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.
Recent studies of a number of volatile aromatic hydrocarbons have suggested that the formation of covalently bound metabolites arises solely through the intermediate formation of phenols. This study further examines the involvement of 1-naphthol in the in vivo and in vitro formation of covalently bound metabolites and pulmonary bronchiolar necrosis by naphthalene. Marked differences were observed in the rate of 1-naphthol formation in lung and liver microsomal incubations without correspondingly large differences between the rates of formation of covalently bound metabolites from naphthalene and 1-naphthol. Glutathione decreased covalent binding in hepatic microsomal incubations containing 14[C]1-naphthol but did not result in the formation of any of the glutathione adducts isolated from identical incubations containing 14[C]naphthalene. Tissue levels of covalently bound radioactivity in mice treated with 14[C]1-naphthol or 14[C]naphthalene were similar; however, in contrast to studies with naphthalene, 1-naphthol administration did not deplete tissue glutathione nor result in detectable tissue injury. These studies indicate that 1-naphthol is not an obligate intermediate in the formation of covalently bound metabolites from naphthalene nor does it appear to be a more proximate lung toxic metabolite.  相似文献   

2.
The marine cyanobacterium Oscillatoria sp. strain JCM oxidized naphthalene predominantly to 1-naphthol. Experiments with [1-2H]naphthalene and [2-2H]naphthalene indicated that 1-naphthol was formed with 68 and 74% retention of deuterium, respectively. No significant isotope effect was observed when the organism was incubated with a 1:1 mixture of naphthalene and [2H8]naphthalene. The results indicate that 1-naphthol is formed through a naphthalene 1,2-oxide intermediate, which rearranges spontaneously via an NIH shift mechanism.  相似文献   

3.
The marine cyanobacterium Oscillatoria sp. strain JCM oxidized naphthalene predominantly to 1-naphthol. Experiments with [1-2H]naphthalene and [2-2H]naphthalene indicated that 1-naphthol was formed with 68 and 74% retention of deuterium, respectively. No significant isotope effect was observed when the organism was incubated with a 1:1 mixture of naphthalene and [2H8]naphthalene. The results indicate that 1-naphthol is formed through a naphthalene 1,2-oxide intermediate, which rearranges spontaneously via an NIH shift mechanism.  相似文献   

4.
The quantification and identification of xenobiotic reactive intermediates is difficult in the absence of highly radiolabeled drug. We have developed a method for identifying these intermediates by measuring the formation of adducts to intracellularly generated radiolabeled glutathione (GSH). Freshly isolated adherent rat and human hepatocytes were incubated overnight in methionine and cystine-free ('thio-free') medium. They were then exposed to 100 microM methionine and 10 microCi 35S-labeled methionine in otherwise thio-free medium to replete cellular GSH pools with intracellularly generated 35S-labeled GSH. After 3 h, acetaminophen was added as a test compound and the cells were incubated for an additional 24 h. Intracellular GSH and its specific activity were quantified after reaction with monobromobimane followed by HPLC analysis with fluorescence and radiochemical detection. Radiolabeled GSH was detectable at 3 h and maintained high specific activity and physiological concentrations for up to 24 h. Incubation medium from acetaminophen treated and nontreated hepatocytes were analyzed for radiolabeled peaks by HPLC using radiochemical detection. Radiolabeled peaks not present in nontreated hepatocytes were identified as acetaminophen GSH adducts by LC-MS. Formation of acetaminophen 35S-GSH adducts by rat hepatocytes containing endogenously synthesized 35S-GSH was increased with acetaminophen concentrations ranging from 500 to 2 mM.  相似文献   

5.
Microsomal epoxide hydrolase activity, determined using benzpyrene 4,5-oxide and styrene 7,8-oxide, increased in cultured hepatocytes compared to freshly isolated cells. In contrast, cytosolic epoxide hydrolase activity, assayed using trans-stilbene oxide, had decreased 80% by 24 hr and was barely detectable after 96 hr in culture. There was no difference in enzyme activity between freshly isolated hepatocytes and the two rat hepatoma cell lines McA-RH 7777 and H4-II-E, when styrene 7,8-oxide was used as substrate. However, benzpyrene 4,5-oxide hydrolase activity of the McA-RH 7777 and H4-II-E cell lines were 55 and 10%, respectively, of freshly isolated hepatocytes. These results show that hepatoma cell lines provide a suitable system for studying the regulation of both the microsomal and cytosolic epoxide hydrolase enzymes.  相似文献   

6.
The metabolism of styrene by prostaglandin hydroperoxidase and horseradish peroxidase was examined. Ram seminal vesicle microsomes in the presence of arachidonic acid or hydrogen peroxide and glutathione converted styrene to glutathione adducts. Neither styrene 7,8-oxide nor styrene glycol was detected as a product in the incubation. Also, the addition of styrene 7,8-oxide and glutathione to ram seminal vesicle microsomes did not yield styrene glutathione adducts. The peroxidase-generated styrene glutathione adducts were isolated by high pressure liquid chromatography and characterized by NMR and tandem mass spectrometry as a mixture of (2R)- and (2S)-S-(2-phenyl-2-hydroxyethyl)glutathione. (1R)- and (1S)-S-(1-phenyl-2-hydroxyethyl)glutathione were not formed by the peroxidase system. The addition of phenol or aminopyrine to incubations, which greatly enhances the oxidation of glutathione to a thiyl radical by peroxidases, increased the formation of styrene glutathione adducts. We propose a new mechanism for the formation of glutathione adducts that is independent of epoxide formation but dependent on the initial oxidation of glutathione to a thiyl radical by the peroxidase, and the subsequent reaction of the thiyl radical with a suitable substrate, such as styrene.  相似文献   

7.
Previous studies indicate that dietary administration of phenolic antioxidants, 2(3)-tert-butyl-4-hydroxyanisole (BHA) and 3,5-di-tert-butyl-4-hydroxytoluene, inhibits the carcinogenic effect of a number of chemical carcinogens including aflatoxin B1 (AFB1). Induction of hepatic enzymes, such as glutathione S-transferase, UDP-glucuronyltransferase, and epoxide hydrolase, has been shown to be responsible for the reduction of AFB1 cytotoxic and carcinogenic effects. The effect of BHA on AFB1 activation was examined in vitro utilizing isolated rat hepatocytes and liver microsomes. In hepatocytes, the total AFB1 content and bound form of AFB1 were 3.4 and 1.4 pmol/10(6) cells, respectively. In the cell-free microsomal activating system, 2.2 pmol were activated per mg of microsomal protein during 60 min of incubation. BHA (0.1-0.5 mM) inhibited AFB1 activation and binding in both systems in a dose-dependent manner; in hepatocytes, 90% inhibition was observed at 0.5 mM. Analyzing various AFB1 adducts, BHA (0.25 mM)-treated hepatocytes contained a significantly reduced amount of AFB1 macromolecular adducts. The antioxidant neither stimulated nor inhibited the cytosolic glutathione S-transferase and microsomal UDP-glucuronyltransferase activities. Analysis of various hydroxylated (aflatoxins M1 and Q1 (AFM1 and AFQ1] and demethylated (aflatoxin P1 (AFP1] metabolites of AFB1 in both the conjugated and unconjugated form indicated that there was a 30-50% reduction of unconjugated AFP1, AFQ1, and AFM1, whereas AFB1 was increased 3-fold. There was no significant change of conjugated metabolites. The effect of BHA on AFB1 activation in hepatocytes was compared with that of other cytochrome P-450 inhibitors; the ED50 values of SKF 525A, BHA, and metyrapone were 9 microM, 40 microM, and 280 microM, respectively. In the cell-free microsomal system, biotransformation of AFB1 to AFP1, AFM1, and AFQ1 was also inhibited. Kinetic analysis of p-nitroanisole O-demethylase activity of rat liver microsomes demonstrated that BHA inhibited noncompetitively with an apparent Ki of 90 microM. In the absence of enzyme induction, the phenolic antioxidant, BHA, blocks the oxidative biotransformation of AFB1 in isolated hepatocytes.  相似文献   

8.
The toxic and carcinogenic effects of arsenic may be mediated by both inorganic and methylated arsenic species. The methylation of arsenic(III) is thought to take place via sequential oxidative methylation and reduction steps to form monomethylarsenic (MMA) and dimethylarsenic (DMA) species, but recent evidence indicates that glutathione complexes of arsenic(III) can be methylated without oxidation. The kinetics of arsenic methylation were determined in freshly isolated hepatocytes from male B6C3F1 mice. Hepatocytes (>90% viability) were isolated by collagenase perfusion and suspended in Williams' Medium E with various concentrations of arsenic(III) (sodium m-arsenite). Aliquots of the lysed cell suspension were analyzed for arsenic species by hydride generation-atomic absorption spectrometry. The formation of MMA(III) from sodium arsenite (1 microM) was linear with respect to time for >90 min. DMA(III) formation did not become significant until 60 min. MMA(V) and DMA(V) were not consistently observed in the incubations. These results suggest that the glutathione complex mechanism of methylation plays an important role in arsenic biotransformation in mouse hepatocytes. Metabolism of arsenic(V) was not observed in mouse hepatocytes, consistent with inhibition of arsenic(V) active cellular uptake by phosphate in the medium. The formation of MMA(III) increased with increasing arsenic(III) concentrations up to approximately 2 microM and declined thereafter. The concentration dependence is consistent with a saturable methylation reaction accompanied by uncompetitive substrate inhibition of the reaction by arsenic(III). Kinetic analysis of the data suggested an apparent K(M) of approximately 3.6 microM arsenic(III), an apparent V(max) of approximately 38.9 microg MMA(III) formed/L/h/million cells, and an apparent K(I) of approximately 1.3 microM arsenic(III). The results of this study can be used in the physiologically based pharmacokinetic model for arsenic disposition in mice to predict the concentration of MMA(III) in liver and other tissues.  相似文献   

9.
Naphthalene is metabolized in the lung and liver to reactive intermediates by cytochrome P450 enzymes. These reactive species deplete glutathione, covalently bind to proteins, and cause necrosis in Clara cells of the lung. The importance of glutathione loss in naphthalene toxicity was investigated by using the glutathione prodrugs (glutathione monoethylester or cysteine-glutathione mixed disulfide) to maintain glutathione pools during naphthalene exposure. Mice given a single intraperitoneal injection of naphthalene (1.5 mmol/kg) were treated with either prodrug (2.5 mmol/kg) 30 min later. Both compounds effectively maintained glutathione levels and decreased naphthalene-protein adducts in the lung and liver. However, cysteine-glutathione mixed disulfide was more effective at preventing Clara cell injury. To study the prodrugs in Clara cells without the influence of hepatic naphthalene metabolism and circulating glutathione, dose-response and time-course studies were conducted with intrapulmonary airway explant cultures. Only the ester of glutathione raised GSH in vitro; however, both compounds limited protein adducts and cell necrosis. In vitro protection was not associated with decreased naphthalene metabolism. We conclude that (1) glutathione prodrugs can prevent naphthalene toxicity in Clara cells, (2) the prodrugs effectively prevent glutathione loss in vivo, and (3) cysteine-glutathione mixed disulfide prevents naphthalene injury in vitro without raising glutathione levels.  相似文献   

10.
Bacillus cereus ATCC 14579 transformed naphthalene predominately to 1-naphthol. Experiments with [14C]naphthalene showed that over a 24 h period, B. cereus oxidized 5.2% of the added naphthalene. 1-Naphthol accounted for approximately 80% of the total metabolites. B. cereus incubated with naphthalene under the presence of 18O2 led to the isolation of 1-naphthol that contained 94% 18O. The metabolism of [1-2H]-and [2-2H]-naphthalene by B. cereus yielded 1-naphthol which retained 95% and 94% deuterium, respectively, as determined by mass spectral analysis. NMR spectroscopic analysis of the deuterated 1-naphthol formed from [1-2H]-naphthalene indicated an NIH shift mechanism in which 19% of the deuterium migrated from the C-1 to the C-2 position. The 18O2 and NIH shift experiments implicate naphthalene-1,2-oxide as an intermediate in the formation of 1-naphthol from naphthalene by B. cereus.Abbreviations HPLC High performance liquid chromatography - NMR nuclear magnetic resonance  相似文献   

11.
Covalent binding of benzo(a)pyrene (BP) metabolites to DNA was investigated in hepatocytes and liver microsomes (MC-microsomes) isolated from 3-methylcholanthrene-treated rats. The major DNA adducts formed during BP metabolism in both hepatocytes and incubations of calf thymus DNA with MC-microsomes were adducts of anti and syn isomers of trans-7,8,-dihydroxy-9,10-epoxy-7,8,9,10-tetrahydrobenzo(a)pyrene (diol-epoxides) and of epoxide derivatives of BP-9-phenol (phenol-oxides). Diol-epoxide adducts predominated over phenol-oxide adducts in hepatocytes, while the reverse was found in microsomal incubations. In hepatocytes, both diol-epoxide and phenol-oxide adducts increased with increasing BP concentration; the ratio of diol-epoxide adduct to phenol-oxide adduct decreased from 6:1 to 3:1 between 30 and 100 μm BP. In microsomal incubations, decreases in DNA concentration or addition of the hepatocyte L15 medium produced larger decreases in phenol-oxide adducts than in diol-epoxide adducts. The effects of the inhibitors salicylamide, diethylmaleate, and 3,3,3,-trichloropropene oxide on formation of BP-DNA adducts are interpreted in terms of changes in precursor formation and metabolism and reductions in hepatocyte glutathione levels. Addition of 1.5 mg/ml exogenous DNA to hepatocyte incubations produced no change in covalent binding to cellular DNA, even though extracellular BP-DNA adducts accounted for 97% of the total adducts formed. Both the relative amounts of diol-epoxide and phenol-oxide adducts and the total adducts per milligram of DNA were indistinguishable with respect to extracellular and intracellular DNA. Modification of extracellular DNA by diol-epoxides was at least as efficient as modification of calf thymus DNA in incubations with MC-microsomes. It is concluded that BP diol-epoxides and phenol-oxides can leave the cell or enter the nucleus with equal facility but are more effective in binding to DNA in the cell in which they are generated.  相似文献   

12.
Absolute configurations of the arene 1,2-oxides formed from napththalene and anthracene by cytochrome P-450c, the predominant isozyme of cytochrome P-450 found in the livers of rats treated with 3-methylcholanthrene, were determined via two different approaches. The first consisted of trapping the arene oxides with N-acetyl-L-cysteine to form S-conjugates, methylation of the conjugates with diazomethane, and separation of the resulting diastereomeric esters by reversed phase high performance liquid chromatography. Analysis by this procedure of the arene oxides formed from radioactive naphthalene and anthracene by a highly purified and reconstituted monooxygenase system containing cytochrome P-450c indicated that 73 and greater than or equal to 95%, respectively, of the metabolically formed arene oxides consisted of the (+)-(1R,2S)-enantiomer. In the second approach, each hydrocarbon was incubated with a reconstituted system containing both cytochrome P-450c and epoxide hydrolase. Under these conditions, the predominant metabolites are trans-1,2-dihydrodiols formed by epoxide hydrolase catalyzed trans-addition of water to the arene oxide intermediates. In both cases, the (-)-(1R,2R)-dihydrodiols predominated; 92% for naphthalene and 99% for anthracene. Enzyme-catalyzed addition of water to (+)- and (-)-anthracene 1,2-oxide and (+)-napthalene 1,2-oxide occurred exclusively (greater than 99%) at the allylic 2-position. The (-)-(1S,2R)-naphthalene 1,2-oxide, however, is converted to a 40:60 mixture of the (-)-(1R,2R)- and (+)-(1S,2S)-dihydrodiols by benzylic and allylic attack, respectively, resulting in increased enantiomeric purity of the dihydrodiol relative to the oxide. Thus, qualitatively and quantitatively both approaches indicate that the (+)-arene (1R,2S)-oxides predominate. The results are discussed in terms of the steric constraints of a proposed model for the catalytic binding site of cytochrome P-450c.  相似文献   

13.
The enzymatic mechanisms involved in the degradation of phenanthrene by the white rot fungus Pleurotus ostreatus were examined. Phase I metabolism (cytochrome P-450 monooxygenase and epoxide hydrolase) and phase II conjugation (glutathione S-transferase, aryl sulfotransferase, UDP-glucuronosyltransferase, and UDP-glucosyltransferase) enzyme activities were determined for mycelial extracts of P. ostreatus. Cytochrome P-450 was detected in both cytosolic and microsomal fractions at 0.16 and 0.38 nmol min(sup-1) mg of protein(sup1), respectively. Both fractions oxidized [9,10-(sup14)C]phenanthrene to phenanthrene trans-9,10-dihydrodiol. The cytochrome P-450 inhibitors 1-aminobenzotriazole (0.1 mM), SKF-525A (proadifen, 0.1 mM), and carbon monoxide inhibited the cytosolic and microsomal P-450s differently. Cytosolic and microsomal epoxide hydrolase activities, with phenanthrene 9,10-oxide as the substrate, were similar, with specific activities of 0.50 and 0.41 nmol min(sup-1) mg of protein(sup-1), respectively. The epoxide hydrolase inhibitor cyclohexene oxide (5 mM) significantly inhibited the formation of phenanthrene trans-9,10-dihydrodiol in both fractions. The phase II enzyme 1-chloro-2,4-dinitrobenzene glutathione S-transferase was detected in the cytosolic fraction (4.16 nmol min(sup-1) mg of protein(sup-1)), whereas aryl adenosine-3(prm1)-phosphate-5(prm1)-phosphosulfate sulfotransferase (aryl PAPS sulfotransferase) UDP-glucuronosyltransferase, and UDP-glucosyltransferase had microsomal activities of 2.14, 4.25, and 4.21 nmol min(sup-1) mg of protein(sup-1), respectively, with low activity in the cytosolic fraction. However, when P. ostreatus culture broth incubated with phenanthrene was screened for phase II metabolites, no sulfate, glutathione, glucoside, or glucuronide conjugates of phenanthrene metabolites were detected. These experiments indicate the involvement of cytochrome P-450 monooxygenase and epoxide hydrolase in the initial phase I oxidation of phenanthrene to form phenanthrene trans-9,10-dihydrodiol. Laccase and manganese-independent peroxidase were not involved in the initial oxidation of phenanthrene. Although P. ostreatus had phase II xenobiotic metabolizing enzymes, conjugation reactions were not important for the elimination of hydroxylated phenanthrene.  相似文献   

14.
Naphthalene is an environmental toxicant to which humans are exposed. Naphthalene causes dose-dependent cytotoxicity to murine airway epithelial cells but a link between exposure and human pulmonary disease has not been established. Naphthalene toxicity in rodents depends on P450 metabolism. Subsequent biotransformation results in urinary elimination of several conjugated metabolites. Glucuronide and sulfate conjugates of naphthols have been used as markers of naphthalene exposure but, as the current studies demonstrate, these assays provide a limited view of the range of metabolites generated from the parent hydrocarbon. Here, we present a liquid chromatography tandem mass spectrometry method for measurement of the glucuronide and sulfate conjugates of 1-naphthol as well as the mercapturic acids and N-acetyl glutathione conjugates from naphthalene epoxide. Standard curves were linear over 2 log orders. On column detection limits varied from 0.91 to 3.4 ng; limits of quantitation from 1.8 to 6.4 ng. The accuracy of measurement of spiked urine standards was -13.1 to + 5.2% of target and intra-day and inter-day variability averaged 7.2 (± 4.5) and 6.8 (± 5.0) %, respectively. Application of the method to urine collected from mice exposed to naphthalene at 15 ppm (4 hrs) showed that glutathione-derived metabolites accounted for 60-70% of the total measured metabolites and sulfate and glucuronide conjugates were eliminated in equal amounts. The method is robust and directly measures several major naphthalene metabolites including those derived from glutathione conjugation of naphthalene epoxide. The assays do not require enzymatic deconjugation, extraction or derivatization thus simplifying sample work up.  相似文献   

15.
As is the case for cytochrome P-450c, arene 1,2-oxides have been identified as initial metabolites when naphthalene and anthracene are oxidized by cytochrome P-450b in a highly purified, reconstituted system. Overall rates of metabolism by cytochrome P-450b are greater than 3-fold and greater than 50-fold lower than the respective rates of metabolism by cytochrome P-450c. For both hydrocarbons, the (-)-(1S,2R)-oxide predominates (74%) with cytochrome P-450b as the terminal oxidant, based on trapping the labile arene oxides as N-acetyl-L-cysteine S-conjugates of known absolute configuration. This result is in marked contrast to data obtained with cytochrome P-450c where the (+)-(1R,2S)-oxides predominate (73-greater than 95%). In the absence of added epoxide hydrolase, the metabolically formed arene oxides rapidly isomerize to phenols. Addition of increasing amounts of epoxide hydrolase to the incubation medium results in the formation of trans-1,2-dihydrodiols at the expense of phenols from the common arene oxide intermediates. Evaluation of the kinetic parameters (Km and kcat) for the hydration of the (+)- and (-)-enantiomers of both arene oxides by epoxide hydrolase has indicated that the (+)-(1R,2S)-enantiomers exhibit lower values of Km (approximately 1 microM) whereas the values of kcat are similar for both enantiomers of a given arene oxide. These parameters have allowed construction of a mathematical model which predicts the enantiomer composition of the dihydrodiols formed from naphthalene in reconstituted systems containing specific epoxide hydrolase concentrations. The data reported argue against a selective functional coupling mechanism between cytochrome P-450c and epoxide hydrolase in the metabolism of naphthalene and anthracene to the 1,2-dihydrodiols.  相似文献   

16.
Solubilized cytochrome P-450 monooxygenase and epoxide hydrase activities from rat liver microsomes have been separated by column chromatography. The highly active epoxide hydrase fraction is still contaminated with cytochrome P-450, which has very low monooxygenase activity. The highly purified cytochrome P-450 fraction possesses high monooxygenase activity and is essentially devoid of epoxide hydrase activity. Purification factors for the epoxide hydrase through four purification steps are similar with [3H]styrene oxide, [3H]naphthalene oxide, [3H]cyclohexene oxide, and benzene oxide as substrates. Failure of benzene oxide to inhibit hydration of styrene or naphthalene oxide in the most purified preparations in indicative of the presence of at least two hydrases. These purified cytochrome monooxygenase and hydrase preparations represent valuable tools for the study of the intermediacy of arene oxides in drug metabolism. Thus, with naphthalene, only naphthol is formed with the monooxygenase, while both naphthol and the dihydrodiol are formed in the presence of monooxygenase and hydrase. A convenient radiochemical synthesis of [3H]naphthalene 1,2-oxide and assays for the measurement of the hydration of [3H]naphthalene oxide and benzene oxide, based on differential extractions and high-pressure liquid chromatography, respectively, are described.  相似文献   

17.
The toxicity of naphthalene in rodents has been attributed to the reactive metabolites naphthalene-1,2-oxide (NPO), 1,2-naphthoquinone (1,2-NPQ) and 1,4-naphthoquinone (1,4-NPQ). Differences in the formation of these reactive metabolites in different species can shed light on the mechanism by which naphthalene exerts its toxicity. Protein adducts allow investigators to study the disposition of reactive metabolites that cannot be measured directly. We measured cysteinyl adducts of the above metabolites in hemoglobin (Hb) and albumin (Alb) from the blood of male Swiss Webster mice dosed with 1.56-200mg naphthalene/kg b.w. Levels of NPO adducts (designated as NPO1-Hb, NPO2-Hb, NPO1-Alb and NPO2-Alb) increased nonlinearly with the administered dose; levels of Alb adducts were higher than those of Hb adducts; levels of NPO1 adducts were higher than those of NPO2 adducts. Levels of NPQ adducts (1,2-NPQ-Alb, 1,4-NPQ-Alb, 1,2-NPQ-Hb and 1,4-NPQ-Hb) were lower than those of NPO. Although NPQ-Alb increased with doses above 12.5 mg naphthalene/kg body wt. (b.w.), levels of NPQ-Hb barely increased above the background levels within the dose range examined. The shapes of the dose response curves for total cysteinyl adducts (combined NPO and NPQ) in Hb and Alb were consistent with previous results of radiobinding experiments in naphthalene-dosed mice. Dose-specific levels of NPO-Alb remained essentially constant in mice over the dose range of 25-200 mg/kg b.w. while those of 1,2- and 1,4-NPQ-Alb diminished over this range. Comparing dose-specific levels of NPO-Alb in Swiss Webster mice with those published previously in F344 rats suggests that glutathione depletion in mice occurred at about 1/8th the administered dose previously observed in rats. This suggests that mice could be more susceptible than rats to the toxic effects of naphthalene due to more pronounced depletion of glutathione at a given dose.  相似文献   

18.
Incubation of benzo[alpha] pyrene 4,5-oxide with poly(G) in neutral aqueous ethanol resulted in the formation of covalent adducts and in the production of free 4-hydroxybenzo[alpha]pyrene. This phenol, which was identified by its UV spectral properties and by its chromatographic characteristics, was also formed but at a much slower rate when the epoxide was incubated with DNA or with GMP. Phenol formation was not detected when benzo[alpha]-pyrene 4,5-oxide was incubated for prolonged periods in the presence of poly(A), poly(C) or poly(U) or in the absence of nucleic acid. Formation of 4-hydroxybenzo[alpha] pyrene from the epoxide in the presence of poly(G) was not accompanied by detectable base modifications or by breakage of phosphodiester linkages.  相似文献   

19.
Oxidative stress is one of the major causes of cellular injury. Various reactive oxygen (ROS) and nitrogen (RNS) species such as superoxide, hydroxyl radical, peroxynitrite, and nitric oxide are involved in the manifestations of different types of organ toxicity and the resultant syndromes, symptoms, or diseases. Hypothermic conditions have been reported to reduce the oxidative stress in various in vitro and in vivo studies. In the present study, we sought to determine the effect of lowered temperatures on oxidative stress-induced cell death in Chinese hamster ovary (CHO) cells. We also investigated the oxidative stress-induced alterations in the expression of anti-apoptotic protein, bcl-2, in CHO cells at lowered temperatures. CHO cells were incubated at four different temperatures of 30, 32, 35, and 37 degrees C (control temperature) from 1 to 4 d. In another set, the cells were incubated with 100 microM hydrogen peroxide (H(2)O(2)) for 30 min before harvesting at different time points. The cells were harvested at 1, 2, 3, and 4 d. Cell survival was significantly higher at 30 degrees C as compared to 37 degrees C over 4 d of incubation. In cells incubated with H(2)O(2), significantly higher cell viability was observed at lower temperatures as compared to the cells incubated at 37 degrees C. The activity of glutathione peroxidase (GSH-Px) also increased significantly at lower temperatures. Lowered temperature also provided a significant increase in the expression of anti-apoptotic protein, bcl-2 after 4 d of incubation. These data suggest that hypothermic conditions lowers the risk of oxidative stress-induced cellular damage and programmed cell death by increasing the activity of GSH-Px and by the induction in the expression of the anti-apoptotic protein, bcl-2.  相似文献   

20.
This preliminary study reports the functional capacities of freshly isolated human hepatocytes in regard to their energetic metabolism and monooxygenase activities. Incubated for 30 or 60 min, isolated cells maintain their membrane integrity, AIP and reduced glutathione content and redox potential estimated by means of lactate to pyruvate and ß-hydroxypyruvate to acetoacetate ratios. Three monooxygenase activities, supported by different isoenzymes of cytochrome P30 are determined by the accumulation of unconjugated metabolites: their relative magnitudes are similar to those observed in microsomes, indicating a good preservation of hydroxylase activities during cell isolation and incubation. Although incubations did not exceed 60 min, one can conclude that human hepatocytes maintain their viability and metabolic capacities after isolation and might be considered in transplantation process for the treatment of acute hepatic failure. Isolated human hepatocytes might be also used as a tool for studying biochemical and toxicological effects of a drug.  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