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 46 毫秒
1.
To characterise the pharmacokinetic properties of each enantiomer of ibuprofen in broiler chickens, a stereospecific HPLC method based on a α1-acid glycoprotein bonded chiral stationary phase has been validated. S-(+)-naproxen was used as internal standard. Enantiomers of ibuprofen and S-(+)-naproxen were baseline separated using a mobile phase consisting of 0.1 M phosphate buffer pH=7 and 0.4% 2-propanol. The method is precise, specific, accurate and reproducible. Recoveries were higher than 80% and the limits of quantification for R-(−)- and S-(+)-ibuprofen were 1.16 and 1.37 μg ml−1, respectively. The method seemed suitable for the pharmacokinetic studies of ibuprofen in chickens.  相似文献   

2.
Candida rugosa lipase (CRL) was applied in a non-solvent esterification reaction to yield twelve wax esters. All products were obtained in nearly 100% yield for 10 h at 50°C when immobilized PEG2000-activated C. rugosa lipase was added to the reaction mixture. The surfactant had also a beneficial effect on the stability of the biocatalytic preparation with 83% of its activity conserved after the seventh run of repeated batch reactions.  相似文献   

3.
In this study, the stereoselective pharmacokinetics of doxazosin enantiomers and their pharmacokinetic interaction were studied in rats. Enantiomer concentrations in plasma were measured using chiral high‐pressure liquid chromatography (HPLC) with fluorescence detection after oral or intravenous administration of (–)‐(R)‐doxazosin 3.0 mg/kg, (+)‐(S)‐doxazosin 3.0 mg/kg, and rac‐doxazosin 6.0 mg/kg. AUC values of (+)‐(S)‐doxazosin were always larger than those of (–)‐(R)‐doxazosin, regardless of oral or intravenous administration. The maximum plasma concentration (Cmax) value of (–)‐(R)‐doxazosin after oral administration was significantly higher when given alone (110.5 ± 46.4 ng/mL) versus in racemate (53.2 ± 19.7 ng/mL), whereas the Cmax value of (+)‐(S)‐doxazosin did not change significantly. The area under the curve (AUC) and Cmax values for (+)‐(S)‐doxazosin after intravenous administration were significantly lower, and its Cl value significantly higher, when given alone versus in racemate. We speculate that (–)‐(R)‐doxazosin increases (+)‐(S)‐doxazosin exposure probably by inhibiting the elimination of (+)‐(S)‐doxazosin, and the enantiomers may be competitively absorbed from the gastrointestinal tract. In conclusion, doxazosin pharmacokinetics are substantially stereospecific and enantiomer–enantiomer interaction occurs after rac‐administration. Chirality 27:738–744, 2015. © 2015 Wiley Periodicals, Inc.  相似文献   

4.
Organotin compounds are the active components of some fungicides, which are potential inhibitors of the F1F0-ATP synthase. The studies about the reaction mechanism might indicate a pathway to understand how these compounds work in biological systems, however, has not been clarified so far. In this line, molecular modeling studies and density functional theory calculations were performed in order to understand the molecular behavior of those compounds when they interact with the active site of the enzyme. Our findings indicate that a strong interaction with His132 can favor a chemical reaction with organotin compounds due to π–π stacking interactions with aromatic rings of organotin compounds. Furthermore, dependence on molecule size is related to possibility of reaction with the amino acid residue His132. Thus, it can also be noticed, for organotin compounds, that substituents with four carbons work by blocking the subunit a, in view of the high energy transition found characterized by steric hindrance.  相似文献   

5.
An interesting observation was found during our continued studies on the hydrolysis of ibuprofen esters by Candida rugosa lipase (CRL). An important role is played by pH in the stereospecific hydrolysis of these esters. The flap region of CRL plays a significant role in the access of the substrate to the active site of the enzyme. At pH 5.6, 48% of the methyl ester and 5% of the butyl ester of ibuprofen were hydrolysed in 5.5 h, whereas at pH 7.2, 9% of methyl ester and 45% of the butyl ester of ibuprofen was hydrolysed in a identical reaction time using CRL. This lead us to assume that CRL prefers the methyl ester of ibuprofen as a substrate at an acidic pH and the butyl ester of ibuprofen at a neutral pH. Therefore, in order to understand the role of pH in the substrate selection by CRL for the esters of ibuprofen we used the crystallographic coordinates of the open form of the CRL (1CRL) for molecular dynamics (MD) simulations under acidic and neutral conditions for 2 ns using GROMACS. The final structures obtained after simulation in acidic and neutral conditions were compared with the energy-minimized structure, and the root-mean-square deviations (r.m.s.ds) were calculated. The r.m.s.d. of the CRL flap at neutral pH was found to be greater than that of the CRL flap at acidic pH. The extent to which the flap opens at neutral pH allowed the bulkier substrate, the butyl ester of ibuprofen, to diffuse into the active site and provides the best enzyme-substrate fit for this specific substrate. At acidic pH there is a decreased opening of the flap thereby accommodating a more compact substrate, namely the methyl ester of ibuprofen. Thus, simulation experiments using MD provide reasonable insight for the pH-dependent substrate selectivity of CRL in aqueous environments.  相似文献   

6.
The pharmacokinetics and metabolic chiral inversion of the S(+)‐ and R(−)‐enantiomers of tiaprofenic acid (S‐TIA, R‐TIA) were assessed in vivo in rats, and in addition the biochemistry of inversion was investigated in vitro in rat liver homogenates. Drug enantiomer concentrations in plasma were investigated following administration of S‐TIA and R‐TIA (i.p. 3 and 9 mg/kg) over 24 hr. Plasma concentrations of TIA enantiomers were determined by stereospecific HPLC analysis. After administration of R‐TIA it was found that 1) there was a time delay of peak S‐TIA plasma concentrations, 2) S‐TIA concentrations exceeded R‐TIA concentrations from ∼2 hr after dosing, 3) Cmax and AUC(0‐∞) for S‐TIA were greater than for R‐TIA following administration of S‐TIA, and 4) inversion was bidirectional but favored inversion of R‐TIA to S‐TIA. Bidirectional inversion was also observed when TIA enantiomers were incubated with liver homogenates up to 24 hr. However, the rate of inversion favored transformation of the R‐enantiomer to the S‐enantiomer. In conclusion, stereoselective pharmacokinetics of R‐ and S‐TIA were observed in rats and bidirectional inversion in rat liver homogenates has been demonstrated for the first time. Chiral inversion of TIA may involve metabolic routes different from those associated with inversion of other 2‐arylpropionic acids such as ibuprofen. Chirality 11:103–108, 1999. © 1999 Wiley‐Liss, Inc.  相似文献   

7.
Whole cells and cell-free extracts derived from Rhodococcus rhodochrous NCIMB 11216 were shown to hydrolyse both aliphatic and aromatic nitriles, when the organism had been grown on either propionitrile or benzonitrile as the source of carbon and nitrogen. Whole cell suspensions and cell-free extracts derived from bacteria grown on either substrate were able to biotransform R-(-),S-(+)-2-methylbutyronitrile. The S-(+) enantiomer was biotransformed more rapidly than the the R-(-) enantiomer. For whole cell biotransformations at 30°C, the maximum enantiomeric excess (ee) of the remaining R-(-)-2-methylbutyronitrile was 93% when 70% of the R-(-) enantiomer had been converted to the product, 2-methylbutyric acid. For the corresponding biotransformation at 4°C, there was an ee of 93% for the residual R-(-) enantiomer of the substrate when only 60% of it had been converted to product. For biotransformations by cell-free extracts at 30°C the 2-methylbutyric acid product had an ee of 17% for the S-(+) enantiomer at the time of optimal ee for the remaining R-(-) enantiomer of the substrate. In contrast, when the reaction was carried out by whole cells, the ee for the product acid was 0.36%. This was probably due to further, non-selective metabolism of the acid, which was especially significant at the beginning of the reaction. At both temperatures, the ee for the S-(+) enantiomer of 2-methylbutyric acid was at a maximum in the early stage of the biotransformation; for example, at 4°C the maximum detectable ee was 100% when the yield was 11%.Abbreviations EDTA Ethylenediaminetetraacetic acid - ee enantiomeric excess - FID flame ionisation detector - GC gas chromatography - 1HNMR H nuclear magnetic resonance - K m Michaelis constant - NCIMB National Collection of Industrial and Marine Bacteria - td doubling time - V max Maximum velocity  相似文献   

8.
Use of lipases in the resolution of racemic ibuprofen   总被引:5,自引:0,他引:5  
Summary Resolution of (R,S)-ibuprofen enantiomers by esterification in different organic solvents was studied using Candida cylindracea lipase. This enzyme preparation had high enantiospecificity for S(+)-ibuprofen in the esterification reaction of a racemic ibuprofen with primary alcohols. The esterification yields of secondary alcohols were much lower than those of primary alcohols. Esterification with tertiary alcohols was not observed. The synthesis of esters was profoundly affected by the amount of water in the reaction mixture. C. cylindracea lipase was active only in very hydrophobic solvents. The esterification activity of the lipase was reduced significantly by addition of water. The R- and S-enantiomers of ibuprofen were determined without derivatization by HPLC using a chiral column.  相似文献   

9.
This review is a journey concerning the investigations of the kinetic resolution of racemic ibuprofen for the last 20 years. The relevancy of the pharmacological uses of the S(?+?) enantiomer along with its higher cost compared with racemic profen are the driving forces of a variety of scientific research studies addressing the enzymatic resolution of ibuprofen through enantiomeric esterification using lipases as biocatalysts. Lipases of fungal sources such as Candida rugosa, Rhizomucor miehei and the lipase B of Candida antarctica have been extensively studied both in homogeneous and heterogeneous (immobilized on solid supports) processes. In this context, the various alcohols and organic co-solvents frequently used in the esterification of racemic ibuprofen are summarized and discussed in this review. Moreover, recent investigations using membranes as reactors coupled with the separation of the desired product and microfluidic devices are presented. Finally, some guidelines about future perspectives regarding the technology of the kinetic resolution of profens and research niches are given.  相似文献   

10.
The chemical mechanism of action of glucose oxidase from Aspergillus niger   总被引:2,自引:0,他引:2  
Glucose oxidase from Aspergillus niger (EC 1.1.3.4) is able to catalyze the oxidation of -D-glucose with p-benzoquinone, methyl-1,4-benzoquinone, 1,2-naphthoquinone, 1,2-naphthoquinone-4-sulfonic acid, potassium ferricyanide, phenazine methosulfate, and 2,6-dichloroindophenol. In this work, the steady-state kinetic parameters, V 1/K B , for reactions of these substrates were collected from pH 2.5–8. Further, the molecular models of the enzyme's active site were constructed for the free enzyme in the oxidized state, the complex of -D-glucose with the oxidized enzyme, the complex of reduced enzyme with methyl-1,4-benzoquinone, the reduced enzyme plus 1,2-naphthoquinone-4-sulfonic acid, oxidized enzyme plus reduced 1,2-naphthoquinone-4-sulfonic acid (hydroquinone anion), and oxidized enzyme plus fully reduced 1,2-naphthoquinone-4-sulfonic acid.Combining the steady-state kinetic and structural data, it was concluded that Glu412 bound to His559, in the active site of enzyme, modulates powerfully its catalytic activity by affecting all the rate constants in the reductive and the oxidative half-reaction of the catalytic cycle. His516 is the catalytic base in the oxidative and the reductive part of the catalytic cycle. It was estimated that the pK a of Glu412 (bound to His559) in the free reduced enzyme is 3.4, and the pK a of His516 in the free reduced enzyme is 6.9.  相似文献   

11.
Chiral discrimination observed in high‐performance liquid chromatography (HPLC) with the novel chiral stationary phase (CSP‐18C6I) derived from (+)‐(R)‐18‐crown‐6 tetracarboxylic acid [(+)‐18C6H4] was investigated by X‐ray crystallographic analysis of the complex composed of the R‐enantiomer of 1‐(1‐naphthyl)ethylamine (1‐NEA) and (+)‐18C6H4. Mixtures of 1‐NEA (the R‐ or S‐enantiomer) and (+)‐18C6H4 were dissolved in methanol‐water (1:1) solution and allowed to stand for crystallization. The R‐enantiomer crystallized with (+)‐18C6H4 as a co‐crystal, although the S‐enantiomer did not. This result was in good agreement with the enantiomer elution order of 1‐NEA in CSP‐18C6I. The apparent binding constants (Ka) of the enantiomers to the (+)‐18C6H4 obtained from 1H‐NMR experiments also supported the above‐mentioned result. The X‐ray crystal structure of the 1:1 complex of the R‐enantiomer and (+)‐18C6H4 indicated the four sets of hydrogen bond association between the naphthylethylammonium cation and oxygen of polyether ring or carbonyl group of (+)‐18C6H4. Chirality 11:173–178, 1999. © 1999 Wiley‐Liss, Inc.  相似文献   

12.
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   

13.
Alpha‐cypermethrin (α‐CP), [(RS)‐a‐cyano‐3‐phenoxy benzyl (1RS)‐cis‐3‐(2, 2‐dichlorovinyl)‐2, 2‐dimethylcyclopropanecarboxylate], comprises a diastereoisomer pair of cypermethrin, which are (+)‐(1R‐cis‐αS)–CP (insecticidal) and (?)‐(1S‐cis‐αR)–CP (inactive). In this experiment, the stereoselective degradation of α‐CP was investigated in rat liver microsomes by high‐performance liquid chromatography (HPLC) with a cellulose‐tris‐ (3, 5‐dimethylphenylcarbamate)‐based chiral stationary phase. The results revealed that the degradation of (?)‐(1S‐cis‐αR)‐CP was much faster than (+)‐(1R‐cis‐αS)‐CP both in enantiomer monomers and rac‐α‐CP. As for the enzyme kinetic parameters, there were some variances between rac‐α‐CP and the enantiomer monomers. In rac‐α‐CP, the Vmax and CLint of (+)‐(1R‐cis‐αS)–CP (5105.22 ± 326.26 nM/min/mg protein and 189.64 mL/min/mg protein) were about one‐half of those of (?)‐(1S‐cis‐αR)–CP (9308.57 ± 772.24 nM/min/mg protein and 352.19 mL/min/mg protein), while the Km of the two α‐CP enantiomers were similar. However, in the enantiomer monomers of α‐CP, the Vmax and Km of (+)‐(1R‐cis‐αS) ‐CP were 2‐fold and 5‐fold of (?)‐(1S‐cis‐αR)‐CP, respectively, which showed a significant difference with rac‐α‐CP. The CLint of (+)‐(1R‐cis‐αS)–CP (140.97 mL/min/mg protein) was still about one‐half of (?)‐(1S‐cis‐αR)–CP (325.72 mL/min/mg protein) in enantiomer monomers. The interaction of enantiomers of α‐CP in rat liver microsomes was researched and the results showed that there were different interactions between the IC50 of (?)‐ to (+)‐(1R‐cis‐αS)‐CP and (+)‐ to (?)‐(1S‐cis‐αR)‐CP(IC50(?)/(+) / IC50(+)/(?) = 0.61). Chirality 28:58–64, 2016. © 2015 Wiley Periodicals, Inc.  相似文献   

14.
15.
β‐cyclodextrin (CD) and its derivatives HP‐β‐CD, DM‐β‐CD, and TM‐β‐CD have been employed as chiral selectors for the separation of three nonsteroidal antiinflammatory drugs (NSAIDs) and anticoagulant at relatively low concentration (8–15 mM) by capillary zone electrophoresis (CZE). In this study, baseline separation was achieved for ibuprofen, ketoprofen, naproxen, and warfarin. It was found that the addition of 0.1% hydroxypropyl methyl cellulose (HPMC) was effective for separation. Under these conditions, the S‐(+) enantiomer eluted before R‐(−) in terms of ibuprofen; the calculated energy values obtained from the molecular modeling correlated well with the elution order. An equation for calculating the pKa values by capillary electrophoresis was introduced, and the pKa values of the four chiral drugs at 25°C were obtained based on the equation. The value pKa + 0.5 is proposed to be the suitable pH of the background electrolyte for the separation of chiral compounds containing a carboxylic group. Chirality 11:56–62, 1999. © 1999 Wiley‐Liss, Inc.  相似文献   

16.
A kinetic study on esterification between d-glucose and l-phenylalanine catalysed by lipases from Rhizomucor miehei (RML) and Candida rugosa (CRL) in organic media investigated in detail showed that both the lipases followed a Ping-Pong Bi-Bi mechanism with two distinct types of competitive inhibitions. Graphical double reciprocal plots and computer simulation studies showed that competitive double substrate inhibition took place at higher concentrations leading to dead-end inhibition in the case of RML and in the case of CRL, inhibition only by d-glucose at higher concentrations leading to dead-end lipase–d-glucose complexes. An attempt to obtain the best fit of these kinetic models through curve-fitting yielded in good approximation, the apparent values of important kinetic parameters, RML: k cat = 2.24 ± 0.23 mM h−1 (mg protein)−1, K m l-phenylalanine = 95.6 ± 9.7 mM, K m d-glucose = 80.0 ± 8.5 mM, K i l-phenylalanine = 90.0 ± 9.2 mM, K i d-glucose = 13.6 ± 1.42 mM; CRL: k cat = 0.51 ± 0.06 mM h−1 (mg protein)−1, K m l-phenylalanine = 10.0 ± 0.98 mM, K m d-glucose = 6.0 ± 0.64 mM, K i d-glucose = 8.5 ± 0.81 mM.  相似文献   

17.
Abstract

Mammalian D1 and D2 dopamine receptors were stably expressed in Drosophila Schneider‐2 (S2) cells and screened for their pharmacological properties. Saturable, dose‐dependent, high affinity binding of the D1‐selective antagonist [3H]SCH‐23390 was detected only in membranes from S2 cells induced to express rat dopamine D1 receptors, while saturable, dose‐dependent, high affinity binding of the D2‐selective antagonist [3H]methylspiperone was detected only in membranes from S2 cells induced to express rat dopamine D2 receptors. No specific binding of either radioligand could be detected in membranes isolated from uninduced or untransfected S2 cells. Both dopamine D1 and D2 receptor subtypes displayed the appropriate stereoselective binding of enantiomers of the nonselective antagonist butaclamol. Each receptor subtype also displayed the appropriate agonist stereoselectivities. The dopamine D1 receptor bound the (+)‐enantiomer of the D1‐selective agonist SKF38393 with higher affinity than the (?)‐enantiomer, while the dopamine D2 receptor bound the (?)‐enantiomer of the D2‐selective agonist norpropylapomorphine with higher affinity than the (+)‐enantiomer. At both receptor subtypes, dopamine binding was best characterized as occurring to a single low affinity site. In addition, the low affinity dopamine binding was also found to be insensitive to GTPγS and magnesium ions. Overall, the pharmacological profiles of mammalian dopamine D1 and D2 receptors expressed in Drosophila S2 cells is comparable to those observed for these same receptors when they are expressed in mammalian cell lines. A notable distinction is that there is no evidence for the coupling of insect G proteins to mammalian dopamine receptors. These results suggest that the S2 cell insect G system may provide a convenient source of pharmacologically active mammalian D1 and D2 dopamine receptors free of promiscuous G protein contaminants.  相似文献   

18.
The H2-splitting active site of [NiFe] hydrogenases is tightly bound to the protein matrix via four conserved cysteine residues. In this study, the nickel-binding cysteine residues of HoxC, the large subunit of the H2-sensing regulatory hydrogenase (RH) from Ralstonia eutropha, were replaced by serine. All four mutant proteins, C60S, C63S, C479S, and C482S, were inactive both in H2 sensing and H2 oxidation and did not adopt the native oligomeric structure of the RH. Nickel was bound only to the C482S derivative. The assembly of the [NiFe] active site is a complex process that requires the function of at least six accessory proteins. Among these proteins, HypC has been shown to act as a chaperone for the large subunit during the maturation process. Immunoblot analysis revealed the presence of a strong RH-dependent HypC-specific complex in extracts containing the C60S, C63S, and C482S derivatives, pointing to a block in maturation for these mutant proteins. The lack of this complex in the extract containing C479S indicates that this specific cysteine residue might be crucial for the interaction between HoxC and HypC.This work is dedicated to Prof. H.G. Schlegel on the occasion of his 80th birthday.  相似文献   

19.
The product of human fragile histidine triad (FHIT) gene is a tumor suppressor protein of still largely unknown cellular background. We have shown previously that it binds protoporphyrin IX (a photosensitizer) which alters its enzymatic activity in vitro. Fhit, diadenosine triphosphate (Ap3A) hydrolase, possesses the active site with histidine triad His-φ-His-φ-His-φφ. So-called histidine Fhit mutants (His94Asn, His96Asn and His98Asn) exhibit highly reduced activity in vitro, however, their antitumor function has not been fully described yet. In this work we have cloned the cDNAs of histidine mutants into pPROEX-1 vector allowing the production of His6-fusion proteins. The mutated proteins: Fhit-H94N, Fhit-H96N and Fhit-H98N, were expressed in Escherichia coli BL21(DE3) and purified (up to 95%) by an improved, one-step affinity chromatography on Ni-nitrilotriacetate resin. The final yield was 2 mg homogenous proteins from 1 g bacteria (wet wt). The activity of purified proteins was assessed by previously described assay. The same purification procedure yielded 0.8 mg/ml and highly active wild-type Fhit protein (K m value for Ap3A of 5.7 μM). Importantly, purified mutant forms of Fhit also interact with a photosensitizer, protoporphyrin IX in vitro.  相似文献   

20.
Although multiple roles of dopamine through D1-like (D1 and D5) and D2-like (D2, D3, and D4) receptors are initiated primarily through stimulation or inhibition of adenylyl cyclase via Gs/olf or Gi/o, respectively, there have been many reports indicating diverse signaling mechanisms that involve alternative G protein coupling. In this study, dopamine-induced Gαq activation in rat brain membranes was investigated. Agonist-induced Gαq activation was assessed by increase in guanosine-5′-O-(3-[35S]thio)triphosphate ([35S]GTPγS) binding to Gαq determined by [35S]GTPγS binding/immunoprecipitation assay in rat brain membranes. Dopamine-stimulated Gαq functionality was highest in cortex as compared to hippocampus or striatum. In cerebral cortical membranes, this effect was mimicked by benzazepine derivatives with agonist properties at dopamine D1-like receptors, that is, SKF83959, SKF83822, R(+)-SKF81297, R(+)-SKF38393, and SKF82958, but not by the compounds with dopamine D2-like receptor agonist properties except for aripiprazole. Against expectation, stimulatory effects were also induced by SKF83566, R(+)-SCH23390, and pergolide. The pharmacological profiling by using a series of antagonists indicated that dopamine-induced response was mediated through dopamine D1-like receptor, which was distinct from the receptor involved in 5-HT-induced response (5-HT2A receptor). Conversely, the responses induced by SKF83566, R(+)-SCH23390, and pergolide were most likely mediated by 5-HT2A receptor, but not by dopamine D1-like receptor. Caution should be paid when interpreting the experimental data, especially in behavioral pharmacological research, in which SKF83566 or R(+)-SCH23390 is used as a standard selective dopamine D1-like receptor antagonist. Also, possible clinical implications of the agonistic effects of pergolide on 5-HT2A receptor has been mentioned.  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