首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Propylisopropyl acetamide (PID), an amide analogue of the major antiepileptic drug valproic acid (VPA), possesses favorable anticonvulsant and CNS properties. PID contains one chiral carbon atom and therefore exists in two enantiomeric forms. The purpose of this work was to synthesize the two PID enantiomers and evaluate their enantiospecific teratogenicity. Enantioselective synthesis of PID enantiomers was achieved by coupling valeroyl chloride with optically pure (4S)‐ and (4R)‐benzyl‐2‐oxazolidinone chiral auxiliaries. The two oxazolidinone enolates were alkylated with isopropyl triflate, hydrolyzed, and amidated to yield (2R)‐ and (2S)‐PID. These two PID enantiomers were obtained with excellent enantiomeric purity, exceeding 99.4%. Unlike VPA, both (2R)‐ and (2S)‐PID failed to exert teratogenic effects in NMRI mice following a single 3 mmol/kg subcutaneous injection. From this study we can conclude that individual PID enantiomers do not demonstrate stereoselective teratogenicity in NMRI mice. Due to its better anticonvulsant activity than VPA and lack of teratogenicity, PID (in a stereospecific or racemic form) has the potential to become a new antiepileptic and CNS drug. Chirality 11:645–650, 1999. © 1999 Wiley‐Liss, Inc.  相似文献   

2.
Three 2,3‐dihydro‐1H‐isoindol‐1‐ones structurally related with piracetam (=2‐oxopyrrolidine‐1‐acetamide) have been synthesized and tested for their nootropic effects in the passive avoidance test in mice. Compounds (RS)‐ 2 , (R,R)‐ 3 , and (R,S)‐ 3 were obtained in good yields in only two steps starting from methyl dl ‐phthaloylalanine. Compound (RS)‐ 2 exhibited nootropic activity at lower doses than piracetam, used as reference drug, but it showed lower efficacy. Whereas diastereoisomers (R,R)‐ 3 and (R,S)‐ 3 were as potent as piracetam to revert amnesia induced by scopolamine, (R,S)‐ 3 showed lower efficacy than (R,R)‐ 3 . Only (R,R)‐ 3 showed myorelaxant effect at doses of 10 and 30 mg/kg; other compounds did not exhibit any anticonvulsant, sedative, myorelaxant, or impaired motor‐coordination effect in mice. These synthesized 2,3‐dihydro‐1H‐isoindol‐1‐one derivatives constitute a new kind of nootropic compounds.  相似文献   

3.
Dimerization of a hydroxycarbazole produces an axially chiral biaryl, BICOL ( 2 ). One enantiomer (R)‐ 2 , is capable of enantioselective binding to different polymorphs of DNA. The biaryl (R)‐ 2 was shown by fluorescence and circular dichroism to induce a shift of Z‐DNA to B‐DNA. The opposite enantiomer (S)‐ 2 shows no specific binding. The significant difference in behaviour between the two enantiomers (S)‐ 2 and (R)‐ 2 is in line with molecular modelling studies which show two very different binding geometries between the enantiomers with each polymorph of DNA. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

4.
Racemic RS‐4‐(4‐hydroxyphenyl)‐2‐butanol (rhododendrol, RD) was used as a topical skin‐whitening agent until it was recently reported to induce leukoderma. We then showed that oxidation of RD with mushroom tyrosinase rapidly produces RD‐quinone, which is quickly converted to RD‐cyclic quinone and RD‐hydroxy‐p‐quinone. In this study, we examined whether either or both of the enantiomers of RD can be oxidized by human tyrosinase. Using a chiral HPLC column, racemic RD was resolved optically to R(?)‐RD and S(+)‐RD enantiomers. In the presence of a catalytic amount of l ‐dopa, human tyrosinase, which can oxidize l ‐tyrosine but not d ‐tyrosine, was found to oxidize both R(?)‐ and S(+)‐RD to give RD‐catechol and its oxidation products. S(+)‐RD was more effectively oxidized than l ‐tyrosine, while R(?)‐RD was less effective. These results support the notion that the melanocyte toxicity of RD depends on its tyrosinase‐catalyzed conversion to toxic quinones and the concomitant production of reactive oxygen species.  相似文献   

5.
The acetylcholinesterase inhibition by enantiomers of exo‐ and endo‐2‐norbornyl‐Nn‐butylcarbamates shows high stereoselelectivity. For the acetylcholinesterase inhibitions by (R)‐(+)‐ and (S)‐(?)‐exo‐2‐norbornyl‐Nn‐butylcarbamates, the R‐enantiomer is more potent than the S‐enantiomer. But, for the acetylcholinesterase inhibitions by (R)‐(+)‐ and (S)‐(?)‐endo‐2‐norbornyl‐Nn‐butylcarbamates, the S‐enantiomer is more potent than the R‐enantiomer. Optically pure (R)‐(+)‐exo‐, (S)‐(?)‐exo‐, (R)‐(+)‐endo‐, and (S)‐(?)‐endo‐2‐norbornyl‐Nn‐butylcarbamates are synthesized from condensations of optically pure (R)‐(+)‐exo‐, (S)‐(?)‐exo‐, (R)‐(+)‐endo‐, and (S)‐(?)‐endo‐2‐norborneols with n‐butyl isocyanate, respectively. Optically pure norborneols are obtained from kinetic resolutions of their racemic esters by lipase catalysis in organic solvent. Chirality 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

6.
Optically active 1,4-thiazane-3-carboxylic acid [TCA] was synthesized from cysteine via optical resolution by preferential crystallization. The intermediate (RS)-2-amino-3-[(2-chloroethyl)sulfanyl]propanoic acid hydrochlo-ride [(RS)-ACS?HCl] was found to exist as a conglomerate based on its melting point, solubility and IR spectrum. (RS)-ACS?HCl was optically resolved by preferential crystallization to yield (R)- and (S)-ACS?HCl. (R)- and (S)-ACS?HCl thus obtained were recrystallized from a mixture of hydrochloric acid and 2-propanol, taking account of the solubility of (RS)-ACS?HCl, efficiently yielding both enantiomers in optically pure forms. (R)- and (S)-TCA were then respectively synthesized by the cyclization of (R)- and (S)-ACS?HCl in ethanol in the presence of triethylamine.  相似文献   

7.
The stereoselective metabolism of the enantiomers of fenoxaprop‐ethyl (FE) and its primary chiral metabolite fenoxaprop (FA) in rabbits in vivo and in vitro was studied based on a validated chiral high‐performance liquid chromatography method. The information of in vivo metabolism was obtained by intravenous administration of racemic FE, racemic FA, and optically pure (−)‐(S)‐FE and (+)‐(R)‐FE separately. The results showed that FE degraded very fast to the metabolite FA, which was then metabolized in a stereoselective way in vivo: (−)‐(S)‐FA degraded faster in plasma, heart, lung, liver, kidney, and bile than its antipode. Moreover, a conversion of (−)‐(S)‐FA to (+)‐(R)‐FA in plasma was found after injection of optically pure (−)‐(S)‐ and (+)‐(R)‐FE separately. Either enantiomers were not detected in brain, spleen, muscle, and fat. Plasma concentration–time curves were best described by an open three‐compartment model, and the toxicokinetic parameters of the two enantiomers were significantly different. Different metabolism behaviors were observed in the degradations of FE and FA in the plasma and liver microsomes in vitro, which were helpful for understanding the stereoselective mechanism. This work suggested the stereoselective behaviors of chiral pollutants, and their chiral metabolites in environment should be taken into account for an accurate risk assessment. Chirality, 2011. © 2011 Wiley‐Liss, Inc.  相似文献   

8.
A survey of linalool enantiomers in kiwifruit (Actinidia) flowers was conducted to determine their potential as sources of these valuable floral fragrances, and revealed a wide range of enantiomeric ratios. While flowers of A. polygama and A. chrysantha contained almost exclusively one enantiomer, most species contained significant amounts of both (R) and (S) isomers. In some species enantiomeric ratios of floral linalool differed between genotypes, full siblings, and in one case clones, and ratios changed from year to year as well as diurnally. Enantioselective biosynthesis of the linalool‐derived furanoid and pyranoid linalool oxides was examined in flowers of an A. chrysantha and an A. polygama genotype. The flowers of both species produced almost exclusively (S)‐linalool. A. chrysantha flowers incubated with racd5‐linalool preferentially processed the (S)‐isomer through to the linalool oxides. However, the A. polygama flowers were less discriminatory in their use of rac‐d5‐linalool and processed significant quantities of d5‐(R)‐linalool as well. Chirality, 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

9.
The racemic structures of (RS)-2-amino-3-chloropropanoic acid [(RS)-ACP] and (RS)-2-amino-3-chloropropanoic acid hydrochloride [(RS-ACP·HCl] were examined to obtain (R)- and (S)-ACP via optical resolution by preferential crystallization. The melting point, infrared spectrum, solubility, and ternary solubility diagram suggested that (RS)-ACP·HCl exists as a conglomerate and that (RS)-ACP forms a racemic compound. Optical resolution by preferential crystallization of (RS)-ACP·HCl was successfully achieved to yield (R)- and (S)-ACP·HCl. Optically pure (R)- and (S)-ACP were obtained from the purified (R)-and (S)-ACP·HCl, respectively. © 1996 Wiley-Liss, Inc.  相似文献   

10.
The aim of the present study was to elucidate the differences in the plasma concentration of two enantiomers of donepezil in Chinese patients with Alzheimer's disease (AD) and investigate in vitro stereoselective metabolism and transport. Donepezil enantiomers were separated and determined by LC‐MS/MS using D5‐donepezil as an internal standard on a Sepax Chiralomix SB‐5 column. In vitro stereoselective metabolism and transport of donepezil were investigated in human liver microsomes and MDCKII‐MDR1 cell monolayer. Pre‐dose (Css‐min) plasma concentrations were determined in 52 patients. The mean plasma level of (R)‐donepezil was 14.94 ng/ml and that of (S)‐donepezil was 23.37 ng/ml. One patient's plasma concentration of (R)‐donepezil was higher than (S)‐donepezil and the ratio is 1.51. The mean plasma levels of (S)‐donepezil were found to be higher than those of (R)‐donepezil in 51 patients and the ratio of plasma (R)‐ to (S)‐donepezil varies from 0.34 to 0.85. In the in vitro microsomal system, (R)‐donepezil degraded faster than (S)‐donepezil. Vmax of (R)‐donepezil was significantly higher than (S)‐donepezil. The P‐gp inhibition experiment shown that the Papp of the two enantiomers was higher than 200 and the efflux ratios were 1.11 and 0.99. The results of the P‐gp inhibition identification experiment showed IC50 values of 35.5 and 20.4 μM, respectively, for the two enantiomers. The results indicate that donepezil exhibits stereoselective hepatic metabolism that may explain the differences in the steady‐state plasma concentrations observed. Neither (R)‐ nor (S)‐donepezil was a P‐gp substance and the two enantiomers are highly permeable through the blood–brain barrier. Chirality 25:498–505, 2013. © 2013 Wiley Periodicals, Inc.  相似文献   

11.
A series of enantiomerically pure new fragrances, derived from 2‐ethylhexanol, have been prepared and their olfactory properties evaluated. The key step of the synthesis is cinchona‐alkaloid‐catalyzed desymmetrization of cyclic meso‐anhydrides with (R)‐ and (S)‐2‐ethylhexanol and proceeded in good to excellent diastereoselectivities (92:8–98:2 dr). Enantiomerically pure alcohols were prepared by lipase‐catalyzed kinetic resolution of 2‐ethylhexanol using vinyl laurate as acyl donor. Chirality 2009. © 2009 Wiley‐Liss, Inc.  相似文献   

12.
First, (RS)-2-chloro-3-phenylpropanoic acid [(RS)-CPP] was optically resolved using ethyl (S)-phenylalaninate as a resolving agent, aiming at preparation of optically active 2-methylamino-3-phenylpropanoic acid (MPP). The (R)-CPP obtained as the sodium salt monohydrate was reacted with methylamine to give (S)-2-methylamino-3-phenylpropanoic acid [(S)-MPP]. Next, the optical resolution of (RS)-MPP was also attempted via molecular compound formation with optically active mandelic acid (MAN). The molecular compound of (R)-MPP with (S)-MAN [(R)-MPP (S)-MAN] was obtained as the less soluble diastereomeric compound, while the (S)-MPP (S)-MAN compound was found to be the more soluble one. Recrystallization of (R)-MPP (S)-MAN compound from water, followed by treatment with acetone, gave optically pure (R)-MPP in 79% yield, based on a half amount of the starting (RS)-MPP. The (S)-MPP obtained from (S)-MPP (S)-MAN compound was again subjected to formation of molecular compound with (R)-MAN to give optically pure (S,)-MPP in 66% yield. Chirality 9:386–389, 1997. © 1997 Wiley-Liss, Inc.  相似文献   

13.
Stereoselective metabolism of propranolol side‐chain glucuronidation was studied for two recombinant human uridine diphosphate glucuronosyltransferases (UGTs), UGT1A9 and UGT2B7. The S‐ and R‐propranolol side‐chain glucuronides produced in the incubation mixtures were assayed simultaneously by RP‐HPLC with fluorescent detector. The excitation and emission wavelengths were set at 310 nm and 339 nm, respectively. UGT1A9 prefers catalyzing S‐enantiomer to R‐enantiomer and the intrinsic clearance (CLint) ratios of S‐enantiomer to R‐enantiomer are 3.8 times and 6.5times for racemic propranolol and individual enantiomers, respectively. UGT2B7, however, catalyzes slightly less S‐enantiomer than R‐enantiomer and the CLint ratio of S‐enantiomer to R‐enantiomer is 0.8 times. The high concentration of racemic propranolol (>0.57 mmol/l) and individual enantiomers (>0.69 mmol/l) exhibited substrate inhibition of glucuronidation for UGT2B7, but only the S‐enantiomer (>0.44 mmol/l) in racemic propranolol exhibited substrate inhibition for UGT1A9. The substrate inhibition constants (Ksi) were all similar (P > 0.05). Drug–drug interactions were also found between S‐ and R‐enantiomer glucuronidation metabolisms by UGT1A9 and UGT2B7. Chirality 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

14.
A bioequivalence study for venlafaxine generic formulation was conducted as an open label, balanced, randomized, two‐way crossover, single‐dose study. In this study, a comparison of various pharmacokinetic parameters of venlafaxine hydrochloride 150 mg modified release capsules of Ranbaxy and EFEXOR®‐XR 150 mg capsules of Wyeth, in healthy, adult, male, human subjects under fasting condition was performed to conclude bioequivalence. Venlafaxine and its major active metabolite O‐desmethylvenlafaxine (ODV) are racemates. The “(S)‐(+)” and “(R)‐(−)” enantiomers of venlafaxine and ODV are established as being active. Hence, subject samples were analyzed using nonstereoselective and stereoselective assay methods. Both (S)‐(+) and (R)‐(−) enantiomers of venlafaxine and ODV showed similar absorption and disposition. The 90% confidence intervals for venlafaxine, (R)‐(−)‐venlafaxine as well as (S)‐(+)‐venlafaxine were within acceptance range concluding bioequivalence. The results obtained by stereoselective assay were comparable to the nonstereoselective analysis, as sum of concentrations of (S)‐(+)‐ and (R)‐(−)‐enantiomers of venlafaxine and ODV. The mean (S)‐(+)/(R)‐(−) ratios of the enantiomers of venlafaxine and ODV at various time points were consistent in the study subjects. Therefore, the estimation of venlafaxine and ODV using nonstereoselective assay method is effective in distinguishing formulation differences (if any) in bioequivalence studies in a cost‐effective manner. Chirality, 2011. © 2011 Wiley‐Liss, Inc.  相似文献   

15.
Interactions between naringenin and the cytochrome P450 (CYP) system have been of interest since the first demonstration that grapefruit juice reduced CYP3A activity. The effects of naringenin on other CYP isoforms have been less investigated. In addition, it is well known that interactions with enzymes are often stereospecific, but due to the lack of readily available pure naringenin enantiomers, the enantioselectivity of its effects has not been characterized. We isolated pure naringenin enantiomers by chiral high‐performance liquid chromatography and tested the ability of (R)‐,(S)‐ and rac‐naringenin to inhibit several important drug‐metabolizing CYP isoforms using recombinant enzymes and pooled human liver microsomes. Naringenin was able to inhibit CYP19, CYP2C9, and CYP2C19 with IC50 values below 5 μM. No appreciable inhibition of CYP2B6 or CYP2D6 was observed at concentrations up to 10 μM. Whereas (S)‐naringenin was 2‐fold more potent as an inhibitor of CYP19 and CYP2C19 than (R)‐naringenin, (R)‐naringenin was 2‐fold more potent for CYP2C9 and CYP3A. Chiral flavanones like naringenin are difficult to separate into their enantiomeric forms, but enantioselective effects may be observed that ultimately impact clinical effects. Inhibition of specific drug metabolizing enzymes by naringenin observed in vitro may be exploited to understand pharmacokinetic changes seen in vivo. Chirality, 2011. © 2011 Wiley‐Liss, Inc.  相似文献   

16.
Enantioseletive toxicities of chiral pesticides have become an environmental concern recently. In this study, we evaluated the enantiomeric separation of salithion on a suite of commercial chiral columns and assessed the toxicity of enantiomers toward butyrylcholinesterase and Daphnia magna. Satisfactory separations of salithion enantiomers could be achieved on all tested columns, that is, Chiralcel OD, Chiralcel OJ, and Chiralpak AD column. However, the Chiralpak AD column offered the best separation and was chosen to prepare micro‐scale of pure salithion enantiomers for subsequent bioassays. The first and second enantiomers eluted on the Chiralpak AD column were further confirmed to be (?)‐S‐salithion and (+)‐R‐salithion, respectively. The half inhibition concentrations to butyrylcholinesterase of racemate, (+)‐R‐salithion, and (?)‐S‐salithion were 33.09, 2.92, and 15.60 mg/l, respectively, showing (+)‐R‐enantiomer being about 5.0 times more potent than its (?)‐S‐form. However, the median lethal concentrations (96 h) of racemate, (+)‐R‐salithion, and (?)‐S‐salithion toward D. magna were 3.54, 1.10, and 0.36 μg/l, respectively, suggesting that (?)‐S‐salithion was about 3.0 times more toxic than (+)‐R‐form. Racemic salithion was less toxic than either of the enantiomers in both bioassays, suggesting that antagonistic interactions might occur between the enantiomers during the toxication action. This work reveals that the toxicity of salithion toward butyrylcholinesterase and D. magna is enantioselective, and this factor should be taken into consideration in the environmental risk assessment of salithion. Chirality 2009. © 2009 Wiley‐Liss, Inc.  相似文献   

17.
The female Sprague‐Dawley (SD) and Dark Agouti (DA) rats are considered the animal counterparts of the human extensive and poor metabolizer cytochrome P450 (CYP) 2D6 phenotypes, respectively. The aim of this work was to study possible rat strain differences in the steady‐state pharmacokinetics of the (+)‐(S)‐ and (−)‐(R)‐enantiomers of citalopram and its demethylated metabolites. A chronic drug treatment regimen (15 mg/kg daily) was implemented for 13 days in separate groups of SD (n = 9) and DA (n = 9) rats by using osmotic pumps. The concentrations of citalopram and two major metabolites in serum and two brain regions were analyzed by an enantioselective high‐performance liquid chromatography assay. Higher serum and brain levels of citalopram and demethylcitalopram, but lower levels of didemethylcitalopram, were observed in DA rats when compared with SD rats. The enantiomeric (S/R) concentrations ratios of citalopram were lower in the DA rats when compared with the SD rats (0.53 ± 0.05 vs. 0.80 ± 0.03, P < 0.001), indicating a possibly decreased capacity in the metabolism of the (−)‐(R)‐enantiomer in the DA rats. This study shows that CYP2D deficiency results in steady‐state pharmacokinetic differences of the enantiomers of citalopram and its metabolites. Chirality, 2011. © 2010 Wiley‐Liss, Inc.  相似文献   

18.
Compounds based on the pyrroloquinoxaline system can interact with serotonin 5‐HT3, cannabinoid CB1, and μ‐opioid receptors. Herein, a chiral pool synthesis of diastereomerically and enantiomerically pure bromolactam (S,R,R,R)‐ 14A is presented. Introduction of the cyclohexenyl ring at the N‐atom of (S)‐proline derivatives 8 or methyl (S)‐pyroglutamate ( 12 ) led to the N‐cyclohexenyl substituted pyrrolidine derivatives 4 and 13 , respectively. All attempts to cyclize the (S)‐proline derivatives 4 with a basic pyrrolidine N‐atom via [3 + 2] cycloaddition, aziridination, or bromolactamization failed. Fast aromatization occurred during treatment of cyclohexenamines under halolactamization conditions. In contrast, reaction of a 1:1 mixture of diastereomeric pyroglutamates (S,R)‐ 13bA and (S,S)‐ 13bB with LiOtBu and NBS provided the tricyclic bromolactam (S,R,R,R)‐ 14A with high diastereoselectivity from (S,R)‐ 13bA , but did not transform the diastereomer (S,S)‐ 13bB . The different behavior of the diastereomeric pyroglutamates (S,R)‐ 13bA and (S,S)‐ 13bB is explained by different energetically favored conformations. Chirality 26:793–800, 2014. © 2014 Wiley Periodicals, Inc.  相似文献   

19.
It was shown that racemic (±)‐ 2 [1′‐benzyl‐3‐(3‐fluoropropyl)‐3H‐spiro[[2]benzofuran‐1,4′‐piperidine], WMS‐1813 ] represents a promising positron emission tomography (PET) tracer for the investigation of centrally located σ1 receptors. To study the pharmacological activity of the enantiomers of 2 , a preparative HPLC separation of (R)‐2 and (S)‐2 was performed. The absolute configuration of the enantiomers was determined by CD‐spectroscopy together with theoretical calculations of the CD‐spectrum of a model compound. In receptor binding studies with the radioligand [3H]‐(+)‐pentazocine, (S)‐2 was thrice more potent than its (R)‐configured enantiomer (R)‐2 . The metabolic degradation of the more potent (S)‐enantiomer was considerably slower than the metabolism of (R)‐2 . The structures of the main metabolites of both enantiomers were elucidated by determination of the exact mass using an Orbitrap‐LC‐MS system. These experiments showed a stereoselective biotransformation of the enantiomers of 2 . Chirality, 2011. © 2010 Wiley‐Liss, Inc.  相似文献   

20.
A new chemoenzymatic route is reported to synthesize acebutolol, a selective β1 adrenergic receptor blocking agent in enantiopure (R and S) forms. The enzymatic kinetic resolution strategy was used to synthesize enantiopure intermediates (R)‐ and (S)‐N‐(3‐acetyl‐4‐(3‐chloro‐2‐hydroxypropoxy)phenyl)butyramide from the corresponding racemic alcohols. The results showed that out of eleven commercially available lipase preparations, two enzyme preparations (Lipase A, Candida antarctica, CLEA [CAL CLEA] and Candida rugosa lipase, 62316 [CRL 62316]) act in enantioselective manner. Under optimized conditions the enantiomeric excess of both (R)‐ and (S)‐N‐(3‐acetyl‐4‐(3‐chloro‐2‐hydroxypropoxy)phenyl)butyramide were 99.9 and 96.8%, respectively. N‐alkylation of both the (R) and (S) intermediates with isopropylamine gave enantiomerically pure (R and S)‐ acebutolol with a yield 68 and 72%, respectively. This study suggests a high yielding, easy and environmentally green approach to synthesize enantiopure acebutolol. Chirality 27:382–391, 2015. © 2015 Wiley Periodicals, 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