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.
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   

2.
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   

3.
Paclobutrazol, with two stereogenic centers, but gives only (2R, 3R) and (2S, 3S)‐enantiomers because of steric‐hindrance effects, is an important plant growth regulator in agriculture and horticulture. Enantioselective degradation of paclobutrazol was investigated in rat liver microsomes in vitro. The degradation kinetics and the enantiomer fraction were determined using a Lux Cellulose‐1 chiral column on a reverse‐phase liquid chromatography–tandem mass spectrometry system. The t1/2 of (2R, 3R)‐paclobutrazol is 18.60 min, while the t1/2 of (2S, 3S)‐paclobutrazol is 10.93 min. Such consequences clearly indicated that the degradation of paclobutrazol in rat liver microsomes was stereoselective and the degradation rate of (2S, 3S)‐paclobutrazol was much faster than (2R, 3R)‐paclobutrazol. In addition, significant differences between the two enantiomers were also observed in enzyme kinetic parameters. The Vmax of (2S, 3S)‐paclobutrazol was more than 2‐fold of (2R, 3R)‐paclobutrazol and the Clint of (2S, 3S)‐paclobutrazol was higher than that of (2R, 3R)‐paclobutrazol after incubation in rat liver microsomes. These results may have potential implications for better environmental and ecological risk assessment for paclobutrazol. Chirality 27:344–348, 2015. © 2015 Wiley Periodicals, Inc.  相似文献   

4.
Tetrahydropalmatine (THP), with one chiral center, is an alkaloid that possesses analgesic and many other pharmacological actives. The aim of the present study is to investigate stereoselective metabolism of THP enantiomers in human liver microsomes (HLM) and elucidate which cytochrome P450 (CYP) isoforms contribute to the stereoselective metabolism in HLM. Additionally, the inhibitions of THP enantiomers on activity of CYP enzymes are also investigated. The results demonstrated that (+)‐THP was preferentially metabolized by HLM. Ketoconazole (inhibitor of CYP3A4/5) inhibited metabolism of (?)‐THP or (+)‐THP at same degree, whereas the inhibition of fluvoxamine (inhibitor of CYP1A2) on metabolism of (+)‐THP was greater than that of (?)‐THP; moreover, the metabolic rate of (+)‐THP was 5.3‐fold of (?)‐THP in recombinant human CYP1A2. Meanwhile, THP enantiomers did not show obvious inhibitory effect on the activity of various CYP isoforms (CYP1A2, 2A6, 2C8, 2C9, 2C19, 2E1, and 3A4/5), whereas (?)‐THP, but not (+)‐THP, significantly inhibited the activity of CYP2D6 with the Ki value of 6.42 ± 0.38 μM. The results suggested that THP enantiomers were predominantly metabolized by CYP3A4/5 and CYP1A2 in HLM, and (+)‐THP was preferentially metabolized by CYP1A2, whereas CYP3A4/5 contributed equally to metabolism of (?)‐THP or (+)‐THP. Besides, the inhibition of CYP2D6 by (?)‐THP may cause drug–drug interaction, which should be considered. Chirality 25:43–47, 2013. © 2012 Wiley Periodicals, Inc.  相似文献   

5.
Four new tirucallane triterpenoids, (21S,23R,24R)‐21,23‐epoxy‐21,24‐dihydroxy‐25‐methoxytirucall‐7‐en‐3‐one ( 2 ), (3S,21S,23R,24S)‐21,23‐epoxy‐21,25‐dimethoxytirucall‐7‐ene‐3,24‐diol ( 8 ), (21S,23R,24R)‐21,23‐epoxy‐24‐hydroxy‐21‐methoxytirucalla‐7,25‐dien‐3‐one ( 11 ), and (21S,23R,24R)‐21,23‐epoxy‐21,24‐dihydroxytirucalla‐7,25‐dien‐3‐one ( 12 ), along with 16 known analogues, 1 , 3  –  7 , 9  –  10 , and 13  –  20 , were isolated from the fruits of Melia azedarach. Their structures were elucidated by spectroscopic methods including 1D‐ and 2D‐NMR techniques and mass spectrometry. These compounds were evaluated for their cytotoxicities against HepG2 (liver), SGC7901 (stomach), K562 (leukemia), and HL60 (leukemia) cancer cell lines. Compound 20 exhibited potent cytotoxicity against HepG2 and SGC7901 cancer cells with the IC50 values of 6.9 and 6.9 μm , respectively.  相似文献   

6.
Beauvericin is a secondary metabolite natural product from microorganisms and has been shown to have a new potential antifungal activity. In this study, the metabolism and inhibition of beauvericin in human liver microsomes (HLM) and rat liver microsomes (RLM) were investigated. The apparent Km and Vmax of beauvericin in HLM were determined by substrate depletion approach and its inhibitory effects on cytochromes P450 (CYP) activities were evaluated using probe substrates, with IC50 and the (Ki) values were 1.2 μM (0.5 μM) and 1.3 μM (1.9 μM), respectively for CYP3A4/5 (midazolam) and CYP2C19 (mephenytoin). Similarly, beauvericin was also a potent inhibitor for CYP3A1/2 (IC50: 1.3 μM) in RLM. Furthermore, the pharmacokinetics of beauvericin in the rat were studied after p.o administration alone and co-administration with ketoconazole, which indicated a pharmacodynamic function may play a role in the synergistic effect on antifungal activity.  相似文献   

7.
In quantitative PET measurements, the analysis of radiometabolites in plasma is essential for determining the exact arterial input function. Diphenyl sulfide compounds are promising PET and SPECT radioligands for in vivo quantification of the serotonin transporter (SERT) and it is therefore important to investigate their radiometabolism. We have chosen to explore the radiometabolic profile of [11C]MADAM, one of these radioligands widely used for in vivo PET-SERT studies. The metabolism of [11C]MADAM/MADAM was investigated using rat and human liver microsomes (RLM and HLM) in combination with radio-HPLC or UHPLC/Q-ToF-MS for their identification. The effect of carrier on the radiometabolic rate of the radioligand [11C]MADAM in vitro and in vivo was examined by radio-HPLC. RLM and HLM incubations were carried out at two different carrier concentrations of 1 and 10 μM. Urine samples after perfusion of [11C]MADAM/MADAM in rats were also analysed by radio-HPLC. Analysis by UHPLC/Q-ToF-MS identified the metabolites produced in vitro to be results of N-demethylation, S-oxidation and benzylic hydroxylation. The presence of carrier greatly affected the radiometabolism rate of [11C]MADAM in both RLM/HLM experiments and in vivo rat studies. The good concordance between the results predicted by RLM and HLM experiments and the in vivo data obtained in rat studies indicate that the kinetics of the radiometabolism of the radioligand [11C]MADAM is dose-dependent. This issue needs to be addressed when the diarylsulfide class of compounds are used in PET quantifications of SERT.  相似文献   

8.
Shenmai injection (SMI), one of the most popular herbal preparations, is widely used for the treatment of coronary atherosclerotic cardiopathy and viral myocarditis. The purpose of this study was to investigate the effect of Shenmai injection (SMI) on the CYP3A-mediated metabolism of midazolam (MDZ). The present study demonstrated that SMI could significantly inhibit MDZ 4-hydroxylation but activate its 1′-hydroxylation in human liver microsomes (HLMs), rat liver microsomes (RLM) and recombinant human CYP3A4 and CYP3A5. The opposing effect of SMI was characterized by the kinetic change of increasing Vmax/Km for MDZ 1′-hydroxylation and decreasing Vmax/Km for MDZ 4-hydroxylation in HLM and RLM. The presence ofSMI enhanced the inhibition of ketoconazole on MDZ 4-hydroxylation but weakened or reversed its inhibition on MDZ 1′-hydroxylation in HLM. After single or multiple pretreatment with SMI, the ratios of AUC4-OH MDZ/AUCMDZ in rats were significantly decreased, while the ratios of AUC1′-OH MDZ/AUCMDZ were increased. Among the major components in SMI, total ginsenoside (TG), ophiopogon total saponins (OTS), ophiopogon total flavone (OTF), ginsenoside Rd, ophiopogonin D and ophiopogonone A exhibited significant inhibition on both 4-hydroxylation and 1′-hydroxylation of MDZ in HLM and RLM, while no activation on MDZ metabolism was observed in the presence of these major constituents alone or together. To further explore the responsible components, 3 mL of SMI was loaded on a solid phase extraction (SPE) C18 cartridge and then separated by different concentrations of methanol. The fractions eluted with 60% and 90% methanol both showed significant activation on MDZ 1′-hydroxylation in HLM, but the fraction eluted with 30% methanol had no such effect. The results indicated that the activation of SMI on MDZ 1′-hydroxylation might be mainly resulted from the lipid-soluble components in SMI.  相似文献   

9.
Benalaxyl (BX), methyl‐N‐phenylacetyl‐N‐2,6‐xylyl alaninate, is a potent acylanilide fungicide and consist of a pair of enantiomers. The stereoselective metabolism of BX was investigated in rat and rabbit microsomes in vitro. The degradation kinetics and the enantiomer fraction (EF) were determined using normal high‐performance liquid chromatography with diode array detection and a cellulose‐tris‐(3,5‐dimethylphenylcarbamate)‐based chiral stationary phase (CDMPC‐CSP). The t1/2 of (?)‐R‐BX and (+)‐S‐BX in rat liver microsomes were 22.35 and 10.66 min of rac‐BX and 5.42 and 4.03 of BX enantiomers. However, the t1/2 of (?)‐R‐BX and (+)‐S‐BX in rabbit liver microsomes were 11.75 and 15.26 min of rac‐BX and 5.66 and 9.63 of BX enantiomers. The consequence was consistent with the stereoselective toxicokinetics of BX in vitro. There was no chiral inversion from the (?)‐R‐BX to (+)‐S‐BX or inversion from (+)‐S‐BX to (?)‐R‐BX in both rabbit and rat microsomes. These results suggested metabolism of BX enantiomers was stereoselective in rat and rabbit liver microsomes. Chirality, 2011. © 2010 Wiley‐Liss, Inc.  相似文献   

10.
The purpose of this study was to elucidate the pharmacokinetics of terazosin enantiomers in healthy Chinese male subjects. After a single oral dose of 2‐mg terazosin, the plasma concentrations of terazosin enantiomers were measured over the course of 48 h in 12 healthy subjects. The plasma concentrations of (+)‐(R)‐terazosin at all time points were higher than those of (?)‐(S)‐terazosin. The area under the plasma concentration–time curve (AUC0–∞) and maximum plasma concentration of (+)‐(R)‐terazosin were significantly greater than those of the (?)‐(S)‐terazosin (P < 0.01, respectively). The R/S ratio of AUC0–∞ of terazosin was 1.68. For the first time, it was proven that the pharmacokinetics of terazosin was stereoselective in healthy Chinese male subjects. Chirality 24:1047–1050, 2012. © 2012 Wiley Periodicals, Inc.  相似文献   

11.
Golo Storch  Oliver Trapp 《Chirality》2018,30(10):1150-1160
We present rhodium catalysts that contain stereodynamic axially chiral biphenol‐derived phosphinite ligands modified with non‐stereoselective amides for non‐covalent interactions. A chirality transfer was achieved with (R)‐ or (S)‐acetylphenylalanine methyl amide, and the interaction mechanism was investigated by NMR measurements. These interactions at the non‐stereoselective interaction sites and the formation of supramolecular complexes result in an enrichment of either the (Rax)‐ or (Sax) enantiomer of the tropos catalysts, which in turn provide the (R)‐ or (S)‐acetylphenylalanine methyl ester in the hydrogenation of (Z)‐methyl‐α‐acetamidocinnamate.  相似文献   

12.
Papulaspora immersa H. H. Hotson was isolated from roots and leaves of Smallanthus sonchifolius (Poepp. and Endl. ) H. Rob. (Asteraceae), traditionally known as Yacon. The fungus was cultured in rice, and, from the AcOEt fraction, 14 compounds were isolated. Among them, (22E,24R)‐8,14‐epoxyergosta‐4,22‐diene‐3,6‐dione ( 4 ), 2,3‐epoxy‐1,2,3,4‐tetrahydronaphthalene‐c‐1,c‐4,8‐triol ( 10 ), and the chromone papulasporin ( 13 ) were new secondary metabolites. The spectral data of the known natural products were compared with the literature data, and their structures were established as the (24R)‐stigmast‐4‐en‐3‐one ( 1 ), 24‐methylenecycloartan‐3β‐ol ( 2 ), (22E,24R)‐ergosta‐4,6,8(14),22‐tetraen‐3‐one ( 3 ), (?)‐(3R,4R)‐4‐hydroxymellein ( 5 ), (?)‐(3R)‐5‐hydroxymellein ( 6 ), 6,8‐dihydroxy‐3‐methylisocoumarin ( 7 ), (?)‐(4S)‐4,8‐dihydroxy‐α‐tetralone ( 8 ), naphthalene‐1,8‐diol ( 9 ), 6,7,8‐trihydroxy‐3‐methylisocoumarin ( 11 ), 7‐hydroxy‐2,5‐dimethylchromone ( 12 ), and tyrosol ( 14 ). Compound 4 showed the highest cytotoxic activity against the human tumor cell lines MDA‐MB435 (melanoma), HCT‐8 (colon), SF295 (glioblastoma), and HL‐60 (promyelocytic leukemia), with IC50 values of 3.3, 14.7, 5.0 and 1.6 μM , respectively. Strong synergistic effects were also observed with compound 5 and some of the isolated steroidal compounds.  相似文献   

13.
Two new sterols 1 and 2 and five known ones 3 – 7 were isolated for the first time from the fruiting bodies of Cortinarius glaucopus. Their structures were established by 1‐ and 2D‐NMR spectra and HR‐FABS‐MS. The relative configuration of 1 was firmly determined by comparison of the observed 1H–1H couplings and NOESY correlations, with those predicted for the computed geometries of the conformers. Calculations were performed by means of DFT with the B3LYP functional at 6‐31 + G(d,p) level of theory, in CHCl3 as the solvent. The structures of the new ergosterol derivatives, called glaucoposterol A ( 1 ) and B ( 2 ), were thus established as (3S,5R,7R,10R,13R,17R,20S,22R,23R,24R)‐5,6‐epoxy‐3,7,23‐trihydroxystrophast‐8‐en‐14‐one and (22E,3S,5S,9S,10R,13R,17R,20R,24R)‐3,5‐dihydroxyergosta‐6,8(14),22‐trien‐15‐one, respectively. Moreover, the configuration of known strophasterol C ( 3 ) was determined as (3S,5R,6S,7R,10R,13R,17R,20S,22S,24R). Glaucoposterol A ( 1 ) and strophasterol C ( 3 ) represent the second finding in nature of steroids with the rare strophastane skeleton.  相似文献   

14.
Triadimefon is a systemic agricultural fungicide of the triazole class whose major metabolite, triadimenol, also a commercial fungicide, provides the majority of the actual fungicidal activity, i.e., inhibition of steroid demethylation. Both chemicals are chiral: triadimefon has one chiral center with two enantiomers while its enzymatic reduction to triadimenol produces a second chiral center and two diastereomers with two enantiomers each. All six stereoisomers of the two fungicides were separated from each other using a chiral BGB‐172 column on a GC‐MS system so as to follow stereospecificity in metabolism by rainbow trout hepatic microsomes. In these microsomes the S‐(+) enantiomer of triadimefon was transformed to triadimenol 27% faster than the R‐(?) enantiomer, forming the four triadimenol stereoisomers at rates different from each other. The most fungi‐toxic stereoisomer (1S,2R) was produced at the slowest rate; it was detectable after 8 h, but below the level of method quantitation. The triadimenol stereoisomer ratio pattern produced by the trout microsomes was very different from that of the commercial triadimenol standard, in which the most rat‐toxic pair of enantiomers (known as “Diastereomer A”) is about 85% of the total stereoisomer composition. The trout microsomes produced only about 4% of “Diastereomer A”. Complementary metabolomic studies with NMR showed that exposure of the separate triadimefon enantiomers and the racemate to rainbow trout for 48 h resulted in different metabolic profiles in the trout liver extracts, i.e., different endogenous metabolite patterns that indicated differences in effects of the two enantiomers. Chirality, 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

15.
Quartz crystal microbalance (QCM) biosensor was used for the chiral recognition of five pairs of enantiomers by using goat serum albumin (GSA) and rabbit serum albumin (RbSA) as chiral selectors. Serum albumin (SA) was immobilized on the QCM through the self‐assembled monolayer technique, and the surface concentration of GSA and RbSA were 8.8 × 10?12 mol cm?2 and 1.2 × 10?11 mol cm?2, respectively. The QCM biosensors showed excellent sensitivity and selectivity. Meanwhile, the chiral recognition of SA sensors was quite species dependent. There were differences between GSA and RbSA sensors in the ability and the preference of chiral recognition. To R,S‐1,2,3,4‐tetrahydro‐1‐naphthylamine (R,S‐1‐TNA), R,S‐1‐(4‐methoxyphenyl)ethylamine (R,S‐4‐MPEA), and R,S‐1‐(3‐methoxyphenyl)ethylamine (R,S‐3‐MPEA), the preference of the stereoselective SA‐drug binding of the two kinds of SA sensors were consistent. However, to R,S‐2‐octanol (R, S‐2‐OT) and R,S‐methyl lactate (R,S‐MEL), the two kinds of SA sensors had opposite chiral recognition preference. Moreover, the interactions of SA and the five pairs of enantiomers have been further investigated through ultraviolet (UV) and fluorescent (FL) spectra. The UV/FL results were in accordance with the consequence of QCM. Chirality 24:804–809, 2012. © 2012 Wiley Periodicals, Inc.  相似文献   

16.
B.V. Milborrow 《Phytochemistry》1975,14(4):1045-1053
A sample of phaseic acid methyl ester (5 mg, isolated from tomato plants fed (±)-abscisic acid, was reduced to a mixture of the epimeric dihydrophaseates which were separated by TLC. The more polar epimer was identical with the dihydrophaseate isolated from beans by Walton et al. [14]. Comparison of the NMR and IR spectra (H-bonding) of the two epimers shows the secondary hydroxyl of the less polar epimer is cis to the oxymethylene group, which is cis to the tertiary hydroxyl group. The absolute configuration of this centre is known so the absolute configuration of phaseic acid can be deduced. Phaseic acid is (−)-3-methyl-5{8[1(R), 5(R)-dimethyl-8(S)-hydroxy-3-oxo-6-oxabicyclo-(3,2,1)-octane]} 2-cis-4-trans-pentadienoic acid and both it and the reduction products exist in chair conformations. The more polar epimer isolated by Walton et al. is (−)-3-methyl-5{8[3(S,8(S)-dihydroxy-1(R,5(R)-dimethyl-6-oxabicyclo-(3,2,1)-octane]}2-cis-4-trans-pentadienoic acid. It is suggested that the less polar epimer should be referred to as epi-dihydrophaseic acid.  相似文献   

17.
In‐depth conformational analyses of 10 known eremophilane (= (1S,4aR,7R,8aR)‐decahydro‐1,8a‐dimethyl‐7‐(1‐methylethyl)napththalene) sesquiterpenes, 1 – 10 , from Petasites hybridus were performed with molecular mechanics as well as density functional theory methods. Electronic transition energies and rotational strengths of these eight eremophilane lactones and two petasins were calculated by time‐dependent density functional theory (B3PW91/TZVP). The absolute configurations of the constituents could be assigned by comparison of their simulated and experimental circular dichroism (CD) spectra in methanol as (4S,5R,8S,10R) ( 1 , 2 ), (2R,4S,5R,8S,10R) ( 3 , 4 , 5 ), (2R,4S,5R,8R,9R,10R) ( 6 ), (2R,4S,5R,8R,10R) ( 7 , 8 ), and (3R,4R,5R) ( 9 , 10 ). Single‐crystal X‐ray diffraction data of 8β‐hydroxyeremophilanolide ((8S)‐8‐hydroxyeremophil‐7(11)‐en‐12,8‐olide) ( 1 ) served as starting point for the theoretical conformational calculations of the 8β‐epimers of the eremophilane lactones. Experimental CD spectra as well as 1H NMR spectra of compound 1 in methanol were considerably dependent on sample concentration. Chirality, 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

18.
The interaction of the nonsteroidal anti‐inflammatory drug flurbiprofen (FBP) with human serum albumin (HSA) hardly influences the fluorescence of the protein's single tryptophan (Trp). Therefore, in addition to fluorescence, heavy atom‐induced room‐temperature phosphorescence is used to study the stereoselective binding of FBP enantiomers and their methyl esters to HSA. Maximal HSA phosphorescence intensities were obtained at a KI concentration of 0.2 M. The quenching of the Trp phosphorescence by FBP is mainly dynamic and based on Dexter energy transfer. The Stern–Volmer plots based on the phosphorescence lifetimes indicate that (R)‐FBP causes a stronger Trp quenching than (S)‐FBP. For the methyl esters of FBP, the opposite is observed: (S)‐(FBPMe) quenches more than (R)‐FBPMe. The Stern–Volmer plots of (R)‐FBP and (R)‐FBPMe are similar although their high‐affinity binding sites are different. The methylation of (S)‐FBP causes a large change in its effect on the HSA phosphorescence lifetime. Furthermore, the quenching constants of 3.0 × 107 M?1 s?1 of the R‐enantiomers and 2.5 × 107 M?1 s?1 for the S‐enantiomers are not influenced by the methylation and indicate a stereoselectivity in the accessibility of the HSA Trp to these drugs. Chirality 24:840–846, 2012. © 2012 Wiley Periodicals, Inc.  相似文献   

19.
Carbofuran is a carbamate pesticide used in agricultural practice throughout the world. Its effect as a pesticide is due to its ability to inhibit acetylcholinesterase activity. Though carbofuran has a long history of use, there is little information available with respect to its metabolic fate and disposition in mammals. The present study was designed to investigate the comparative in vitro metabolism of carbofuran from human, rat, and mouse liver microsomes (HLM, RLM, MLM, respectively), and characterize the specific enzymes involved in such metabolism, with particular reference to human metabolism. Carbofuran is metabolized by cytochrome P450 (CYP) leading to the production of one major ring oxidation metabolite, 3-hydroxycarbofuran, and two minor metabolites. The affinity of carbofuran for CYP enzymes involved in the oxidation to 3-hydroxycarbofuran is significantly less in HLM (Km = 1.950 mM) than in RLM (Km = 0.210 mM), or MLM (Km = 0.550 mM). Intrinsic clearance rate calculations indicate that HLM are 14-fold less efficient in the metabolism of carbofuran to 3-hydroxycarbofuran than RLM or MLM. A screen of 15 major human CYP isoforms for metabolic ability with respect to carbofuran metabolism demonstrated that CYP3A4 is the major isoform responsible for carbofuran oxidation in humans. CYP1A2 and 2C19 are much less active while other human CYP isoforms have minimal or no activity toward carbofuran. In contrast with the human isoforms, members of the CYP2C family in rats are likely to have a primary role in carbofuran metabolism. Normalization of HLM data with the average levels of each CYP in native HLM, indicates that carbofuran metabolism is primarily mediated by CYP3A4 (percent total normalized rate (% TNR) = 77.5), although CYP1A2 and 2C19 play ancillary roles (% TNR = 9.0 and 6.0, respectively). This is substantiated by the fact that ketoconazole, a specific inhibitor of CYP3A4, is an excellent inhibitor of 3-hydroxycarbofuran formation in HLM (IC50: 0.31 μM). Chlorpyrifos, an irreversible non-competitive inhibitor of CYP3A4, inhibits the formation of 3-hydroxycarbofuran in HLM (IC50: 39 μM). The use of phenotyped HLM demonstrated that individuals with high levels of CYP3A4 have the greatest potential to metabolize carbofuran to its major metabolite. The variation in carbofuran metabolism among 17 single-donor HLM samples is over 5-fold and the best correlation between CYP isoform activity and carbofuran metabolism was observed with CYP3A4 (r2 = 0.96). The interaction of carbofuran and the endogenous CYP3A4 substrates, testosterone and estradiol, were also investigated. Testosterone metabolism was activated by carbofuran in HLM and CYP3A4, however, less activation was observed for carbofuran metabolism by testosterone in HLM and CYP3A4. No interactions between carbofuran and estradiol metabolism were observed.  相似文献   

20.
Triadimenol is a widely used triazole fungicide and consists of four stereoisomers with 1R,2S, 1S,2R, 1R,2R, and 1S,2S configurations. The trans‐enantiomeric pair (1R,2S‐isomer and 1S,2R‐isomer) is also called triadimenol‐A and the cis‐enantiomeric pair (1R,2R‐isomer and 1S,2S‐isomer) triadimenol‐B. In this study, the stereoselective degradation and chiral stability of triadimenol in two soils were investigated in details. The dissipation of technical triadimenol, a 6:1 mixture of triadimenol‐A and triadimenol‐B, showed significant epimerization from triadimenol‐A to triadimenol‐B occurred along with the dissipation process. The degradation exhibited some stereoselectivity, resulting in a concentration order of 1S,2S > 1R,2R > 1R,2S > 1S,2R or 1S,2S > 1R,2R > 1S,2R > 1R,2S at the end of the 100 days incubation for Baoding soil or Wuhan soil, respectively. Further incubation of triadimenol‐B revealed no epimerization, i.e. triadimenol‐B was configurationally stable in soil, and 1R,2R‐triadimenol degraded slightly slower in the former part and slightly faster in the later part of the incubation than 1S,2S‐triadimenol. Moreover, by incubation of enantiopure 1S,2R‐triadimenol and 1R,2S‐triadimenol, the results documented the epimerization for each enantiomer occurred at both C‐1 and C‐2 positions. Finally, the present work also documented that the enantiomerization reaction for all the four stereoisomers was nearly negligible in the soils. Chirality 25:355‐360:, 2013. © 2013 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