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.
The alkylaminoalkylnaphthalene 3 shows interesting opioid-like analgesic properties, μ-selective ligand competition, and enkephalin hydrolyzing enzyme inhibition. 3 possesses two chiral centers and can exist as two racemic pairs and four diastereomers. Since the binding of opioids with the receptor is stereoselective, it was important to have the two racemic pairs as well as the four diastereomers. In this paper the synthesis of the (1R,2R/1S,2S)- and (1R,2S/1S,2R)-racemates and the (1R,2R)- and (1S,2S)-enantiomers of the 1-ethyl-1-hydroxy-1-[2-(6-hydroxynaphthyl)]-2-methyl-3-dimethylaminopropane 3 is considered and the determination of absolute configuration is described. The (1R,2R/1S,2S)- 3 and (1R,2S/1S,2R)- 3 racemates and the (1R,2R)- 3 and (1S,2S)- 3 enantiomers were prepared by reaction of the racemic and optically active 1-dimethylamino-2-methylpentan-3-one 2 , respectively, with the lithiation product obtained from 2-bromo-6-tetrahydropyranyloxynaphthalene and acidic hydrolysis. The optical resolution of aminoketone 2 was carried out via fractional crystallization of salts (+)- and (?)-dibenzoyltartrates. The configuration of the optically active compounds was determined by X-ray analysis of a crystal of (+)-(1R,2R)- 3 · HCl · H2O. Preliminary pharmachological tests showed that (+)-(1R,2R)- 3 enantiomer is able to induce opioid-like analgesia with a relative potency 2.5 times that of (1R,2R/1S,2S)- 3 and about 4 times that of morphine. © 1994 Wiley-Liss, Inc.  相似文献   

2.
The (S,S,S,S) and (R,R,R,R) enantiomers of tetramethyl‐bis(ethylenedithio)‐tetrathiafulvalene (TM‐BEDT‐TTF) show equatorial conformation for the four methyl groups in the solid state, according to the single‐crystal X‐ray analyses. Theoretical calculations at the Density Functional Theory (DFT) and time‐dependent (TD) DFT levels indicate higher gas phase stability for the axial conformer than the equatorial one by 1.25 kcal · mole‐1 and allow the assignment of the UV–vis and circular dichroism transitions. A complete series of radical cation salts of 1:1 stoichiometry with the triiodide anion I3 was obtained by electrocrystallization of both enantiopure and racemic forms of the donor. In the packing the donors are organized in dimers that further interact through S · · · S intermolecular contacts and the triiodide anions lie parallel to pairs of oxidized donors. The conductivity of the racemate, which adopts the same, but disordered, structural type, is considerably lower, with much higher activation energy. Chirality 25:466–474, 2013.© 2013 Wiley Periodicals, Inc.  相似文献   

3.
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   

4.
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   

5.
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   

6.
To determine the stereospecific pharmacokinetics and gastrointestinal permeability (GI) changes (surrogate measures of toxicity) in the rat following oral administration of S, R, and racemic ketorolac (KT), optically pure enantiomers (S and R 2.5 mg/kg), and racemic KT (5 mg/kg) were administered orally to male Sprague-Dawley rats and plasma samples were collected for 6 h post-dose for pharmacokinetic assessments. KT-induced changes in GI permeability were assessed using sucrose and 51Cr-EDTA as markers of gastroduodenal and distal intestinal permeability, respectively. After the racemate, R-KT was predominant in plasma (AUC S/R, 0.45). No significant differences in pharmacokinetic indices were evident following administration of the racemate as compared with individual enantiomers. In plasma, there was only negligible S-KT after administration of R-KT. After S-KT, on the other hand, AUC of R-KT was found to be 6.7% of that of S-KT. Both permeability markers showed considerable interanimal variability. Gastroduodenal permeability was significantly increased from baseline by the racemate but not by either of the two enantiomers administered alone. Permeability to 51Cr-EDTA was not significantly increased above baseline for any of the treatments. The plasma concentration of R-KT found after administration of S-KT may be from the < 2% chiral impurity which appears magnified due to its slower clearance as compared with its antipode. There is no evidence of a pharmacokinetic interaction between the enantiomers. Since 2.5 mg/kg S-KT is somewhat less toxic on the gastroduodenum than 5 mg/kg racemate, it may be a safer alternative to the latter, at least in the rat model.  相似文献   

7.
The steady-state pharmacokinetics in serum and urine of the enantiomers of citalopram and its metabolites, demethylcitalopram (DCT) and didemethylcitalopram (DDCT), were investigated after multiple doses of rac-citalopram for 21 consecutive days (40 mg per day) to healthy human subjects who were extensive metabolisers of sparteine and mephenytoin. Comparable pharmacokinetic variability was noted for (+)-(S)-, (−)-(R)- and rac-citalopram. Enantiomeric (S/R) serum concentration ratios for citalopram were always less than unity and were constant during the steady-state dosing interval. A modest, but statistically significant, stereoselectivity in the disposition of citalopram and its two main metabolites was observed. Serum levels of the (+)-(S)-enantiomers of citalopram, DCT, and DDCT throughout the steady-state dosing interval investigated were 37 ± 6%, 42 ± 3% and 32 ± 3%, respectively, of their total racemic serum concentrations. The (+)-(S)-enantiomers of citalopram, DCT, and DDCT were eliminated faster than their antipodes. For (−)-(R)- and (+)-(S)-citalopram, respectively, the serum t½ averaged 47 ± 11 and 35 ± 4 h and AUCss averaged 4,193 ± 1,118 h · nmol/l and 2,562 ± 1,190 h · nmol/l. The observed enantiospecificities were apparently more related to clearance, rather than to distributional mechanisms. Chirality 9:686–692, 1997. © 1997 Wiley-Liss, Inc.  相似文献   

8.
Chiral high‐performance liquid chromatography (HPLC) separation of trans‐bis[2‐(2‐pyridyl)aminophenolato] dichlorocyclotriphosphazene 1 was achieved and the absolute configuration of (+)-1 was assigned to be S,S by single‐crystal X‐ray structural analysis. The optically pure 1,2‐diphenyl‐1,2‐ethanediolate derivatives (+)‐ 2a and (?)‐ 2b were synthesized by the reactions of (+)-1 and (-)-1 with (R,R)‐hydrobenzoin, respectively, in refluxing toluene in the presence of an excess amount of triethylamine and a catalytic amount of 4‐(dimethylamino)pyridine. The racemization of the enantiomers of 1 and the epimerization of diastereomers of 2 were not observed in refluxing toluene neither under acidic nor basic conditions. The stereochemistry of (+)-1 was confirmed by the crystal structure of (+)‐ 2a and bis[(4‐methyl‐2‐pyridyl)oxy]cyclotriphosphazene (+)-3 derived from (+)-1 . Chirality 28:556–561, 2016. © 2016 Wiley Periodicals, Inc.  相似文献   

9.
The phenomenon of enantioselectivity in the metabolism of mexiletine (MEX) conjugation was investigated in eight female patients with the arrhythmic form of chronic Chagas' heart disease treated with racemic mexiletine hydrochloride (two 100 mg capsules every 8 hr). Blood samples were collected up to 24 hr after the administration of the morning dose, with discontinuation of the subsequent doses during the study period. Plasma concentrations of N‐hydroxymexiletine glucuronide were calculated as the difference between the concentrations of unchanged and total (unchanged + conjugated) MEX enantiomers. Total plasma MEX concentrations were analyzed by HPLC after enzymatic hydrolysis with β‐glucuronidase, the formation of diastereomeric derivatives with the chiral reagent N‐acetyl‐l ‐cysteine/o‐phthalaldehyde, and fluorescence detection. The differences in the pharmacokinetic parameters of the enantiomers were evaluated by the paired t‐test. The plasma concentrations of the (+)‐(S)‐MEX did not differ before and after enzymatic hydrolysis. The pharmacokinetic parameters calculated for (−)‐(R)‐N‐hydroxymexiletine glucuronide are presented as means (95% confidence interval): maximum plasma concentration Cmax = 194.0 ng · ml−1 (154.3–233.7), time to maximum plasma concentration tmax = 1.4 hr (0.3–2.5), area under the plasma concentration versus time curve AUC0–24 = 2099.2 ng · h · ml−1 (1585.6–2612.6), elimination half‐life t1/2β = 12.8 hr (9.9–15.6) and extent of conjugation of 31.6% (24.3–38.9%). The present data indicate stereospecific conjugation of (−)‐(R)‐N‐hydroxymexiletine in the female patients with the arrhythmic form of Chagas' heart disease. Chirality 11:29–32, 1999. © 1999 Wiley‐Liss, Inc.  相似文献   

10.
ML-1035, 4-amino-5-chloro-2-[2-(methylsulfinyl)ethoxy]-N-[2-(diethylamino)ethyl]benzamide, is a sulfoxide compound and a racemic gastroprokinetic agent with a chiral center at the sulfur atom. We have investigated the disposition kinetics of (R)-ML-1035 sulfoxide (R) and (S)-ML-1035 sulfoxide (S) after the single enantiomers and the racemic mixture were administered to rats in separate experiments. There was no noticeable chiral inversion after either enantiomer dose. Both enantiomers were rapidly absorbed. After dosing with enantiomers or with the racemate, the resulting plasma concentration-time curve of R was closely parallel to that of S in both intravenous and oral experiments, suggesting that the two enantiomers have approximately the same disposition kinetics. After intravenous enantiomer doses, only S underwent conversion to sulfide, suggesting that sulfidation in the liver is enantioselective. However, the enantioselective sulfidation after intravenous dosing did not introduce a difference in the global plasma disposition profiles between R and S, since the reduction reaction is a minor metabolic process. Other metabolic reactions such as sulfonation and mono-N-desethylations were not enantioselective. After oral administration, conversion to sulfide was observed for both enantioners, implicating the existence of a nonhepatic pathway in sulfidation. Administration of a prochiral sulfide dose was associated with an enantioselective sulfoxidation, in which the R/S concentration ratios increased as a function of time. In addition, enantiomeric interaction causing changes in pharmacokinetic parameters was observed after the oral racemate dose, while the interaction is negligible after an intravenous racemate dose, indicating a route dependency in enantiomeric interaction. © 1993 Wiley-Liss, Inc.  相似文献   

11.
《Chirality》2017,29(9):522-535
New enantiopure dimethyl‐substituted acridino‐18‐crown‐6 and acridino‐21‐crown‐7 ethers containing a carboxyl group at position 9 of the acridine ring [(S,S )‐ 8 , (S,S )‐ 9 , (R,R )‐ 10 ] were synthesized. The pK a values of the new crown ethers [(S,S )‐ 8 , (S,S )‐ 9 , (R,R )‐ 10 ] and of an earlier reported macrocycle [(R,R )‐ 2 ] were determined by UV‐pH titrations. Crown ether (S,S )‐ 8 was attached to silica gel by covalent bonds and the enantiomeric separation ability of the newly prepared chiral stationary phase [(S,S )‐CSP‐ 12 ] was studied by high‐performance liquid chromatography (HPLC). Homochiral preference was observed and the best separation was achieved for the enantiomers of 1‐NEA. Ligands (S,S )‐ 9 and (R,R )‐ 10 are precursors of enantioselective sensor and selector molecules for the enantiomers of protonated primary amines, amino acids, and their derivatives.  相似文献   

12.
Racemic 2-(5,6-dichloro-3-indolyl)propionic acid (5,6-Cl2-2-IPA) was synthesized from 5,6-dichloroindole-3-acetic acid (5,6-Cl2-IAA) by successive esterification, methoxycarbonylation, methylation, and double hydrolysis. The racemate was converted to the diastereomeric esters of (S)-(-)-1-phenylethyl alcohol. These were separated by HPLC into two optically active diastereomers and then hydrolyzed with p-TsOH to the optically active enantiomers of 5,6-Cl2-2-IPA. The absolute configurations of both the 5,6-Cl2-2-IPA enantiomers were determined by comparing the 1H-NMR spectra of their diastereomeric (S)-(-)-1-phenylethyl esters with those of the diastereomeric (S)-(-)-1-phenylethyl esters of 2-(3-indolyl)propionic acid (2-IPA) whose absolute configurations are already known. There was no essential difference between (S)-(+)- and (R)-(-)-5,6-Cl2-2-IPA in hypocotyl growth-inhibiting activity toward Chinese cabbage, but their inhibitory activities were stronger than that of the potent mother auxin, 5,6-Cl2-IAA. No essential difference in the coleoptile elongating activity of Avena sativa was apparent for the enantiomers, this activity being about one-third that of 5,6-Cl2-IAA.  相似文献   

13.
Racemic 2-(5,7-dichloro-3-indolyl)propionic acid (5,7-Cl(2)-2-IPA) was synthesized from 5,7-dichloroindole-3-acetic acid by successive esterification, methoxycarbonylation, methylation, and double hydrolysis. The racemate was converted to diastereomeric esters of l-menthol; these were separated by recycling HPLC into two optically active diastereomers that were then hydrolyzed with p-TsOH to two optically active enantiomers of 5,7-Cl(2)-2-IPA. The absolute configurations of both these enantiomers were determined by comparing the (1)H-NMR spectra of their diastereomeric l-menthyl esters with those of the diastereomeric l-menthyl esters of 2-(3-indolyl)propionic acid (2-IPA) of known absolute configurations.An assay by the coleoptile elongation of Avena sativa showed the (S)-(+)-enantiomer of 5,7-Cl(2)-2-IPA to have weak auxin activity, whereas the (R)-(-)-antipode had no auxin activity at any concentration tested. Interestingly, the (R)-(-)-enantiomer had antiauxin activity very close to that of 2-(5,7-dichloro-3-indolyl)isobutyric acid (5,7-Cl(2)-IIBA), a strong antiauxin. These data indicate that, of the two methyl groups in its molecule, the antiauxin activity of 5,7-Cl(2)-IIBA was due only to the (R)-methyl group.  相似文献   

14.
The binding of the (R)‐ and (S)‐enantiomers of amlodipine to bovine serum albumin (BSA), human serum albumin (HSA), α1‐acid glycoprotein (AGP), and human plasma (HP) was studied by equilibrium dialysis over the concentration range of 75–200 μM at a protein concentration of 150 μM. Unbound drug concentrations were determined by enantioselective capillary electrophoresis using 50 mM phosphate buffer, pH 2.5, containing 18 mM α‐cyclodextrin as background electrolyte. Saturation of the protein binding sites was not observed over the concentration range tested. Upon application of racemic amlodipine besylate, (S)‐amlodipine was bound to a higher extend by HSA and HP compared with (R)‐amlodipine, whereas the opposite binding of the enantiomers was observed for BSA and AGP. Scatchard analysis was used to illustrate the different binding affinities of amlodipine besylate enantiomers to BSA, HSA and AGP. Chirality, 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

15.
Evidence is accumulating to suggest that 3,4‐methylenedioxymethamphetamine (MDMA) has neurotoxic and neuroinflammatory properties. MDMA is composed of two enantiomers with different biological activities. In this study, we evaluated the in vivo effects of S(+)‐MDMA, R(?)‐MDMA, and S(+)‐MDMA in combination with R(?)‐MDMA on microglial and astroglial activation compared with racemic MDMA, by assessment of complement type 3 receptor (CD11b) and glial fibrillary acidic protein (GFAP) immunoreactivity in the mouse striatum, nucleus accumbens, motor cortex, and substantia nigra. Motor activity and body temperature were also measured, to elucidate the physiological modifications paired with the observed glial changes. Similar to racemic MDMA (4 × 20 mg/kg), S(+)‐MDMA (4 × 10 mg/kg) increased both CD11b and GFAP in the striatum, although to a lower degree, whereas R(?)‐MDMA (4 × 10 mg/kg) did not induce any significant glial activation. Combined administration of S(+) plus R(?)‐MDMA did not induce any further activation compared with S(+)‐MDMA. In all other areas, only racemic MDMA was able to slightly activate the microglia, but not the astroglia, whereas enantiomers had no effect, either alone or in combination. Racemic MDMA and S(+)‐MDMA similarly increased motor activity and raised body temperature, whereas R(?)‐MDMA affected neither body temperature nor motor activity. Interestingly, the increase in body temperature was correlated with glial activation. The results show that no synergism, but only additivity of effects, is caused by the combined administration of S(+)‐ and R(?)‐MDMA, and underline the importance of investigating the biochemical and behavioral properties of the two MDMA enantiomers to understand their relative contribution to the neuroinflammatory and neurotoxic effects of MDMA.  相似文献   

16.
The pharmacokinetics of diniconazole enantiomers in quail (Coturnix coturnix japonica) were investigated by liquid chromatography‐tandem mass spectrometry (LC‐MS/MS). Quails were exposed to racemic diniconazole in capsule by oral at dose of 10 mg/kg (body weight). The maximal concentrations observed in blood, heart, liver, and kidney were 3.18, 11.35, 12.32, 15.03 µg/g for S‐diniconazole, and 1.13, 3.70, 6.00, 2.60 µg/g for R‐diniconazole. The elimination of enantiomers all met the one‐compartment model in blood, heart, liver, and kidney well. The elimination half‐lives (T1/2) of S‐diniconazole were 2.87, 3.85, 5.29, and 4.42 h in blood, heart, liver, and kidney, respectively; the T1/2 of R‐diniconazole were 2.44, 3.42, 146.23, and 74.02 h in blood, heart, liver, and kidney, respectively. The enantiomer fractions (EFs) steadily increased from 0.50 to 0.92 in blood samples and 0.91 in heart samples. Meanwhile, the values increased to 0.70 and 0.80 in liver and kidney initially, and then decreased to 0.33 and 0.44 at the end of the experiment. Metabolism was examined as well and it was found that diniconazole was metabolized to 1, 2, 4‐triazole, (E)‐3‐(1H‐1, 2, 4‐triazol‐1‐yl) acrylaldehyde, (E, S)‐(R, S)‐4‐(2, 4‐dichlorophenyl)‐2, 2‐dimethyl‐5‐(1H‐1, 2, 4‐triazol‐1‐yl) pent‐4‐ene‐1, 3‐diol, (E)‐4‐(2, 4‐dichlorophenyl)‐3‐hydroxy‐2, 2‐dimethyl‐5‐(1H‐1, 2, 4‐triazol‐1‐yl) pent‐4‐enoic acid, and 1, 3‐dichlorobenzen in all samples of quail. Chirality 25:910–916, 2013. © 2013 Wiley Periodicals, Inc.  相似文献   

17.
A new racemic pyrazoline derivative was synthesized and resolved to its enantiomers using analytic and semipreparative high‐pressure liquid chromatography. The absolute configuration of both fractions was established using vibrational circular dichroism. The in vitro monoamine oxidase (MAO) inhibitory profiles were evaluated for the racemate and both enantiomers separately for the two isoforms of the enzyme. The racemic compound and both enantiomers were found to inhibit hMAO‐A selectively and competitively. In particular, the R enantiomer was detected as an exceptionally potent and a selective MAO‐A inhibitor (Ki = 0.85 × 10?3 ± 0.05 × 10?3 μM and SI: 2.35 × 10?5), whereas S was determined as poorer compound than R in terms of Ki and SI (0.184 ± 0.007 and 0.001). The selectivity of the enantiomers was explained by molecular modeling docking studies based on the PDB enzymatic models of MAO isoforms.  相似文献   

18.
The enantioselective pharmacokinetics of TJ0711 hydrochloride were studied in rats given different doses of rac‐TJ0711 hydrochloride via intravenous and oral routes. R‐ and S‐TJ0711 hydrochloride were both rapidly absorbed, and the average AUC0‐∞ of R‐TJ0711 hydrochloride was greater than that of S‐TJ0711 hydrochloride after intragastric administration, with an R/S AUC ratio 1.11 and 1.35 for 30 and 50 mg/kg dose group, respectively. In contrast, the average AUC0‐∞ of R‐TJ0711 hydrochloride was smaller than that of S‐TJ0711 hydrochloride after intravenous injection, with an R/S AUC ratio 0.57 and 0.73 for 10 and 20 mg/kg dose group, respectively. R‐TJ0711 hydrochloride plasma half‐lives were shorter than those of S‐TJ0711 hydrochloride for all groups. AUC0‐4h and Cmax between the two enantiomers were significantly different after oral administration of 50 mg/kg dose of the racemate, while no significant differences between the two enantiomers were found for all the pharmacokinetic parameters of the 30 mg/kg dose group. Significant differences between the two enantiomers were detected for nearly all the pharmacokinetic parameters after intravenous administration, except for the VZ of 20 mg/kg dose group. This study suggests that dose and route of administration will influence the enantioselectivity in the pharmacokinetics of TJ0711 hydrochloride in rats. Chirality 27:53–57, 2015. © 2014 Wiley Periodicals, Inc.  相似文献   

19.
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   

20.
《Chirality》2017,29(7):348-357
Imazethapyr (IM) is a chiral herbicide composed of an (−)‐R‐enantiomer and an (+)‐S‐enantiomer with differential herbicidal activity. In this study, the effects of microbial organisms, humidity, and temperature on the selective degradation of the (−)‐R‐ and (+)‐S‐enantiomers of IM were determined in silty loam (SL) and clay loam (CL) soil with different pH values. The (−)‐R‐enantiomer of IM was preferentially degraded in two soils under different microorganism, humidity, and temperature conditions. The average half‐lives of R‐IM ranged from 43 to 66.1 days and were significantly shorter (P <  0.05) than those of S‐IM, which ranged from 51.4 to 79.8 days. The enantiomer fraction (EF = (+)‐S‐enantiomer/((−)‐R‐enantiomer + (+)‐S‐enantiomer)) values were used to describe the enantioselectivity of degradation of IM were >0.5 (P <  0.05) in two unsterilized soils under different humidity and temperature conditions. The highest EF values were observed at unsterilized CL soil samples under 50% maximum water‐holding capacity (MWHC) and 25 °C environmental conditions. The EF values of the IM enantiomers were significantly higher (P <  0.05) in CL soils (higher pH = 5.81) and were 0.581 (unsterilized) and 0.575 (50% MWHC; 25 °C) compared with those recorded in SL soil (lower pH = 4.85). In addition, this study revealed that microbial organisms preferentially utilized the more herbicidal active IM enantiomer.  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