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.
Hydrochlorofluorocarbons (HCFCs) that are structural analogues of the anesthetic agent halothane may follow a common pathway of bioactivation and formation of adducts to cellular targets of distinct tissues. Exposure of rats to a single dose of HCFC 123 (2,2-dichloro- 1,1,1-trifluoroethane) or its structural analogue halothane (2-bromo-2-chloro-1,1,1-trifluoroethane) in vivo resulted in the formation of one prominent trifluoroacetylated protein adduct (TFA-protein adduct) in the heart. In contrast, a variety of distinct TFA-protein adducts were formed in the liver and the kidney of the same animals. The TFA-protein adduct in the heart was processed rapidly; t1/2 of the intact TFA-protein adduct was less than 12 h.  相似文献   

2.
Cytochrome P-450 from liver microsomes of phenobarbital-treated rabbits catalyzed anaerobic dehalogenation of halothane (2-bromo-2-chloro-1,1,1-trifluoroethane) when combined with NADPH and NADPH-cytochrome P-450 reductase. Cytochromes P-450B1 and P-448 from liver microsomes of untreated rabbits were less active. Triton X-100 accelerated the reaction. Unlike anaerobic dehalogenation of halothane in microsomes, the major product was 2-chloro-1,1,1-trifluoroethane and 2-chloro-1,1-difluoroethylene was negligible. These products were not detected under aerobic conditions, and dehalogenation activity was inhibited by carbon monoxide, phenyl isocyanide and metyrapone.  相似文献   

3.
A monospecific antibody (anti-CF3CO antibody) was obtained by affinity chromatography on a N epsilon-trifluoroacetyl-L-lysine (CF3CO-Lys) matrix of a rabbit polyclonal antiserum, directed against trifluoroacetylated protein adducts (CF3CO-proteins). The anti-CF3CO antibody recognized distinct CF3CO-proteins on immunoblots of a liver biopsy obtained from a human individual 10 h after halothane anaesthesia. Cross-reactive proteins of 52 kDa and 64 kDa were recognized on immunoblots of livers obtained from human individuals not exposed to halothane. Recognition of both CF3CO-proteins and the 52-kDa and 64-kDa cross-reactive proteins was abolished in the presence of 1 mM CF3CO-Lys. Anti-CF3CO antibody, affinity-adsorbed to the 52-kDa or the 64-kDa cross-reactive proteins of human liver, recognized the majority of target CF3CO-proteins on immunoblots of the human liver biopsy of an individual exposed to halothane. Liver biopsies of 5 out of 7 (71%) patients with halothane hepatitis exhibited an absence or low amounts of immunorecognizable 52-kDa and/or 64-kDa cross-reactive proteins. In contrast, of 22 control human individuals tested, all liver tissue samples were positive for the 52-kDa and/or the 64-kDa cross-reactive proteins. These data indicate that epitopes on the cross-reactive proteins of 52 kDa and 64 kDa of human liver bear strong immunochemical resemblance to epitopes on human liver CF3CO-proteins. Low-level expression of the cross-reactive proteins of 52 kDa and 64 kDa is discussed as one possible factor in human susceptibility to halothane hepatitis.  相似文献   

4.
To study the modulation of the reductive metabolism of halothane (2-bromo-2-chloro-1,1,1-trifluoroethane) by microsomal cytochrome b5, formation of 2-chloro-1,1,1-trifluoroethane (CTE) and 2-chloro-1,1-difluoroethylene (CDE), major reduced metabolites of halothane, was analyzed in vivo and in vitro. Rats were pretreated with both malotilate (diisopropyl-1,3-dithiol-2-ylidenemalonate) and sodium phenobarbital (malotilate-treated rats) or only with sodium phenobarbital (control rats). The microsomes of malotilate-treated rats had significantly more cytochrome b5 than the controls, whereas the cytochrome P-450 content was not different between the two groups. At the end of 2-h exposure to 1% halothane in 14% oxygen, the ratio of CDE to CTE in arterial blood was significantly higher in malotilate-treated rats than in the controls. Under anaerobic conditions, the formation of CDE and the ratio of CDE to CTE were significantly greater in microsomal preparations of malotilate-treated rats than those of the controls. In a reconstituted system containing cytochrome P-450PB purified from rabbit liver, addition of cytochrome b5 to the system enhanced the formation of CDE and increased the ratio of CDE to CTE. These results suggested that cytochrome b5 enhances the formation ratio of CDE to CTE by stimulating the supply of a second electron to cytochrome P-450, which might reduce radical reactions in the reductive metabolism of halothane.  相似文献   

5.
The source of the hydrogen atoms in reduced metabolites of carbon tetrachloride and halothane has been studied. This was approached by measuring deuterium incorporation into chloroform and 2-chloro-1,1,1-trifluoroethane formed as microsomal metabolites of carbon tetrachloride and halothane, respectively, in a medium enriched in deuterium oxide. GC/MS analysis showed no deuterium enrichment of chloroform when hepatic microsomal fractions from control rats were used; however, small increases in enrichment were seen when microsomes from phenobarbitalor benzpyrene-treated rats were employed. No detectable deuterium incorporation into 2-chloro-1,1,1-trifluoroethane was observed. These results suggest that carbanions are not formed as major intermediates and suggest that one-electron transfer reactions predominate in the reductive metabolism of carbon tetrachloride and halothane.  相似文献   

6.
A simple and sensitive gas chromatographic method for the determination of 2-chloro-1, 1-difluoroethylene (CDE) and 2-chloro-1,1,1-trifluoroethane (CTE), two highly volatile metabolites of halothane, in blood, liver and isolated hepatic microscomes is described. The entire head-space in equilibrium with a known volume or weight of the sample is injected into the gas chromatograph equipped with a flame ionization detector. Quantification is accomplished with standards prepared by fortifying blank samples with known concentrations of CDE and CTE which are treated under the same conditions as the samples. Detection limits for CDE and CTE were 2 pmole/ml in blood and 10 pmole/g in liver and the mean relative standard deviations are no greater than ± 6% except for CTE in hepatic microsomes (± 9%). A preliminary study of blood CDE and CTE levels in humans anesthetized with halothane is reported.  相似文献   

7.
Formation of Br? and, under certain conditions also F? ions has been observed in the radiation chemically induced one-electron reduction of the anesthetic halothane (2-bromo-2-chloro-1,1,1-trifluoroethane) in aqueous solutions. The initial step is the release of Br? and formation of the 2-chloro-1,1,1-trifluoroethyl radical. The latter can react via competing pathways including H-atom abstraction, addition of molecular oxygen and further reduction by an antioxidant. All of these three competitive routes lead to different product patterns. High yields of F? ions are observed under anaerobic conditions in the presence of antioxidants such as ascorbate, propylgallate, etc. The fluoride elimination is strongly pH-dependent and seems to occur in various steps after initiation through reduction of the (CF3CHCl) radical. The implication for biochemical studies on the metabolism of halothane under different oxygen concentrations is discussed.  相似文献   

8.
NADPH reduced rabbit liver microsomal enzymes catalyzed anaerobic dehalogenation of halothane (2-bromo-2-chloro-1,1,1-trifluoroethane) to produce CF2CHCl and CF3CH2Cl. Anaerobic dehalogenation was optimal at pH7.4 and was blocked by either oxygen or carbon monoxide. The degree of inhibition of anaerobic dehalogenation by carbon monoxide was closely correlated to the proportion of carbon monoxide complex of cytochrome P450. Anaerobic dehalogenation was enhanced by pretreatment of the animals with phenobarbital but not with methylcholanthrene.  相似文献   

9.
The effects of an inhalation anesthetic, halothane (2-bromo-2-chloro-1,1,1-trifluoroethane) on the formation of 5-lipoxygenase metabolites such as leukotriene B4, 5(S)-hydroxyeicosatetraenoic acid (5-HETE), 6-trans-isomers of leukotriene B4 and leukotriene C4 were studied in human leukocytes stimulated with calcium ionophore A23187. Halothane inhibited the formation of all these metabolites dose dependently and the formation was restored by removal of the drug. The anesthetic also reversibly inhibited the release of [3H]arachidonic acid from neutrophils with a half-inhibition concentration of less than 0.19 mM. The formation of 5-lipoxygenase metabolites was not inhibited by the anesthetic when leukocytes were stimulated with the ionophore in the presence of exogenous arachidonic acid. These observations indicate that the inhibitory effect of halothane on the formation of 5-lipoxygenase metabolites in leukocytes is mainly due to the inhibition of arachidonic acid release.  相似文献   

10.
The ability of microorganisms to degrade trace levels of the hydrochlorofluorocarbons HCFC-21 and HCFC-123 was investigated. Methanotroph-linked oxidation of HCFC-21 was observed in aerobic soils, and anaerobic degradation of HCFC-21 occurred in freshwater and salt marsh sediments. Microbial degradation of HCFC-123 was observed in anoxic freshwater and salt marsh sediments, and the recovery of 1,1,1-trifluoro-2-chloroethane indicated the involvement of reductive dechlorination. No degradation of HCFC-123 was observed in aerobic soils. In some experiments, HCFCs were degraded at low (parts per billion) concentrations, raising the possibility that bacteria in nature remove HCFCs from the atmosphere.  相似文献   

11.
The inhibition of tetrachloroethene (PCE) degradation in anaerobic, ethanol-fed PCE-enrichment cultures by chlorofluorocarbon 113 (CFC113) was a function of the initial CFC113 concentration. Typically, aqueous CFC113 concentrations up to 1 mg/L slowed, but did not stop PCE-degradation, but cis-1,2-dichloroethene (cDCE) degradation was inhibited by 0.2 mg/L CFC113. In some cultures, however, PCE degradation was stopped by as little as 0.15 mg/L CFC113. CFC113 also slowed the consumption of hydrogen and the concurrent methane production. CFC113 slowly degraded in PCE-enrichment cultures to hydrochlorofluorocarbon 123a (HCFC123a). Chlorotrifluoroethene was also detected. Although relatively non-toxic, CFC113 may nevertheless pose remediation challenges when present at sites that also contain PCE.  相似文献   

12.
The effect of halothane (2-bromo-2-chloro-1,1,1-trifluoroethane) on the uptake of 14C-labelled 5-hydroxytryptamine (5-HT) and its metabolism to 5-hydroxyindol-3-ylacetic acid (5-HIAA) was investigated in rat lungs perfused in situ. The rate of accumulation of 14C-labelled 5-HIAA in the tissue, monitored as an index of 5-HT metabolism, was linear with time, displayed saturation kinetics and remained stable for at least 180 min of perfusion. Exposure of the lungs to halothane (4%) for 60 min reversibly reduced production of 5-HIAA through an increase in the apparent Km for metabolism of the amine from 1.45 to 3.52 microM (P less than 0.001); the anaesthetic had no effect on the Vmax. of the process. The magnitude of the inhibition increased with time of exposure to the anaesthetic. Halothane exposure did not alter the distribution of [3H]sorbitol or [14C]5-HT, pulmonary vascular resistance, levels of ATP or the kinetics of amino acid transport in the tissue. Inhibition of protein synthesis by cycloheximide did not mimic the effect of the anaesthetic. These observations, together with those made in lungs exposed to inhibitors of 5-HT uptake and metabolism, were consistent with a halothane-mediated inhibition of 5-HT uptake, which did not appear to involve non-specific changes in membrane permeability.  相似文献   

13.
The inhibition of tetrachloroethene (PCE) degradation in anaerobic, ethanol-fed PCE-enrichment cultures by chlorofluorocarbon 113 (CFC113) was a function of the initial CFC113 concentration. Typically, aqueous CFC113 concentrations up to 1 mg/L slowed, but did not stop PCE-degradation, but cis-1,2-dichloroethene (cDCE) degradation was inhibited by 0.2 mg/L CFC113. In some cultures, however, PCE degradation was stopped by as little as 0.15 mg/L CFC113. CFC113 also slowed the consumption of hydrogen and the concurrent methane production. CFC113 slowly degraded in PCE-enrichment cultures to hydrochlorofluorocarbon 123a (HCFC123a). Chlorotrifluoroethene was also detected. Although relatively non-toxic, CFC113 may nevertheless pose remediation challenges when present at sites that also contain PCE.  相似文献   

14.
Halothane-induced lipid peroxidation in NADPH-reduced liver microsomes from phenobarbital-pretreated male rats was studied under defined steady state oxygen partial pressures (Po2). Under anaerobic conditions, as well as at a Po2 above 10 mm Hg no halothane-induced formation of malondialdehyde was detected. At a Po2 below 10 mm Hg, however, with a maximum near 1 mm Hg oxygen, significant halothane-induced malondialdehyde formation was found. This evidence supports the hypothesis that halothane can induce lipid peroxidation. The Po2 (i) must be low enough to permit the reductive formation of . CF3 CHCl-radicals but (ii), it must be high enough to promote formation of lipid peroxides.  相似文献   

15.
The rate of superoxide generation of guinea pig intraperitoneal neutrophils by a chemotactic peptide or 12-O-tetradecanoylphorbol-13-acetate (TPA) was increased by 2-bromo-2-chloro-1,1,1,-trifluoroethane (halothane), an inhalation anesthetic. This increase was inhibited by 1-(5-isoquinolinesulfonyl)methylpiperazine dihydrochloride (H-7), a specific inhibitor of Ca2+- and phospholipid-dependent protein kinase C (PKC). Halothane was found to significantly activate partially purified PKC. The activation required phosphatidylserine (PS) and Ca2+. Dioleoylglycerol- or TPA-activated PKC activity was further increased by halothane. The cytoplasmic proteins of guinea pig neutrophils phosphorylated by halothane-activated PKC were similar to those phosphorylated by PMA-activated PKC. The phosphorylation of a 48 kDa protein, a phosphorylated protein required for NADPH oxidase activation, was also increased by halothane. These data suggest that the increase of superoxide production by halothane is correlated with its activation of PKC.  相似文献   

16.
The metabolism of N-methyl substituted 7H-dibenzo[c,g]carbazole (N-Me DBC) was investigated in vitro using liver microsomes from 3-methylcholanthrene (MC)-, benzo[c]carbazole (BC) and Arochlor-pretreated mice and rats. N-Me DBC is a potent sarcomatogen devoid of hepatotoxicity and liver carcinogenic activity. The ethyl acetate-extractable metabolites were separated by high performance liquid chromatography (HPLC) and most of them were identified by proton magnetic resonance (PMR), mass spectrometry (MS) and comparison with synthetically prepared specimens. Mouse and rat microsomes gave rise to the same metabolites. The major metabolites were 5-OH-N-Me DBC (50%), N-hydroxymethyl (HMe) DBC (25-30%) and 3-OH-N-Me DBC (10%). Addition of 1,1,1-trichloropropene-2,3-oxide (TCPO) to the standard incubation medium permitted the identification of two dihydrodiols among the minor metabolites. No metabolite of DBC was observed after incubation of N-Me DBC, or its major metabolite N-HMe DBC, with either mouse or rat microsomes, but the possibility of a slight demethylation cannot be totally excluded. The lack of biotransformation at the nitrogen atom site may explain the lack of hepatotoxicity and liver carcinogenic activity of N-Me DBC. The modulation of metabolism by epoxide hydrolase, cytosol and glutathione was also investigated. The results are discussed in the light of data previously obtained with hepatotoxic and hepatocarcinogenic DBC.  相似文献   

17.
Numerous xenobiotics are known to be bioactivated and to covalently bind to proteins, but the resulting amino acid adducts (AAAs) are unknown. In this study the AAAs of twelve 14C-labeled aliphatic halides were examined after formation in an in vitro microsomal system. After exhaustive solvent extraction of the precipitated microsomal protein, the AAAs were isolated by Pronase digestion, followed by filtration through a 500 mol. wt. exclusion membrane. The liberated AAAs were applied to a constant flow DC-4A cation exchange column, resolved by stepwise buffer elution, collected and counted for radioactivity. Column recovery for applied radioactivity was 100 ± 4%. Generally, 1–4 different AAAs (defined by eluting radioactivity) were resolved, with each organohalogen displaying a characteristic elution profile. Methyl iodide, trichloroethylene and 1,2-dichloroethylene had a single major AAA while bromotrichloromethane, 1,2-dibromoethane, 1,1,1-trichloroethane, 1,2-dichloroethane, 1,1,2-trichloroethane, 2-bromo-2-chloro-1,1,1-trifluoroethane, chloroform and carbon tetrachloride had up to 4 AAAs or more, indicating combinations of binding site(s) and reactive intermediate(s). The single AAA formed following incubation of methyl iodide with the microsomes was identified as S-methylcysteine. Thus, this method appears capable of resolving binding sites and is the initial isolation step for identifying specific adducts to proteins.  相似文献   

18.
The pharmacokinetics of the antimalarial drug (+/-)-halofantrine are stereoselective in humans and rats. To better understand the stereoselective metabolism of the drug to its primary metabolite, desbutylhalofantrine (DHF), a series of in vitro and in vivo experiments were undertaken in the rat. Formation of (-)-DHF exceeded that of (+)-DHF in liver microsomes [(-):(+) ratio of intrinsic formation clearances = 1.4]. In contrast, in intestinal microsomes no significant stereoselectivity was noted in the formation of the DHF enantiomers. Intestinal microsomes were also less efficient at producing the DHF enantiomers than were liver microsomes. Based on kinetic analysis of the DHF formation, there appeared to be more than one enzyme involved in the biotransformation. (+/-)-Ketoconazole (KTZ) effectively inhibited the formation of both DHF enantiomers by both liver and intestinal microsomes, although the reduction was more marked in liver microsomes. Through a combination of the use of CYP antibodies and recombinant CYP isoenzymes, the involvement of CYP 2B1/2, 3A1, 3A2, 1A1, 2C11, 2C6, 2D1, and 2D2 were implicated in the metabolism of halofantrine to DHF. Of these, CYP3A1/2 and CYP2C11 appeared to be the primary isoenzymes involved, although CYP2C11 showed greater (+)-DHF than (-)-DHF formation, whereas for CYP3A1 it was similar to the isolated rat liver microsomes. In vivo, oral (+/-)-KTZ caused significant increases in plasma halofantrine and decreases in DHF enantiomer plasma concentrations.  相似文献   

19.
In vitro monocyte-macrophage colony-forming cells (CFC) have been detected in the thymus (30/10(6) cells) and in the cervical (22/10(6)) and mesenteric (20/10(6)) lymph nodes (LN) of the mouse. Thymus and LN derived CFC differed from bone marrow derived CFU-c in several characteristics parameters: (1) sole specificity of PMUE to induce colony formation (CF), (2) apparent singular line of monocyte-macrophage differentiation, (3) a marked 6- to 10-day lag period prior to initiation of CF, and (4) significantly slower rates of appearance of colonies in culture after initiation of CF. Two of these parameters are shared with those CFC detected within alveolar space, peritoneal exudate and pleural effusion. These are the delay prior to CF and the singular monocyte-macrophage differentiation. These similarities suggested that T-CFC and LN-CFC are probably of similar origin and represent, as suggested by Lin and Stewart ('74), a population of progenitor cells exclusively for monocyte-macrophages.  相似文献   

20.
The metabolism of [4-14C]estrone (E1) was examined in liver and kidney microsomes of adult castrated male and ovariectomized female hamsters and rats and in neonatal and immature hamster renal microsomes. In castrated male hamster liver microsomes, E1 was metabolized extensively to six major metabolites; 15 beta-hydroxyestrone, 7 alpha-hydroxyestrone, 6 alpha-hydroxyestrone, 6 beta-hydroxyestrone, 2-hydroxyestrone, and delta(9,11)-dehydroestrone, and a nonpolar fraction. Six minor metabolites of E1 were also detected. In contrast, kidney microsomes derived from castrated male hamsters metabolized E1 to mainly 17 beta-estradiol, 2- and 4-hydroxyestrone, 6 alpha-hydroxyestrone, 6 beta-hydroxyestrone and one monohydroxyestradiol metabolite. However, 16 alpha-hydroxyestrone was not detected. A variable, but low amount of estriol was also found. Interestingly, the quantity of 2-hydroxyestrone found in kidney microsomes of the hamster represented 26% of the total amount of metabolites formed, whereas in liver microsomes, only 9% of the overall metabolism resulted in the formation of 2-hydroxyestrone. The ability of kidney microsomes of female ovariectomized hamsters and two different rat strains to metabolize E1 was 5.9- and 9.4-fold lower, respectively, compared to renal microsomes of male castrated hamsters. The onset of oxidative metabolism in newborn hamster kidneys during development was also assessed. The results indicate that the oxidative metabolism of [14C]E1 in renal microsomes of newborn hamsters was 20-fold less than in kidney microsomes of adult hamsters. While catechol E1 metabolites were essentially negligible in hamster kidneys of these ages, it was evident that the conversion of E1 to estradiol via 17 beta-hydroxysteroid dehydrogenase resembles levels seen in the adult animals. Between the age of one and two months, the male hamster kidney exhibited the capacity to metabolize E1 at levels seen in fully mature adult hamsters.  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