibition by enantiomers of aminoglutethimide (AG), rogletimide (RG), and cyclohexylaminoglutethimide (ChAG) were studied by analysing the [3H]H2O released when [1β-3H]androstenedione was converted to estrone. Maximum velocity (Vmax) and the Michaelis-Menten constant (Km) of the microsomal aromatase enzyme were 17.40 ± 0.45 pmol/ml/mg protein/min and 1.02 ± 0.06 μM, respectively. The IC50s for the enantiomers were similar for (+)-R-AG and (?)-R-ChAG (0.86 ± 0.06 and 0.89 ± 0.15 μM, respectively). (+)S-ChA'G was most potent with IC50 of 0.075 ± 0.003 μM. The IC50s for (?)-S-AG, (+)-R-RG, and (?)-S-RG were in the same range (23.15 ± 2.74, 24.58 ± 2.46, and 24.43 ± 2.20 μM, respectively). © 1994 Wiley-Liss, Inc.  相似文献   

14.
The rice koji for sake making usually shows white, but sometimes, during storage in the air it turns brown. This turn is accelerated by some wet conditions.

Three kinds of precursor of the pigment in browned rice-koji were isolated in the present work by column chromatography. A main precursor was adsorbed on alumina at pH 8.5 and eluted with 0.3 n acetic acid. The eluate was then treated by Dowex 50-X8 (H) column, eluted with water and 1 n HCl, and then identified as l-dopa after derived to 3,4-di-O-benzoyl-N-benzoyl-l-dopa ethyl ester. The second precursor was isolated from ethylether soluble fraction of the water eluate described above and identified as protocatechuic acid by some chemical properties. The third precursor was not isolated owing to its chemical unstableness, though it was clearly different from these two precursors.

Isolation of l-dopa from the rice koji made by culture of Aspergillus and also from sake and sake cake was first carried out in the present work.  相似文献   

15.
A liquid chromatographic method for the quantitative analysis of S-(+)- and R-(−)-citalopram in human plasma has been developed and validated. The enantiomers of citalopram and the internal standard, R-(+)-propranolol, were extracted from alkaline plasma with 2% n-butanol in n-hexane. After a clean-up step, the organic phase was evaporated and the residues dissolved in 50–100 μl of 0.001 M HCl. The separation was performed on a Chiral-AGP column with 3.0 mM N-dodecyl-N,N-dimethylammonio-3-propanesulfonate and 10 mM hexanoic acid in phosphate buffer pH 6.5 as the mobile phase. The limit of detection was estimated to be 1 ng/ml (S/N≈3) for each enantiomer monitoring UV absorption at 240 nm. In the range studied, 2.31–191 ng/ml, the recoveries were quantitative and the coefficients of variations were between 2.47% and 11.5%.  相似文献   

16.
The relative configuration of the enantiomers of thioridazine was defined to explore the stereochemistry associated with the selective binding of (?)-thioridazine to dopamine D-1 receptors and (+)-thioridazine to D-2 receptors. Using a seven-step stereoconservative synthesis, (?)-(S)-2-piperidinecarboxylic acid was converted to (?)-(S*)-2-(2-chloroethyl)-1-methylpiperidine, a literature (?)-thioridazine synthetic precursor. Accordingly, (?)- and (+)-thioridazine are the (S)- and (R)-enantiomers, respectively.  相似文献   

17.
Myocardial uptake of thiopental enantiomers by an isolated perfused rat heart preparation was examined after perfusion with protein-free perfusate. Outflow perfusate samples were collected at frequent intervals for 20 min during single-pass perfusion with 10 μg/ml racemic thiopental (washin phase) and for another 45 min during perfusion with drug-free perfusate (washout phase). (+)- and (−)-thiopental concentrations were assayed by chiral high-performance liquid chromatography. Heart rate, perfusion pressure, and electrocardiogram were also monitored. During the washin phase, there was no significant difference between the mean values of the equilibration rate constants of (+)- and (−)-thiopental enantiomers (0.44 ± 0.07 min−1 and 0.43 ± 0.09 min−1, respectively, P > 0.05). Mean volumes of distribution of (+)- and (−)-thiopental enantiomers were similar (6.34 ± 1.20 and 6.45 ± 1.29 ml/g for the washin phase and 7.22 ± 0.71 and 7.47 ± 0.81 ml/g for the washout phase, respectively, P > 0.05). This indicates that tissue accumulation of thiopental enantiomers in the isolated perfused rat heart was not stereoselective. Uptake of thiopental by the heart was perfusion flow rate-limited and independent of capillary permeability. These findings suggest that myocardial tissue concentration of racemic thiopental should be an accurate predictor of myocardial drug effect. © 1996 Wiley-Liss, Inc.  相似文献   

18.
The enantiomers of the aromatase inhibitors 3-(4-aminophenyl)-pyrrolidine-2,5-dione (WSP-3, II ), its N-pentyl derivative ( III ), and the antifungal econazole ( IV ), all possessing a benzylic proton at the chiral centre, are rapidly racemised in vitro in phosphate buffer (0.01 M) at pH 7.4 and 23°C with t½ values of 7, 6, and 5 h respectively. In vivo studies in rats show that (+)-econazole is racemised after intraperitoneal injection with t½ = 1.24h. The enantiomers of the antifungal 1-[(benzofuran-2-yl)-4-chlorophenylmethyl] imidazole ( V ) were stable at pH 7.4, attributable to steric hindrance to carbanion formation in the racemisation step. © 1994 Wiley-Liss, Inc.  相似文献   

19.
The enantiomeric separation of eight pesticides including bitertanol ( 1 ), diclobutrazol ( 2 ), fenbuconazole ( 3 ), triticonazole ( 4 ), imazalil ( 5 ), triapenthenol ( 6 ), ancymidol ( 7 ), and carfentrazone‐ethyl ( 8 ) was achieved, using normal‐phase high‐performance liquid chromatography on two cellulosed‐based chiral columns. The effects of isopropanol composition from 2% to 30% in the mobile phase and column temperature from 5 to 40 °C were investigated. Satisfactory resolutions were obtained for bitertanol ( 1 ), triticonazole ( 4 ), imazalil ( 5 ) with the (+)‐enantiomer eluted first and fenbuconazole ( 3 ) with the (—)‐enantiomer eluted first on Lux Cellulose‐2 and Lux Cellulose‐3. (+)‐Enantiomers of diclobutrazol ( 2 ) and triapenthenol ( 6 ) were first eluted on Lux Cellulose‐2. (—)‐Carfentrazone‐ethyl ( 8 ) were eluted first on Lux Cellulose‐2 and Lux Cellulose‐3 with incomplete separation. Reversed elution orders were obtained for ancymidol (7). (+)‐Ancymidol was first eluted on Lux Cellulose‐2 while on Lux Cellulose‐3 (—)‐ancymidol was first eluted. The results of the elution order at different column temperatures suggested that column temperature did not affect the optical signals of the enantiomers. These results will be helpful to prepare and analyze individual enantiomers of chiral pesticides. Chirality 27:32–38, 2015. © 2014 Wiley Periodicals, Inc.  相似文献   

20.
Competition between the (+)- and (?) enantiomers of 2,2,2-trifluoro-1-(9-anthryl) ethanol as mobile phase additives was indicated by the chromatographic behavior of their system peaks. Two types of chiral stationary phases were used, one based on dinitrobenzoylphenylglycine and the other on dinitrobenzylphenylethylamine plus tartaric acid. The racemic mixture was used as the mobile phase additive and k′ of their system peaks was studied as a function of the mixture concentration in the mobile phase in both cases. A shift in k′ of the two system peaks was observed and considered as an indication that competition occurred. The areas of the two system peaks were also studied as a function of the concentration of the enantiomers in the samples, using two different compositions of the mobile phase. The dependency of system peaks' area on the sample composition indicated whether competition between the enantiomers occurred. One mobile phase contained 0.1 mM of the racemic mixture, where the area of the two retained system peaks behaved independently, i.e., only the peak corresponding to the enantiomer was affected by its presence in the sample. The other mobile phase contained 0.75 mM of the racemic mixture, and both peaks were affected by the injection of any one of the enantiomers. The interdependency of the system peaks' area on both the enantiomers indicated that their distribution in the chiral system was interrelated due to mutual interactions. A quantitative treatment of the interdependency and competition was excluded, due to the irreversible adsorption of the two enantiomers on the chiral stationary phase after using overloading concentrations. This irreversible adsorption was visualized by the appearance of two retained system peaks of the two residual enantiomers. These system peaks were detected only when the sample contained pure enantiomers due to competition between the enantiomer in the sample with the residual enantiomers in the stationary phase. © 1994 Wiley-Liss, 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