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.
Halohydrin dehalogenase (HheC) from Agrobacterium radiobacter AD1 catalyzes the reversible intramolecular nucleophilic displacement of a halogen by a hydroxyl group in vicinal haloalcohols, producing the corresponding epoxides. The enzyme displays high enantioselectivity toward some aromatic halohydrins. To understand the kinetic mechanism and enantioselectivity of the enzyme, steady-state and pre-steady-state kinetic analysis was performed with p-nitro-2-bromo-1-phenylethanol (PNSHH) as a model substrate. Steady-state kinetic analyses indicated that the k(cat) of the enzyme with the (R)-enantiomer (22 s(-1)) is 3-fold higher than with the (S)-enantiomer and that the K(m) for the (R)-enantiomer (0.009 mM) is about 45-fold lower than that for the (S)-enantiomer, resulting in a high enantiopreference for the (R)-enantiomer. Product inhibition studies revealed that HheC follows an ordered Uni Bi mechanism for both enantiomers, with halide as the first product to be released. To identify the rate-limiting step in the catalytic cycle, pre-steady-state experiments were performed using stopped-flow and rapid-quench methods. The results revealed the existence of a pre-steady-state burst phase during conversion of (R)-PNSHH, whereas no such burst was observed with the (S)-enantiomer. This indicates that a product release step is rate-limiting for the (R)-enantiomer but not for the (S)-enantiomer. This was further examined by doing single-turnover experiments, which revealed that during conversion of the (R)-enantiomer the rate of bromide release is 21 s(-1). Furthermore, multiple turnover analyses showed that the binding of (R)-PNSHH is a rapid equilibrium step and that the rate of formation of product ternary complex is 380 s(-1). Taken together, these findings enabled the formulation of an ordered Uni Bi kinetic mechanism for the conversion of (R)-PNSHH by HheC in which all of the rate constants are obtained. The high enantiopreference for the (R)-enantiomer can be explained by weak substrate binding of the (S)-enantiomer and a lower rate of reaction at the active site.  相似文献   

2.
RS-8359, (+/-)-4-(4-cyanoanilino)-5,6-dihydro-7-hydroxy-7H-cyclopenta[d]-pyrimidine is a racemic compound with a selective and reversible monoamine oxidase A (MAO-A) inhibition activity. The substrate and product enantioselectivity with respect to 2-hydroxylation of RS-8359 enantiomers was studied using mouse and rat liver microsomes. In mice, the (S)-enantiomer was transformed to the cis-diol metabolite, whereas the (R)-enantiomer to the trans-diol metabolite. The Vmax/Km value for the formation of the cis-diol metabolite from the (S)-enantiomer was sevenfold greater than that for the formation of the trans-diol metabolite from the (R)-enantiomer. The greater Vmax/Km value for the (S)-enantiomer was due to the tenfold smaller Km value compared to that for the (R)-enantiomer. The results were in fair agreement with the previously reported low plasma concentrations of the (S)-enantiomer and the high recovery of the cis-diol metabolite derived from the (S)-enantiomer in urine after oral administration of RS-8359 to mice. Similarly to mice, in rats the (R)-enantiomer was transformed to the trans-diol metabolite, whereas the (S)-enantiomer yielded the cis-diol and trans-diol metabolites. The Vmax/Km value for the (R)-enantiomer was larger than that for the (S)-enantiomer in rats, indicating that the low plasma concentration of the (S)-enantiomer in rats might be caused by a metabolic reaction other than P450-dependent hydroxylation. CYP3A was shown to be responsible for the trans-diol formation from the (R)-enantiomer.  相似文献   

3.
Wen YZ  Yuan YL  Shen CS  Liu HJ  Liu WP 《Chirality》2009,21(3):396-401
The enantioselective interaction between penicillium expansum alkaline lipase and chiral phenoxypropionic acid herbicide dichlorprop was studied by using UV differential spectrophotometry and fluorescence spectrophotometry in the presence of a pH 8, phosphate buffer solution. Chiral differences in the UV absorption and fluorescence spectra of lipase with dichlorprop were detected. (R)-Dichlorprop interacted the strongest with lipase as measured by both UV absorption and fluorescence spectrophotometry, followed by (Rac)-dichlorprop, while (S)-dichlorprop had the weakest interaction. The hydrophobic interaction seem to play the dominant role in the interactions and the (R)-enantiomer needed the minimum put of energy to drive the endothermic reaction, while the Rac-type and S-type compounds needed more for the reaction to take place. In the meantime, the catalytic hydrolysis of FDA with lipase show that (R)-DCPP could inhibit lipase the most strongly relatively at the same condition, perhaps because (R)-DCPP had a stronger combining effect and high enantiomeric selectivity on lipase than (Rac)-DCPP and (S)-DCPP.  相似文献   

4.
Liu Y  Wang F  Tan T 《Chirality》2009,21(3):349-353
Extracellular lipase LIP prepared in our lab from the yeast Yarrowia lipolytica was used for the resolution of racemic ibuprofen. The (S)-enantiomer was preferred by lipase LIP, and the unreacted (R)-enantiomer was extracted and racemized in basic solvent-water medium to be re-resolved. Solvent, content of solvent, base concentration, and temperature have a strong effect on racemization. The (S)-ester was separated and hydrolyzed to (S)-ibuprofen in acidic dimethyl sulfoxide-water mixture containing 70% dimethyl sulfoxide. The high purity (S)-ibuprofen (ee = 0.98) was obtained using lipase LIP to catalyze hydrolysis of (S)-ester in 0.1 M phosphate buffer (pH = 8).  相似文献   

5.
The influence of endotoxin-induced inflammation on the enantioselective pharmacokinetics of propranolol, oxprenolol, and verapamil, which bind to α1-acid glycoprotein, was studied in the rat. The racemic mixtures were given orally. In the control animals, for propranolol and oxprenolol, the plasma concentrations of the (R)-enantiomer were higher than those of the (S)-enantiomer, while for verapamil the reverse was true. Protein binding and intrinsic clearance are the main factors responsible for this enantioselectivity. After endotoxin treatment, for the three drugs tested the plasma concentrations and the plasma binding of both enantiomers were significantly increased. This effect was more pronounced for (R)-propranolol, (R)-oxprenolol, and (S)-verapamil than for their respective antipodes. The enantioselective effect of endotoxin on the plasma concentrations of the drugs studied seems mainly due to the enantioselective increase in binding to α1-acid glycoprotein. © 1994 Wiley-Liss, Inc.  相似文献   

6.
Wang X  Wang XJ  Ching CB 《Chirality》2002,14(4):318-324
Characterization of the racemic species, which can be a racemic compound, a racemic conglomerate, or a pseudoracemate (solid solution), is a prerequisite for the design of crystallization resolution processes. It is useful to determine the solid/liquid equilibrium solubility of the enantiomer mixtures for crystallization operation. For the beta-blocker drug propranolol hydrochloride, Gibbs free energy of formation of racemic compound and entropy of mixing of the (R)- and (S)- enantiomers in the liquid state for racemic conglomerate were calculated. The structural differences between (R, S)-propranolol hydrochloride and its (S)-enantiomer were further investigated by powder X-ray diffraction patterns, infrared spectra, and solid-state NMR spectra. The solubility and metastable zone width of (R, S)- propranolol hydrochloride in a mixed solvent of methanol and acetone were determined by cooling crystallization over the temperature range 3.5-42.5 degrees C. The ternary solubility diagram of (R)-, (S)-propranolol hydrochloride was constructed using the same mixed solvent. The diagram will be useful as a guide for choosing crystallization operation conditions to produce pure enantiomers.  相似文献   

7.
RS-8359, (+/-)-4-(4-cyanoanilino)-5,6-dihydro-7-hydroxy-7H-cyclopenta[d]-pyrimidine is a selective and reversible MAO-A inhibitor. The (S)-enantiomer of RS-8359 has been demonstrated to be inverted to the (R)-enantiomer after oral administration to rats. In the current study, we investigated the chiral inversion mechanism and the properties of involved enzymes using rat liver subcellular fractions. The 7-hydroxy function of RS-8359 was oxidized at least by the two different enzymes. The cytosolic enzyme oxidized enantiospecifically the (S)-enantiomer with NADP as a cofactor. On the other hand, the microsomal enzyme catalyzed more preferentially the oxidation of the (S)-enantiomer than the (R)-enantiomer with NAD as a cofactor. With to product enantioselectivity of reduction of the 7-keto derivative, it was found that only the alcohol bearing (R)-configuration was formed by the cytosolic enzyme with NADPH and the microsomal enzyme with NADH at almost equal rate. The reduction rate was much larger than the oxidation rate of 7-hydroxy group. The results suggest that the chiral inversion might occur via an enantioselectivity of consecutive two opposing reactions, oxidation and reduction of keto-alcohol group. In this case, the direction of chiral inversion from the (S)-enantiomer to the (R)-enantiomer is governed by the enantiospecific reduction of intermediate 7-keto group to the alcohol with (R)-configuration. The enzyme responsible for the enantiospecific reduction of the 7-keto group was purified from rat liver cytosolic fractions and identified as 3alpha-hydroxysteroid dehydrogenase (3alpha-HSD) via database search of peptide mass data obtained by nano-LC/MS/MS.  相似文献   

8.
Clift MD  Ji H  Deniau GP  O'Hagan D  Silverman RB 《Biochemistry》2007,46(48):13819-13828
Gamma-aminobutyric acid aminotransferase (GABA-AT), a pyridoxal 5'-phosphate dependent enzyme, catalyzes the degradation of the inhibitory neurotransmitter gamma-aminobutyric acid (GABA) to succinic semialdehyde with concomitant conversion of pyridoxal 5'-phosphate (PLP) to pyridoxamine 5'-phosphate (PMP). The enzyme then catalyzes the conversion of alpha-ketoglutarate to the excitatory neurotransmitter L-glutamate. Racemic 4-amino-3-fluorobutanoic acid (3-F-GABA) was shown previously to act as a substrate for GABA-AT, not for transamination, but for HF elimination. Here we report studies of the reaction catalyzed by GABA-AT on (R)- and (S)-3-F-GABA. Neither enantiomer is a substrate for transamination. Very little elimination from the (S)-enantiomer was detected using a coupled enzyme assay; The rate of elimination of HF from the (R)-enantiomer is at least 10 times greater than that for the (S)-enantiomer. The (R)-enantiomer is about 20 times more efficient as a substrate for GABA-AT catalyzed HF elimination than GABA is a substrate for transamination. The (R)-enantiomer also inhibits the transamination of GABA 10 times more effectively than the (S)-enantiomer. Using a combination of computer modeling and the knowledge that vicinal C-F and C-NH3+ bonds have a strong preference to align gauche rather than anti to each other, it is concluded that on binding of free 3-F-GABA to GABA-AT the optimal conformation places the C-NH3+ and C-F bonds gauche in the (R)-enantiomer but anti in the (S)-enantiomer. Furthermore, the dynamic binding process and the bioactive conformation of GABA bound to GABA-AT have been inferred on the basis of the different biological behavior of the two enantiomers of 3-F-GABA when they bind to the enzyme. The present study suggests that the C-F bond can be utilized as a conformational probe to explore the dynamic binding process and provide insight into the bioactive conformation of substrates, which cannot be easily determined by other biophysical approaches.  相似文献   

9.
8(S),9(R)-epoxyeicosatrienoic acid, a major product of the renal cortex, was found to be a substrate for cyclooxygenase from human platelets and ram seminal vesicles. 11(R)-hydroxy-8(S),9(R)-epoxyeicosatrienoic acid was the sole metabolic product. The 8(R),9(S)-enantiomer formed both C-11 and C-15 hydroxylated metabolites. These novel findings suggest that the cyclooxygenase-dependent renal vasoconstrictor activity of 8(S),9(R)-epoxyeicosatrienoic acid may be due to the 11(R)-hydroxy metabolite.  相似文献   

10.
The influence of a single oral dose of 30 mg nicardipine on the pharmacokinetics of (R)- and (S)-propranolol, given orally as rac-propranolol 80 mg, was studied in 12 healthy volunteers. The plasma concentrations were higher for the (S)-enantiomer than for the (R)-enantiomer. The Clo and the Cl′intr of (S)-propranolol were significantly lower than the Clo and Cl′intr of (R)-propranolol. The unbound fraction of (R)-propranolol was significantly higher than that of (S)-propranolol. Coadministration of nicardipine significantly increased the AUC and Cmax and significantly decreased the Clo and Cl′intr for unbound drug of (R)- and (S)-propranolol. These changes were more important for (R)- than for (S)-propranolol. The protein binding was not altered by nicardipine. The enantioselective effect of nicardipine on the metabolic clearance of propranolol appears to be due to an interaction at the level of the metabolizing enzymes. The effect on blood pressure of rac-propranolol was little affected when nicardipine was coadministered with rac-propranolol, and its bradycardic effect was reduced. © 1994 Wiley-Liss, Inc.  相似文献   

11.
Conscious male Wistar SPF Riv:TOX rats were dosed intravenously with 2.5, 5, or 10 mg/kg rac-propranolol·HCl, or with 5 mg/kg of either (-)-(S)- or (+)-(R)-propranolol·HCl. Disposition of (-)-(S)- and (+)-(R)-propranolol after dosing of rac-propranolol was linear in the dose range examined. Total plasma clearance was not changed in animals dosed with the individual enantiomers compared to the animals that were dosed with rac-propranolol. However, for (-)-(S)-propranolol both volume of distribution and elimination half-life decreased, whereas for (+)-(R)-propranolol increases were observed for these characteristics, in animals dosed with the individual enantiomers. Our observations suggest that the (+)-(R)-enantiomer competes with (-)-(S)-propranolol for plasma protein binding sites, resulting in lower plasma protein binding of the (-)-(S)-enantiomer when the racemate is administered. From recent toxicological experiments, it was concluded that rac-propranolol is more toxic than the individual enantiomers in the rat, when dosed iv at the same total mass. It is concluded that the observed potentiation of toxic effects of propranolol enantiomers when administered as a racemate can at least partly be explained by a pharmacokinetic interaction. © 1995 Wiley-Liss, Inc.  相似文献   

12.
ImmobilizedCandida antarctica lipase was used to catalyze the separation of ketoprofen into its components by means of esterification followed by the enzymatic hydrolysis of the ester product. In this study, ketoprofen underwent esterification to ethanol in the presence of isooctane. When the reaction was complete, 58.3% of the ketoprofen had been transformed into an ester. The ketoprofen remaining in solution after the reation was complete consisted primarily of itsS-enantiomer (83.0%), while the 59.4% of the ketoprofen component of the ester consisted of itsR-enantiomer. We then subjected the ester product to enzymatic hydrolysis in the presence of the same enzyme and produced a ketoprofen product rich in theR-enantiomer; 77% of this product consisted of theR-enantiomer when 50% of the ester had been hydrolyzed, and 90% of it consisted of theR-enantiomer when 30% of the ester had been hydrolyzed. By contrast, theR-enantiomer levels only reached approximately 42 and 65%, respectively, when 50 and 30% of the racemic ester was hydrolyzed under the same conditions.  相似文献   

13.
Synthesis of both enantiomers of methyl 4,5-didehydrojasmonate (1, Delta(4,5)-MJA; >99.8% ee), a constituent of jasmin absolute, established the absolute configuration of the natural product, and their odor quality was evaluated. The fragrance of the natural (3S,7R)-enantiomer (a fresh natural, sweet floral fruity odor, reminiscent of Jasmin and Ylang Ylang flower, more intensive and tenacious) was superior to that of the unnatural (3R,7S)-enantiomer (a floral green odor with slight metallic green aspect, less intensive than the natural form) and the racemate (green-floral note, having weak and less volume than methyl jasmonate). Odor difference between natural and unnatural enantiomers of methyl jasmonate (2) is also reported.  相似文献   

14.
The enantioselective esterification of racemic ibuprofen, catalyzed by a Candida cylindracea lipase, was studied in a water-in-oil microemulsion (AOT/isooctane). By using n-propanol as the alcohol, an optimal W(0) ([H(2)O]/[AOT] ratio) of 12 was found for the synthesis of n-propyl-ibuprofenate at room temperature. The lipase showed high preference for the S(+)-enantiomer of ibuprofen, which was esterified to the corresponding S(+)-ibuprofen ester. The R(-)-ibuprofen remained unesterified in the microemulsion. The calculated enantioselectivity value (E) for S-ibuprofen ester was greater than 150 (conversion 0.32). The enzyme activities of n-alcohols with different chain lengths (3-12) were compared, and it appeared that short- (propanol and butanol) and long-chained (decanol and dodecanol) alcohols were better substrates than the intermediate ones (pentanol, hexanol, and octanol). However, unlike secondary and tertiary alcohols, all of the tested primary alcohols were substrates for the lipase. The reversible reaction (i.e., the hydrolysis of racemic ibuprofen ester in the microemulsion) was also carried out enantioselectively by the enzyme. Only the S form of the ester was hydrolyzed to the corresponding S-ibuprofen. The reaction yield was, however, only about 4% after 10 days of reaction. The corresponding yield for the esterification of ibuprofen was about 35% (10 days). The high enantioselectivity displayed by the lipase in the microemulsion system was seen neither in a similar esterification reaction in a pure organic solvent system (isooctane) nor in the hydrolysis reaction in an aqueous system (buffer). The E value for S-ibuprofen ester in the isooctane system was 3.0 (conversion 0.41), and only 1.3 for S-ibuprofen in the hydrolysis reaction (conversion 0.32). The differences in enantioselectivity for the lipase in various systems are likely due to interfacial phenomena. In the microemulsion system, the water in which the enzyme is dissolved is separated from the solvent by a layer of surfactant molecules, thus creating an interface with a relatively large area. Such interfaces are not present in the pure organic solvent systems (no surfactant) nor in aqueous systems. (c) 1993 John Wiley & Sons, Inc.  相似文献   

15.
Coenzyme B(12)-dependent ethanolamine ammonia-lyase acts on both enantiomers of the substrate 2-amino-1-propanol [Diziol, P., et al. (1980) Eur. J. Biochem. 106, 211-224]. To rationalize this apparent lack of stereospecificity and the enantiomer-specific stereochemical courses of the deamination, we analyzed the X-ray structures of enantiomer-bound forms of the enzyme-cyanocobalamin complex. The lower affinity for the (R)-enantiomer may be due to the conformational change of the Valα326 side chain of the enzyme. In a manner consistent with the reported experimental results, we can predict that the pro-S hydrogen atom on C1 is abstracted by the adenosyl radical from both enantiomeric substrates, because it is the nearest one in both enantiomer-bound forms. We also predicted that the NH(2) group migrates from C2 to C1 by a suprafacial shift, with inversion of configuration at C1 for both enantiomeric substrates, although the absolute configuration of the 1-amino-1-propanol intermediate is not yet known. Reported labeling experiments demonstrate that (R)-2-amino-1-propanol is deaminated by the enzyme with inversion of configuration at C2, whereas the (S)-enantiomer is deaminated with retention. By taking these results into consideration, we can predict the rotameric radical intermediate from the (S)-enantiomer undergoes flipping to the rotamer from the (R)-enantiomer before the hydrogen back-abstraction. This suggests the preference of the enzyme active site for the rotamer from the (R)-enantiomer in equilibration. This preference might be explained in terms of the steric repulsion of the (S)-enantiomer-derived product radical at C3 with the Pheα329 and Leuα402 residues.  相似文献   

16.
Both hitherto unknown (+)-(R)- and (?)-(S)-thioglycidyl esters, (R)-( 2 ) and (S)-( 2 ), have been synthesized with different high enantiomeric excesses (ee) by two routes from the corresponding rac-glycidyl esters rac-( 1 ). The first includes a porcine pancreatic lipase (PPL)-mediated kinetic resolution of these esters followed by sulfuration with practically complete inversion to the (+)-(R)-enantiomer (+)-(R)-( 2 ) (36–86% ee). (?)-(S)-Thioglycidyl esters (?)-(S)-( 2 ) are obtained by the reverse reaction sequence (43–80% ee). In the latter case the hydrolysis rate is lower than that of analogous glycidyl esters. Moreover, the dependence of enantiomeric excess on the size of the acyl-group is of the opposite tendency. Therefore, in both cases suitable selection of the acid residue gives rise to maximum enantioselectivity. The irreversible lipase-catalyzed acylation of rac-glycidol and rac-thioglycidol, however, was found to be a less suitable alternative. The enantiomeric excess of recovered homochiral esters was determined by chiral chromatography using modified cellulose stationary phases (OB, OD). © 1993 Wiley-Liss, Inc.  相似文献   

17.
The influence of aging on the pharmacokinetics and the tissue distribution of (R)- and of (S)-propranolol was studied in 3-, 12-, and 24-month-old rats. After both iv and oral administration of rac-propranolol, the plasma concentrations were higher for the (R)- than for the (S)-enantiomer. For the tissue concentrations, the reverse was true. The free fraction of (S)-propranolol in plasma was about 4 times larger than that of (R)-propranolol, and this is the main factor responsible for the differences in kinetics between the two enantiomers. There was a suggestion for a difference in tissue binding between the two enantiomers. With aging, the plasma and tissue concentrations of both enantiomers increase, probably due to a decrease in blood clearance. Tissue binding did not change much with aging. Notwithstanding the marked differences between the kinetics of the propranolol enantiomers, the changes which occur with aging affect both enantiomers to the same degree.  相似文献   

18.
Enantiomerically pure (R)- and (S)-1-(1H-1,2,4-triazol-1-yl)-2-(4-fluorophenyl)-3-trimethylsilylpropan-2-ol 1 were prepared via an enantioselective Grignard reaction. The absolute stereochemistry of 1 was determined by X-ray analysis. In a comparison of in vitro antifungal activities of the enantiomers, the (-)-enantiomer with the R-absolute configuration was far more potent than the (+)-enantiomer.  相似文献   

19.
(2S)-Flavanone 3-hydroxylase from flowers of Petunia hybrida catalyses the conversion of (2S)-naringenin to (2R,3R)-dihydrokaempferol. The enzyme could be partially stabilized under anaerobic conditions in the presence of ascorbate. For purification, 2-oxoglutarate and Fe2+ had to be added to the buffers. The hydroxylase was purified about 200-fold by a six-step procedure with low recovery. The Mr of the enzyme was estimated by gel filtration to be about 74,000. The hydroxylase reaction has a pH optimum at pH 8.5 and requires as cofactors oxygen, 2-oxoglutarate, Fe2+ and ascorbate. With 2-oxo[1-14C]glutarate in the enzyme assay dihydrokaempferol and 14CO2 are formed in a molar ratio of 1:1. Catalase stimulates the reaction. The product was unequivocally identified as (+)-(2R,3R)-dihydrokaempferol. (2S)-Naringenin, but not the (2R)-enantiomer is a substrate of the hydroxylase. (2S)-Eriodictyol is converted to (2R,3R)-dihydroquercetin. In contrast, 5,7,3',4',5'-pentahydroxy-flavanone is not a substrate. Apparent Michaelis constants for (2S)-naringenin and 2-oxoglutarate were determined to be respectively 5.6 mumol X l-1 and 20 mumol X l-1 at pH 8.5. The Km for (2S)-eriodictyol is 12 mumol X l-1 at pH 8.0. Pyridine 2,4-dicarboxylate and 2,5-dicarboxylate are strong competitive inhibitors with respect to 2-oxoglutarate with Ki values of 1.2 mumol X l-1 and 40 mumol X l-1, respectively.  相似文献   

20.
Here we report a detailed procedure for the enzymatic kinetic resolution of 4-chloro-2-(1-hydroxyalkyl)pyridines, valuable precursors for the preparation of enantiomerically pure catalysts derived from 4-(N,N-dimethylamino)pyridine. Pseudomonas cepacia lipase shows excellent enantioselectivity in the acylation of the (R)-enantiomer at 30 degrees C and 250 r.p.m., with vinyl acetate as the acyl donor and tetrahydrofuran as the solvent. The reaction times for resolution of the pyridine derivatives depend on the structure of the selected substrate.  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