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 13 毫秒
1.
Xu X  Shen Z  Diao J  Zhang P  Jiang J  Zhou Z 《Chirality》2011,23(6):472-478
We investigated the stereoselective degradation kinetics of fluroxypyr methylheptyl ester (FPMH) in rabbits using a chiral high-performance liquid chromatographic method. In 20% rabbit plasma, the half lives of (+)-FPMH and (-)-FPMH were 2.5 and 10.9 min, respectively. Thus, the enantioselective degradation was faster for (+)-FPMH than for (-)-FPMH in rabbit plasma in vitro, and there was no chiral conversion or transformation during incubation of the plasma. The degradation of (+)-FPMH was also much faster than that of the (-)-FPMH in the kidney, lung, and muscle after the intravenous administration of 50 mg/kg racemic FPMH (rac-FPMH), whereas the concentrations of FPMH were below the limit of quantification in other tissues. Furthermore, 98% rac-FPMH was quickly (within 10 min) hydrolyzed to fluroxypyr (FP) in rabbit liver microsomes. Therefore, we examined FP in rabbit plasma and tissues in vivo. We detected FP in all tissues; its concentration was higher in the urine than in the other tissues. FP was rapidly excreted unchanged, principally in the urine. The data presented here are important for a more thorough understanding of this pesticide and should be useful for its full environmental assessment.  相似文献   

2.
The present study was an attempt to elucidate the relationship between stereoselective pharmacokinetics and protein binding of KE-298 and its active metabolites, deacetyl-KE-298 (M-1) and S-methyl-KE-298 (M-2). Metabolic chiral inversion was also investigated. The levels of unchanged KE-298 in plasma after oral administration of (+)-(S)-KE-298 to rats were lower than those of (−)-(R)-KE-298, whereas the levels of M-1 and M-2 after administration of (+)-(S)-KE-298 were higher than after (−)-(R)-KE-298. In vitro, rat plasma protein binding of (+)-(S)-KE-298 was lower than that of (−)-(R)-KE-298. In contrast, the binding of (+)-(S)-M-1 and (+)-(S)-M-2 was higher than that of (−)-(R)-M-1 and (−)-(R)-M-2. Displacement studies revealed that the (+)-(S) and (−)-(R)-enantiomers of KE-298 and their metabolites bound to the warfarin binding site on rat serum albumin. These results suggest that the stereoselective plasma levels in KE-298 and its metabolites were closely related to enantiomeric differences in protein binding, attributed to quantitative differences in binding to albumin rather than to the different binding sites. Unidirectional chiral inversion was detected after oral administration of either (−)-(R)-KE-298 or (−)-(R)-M-2 to rats both yielding (+)-(S)-M-2. Chirality 9:22–28, 1997 © 1997 Wiley-Liss, Inc.  相似文献   

3.
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   

4.
The drug chirality is attracting increasing attention because of different biological activities, metabolic pathways, and toxicities of chiral enantiomers. The chiral separation has been a great challenge. Optimized high‐performance liquid chromatography (HPLC) methods based on vancomycin chiral stationary phase (CSP) were developed for the enantioseparation of propranolol, atenolol, metoprolol, venlafaxine, fluoxetine, and amlodipine. The retention and enantioseparation properties of these analytes were investigated in the variety of mobile phase additives, flow rate, and column temperature. As a result, the optimal chromatographic condition was achieved using methanol as a main mobile phase with triethylamine (TEA) and glacial acetic acid (HOAc) added as modifiers in a volume ratio of 0.01% at a flow rate of 0.3 mL/minute and at a column temperature of 5°C. The thermodynamic parameters (eg, ΔH, ΔΔH, and ΔΔS) from linear van 't Hoff plots revealed that the retention of investigated pharmaceuticals on vancomycin CSP was an exothermic process. The nonlinear behavior of lnk′ against 1/T for propranolol, atenolol, and metoprolol suggested the presence of multiple binding mechanisms for these analytes on CSP with variation of temperature. The simulated interaction processes between vancomycin and pharmaceutical enantiomers using molecular docking technique and binding energy calculations indicated that the calculated magnitudes of steady combination energy (ΔG) coincided with experimental elution order for most of these enantiomers.  相似文献   

5.
Stereoselective degradation kinetics of tebuconazole in rabbits   总被引:2,自引:0,他引:2  
Zhu W  Qiu J  Dang Z  Lv C  Jia G  Li L  Zhou Z 《Chirality》2007,19(2):141-147
Tebuconazole[(RS)-1-p-chlorophenyl-4,4-dimethyl-3-(1H-1,2,4-triazol-1-ylmethyl)pentan-3-ol] is a potent triazole fungicide and consists of a pair of enantiomers. The enantioselective degradation kinetics of tebuconazole was investigated in rabbits by intravenous (iv) injection. The concentrations of (-)-(R)-tebuconazole and (+)-(S)-tebuconazole in plasma and tissues were determined by HPLC with a cellulose tris(3,5-dimethylphenylcarbamate)-based chiral stationary phase. Enantioselective analysis methods for this fungicide in plasma and tissues were developed and validated. Good linearities were obtained over the concentration range of 0.25-25 mg/l for both enantiomers. The degradation followed pseudo-first-order kinetics and the degradation of the (+)-(S)-tebuconazole was much faster than that of the (-)-(R)-tebuconazole in plasma after administration of racemic tebuconazole. This study also indicated that environmental assessment of enantiomeric degradation may be needed to fully evaluate risks of tebuconazole use.  相似文献   

6.
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   

7.
The chromatographic chiral resolution of racemic methyl 1,4‐dihydro‐2,6‐dimethyl‐5‐nitro‐4‐benzofurazanyl‐3‐carboxylates 1 and 2 and their benzofuroxanyl analogues 3 and 4 were studied on Whelk‐O1, Chiralcel OD‐H, Chiralcel OJ, and Chiralpak AD and AS. These CSPs were selected on the basis of the results of structural searches in Chirbase. Examination of the data and cluster analysis pointed out the influence of benzofurazane–benzofuroxane change versus α–β connection change on retention and enantioselectivity, respectively. The major contribution to the retention change arose from the type of heterocycle, whereas the major contribution to the enantioselectivity change came from the mode of connection (α or β) almost irrespective of the nature of the heterocycle. It resulted in a similarity of behaviour between 1 and 2 on one hand and 3 and 4 on the other as far as capacity factors were concerned, and in a similarity of behaviour between 1 and 3 on the one hand and 2 and 4 on the other as far as enantioselectivities were concerned. Chiralpak AS was selected for semipreparative resolution of the enantiomers. The study of several CSPs allowed us to obtain correlations of structure with retention and enantioselectivity as well as the choice of a semipreparative support to provide the quantities for biological tests. Chirality 11:602–608, 1999. © 1999 Wiley‐Liss, Inc.  相似文献   

8.
The stereoselectivity of muscarinic receptors was quantified by the Free‐Wilson method. The analysis was performed with the affinity data, published by Gualtieri for certain muscarinic ligands (1,3‐oxathiolanes) in three different tissues. The 2 and 3 positions in the 1,3‐oxathiolane ring were important to differentiate between the receptors in those tissues. The results also showed that the three chiral centers in the 1,3‐oxathiolanes displayed different stereoselective discrimination. The results of the Free‐Wilson analysis were compared with those of a previous study on the same compounds obtained by epimeric eudismic analysis. The agreements and differences between the results obtained by these two methods are discussed in depth. Chirality 11:139–143, 1999. © 1999 Wiley‐Liss, Inc.  相似文献   

9.
Tetrahydropalmatine (THP) is one of the active alkaloid ingredients of Rhizoma Corydalis. THP has a chiral center, and the stereoselective pharmacokinetics and tissue distribution have been reported. The aim of the present article is to study the stereoselective protein binding of THP using equilibrium dialysis followed by HPLC‐UV analysis. The results showed that THP stereoselectively binds to human serum albumin (HSA), α1‐acid glycoprotein (AGP), and proteins in human plasma. The fraction binding of (+)‐THP was significantly higher than that of (?)‐THP, whereas such stereoselectivity was not found in rat plasma. The affinity of HSA and AGP to (+)‐THP, expressed as nKA, were 9.0 × 103 M?1 and 2.34 × 105 M?1, respectively, which were notablely higher than to (?)‐THP, with the nKA of 3.4 × 103 M?1 and 1.44 × 105 M?1, respectively. The binding site of HSA for (?)‐THP was Site I, whereas for (+)‐THP was both Site I and Site II. The F1/S variants of AGP were proved to be the key variants (?)‐ and (+)‐THP binding to both. Finally, the AGP binding drugs, such as mifepristone, were demonstrated to reduce the fraction binding of (?)‐ and (+)‐THP with pure AGP (1 mg/ml) but did not affect the fraction binding of both (?)‐ and (+)‐THP with proteins in human plasma. It can be concluded that protein binding of THP is species dependent and stereoselective, both HSA and AGP contribute to the stereoselective binding to THP enatiomers, and AGP binding drugs may not cause the drug–drug interaction on THP in healthy human plasma. Chirality, 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

10.
Flecainide, an antiarrythmic agent, and its analogs were resolved on a high performance liquid chromatographic chiral stationary phase (CSP) based on (+)‐(18‐crown‐6)‐2,3,11,12‐tetracarboxylic acid with the use of a mobile phase consisting of methanol‐acetonitrile‐trifluoroacetic acid‐triethylamine (80/20/0.1/0.3, v/v/v/v). The chiral resolution was quite successful, the separation factors (α) and the resolutions (RS) for 20 analytes including flecainide being in the range of 1.19–1.82 and 1.73–6.80, respectively. The ortho‐substituent of the benzoyl group of analytes was found to cause decrease in the retention times of analytes probably because of the conformational deformation of analytes originated from the steric hindrance exerted by the ortho‐substituent. Chirality, 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

11.
Mono‐N‐ethylated α‐amino acid esters are obtained in high yields using reductive amination procedures. Formation of imine is achieved by excess of acetaldehyde, followed by removal of acetaldehyde and reduction by NaBH(OAc)3. The elaborated one‐pot synthesis allows for the efficient synthesis of side‐chain protected amino acid derivatives. Copyright © 1999 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

12.
Wang SH  Yin LN  Liang ZH  Lu SJ  Zeng S 《Chirality》2007,19(10):769-774
The stereoselectivity of release of ketoprofen (KET) enantiomers from a biodegradable injectable implant containing racemic KET (rac-KET) was investigated in vivo. A pre-column chiral derivatization RP-HPLC method was employed to assay diastereoisomeric derivatives of R- and S-KET. The rac-KET injectable implant, once injected subcutaneously in rats, produced long-lasting plasma levels of S-KET, which were always greater than those of R-KET. The difference in enantiomer concentration was to be related to stereoselective release, due to stereoselective interaction between D,L-PLG in the implant and KET enantiomers, as well as the chiral inversion of KET in vivo. The rac-KET injectable implant provided the sustained release of S-KET with effective plasma levels maintained for about 8 wk after a single injection.  相似文献   

13.
14.
Metabolic chiral inversion of ibuprofen in isolated rat hepatocytes   总被引:1,自引:0,他引:1  
Ibuprofen was used to demonstrate that isolated rat hepatocytes offer a suitable in vitro model to investigate the metabolic chiral inversion of anti-inflammatory 2-arylpropionic acids (profens). The inversion of the pharmacologically inactive (-)-(R)-ibuprofen to the active (+)-(S)-ibuprofen was shown to obey apparent first-order kinetics during 5 h and to increase linearly with increasing hepatocyte concentration up to 4 x 10(5) cells/ml. No elimination of (R)-ibuprofen by routes other than inversion was seen, whereas the elimination of (S)-ibuprofen appeared to be saturable.  相似文献   

15.
Metal–organic frameworks (MOFs) are excellent porous materials with nanoscale cavities and high surface areas, which make them promising as novel adsorbents in solid‐phase extraction (SPE). In this article we report a new application of the chiral MOF [Zn2(D‐Cam)2(4,4′‐bpy)]n in SPE used for the measurement of the enantiomeric excess (ee) of (±)‐1,1′‐bi‐2‐naphthol. Several important experimental parameters that may influence the extraction efficiency were investigated and optimized. Under the optimum conditions, a good linearity (R2 > 0.999) was found between the ee value and the reciprocal of the peak areas. When compared with the actual ee measured using chiral HPLC, the SPE‐based assay also showed good accuracy and precision. The results showed that SPE based on chiral MOFs as adsorbents is a simple and effective method for the determination of the ee values of chiral compounds.  相似文献   

16.
Racemic cyclohexylaminoglutethimide (±ChAG) and its acetylated metabolite (±ChAG) were resolved by a direct chromatographic method using a Chiracel OD column without derivatization. Maximum resolutions (R) of 4.89 and 0.74 were obtained for the enantiomers of cyclohexylaminoglutethimide and its acetylated metabolite, respectively.  相似文献   

17.
Work from this paper details a novel walk‐up open‐access (OA) approach to enable chiral analytical method development and preparative separation of enantiomers in early discovery chemistry using supercritical fluid chromatography (SFC). We have demonstrated the success of this OA approach using immobilized chiral stationary phases (CSPs). After screening a diverse set of racemic drug candidates, we have concluded that a simplified OA chiral SFC platform can successfully purify approximately 60% of the analysed racemates. This streamlined OA workflow enables medicinal chemists with limited expertise in chiral method development to successfully and rapidly purify enantiomers for their projects using Waters UPC2 and Prep100‐SFC instrumentation.  相似文献   

18.
The title compound of 3‐benzothiazole‐9‐ethyl carbazole was synthesized by the reaction of 3‐aldehyde‐9‐ethyl carbazole and 2‐aminothiophenol. The compound was characterized by 1H nuclear magnetic resonance (NMR) and mass spectrometry (MS). Its crystal structure was obtained and determined by single crystal X‐ray diffraction. The results showed that the crystal belongs to the orthorhombic crystal system and the cell parameters of space group P2(1)2(1)2(1) were a = 5.6626 (12) Å, b = 12.606 (3) Å, c = 22.639 (5) Å, α = 90°, β = 90°, γ = 90°, V = 1616.0 (6) Å3, Z = 4, Dc = 1.350 mg/m3. The UV–vis and fluorescence spectra were also studied preliminarily. The fluorescence spectra of the title compound with bovine serum albumin (BSA) showed that BSA could be marked with the compound and the stability constant between them was 0.82 × 107 M?1. Meanwhile, the crystal and molecule were theoretically surveyed by density functional tight‐binding (DFTB). The results showed that there was an orbital overlap for lowest unoccupied molecular orbital (LUMO) between the neighbouring molecules for the crystal, which is different from the molecule structure. It was also showed that the crystal structure is a non‐conductor. Copyright © 2016 John Wiley & Sons, Ltd.  相似文献   

19.
A novel chiral sensor based on the self‐assembled monolayer of (6A‐ω‐mercaptoethylureado‐6A‐deoxy)heptakis(2,3‐di‐o‐phenylcarbamoyl)‐6B, 6C, 6D, 6E, 6F, 6G‐ hexa‐o‐phenylcarbamoyl‐β‐cyclodextrin (Ph‐β‐CD‐SH) on a quartz crystal transducer for chiral recognition was set up. (R,S)‐(±)‐(3‐Methoxyphenyl)ethylamine were recognized by this QCM chiral sensor with a QCM chiral discrimination factor of 1.33. Furthermore, UV spectroscopy was used to investigate the mechanism of host‐guest interactions between (6A‐azido‐6A‐deoxy)heptakis(2,3‐di‐o‐phenylcarbamoyl)‐6B, 6C, 6D, 6E, 6F, 6G‐hexa‐o‐phenylcarbamoyl‐β‐cyclodextrin (Ph‐β‐CD) and (R,S)‐(±)‐(3‐methoxyphenyl) ethylamine. The UV discrimination factor was determined to be 0.066. Chirality, 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

20.
The two enantiomers of ethyl 3‐hydroxybutyrate are important intermediates for the synthesis of a great variety of valuable chiral drugs. The preparation of chiral drug intermediates through kinetic resolution reactions catalyzed by esterases/lipases has been demonstrated to be an efficient and environmentally friendly method. We previously functionally characterized microbial esterase PHE21 and used PHE21 as a biocatalyst to generate optically pure ethyl (S)‐3‐hydroxybutyrate. Herein, we also functionally characterized one novel salt‐tolerant microbial esterase WDEst17 from the genome of Dactylosporangium aurantiacum subsp. Hamdenensis NRRL 18085. Esterase WDEst17 was further developed as an efficient biocatalyst to generate (R)‐3‐hydroxybutyrate, an important chiral drug intermediate, with the enantiomeric excess being 99% and the conversion rate being 65.05%, respectively, after process optimization. Notably, the enantio‐selectivity of esterase WDEst17 was opposite than that of esterase PHE21. The identification of esterases WDEst17 and PHE21 through genome mining of microorganisms provides useful biocatalysts for the preparation of valuable chiral drug intermediates.  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