首页 | 官方网站   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.
Mammalian soluble and microsomal epoxide hydrolases have been proposed to belong to the family of alpha/beta-hydrolase-fold enzymes. These enzymes hydrolyse their substrates by a catalytic triad, with the first step of the enzymatic reaction being the formation of a covalent enzyme-substrate ester. In the present paper, we describe the direct visualization of the ester formation between rat microsomal epoxide hydrolase and its substrate. Microsomal epoxide hydrolase was precipitated with acetone after brief incubation with [1-(14)C]epoxystearic acid. After denaturing SDS gel electrophoresis the protein-bound radioactivity was detected by fluorography. Pure epoxide hydrolase and crude microsomes showed a single radioactive signal of the expected molecular mass that could be suppressed by inclusion of the competitive inhibitor 1,1,1-trichloropropene oxide in the incubation mixture. In a similar manner, 4-fluorochalcone-oxide-sensitive binding of epoxystearic acid to rat soluble epoxide hydrolase could be demonstrated in rat liver cytosol. Under similar conditions, no covalent binding of [26-(14)C]cholesterol-5alpha,6alpha-epoxide to microsomal proteins or solubilized fractions tenfold enriched in cholesterol epoxide hydrolase activity could be observed. Our data provide definitive proof for the formation of an enzyme-substrate-ester intermediate formed in the course of epoxide hydrolysis by microsomal epoxide hydrolase, show no formation of a covalent intermediate between cholesterol epoxide hydrolase and its substrate under the same conditions as those under which an intermediate was shown for both microsomal and soluble epoxide hydrolases and therefore indicate that the cholesterol epoxide hydrolase apparently does not act by a similar mechanism and is probably not structurally related to microsomal and soluble epoxide hydrolases.  相似文献   

2.
Chemically reactive epoxide metabolites have been implicated in various forms of drug and chemical toxicity. Naphthalene, which is metabolized to a 1,2-epoxide, has been used as a model compound in this study in order to investigate the effects of perturbation of detoxication mechanisms on the in vitro toxicity of epoxides in the presence of human liver microsomes. Naphthalene (100 microM) was metabolized to cytotoxic, protein-reactive and stable, but not genotoxic, metabolites by human liver microsomes. The metabolism-dependent cytotoxicity and covalent binding to protein of naphthalene were significantly higher in the presence of phenobarbitone-induced mouse liver microsomes than with human liver microsomes. The ratio of trans-1,2-dihydrodiol to 1-naphthol was 8.6 and 0.4 with the human and the induced mouse microsomes, respectively. The metabolism-dependent toxicity of naphthalene toward human peripheral mononuclear leucocytes was not affected by the glutathione transferase mu status of the co-incubated cells. Trichloropropene oxide (TCPO; 30 microM), an epoxide hydrolase inhibitor, increased the human liver microsomal-dependent cytotoxicity (19.6 +/- 0.9% vs 28.7 +/- 1.0%; P = 0.02) and covalent binding to protein (1.4 +/- 0.3% vs 2.8 +/- 0.2%; P = 0.03) of naphthalene (100 microM), and reversed the 1,2-dihydrodiol to 1-naphthol ratio from 6.6 (without TCPO) to 2.6, 0.6 and 0.1 at TCPO concentrations of 30, 100 and 500 microM, respectively. Increasing the human liver microsomal protein concentration reduced the cytotoxicity of naphthalene, while increasing its covalent binding to protein and the formation of the 1,2-dihydrodiol metabolite. Co-incubation with glutathione (5 mM) reduced the cytotoxicity and covalent binding to protein of naphthalene by 68 and 64%, respectively. Covalent binding to protein was also inhibited by gestodene, while stable metabolite formation was reduced by gestodene (250 microM) and enoxacin (250 microM). The study demonstrates that human liver cytochrome P450 enzymes metabolize naphthalene to a cytotoxic and protein-reactive, but not genotoxic, metabolite which is probably an epoxide. This is rapidly detoxified by microsomal epoxide hydrolase, the efficiency of which can be readily determined by measurement of the ratio of the stable metabolites, naphthalene 1,2-dihydrodiol and 1-naphthol.  相似文献   

3.
Incubations of butadiene monoxide (BMO) with mouse, rat, and human liver microsomes or cDNA-expressed human microsomal epoxide hydrolase led to 3-buten-1,2-diol (BDD) detection; the BDD peak exhibited a GC/MS fragmentation pattern similar to that of reference material. Incubations with rat liver cytosol did not lead to BDD detection; however, when mouse or human liver cytosol was used, BDD was detected but at levels lower than those detected with the liver microsomes. The catalytic efficiency (V(max)/K(m) ratio) of BDD formation in rat liver microsomes was nearly 3-fold higher than the ratio obtained with mouse liver microsomes. Among two human liver microsomal samples, one sample exhibited a ratio that was nearly 3-fold higher than that of rat liver microsomes, and the second sample exhibited a ratio that was similar to that of rat liver microsomes. Although these results suggest epoxide hydrolases may play a role in BMO metabolism in vivo, rats and mice given BMO (71.3-285 micromol/kg) excreted <1% of the dose as BDD into urine within 24 hr. Thus, further studies into the role of epoxide hydrolases in BMO metabolism and disposition and the fate of BDD are warranted.  相似文献   

4.
The carcinogenic effects of acrylonitrile in rats are believed to be mediated by its DNA-reactive epoxide metabolite, 2-cyanoethylene oxide (CEO). Previous studies have shown that conjugation with glutathione is the major detoxication pathway for both acrylonitrile and CEO. This study investigated the role of epoxide hydrolase in the hydrolysis of CEO by HPLC analysis of the products from [2,3-14C]CEO. CEO is a relatively stable epoxide with a half-life of 99 min at 37 degrees C in sodium phosphate buffer (0.1 M), pH 7.3. Incubation with hepatic microsomes or cytosols from male F-344 rats or B6C3F1 mice did not enhance the rate of hydrolysis of CEO (0.69 nmol/min). Human hepatic microsomes significantly increased the rate of hydrolysis of CEO, whereas human hepatic cytosols did not. Human hepatic microsomal hydrolysis activity was heat-sensitive and potently inhibited by 1,1,1-trichloropropene oxide (IC50 of 23 microM), indicating that epoxide hydrolase was the catalyst. The hydrolysis of CEO catalyzed by hepatic microsomes from six individuals exhibited normal saturation kinetics with KM ranging from 0.6 to 3.2 mM and Vmax from 8.3 to 18.8 nmol hydrolysis products/min/mg protein. Pretreatment of rodents with phenobarbital or acetone induced hepatic microsomal hydrolysis activity toward CEO, whereas treatment with beta-naphthoflavone, dexamethasone or acrylonitrile itself was without effect. These data show that humans possess an additional detoxication pathway for CEO that is not active in rodents (but is inducible). The presence of an active epoxide hydrolase hydrolysis activity toward CEO in humans should be considered in assessments of cancer risk from acrylonitrile exposure.  相似文献   

5.
We describe the first cDNA sequence encoding a juvenile hormone-specific epoxide hydrolase from an insect. A full-length cDNA clone revealed a 462-amino-acid open reading frame encoding an amino acid sequence with 44% identity and 64% similarity to human microsomal epoxide hydrolase. All residues in the catalytic triad (residues Asp227-His428-Asp350 in the M. sexta protein) were present, as was the conserved Trp154 corresponding to the oxyanion hole. The surprising similarity of insect juvenile hormone epoxide hydrolase to vertebrate microsomal epoxide hydrolases, coupled with the ancient lineage of the epoxide hydrolases and haloalkane dehalogenases, suggests that this catabolic enzyme evolved from an original ubiquitous detoxication function to a more recent role in hormonal regulation.  相似文献   

6.
The majority of the JH III epoxide hydrolase activity in last stadium day 3 (gate 1) wandering Trichoplusia ni was membrane bound with approximately 9% of the activity found in the cytosol. Both the microsomal and cytosolic JH epoxide hydrolases were stable, retaining 30% of their original activity after incubation at 4 degrees C for 15 days. 18O-labeled water underwent enzyme catalyzed regioselective addition to the least substituted C10 position of JH III. In multiple turnover reactions with JH epoxide hydrolase in 97.9% 18O-labeled water, only 91.3% 18O incorporation was observed. This is consistent with an SN2 reaction likely involving a carboxylate in the active site of JH epoxide hydrolase. The DNA amplification cloning of a fragment of a putative T. ni epoxide hydrolase is reported. The deduced amino acid sequence shares 67% similarity to the rat microsomal epoxide hydrolase.  相似文献   

7.
Diclofenac antiserum was previously developed and used to detect protein adducts of metabolites of dichlofenac in livers of mice and rats. In this study, the antibody has been used to facilitate the purification of a major 51 kDa microsomal adduct of diclofenac from the liver microsomes of male rats that were treated with diclofenac. The adduct was identified as male-specific cytochrome P4502C11 based on its N-terminal amino acid sequence, reaction with a cytochrome P4502C11 antibody, and by its absence from liver microsomes of diclofenac-treated female rats. When diclofenac was incubated with liver microsomes of control rats in the presence of NADPH, only the 51 kDa adduct was produced. The formation of the adduct was inhibited by a cytochrome P4502C11 monoclonal antibody, but not by reduced glutathione or N-alpha-acetyl-L-lysine. No adduct was detected when diclofenac was incubated with liver microsomes from female rats. Moreover, adduct formation in vivo appeared to lead to a 72% decrease in the activity of cytochrome P4502C11. The results indicate that cytochrome P4502C11 metabolizes diclofenac into a highly reactive product that covalently binds to this enzyme before it can diffuse away and react with other proteins.  相似文献   

8.
The effect of age and gender on the in vitro biotransformation of 2-methylpropene, an alkene metabolized to 2-methyl-1,2-epoxypropane, was studied. The epoxide concentration and the epoxide metabolizing enzymatic activities were investigated in male and female Brown Norway rats of different ages. Liver tissue of senescent rats was exposed to smaller 2-methyl-1,2-epoxypropane concentrations than that of young animals, although changes during ageing were rather modest. With advancing age a feminization of male glutathione S-transferase and cytosolic epoxide hydrolase activities was found, as well as a significant decline of the female microsomal epoxide hydrolase activity and an increase of the cytochrome P-450 content in the oldest female rats.  相似文献   

9.
The epoxide hydrolase gene from Agrobacterium radiobacter AD1, a bacterium that is able to grow on epichlorohydrin as the sole carbon source, was cloned by means of the polymerase chain reaction with two degenerate primers based on the N-terminal and C-terminal sequences of the enzyme. The epoxide hydrolase gene coded for a protein of 294 amino acids with a molecular mass of 34 kDa. An identical epoxide hydrolase gene was cloned from chromosomal DNA of the closely related strain A. radiobacter CFZ11. The recombinant epoxide hydrolase was expressed up to 40% of the total cellular protein content in Escherichia coli BL21(DE3) and the purified enzyme had a kcat of 21 s-1 with epichlorohydrin. Amino acid sequence similarity of the epoxide hydrolase with eukaryotic epoxide hydrolases, haloalkane dehalogenase from Xanthobacter autotrophicus GJ10, and bromoperoxidase A2 from Streptomyces aureofaciens indicated that it belonged to the alpha/beta-hydrolase fold family. This conclusion was supported by secondary structure predictions and analysis of the secondary structure with circular dichroism spectroscopy. The catalytic triad residues of epoxide hydrolase are proposed to be Asp107, His275, and Asp246. Replacement of these residues to Ala/Glu, Arg/Gln, and Ala, respectively, resulted in a dramatic loss of activity for epichlorohydrin. The reaction mechanism of epoxide hydrolase proceeds via a covalently bound ester intermediate, as was shown by single turnover experiments with the His275 --> Arg mutant of epoxide hydrolase in which the ester intermediate could be trapped.  相似文献   

10.
An epoxide hydrolase from Rhodococcus erythropolis DCL14 catalyzes the hydrolysis of limonene-1,2-epoxide to limonene-1,2-diol. The enzyme is induced when R. erythropolis is grown on monoterpenes, reflecting its role in the limonene degradation pathway of this microorganism. Limonene-1,2-epoxide hydrolase was purified to homogeneity. It is a monomeric cytoplasmic enzyme of 17 kDa, and its N-terminal amino acid sequence was determined. No cofactor was required for activity of this colorless enzyme. Maximal enzyme activity was measured at pH 7 and 50 degrees C. None of the tested inhibitors or metal ions inhibited limonene-1,2-epoxide hydrolase activity. Limonene-1,2-epoxide hydrolase has a narrow substrate range. Of the compounds tested, only limonene-1,2-epoxide, 1-methylcyclohexene oxide, cyclohexene oxide, and indene oxide were substrates. This report shows that limonene-1,2-epoxide hydrolase belongs to a new class of epoxide hydrolases based on (i) its low molecular mass, (ii) the absence of any significant homology between the partial amino acid sequence of limonene-1,2-epoxide hydrolase and amino acid sequences of known epoxide hydrolases, (iii) its pH profile, and (iv) the inability of 2-bromo-4'-nitroacetophenone, diethylpyrocarbonate, 4-fluorochalcone oxide, and 1, 10-phenanthroline to inhibit limonene-1,2-epoxide hydrolase activity.  相似文献   

11.
A series of substituted chalcone oxides (1,3-diphenyl-2-oxiranyl propanones) and structural analogs was synthesized to investigate the mechanism by which they inhibit soluble epoxide hydrolases (sEH). The inhibitor potency and inhibition kinetics were evaluated using both murine and human recombinant sEH. Inhibition kinetics were well described by the kinetic models of A. R. Main (1982, in Introduction to Biochemical Toxicology, pp. 193-223, Elsevier, New York) supporting the formation of a covalent enzyme-inhibitor intermediate with a half-life inversely proportional to inhibitor potency. Structure-activity relationships describe active-site steric constraints and support a mechanism of inhibition consistent with the electronic stabilization of the covalent enzyme-inhibitor intermediate. The electronic effects induced by altering the ketone functionality and the para-substitution of the phenyl attached to the epoxy C1 (i.e., the alpha-carbon) had the greatest influence on inhibitor potency. The direction of the observed influence was reversed for the inhibitory potency of glycidol (1-phenyl-2-oxiranylpropanol) derivatives. Recent insights into the mechanism of epoxide hydrolase activity are combined with these experimental results to support a proposed mechanism of sEH inhibition by chalcone oxides.  相似文献   

12.
13.
1. The mono-oxygenase activity that forms epoxides has been studied in rat liver microsomes using as substrates carbamazepine and cyclobenzaprine, tricyclic drugs which form stable epoxides in vivo and in vitro. 2. A simple gas chromatographic method has been used to determine the amount of epoxide formed and the linearity of the enzymic reaction with time and protein concentration has been demonstrated. 3. Pre-treatment with carbamazepine increases the rate of formation of carbamazepine epoxide in rat liver microsomal preparations. 4. The effect of SKF 525-A on the formation of these epoxides has been studied.  相似文献   

14.
In mouse liver homogenate with an intact microsomal metabolism covalent binding of [14C]-paracetamol amounted to 1 nmol/mg protein. 65% of the total radioactivity were bound to soluble protein and 35% to microsomes. In the soluble fraction the major radioactivity peak co-chromatographed with glutathione S-transferase activity on Sephacryl S-300. Two different minor labelled fractions with apparent molecular weights of 130 000 and 25 000 daltons were also found. In a second experiment in a reconstituted system of microsomes and supernatant, 86% of the radio-activity was bound to supernatant and 14% by of microsomes. Following ion exchange chromatography of the supernatant on DEAE-Sepharose, the two major radioactivity-containing fractions coincided with GSH-S-transferase activities, but not with selenium-dependent or non-selenium-dependent glutathione peroxidase. The data show that irreversible binding of paracetamol metabolites in mouse liver occurs preferentially to GSH-S-transferases.  相似文献   

15.
Neural tube defects (NTDs) are among the most common congenital malformations, affecting approximately 1 per 1,000 liveborn infants in the United States [Nakano, 1973; Richards et al., 1972]. Maternal exposure to hyperthermia, either through recreational sources or due to an infectious agent, is thought to account for approximately 10% of observed NTD cases. The specific genes conferring susceptibility or resistance to hyperthermia-induced NTDs have not been identified. This study used differential display-polymerase chain reaction (DD-PCR) to characterize alterations in gene expression in the anterior embryonic neural tube of two highly inbred murine strains (SWV/Fnn, LM/Bc/Fnn) known to differ in their genetically determined susceptibility to heat-induced NTDs. Herein, we report the neural tube-specific differential expression of the growth arrest specific (gas 5) gene in the highly susceptible SWV/Fnn strain during neural tube closure (NTC). Although the expression of gas 5 did not appear to be altered by the teratogenic heat treatment, its spatial and strain-specific pattern of expression makes it an excellent candidate gene responsible for the observed genetic differences in NTD susceptibility between these two inbred murine strains.  相似文献   

16.
17-(Allylamino)-17-demethoxygeldanamycin (17AAG), a compound that is proposed for clinical development, shares the ability of geldanamycin to bind to heat shock protein 90 and GRP94, thereby depleting cells of p185erbB2, mutant p53, and Raf-1. Urine and plasma from mice treated i.v. with 17AAG contained six materials with absorption spectra similar to that of 17AAG. Therefore, in vitro metabolism of 17AAG by mouse and human hepatic preparations was studied to characterize: (a) the enzymes responsible for 17AAG metabolism; and (b) the structures of the metabolites produced. These materials had retention times on high-performance liquid chromatography of approximately 2, 4, 5, 6, 7, and 9 min. When incubated in an aerobic environment with 17AAG, murine hepatic supernatant (9000 x g) produced each of these compounds; the 4-min metabolite was the major product. This metabolism required an electron donor, and NADPH was favored over NADH. Metabolic activity resided predominantly in the microsomal fraction. Metabolism was decreased by approximately 80% in anaerobic conditions and was essentially ablated by CO. Microsomes prepared from human livers produced essentially the same metabolites as produced by murine hepatic microsomes, but the 2-min metabolite was the major product, and the 4-min metabolite was next largest. There was no metabolism of 17AAG by human liver cytosol. Metabolism of 17AAG by human liver microsomes also required an electron donor, with NADPH being preferred over NADH, was inhibited by approximately 80% under anaerobic conditions, and was essentially ablated by CO. Liquid chromatography/mass spectrometry analysis of human and mouse in vitro reaction mixtures indicated the presence of materials with molecular weights of 545, 601, and 619, compatible with 17-(amino)-17-demethoxygeldanamycin (17AG), an epoxide, and a diol, respectively. The metabolite with retention time of 4 min was identified as 17AG by cochromatography and mass spectral concordance with authentic standard. Human microsomal metabolism of 17AAG was inhibited by ketoconazole, implying 3A4 as the responsible cytochrome P450 isoform. Incubation of 17AAG with cloned CYP3A4 produced metabolites 4 and 6. Incubation of 17AAG with cloned CYP3A4 and cloned microsomal epoxide hydrolase produced metabolites 2 and 4, with greatly decreased amounts of metabolite 6. Incubation of 17AAG with human hepatic microsomes and cyclohexene oxide, a known inhibitor of microsomal epoxide hydrolase, did not affect the production of metabolite 4 but decreased the production of metabolite 2 while increasing the production of metabolite 6. These data imply that metabolite 2 is a diol and metabolite 6 is an epoxide. Mass spectral fragmentation patterns and the fact that 17AG is not metabolized argue for the epoxide and diol being formed on the 17-allylamino portion of 17AAG and not on its ansamycin ring. These data have implications with regard to preclinical toxicology and activity testing of 17AAG as well as its proposed clinical development because: (a) production of 17AG requires concomitant production of acrolein from the cleaved allyl moiety; and (b) 17AG, which was not metabolized by microsomes, has been described as being as active as 17AAG in decreasing cellular p185erbB2.  相似文献   

17.
Styrene 7,8-oxide and ethylene oxide are widely used genotoxic bulk chemicals, which have been associated with potential carcinogenic hazard for occupationally exposed workers. Both epoxides alkylate DNA preferentially at the N-7 position of guanine and consequently produce single-strand breaks and alkali labile sites in the DNA of exposed cells. In order to study the role of human microsomal epoxide hydrolase (hmEH) in protecting cells against genotoxicity of styrene 7,8-oxide and ethylene oxide, we expressed the cDNA of hmEH in V79 Chinese hamster cells. We obtained a number of cell clones that expressed functionally active epoxide hydrolase. Among these, the clone 92hmEH-V79 revealed an especially high enzymatic mEH activity toward styrene 7,8-oxide (10 nmol converted per mg of protein per min, measured in the 9,000 x g supernatant of the cell homogenate), that was 100 times higher than that determined in mock-transfected cells and within the range of mEH activity in human liver. Styrene 7,8-oxide-induced DNA single-strand breaks/alkali labile sites (dose range 10 microM to 1 mM styrene 7,8-oxide) measured by the alkaline elution technique were significantly lower in the 92hmEH-V79 cells as compared to the mock-transfected cells. The protection against styrene 7,8-oxide genotoxicity in 92hmEH-V79 cells could be abolished by addition of valpromide, a selective inhibitor of microsomal epoxide hydrolase. These results clearly show that the metabolism of styrene 7,8-oxide by hmEH in 92hmEH-V79 cells was responsible for the protection against styrene 7,8-oxide genotoxicity. On the other hand, no protective effect of epoxide hydrolase expression could be observed on ethylene oxide-induced DNA damage with the recombinant cell line over a dose range of 0.5-2.5 mM ethylene oxide. This selectivity of the protective effect on epoxide genotoxicity thus appears to be an important factor that must be taken into account for the prediction of the genotoxic risk of epoxides themselves or compounds that can be metabolically activated to epoxides.  相似文献   

18.
The activation of microsomal glutathione S-transferase in oxidative stress was investigated by perfusing isolated rat liver with 1 mM tert-butyl hydroperoxide (t-BuOOH). When the isolated liver was perfused with t-BuOOH for 7 min and 10 min, microsomal, but not cytosolic, glutathione S-transferase activity was increased 1.3-fold and 1.7-fold, respectively, with a concomitant decrease in glutathione content. A dimer protein of microsomal glutathione S-transferase was also detected in the t-BuOOH-perfused liver. The increased microsomal glutathione S-transferase activity after perfusion with t-BuOOH was reversed by dithiothreitol, and the dimer protein of the transferase was also abolished. When the rats were pretreated with the antioxidant alpha-tocopherol or the iron chelator deferoxamine, the increases in microsomal glutathione S-transferase activity and lipid peroxidation caused by t-BuOOH perfusion of the isolated liver was prevented. Furthermore, the activation of microsomal GSH S-transferase by t-BuOOH in vitro was also inhibited by incubation of microsomes with alpha-tocopherol or deferoxamine. Thus it was confirmed that liver microsomal glutathione S-transferase is activated in the oxidative stress caused by t-BuOOH via thiol oxidation of the enzyme.  相似文献   

19.
The purpose of this study was to determine whether lower chlorinated biphenyls would be bioactivated to electrophilic metabolites by microsomes alone or in combination with peroxidase. Monochloro- and dichlorobiphenyls were incubated with liver microsomes of rats treated with phenobarbital and beta-naphthoflavone, an NADPH-regenerating system, and deoxyguanosine 3'-monophosphate (dGp). The resultant adducts were analyzed by 32P-postlabeling either following microsomal incubation alone ("preoxidized") or coupled with subsequent oxidation with horseradish peroxidase/H2O ("oxidized"). The incubation of 4-monochlorobiphenyl (4-MCB) resulted in the formation of two minor adducts by microsomal activation alone. However, the oxidized sample showed two additional major adducts. Formation of the latter adducts was almost completely (> 80%) inhibited when the oxidation reaction was performed in the presence of ascorbic acid. The other test mono- and dichlorobiphenyls also formed 1-3 major adducts. Compared with microsomal activation alone, these adducts were enhanced after the oxidation reaction or detected only in the oxidized samples. These data suggest that (1) some adducts of the lower chlorinated biphenyls are derived from arene oxides and (2) many adducts may be formed by metabolism of the parent compounds to catechol and p-hydroquinone species, which are oxidized to semiquinones and/or quinones. The involvement of quinones and/or semiquinones was supported by UV/vis spectroscopic measurements, which showed that metabolites of 4-MCB can be oxidized to products with spectra characteristic of quinones. These data raise the possibility that lower chlorinated biphenyls may be genotoxic and may explain the fact that commercial polychlorinated biphenyl mixtures are complete rodent carcinogens.  相似文献   

20.
1. During anaerobic reductive incubation of liver microsomes, from either the pyridine- or phenobarbital-treated rat, with 1,1-dichloro-1-fluoroethane (HCFC-141b) in the presence of a NADPH-regenerating system, a time- and dose-dependent formation of reactive metabolites was detected as indicated by a depletion of added exogenous glutathione. 2. A statistically significant, dose-dependent loss of both cytochrome P450 and microsomal haem was also observed under these experimental conditions. Furthermore, a statistically significant decrease of p-nitrophenol hydroxylase and pentoxyresorufin O-depentylase activity was measured in microsomes from the pyridine- and phenobarbital-induced rat, respectively indicating that both P4502E1 and P4502B undergo substrate-dependent inactivation. 3. Both reactive metabolite formation and P450 inactivation were almost completely inhibited by previous bubbling of the incubation mixture with carbon monoxide, indicating that interaction of the substrate with a free and reduced P450 haem iron is required for substrate bioactivation and enzyme loss. 4. The presence in the incubation mixture of the spin-trap N-t-butyl-alpha-phenylnitrone (PBN) and the carbene trap 2,3-dimethyl-2-butene (DMB) largely prevented both glutathione depletion and P450 loss. This suggests that free radical and carbene intermediates formed by the metabolic activation of the substrate are involved in the inactivation of P450 and the loss of its prosthetic haem group.  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-23

京公网安备 11010802026262号