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 335 毫秒
1.
1. The chief sulphur-containing metabolite of styrene and sytrene oxide in the rabbit and rat is chromatographically identical with N-acetyl-S-(beta-hydroxyphenethyl)-l-cysteine and this compound is also formed, together with N-acetyl-S-phenethyl-l-cysteine, as a metabolite of phenethyl bromide. 2. The amounts of the phenethylmercapturic acid and its hydroxy derivative excreted in the urine of animals dosed with phenethyl bromide, styrene, styrene oxide, phenyl glycol, S-phenylethylcysteine and phenethylmercapturic acid have been determined. 3. Liver slices convert phenethylcysteine and phenethylmercapturic acid into N-acetyl-S-(beta-hydroxyphenethyl)-l-cysteine. 4. Methods for the determination by gas-liquid chromatography of mandelic acid and hippuric acid, which are metabolites of some of the compounds studied, are described.  相似文献   

2.
When styrene oxide was incubated with glutathione in the presence of rat liver cytosolic fraction, two conjugates were formed. Structural investigation by mass spectrometry (MS), proton magnetic resonance (PMR) analysis and chemical fragmentation showed the presence of two positional isomers, namely S-(1-phenyl-2-hydroxyethyl)glutathione and S-(2-phenyl-2-hydroxyethyl)glutathione in a ratio of approx. 60 : 40.  相似文献   

3.
The kinetics of the urinary excretion of mandelic and phenylglyoxylic acids were studied in volunteers exposed to the known concentrations of styrene vapour. The level and the time of exposure were suitably changed to simulate situations in the industrial environment. The aim was to find out the reasons for the contradictory reports in the literature and to verify parameters characterizing the course of excretion of both metabolites. It was found that the course of mandelic acid excretion might be influenced by the length of styrene exposure. If exposure was longer than 4 hours the maximum of excretion was at the end of the exposure time; after short-term exposures (4 h or less) it was somewhat delayed. Maximum excretion of phenylglyoxylic acid was delayed both after short-term and 8-hour exposures. Excretion of the metabolites was diphasic (biexponential). The effective half-lives were found to be independent of the level of exposure. The apparent half-lives (determined in the post-exposure time of 0-16 hours) tended to become prolonged at daily repeated exposures. The ratio of mandelic to phenylglyoxylic acid changed considerably with the level of exposure. In biological monitoring it is advisable to determine both metabolites.  相似文献   

4.
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   

5.
A high-performance liquid chromatographic (HPLC) assay for measuring cytosolic glutathione S-transferase activity with styrene oxide is described. After incubating lung or liver cytosol with reduced glutathione and styrene oxide, unreacted styrene oxide is extracted into ethyl acetate. An aliquot of the aqueous phase is evaporated to dryness and reconstituted in the mobile phase for HPLC analysis. The two glutathione conjugates of styrene oxide [S-(1-phenyl-2-hydroxyethyl)glutathione and S-(2-phenyl-2-hydroxyethyl)glutathione] are separated in less than 10 min; quantitation of transferase activity is based on the comparison of the UV absorbance of the two conjugates at 254 nm with synthetic conjugate standards. As little as 1 nmole of either conjugate can be quantitated with good precision. This assay has advantages over previously published methods for measuring styrene oxide glutathione S-transferase activity as it does not depend on the use of relatively unstable and expensive radiolabelled substrates.  相似文献   

6.
Acrylamide (AA), a widely used industrial monomer which is categorised to be carcinogenic, was found to be generated in starch-containing foods during the heating process. This discovery has caused reasonable concern about possible health risks to humans due to dietary acrylamide uptake. In order to gain more information on human metabolism of acrylamide and to contribute to the assessment of the human carcinogenic risk due to AA uptake we measured the mercapturic acid of AA and its epoxide glycidamide (GA) i.e. N-acetyl-S-(2-carbamoylethyl)-L-cysteine (AAMA) and N-(R,S)-acetyl-S-(2-carbamoyl-2-hydroxyethyl)-L-cysteine (GAMA) in human urine. The relation between AAMA and GAMA is important in this context because GA is thought to be the ultimate carcinogenic metabolite of AA. The median levels in smokers (n=13) were found to be about four times higher than in non-smokers (n=16) with median levels of 127 microg/l versus 29 microg/l for AAMA and 19 microg/l versus 5 microg/l for GAMA. Therefore cigarette smoke proved to be an important source of acrylamide exposure. The level of AAMA in the occupationally non-exposed collective (n=29) ranged from 3 to 338 microg/l, the level of GAMA from 相似文献   

7.
In this investigation Salmonella typhimurium strain TA 1530 and TA 1535 were combined with isolated perfused rat liver. Samples of perfusate and bile produced were tested for mutagenicity after treatment with 1,2-dichloroethane (DCE), 1,2-dibromoethane (DBE) or 2-chloroethanol. The results are in good agreement with our previous experiments which indicate that both DEC and DBE are activated through conjugation with glutathione (GSH). Most GSH conjugates are normally excreted in bile. Following liver perfusion the bile was highly mutagenic after DCE and DBE treatments, while 2-chloroethanol did not have this effect. The highest mutagenic effect was seen 15--30 min after the addition of DCE or DBE. The production of mutagenic bile also occurred in mice treated in vivo with DCE. One possible metabolic endproduct of a GSH conjugate is the corresponding mercapturic acid. Thus synthetic N-acetyl-S-(2-chloroethyl)-L-cysteine was tested on TA 1535 and found to be as mutagenic as S-(2-chloroethyl)-L-cysteine in the concentration range 0.2--0.6 mumol/plate. Differences and similarities in the metabolism of DCE and vinyl chloride are discussed on the basis of these results.  相似文献   

8.
Metabolic disposition of ethylene oxide, dibromoethane, and acrylonitrile in rats after acute exposure was studied by examining the relationship between dose and urinary metabolites, and by establishing the influence of a glutathione precursor, L-2-oxothiazolidine-4-carboxylic acid (OTCA), on the above relationship. Respective urinary metabolites, hydroxyethylmercapturic acid, cyanoethylmercapturic acid, thiocyanate, and ethylene glycol, were quantified to estimate the extent to which each compound was metabolized. The animals were given either ethylene oxide (0.34, 0.68, or 1.36 mmol/kg), dibromoethane (0.2, 0.4, or 0.6 mmol/kg), or acrylonitrile (0.10, 0.38, or 0.76 mmol/kg). Urine samples were collected at 24 h. The metabolic biotransformation of all three chemicals to their respective mercapturic acids was strongly indicative of saturable metabolism. Administration of OCTA (4-5 mmol/kg) enhanced gluthathione availability and increased excretion of urinary mercapturic acids at the higher doses of the chemicals. This study indicates that OTCA increases the capacity for detoxification via the glutathione pathway thereby partially correcting the nonlinearity between the administered dose of ethylene oxide, dibromoethane, and acrylonitrile and the amount of certain urinary metabolites.  相似文献   

9.
Because of the low hepatic activity of gamma-glutamyl-transferase in the rat, the liver is generally considered to play only a minor role in the degradation of glutathione conjugates, a limiting step in mercapturic acid formation. Recent findings indicate, however, that the liver has a prominent role in glutathione catabolism, particularly in species other than rat. To examine the contributions of liver to mercapturic acid biosynthesis, mercapturate formation was compared in isolated perfused livers from rats and guinea pigs dosed with either 0.3 or 3.0 mumol of 1-chloro-2,4-dinitrobenzene (CDNB). Chemically synthesized glutathione conjugate, mercapturic acid, and intermediary metabolites of CDNB were used as standards in the high performance liquid chromatography analysis of bile and perfusate samples. Biliary excretion accounted for almost all of the recovered metabolites. A marked species difference was observed in the pattern of CDNB metabolism. Rat livers dosed with 0.3 mumol of CDNB excreted 55% of total biliary metabolites as the glutathione conjugate and 8.2% as the mercapturic acid, whereas guinea pig livers excreted only 4.8% as the glutathione conjugate and 47% as the mercapturate. Mercapturic formation was also dose-dependent, with a larger fraction formed at the 0.3- versus the 3.0-mumol dose (8.2 versus 3.7% in the rat; 47 versus 19% in the guinea pig). Hepatic conversion of the glutathione conjugate to the mercapturic acid was markedly inhibited in both species after retrograde intrabiliary infusion of acivicin, an inhibitor of gamma-glutamyltransferase activity. These findings provide direct evidence for intrahepatic biosynthesis of mercapturic acids. Thus, glutathione conjugates synthesized within hepatocytes are secreted into bile and broken down to cysteine conjugates; the latter are then presumably reabsorbed by the liver, N-acetylated to form the mercapturic acid and re-excreted into bile.  相似文献   

10.
Using dietary administration, mice were exposed to eight substances known to cause peroxisome proliferation (i.e. clofibrate clofibric acid, 2,4-dichlorophenoxyacetic acid, 2,4,5-trichlorophenoxyacetic acid, nafenopin, ICI-55.897, S-8527 and Wy-14.643) or the related substance p-chlorophenoxyacetic acid (group A). Other animals received di(2-ethylhexyl)phthalate, mono(2-ethylhexyl)phthalate, 2-ethylhexanoic acid, or one of 12 other metabolically and/or structurally related compounds (group B). The effects of these treatments on liver cytosolic and microsomal epoxide hydrolases, microsomal cytochrome P-450, cytosolic glutathione transferase activity, the liver-somatic index and the protein contents of the microsomal and cytosolic fractions prepared from liver were subsequently monitored. In general, peroxisome proliferation was accompanied by increases in cytosolic epoxide hydrolase activity. Many peroxisome proliferators also caused increases in microsomal epoxide hydrolase activity, although the correlation was poorer in this case. Immunochemical quantitation by radial immunodiffusion demonstrated that the increases observed in both of these enzyme activities reflected equivalent increases in enzyme protein, i.e. that induction truly occurred. Induction of total microsomal cytochrome P-450 was obtained after dietary exposure to clofibrate, clofibric acid, 2,4-dichlorophenoxyacetic acid, 2,4,5-trichlorophenoxyacetic acid, nafenopin, Wy-14.643, di(2-ethylhexyl)phthalate and di(2-ethylhexyl)phosphate. The most pronounced effects on cytosolic glutathione transferase activity were the decreases obtained after treatment with clofibrate, clofibric acid and Wy-14.643. Our results, together with those reported by others, suggest that the processes of peroxisome proliferation and induction of cytosolic epoxide hydrolase are intimately related. One possible explanation for this is presented.  相似文献   

11.
Recent studies have described lipid peroxidation to be an early and sensitive consequence of cadmium exposure, and free radical scavengers and antioxidants have been reported to attenuate cadmium-induced toxicity. These observations suggest that cadmium produces reactive oxygen species that may mediate many of the untoward effects of cadmium. Therefore, the effects of cadmium (II) chloride on reactive oxygen species production were examined following a single oral exposure (0.50 LD50) by assessing hepatic mitochondrial and microsomal lipid peroxidation, glutathione content in the liver, excretion of urinary lipid metabolites, and the incidence of hepatic nuclear DNA damage. Increases in lipid peroxidation of 4.0- and 4.2-fold occurred in hepatic mitochondria and microsomes, respectively, 48 h after the oral administration of 44 mg cadmium (II) chloride/kg, while a 65% decrease in glutathione content was observed in the liver. The urinary excretion of malondialdehyde (MDA), formaldehyde (FA), acetaldehyde (ACT), and acetone (ACON) were determined at 0–96 h after Cd administration. Between 48 and 72 h posttreatment maximal excretion of the four urinary lipid metabolites was observed with increases of 2.2- to 3.6-fold in cadmium (II) chloride-treated rats. Increases in DNA single-strand breaks of 1.7-fold were observed 48 h after administration of cadmium. These results support the hypothesis that cadmium induces production of reactive oxygen species, which may contribute to the tissue-damaging effects of this metal ion.  相似文献   

12.
The chemical reaction between (±)-styrene oxide and N-acetylcysteine produces both positional isomers (1 and 2) as a mixture of diastereoisomers with a preference for the benzylic thioether isomer 1 (2 : 1). Synthesis of the mercapturic acid conjugates from either (+)- or (?)-styrene oxide produces only two of the four possible stereoisomers. The single diastereoisomers of 1 and 2 were separated by high pressure liquid chromatography (HPLC) and identified by 1H- and 13C-nuclear magnetic resonance (NMR). The relative stereochemistry at the benzylic carbon center of the mercapturic acid conjugates was assigned on the basis of the established chemical correlation between optically pure styrene oxide and its precursor mandelic acid, and considerations on the mechanism of ring opening of epoxides by sulfur nucleophiles. The stereochemical definition of the isomers 36 should prove useful in investigations of the biotransformation of the glutathione (GSH) conjugates of styrene oxide.  相似文献   

13.
The effects of levonorgestrel treatment (4 micrograms/day per kg body weight 0.75 for 18 days) on rate-limiting enzymes of hepatic triacylglycerol synthesis, namely glycerol-3-phosphate acyltransferase and phosphatidic acid phosphatase were investigated in microsomal, mitochondrial and cytosolic fractions of rat liver. Levonorgestrel treatment resulted in a significant reduction (26%) of hepatic microsomal glycerol-3-phosphate acyltransferase specific activity. Hepatic mitochondrial glycerol-3-phosphate acyltransferase specific activity was unchanged. Levonorgestrel treatment also significantly reduced (by 20%) the specific activity of hepatic microsomal magnesium-independent phosphatidic acid phosphatase. However, magnesium-dependent phosphatic acid phosphatase specific activities in microsomal and cytosolic fractions were unaffected. Cytosolic magnesium-independent phosphatidic acid phosphatase activity was also unchanged. These studies are consistent with the view that levonorgestrel lowers serum triacylglycerol levels, at least in part, by inhibition of the glycerol-3-phosphate acyltransferase (EC 2.3.1.15) step in hepatic triacylglycerol synthesis.  相似文献   

14.
The primary sensory neurons of the olfactory system are chronically exposed to the ambient environment and may therefore be susceptible to damage from occupational exposure to many volatile chemicals. To investigate whether occupational exposure to styrene was associated with olfactory impairment, we examined olfactory function in 2 groups: workers in a German reinforced-plastics boat-manufacturing facility having a minimum of 2 years of styrene exposure (15-25 ppm as calculated from urinary metabolite concentrations, with historical exposures up to 85 ppm) and a group of age-matched workers from the same facility with lower styrene exposures. The results were also compared with normative data previously collected from healthy, unexposed individuals. Multiple measures of olfactory function were evaluated using a standardized battery of clinical assessments from the Monell-Jefferson Chemosensory Clinical Research Center that included tests of threshold sensitivity for phenylethyl alcohol (PEA) and odor identification ability. Thresholds for styrene were also obtained as a measure of occupational olfactory adaptation. Styrene exposure history was calculated through the use of past biological monitoring results for urinary metabolites of styrene (mandelic acid [MA], phenylglyoxylic acid [PGA]); current exposure was determined for each individual using passive air sampling for styrene and biological monitoring for styrene urinary metabolites. Current mean effective styrene exposure during the day of olfactory testing for the group of workers who worked directly with styrene resins was 18 ppm styrene (standard deviation [SD] = 14), 371 g/g creatinine MA + PGA (SD = 289) and that of the group of workers with lower exposures was 4.8 ppm (SD = 5.2), 93 g/g creatinine MA+PGA (SD = 100). Historic annual average exposures for all workers were greater by a factor of up to 6x. No differences unequivocally attributable to exposure status were observed between the Exposed and Comparison groups or between performance of either group and normative population values on thresholds for PEA or odor identification. Although odor identification performance was lower among workers with higher ongoing exposures, performance on this test is not a pure measure of olfactory ability and is influenced by familiarity with the stimuli and their sources. Consistent with exposure-induced sensory adaptation, however, elevated styrene thresholds were significantly associated with higher occupational exposures to styrene. In summary, the present study found no evidence among a cross-section of reinforced-plastics workers that current or historical exposure to styrene was associated with a general impairment of olfactory function. When taken together with prior studies of styrene-exposed workers, these results suggest that styrene is not a significant olfactory toxicant in humans at current exposure levels.  相似文献   

15.
Styrene's capacity to induce chromosomal aberrations was studied in bone marrow cells of CD1 male mice. No mutagenic effect could be detected after either a 4-day treatment course with daily oral doses of 500 mg/kg or a 70-day course with daily oral doses of 200 mg/kg. Urinary elimination of styrene metabolites related to styrene-7,8-oxide formation (i.e. phenylethylene glycol, mandelic acid, benzoic acid, phenylglyoxylic acid and total mercapturic acids) was quantitatively evaluated in the group of mice given the 200 mg/kg dose. In parallel, kinetic studies were made on styrene and styrene-7,8-oxide blood concentrations in the same group of animals. These determinations were carried out on days 1 and 70 of treatment by spectrophotometric, gas chromatographic and mass fragmentographic procedures.Not even nanograms of styrene-7,8-oxide were found in the blood of styrene-treated mice. This suggests that the metabolite does not migrate from the cellular compartment where it is formed being immediately metabolized or irreversibly bound to cellular structures.This observation could well explain the lack of mutagenic effects observed.  相似文献   

16.
The nephrotoxic gas chlorotrifluoroethylene is a substrate for glutathione S-transferase activity in rat hepatic cytosolic and microsomal fractions. The rates of reaction, determined by measuring glutathione disappearance, were 5–15 or 35–70 nmol/min/mg of cytosolic or microsomal protein, respectively. Glutathione disappearance was completely abolished by heat-denaturing the subcellular fractions. A product of the cytosolcatalyzed reaction between chlorotrifluoroethylene and glutathione was isolated and shown by amino acid analysis and 1H- and 19F-NMR to be S-(2-chloro-1,1,2-trifluoroethyl)glutathione. This appears to be the first demonstration of a glutathione S-transferase-catalyzed addition reaction with a halogenated olefin, and this reaction may be of toxicological significance.  相似文献   

17.
Microsomal glutathione S-transferase, UDP-glucuronyl transferase, and aniline hydroxylase activities were determined in liver, renal cortex, and small intestine of control, streptozotocin-diabetic, alloxan-diabetic, and untreated insulin-injected male Wistar rats. Renal microsomal glutathione S-transferase activity showed a direct linear relationship with insulin blood levels, in agreement with our previous report on cytosolic glutathione S-transferase. This result suggests a possible regulatory mechanism of insulin that needs to be further examined. The hepatic microsomal UDP-glucuronyl transferase was only decreased in streptozotocin-diabetic rats and was not restored by insulin treatment. Intestinal UDP-glucuronyl transferase exhibited an opposite response in streptozotocin-treated animals that was not normalized by the administration of insulin. Hepatic aniline hydroxylase showed the same behaviour as intestinal UDP-glucuronyl transferase. These results suggest that streptozotocin and (or) its metabolites have a direct effect on hepatic and intestinal UDP-glucuronyl transferase activity and on hepatic aniline hydroxylase activity. On the other hand, insulin regulation of enzyme activity varies from one organ to another.  相似文献   

18.
In vitro studies have suggested that sporidesmin hepatotoxicity may be related to thiol oxidation and generation of cytotoxic oxygen species. After a single i.p. injection of 2.8 mg/kg bw sporidesmin in guinea-pigs, hepatic and plasma zinc, hepatic metallothionein, cytochromes P-450 and b5, total glutathione and proteins (total, microsomal and cytosolic) were monitored for 21 days. The only variations observed were significant increases in liver concentrations of zinc (cytosolic and total), metallothionein, and cytochromes, which peaked on day 8 after the sporidesmin challenge (+45, 55, 50, 376 and 413%, respectively) and, except for cytochrome b5, went back to control levels before the 21st day. These results suggest that cytochromes P-450 and b5 may be involved in sporidesmin cellular damage.  相似文献   

19.
The streptozotocin-induced short-term (2 week) diabetic rats showed an increase in susceptibility to carbon tetrachloride (CCl4)-induced hepatocellular damage. This diabetes-induced change was associated with a marked impairment in the hepatic glutathione antioxidant/detoxification response to CCl4 challenge, as indicated by the abrogation of the increases in hepatic reduced glutathione (GSH) level, glucose-6-phosphate dehydrogenase and microsomal glutathione S-transferases (GST) activities upon challenge with increasing doses of CCl4. While the hepatic GSH level was increased in diabetic rats, the hepatic mitochondrial GSH level and Se-glutathione peroxidase activity were significantly reduced. Insulin treatment could reverse most of the biochemical alterations induced by diabetes. Both insulin and schisandrin B (Sch B) pretreatments protected against the CCl4 hepatotoxicity in diabetic rats. The hepatoprotection was associated with improvement in hepatic glutathione redox status in both cytosolic and mitochondrial compartments, as well as the increases in hepatic ascorbic acid level and microsomal GST activity. The ensemble of results suggests that the diabetes-induced impairment in hepatic mitochondrial glutathione redox status may at least in part be attributed to the enhanced susceptibility to CCl4 hepatotoxicity. Sch B may be a useful hepatoprotective agent against xenobiotics-induced toxicity under the diabetic conditions. (Mol Cell Biochem 175: 225–232, 1997)  相似文献   

20.
Clinical efficacy of alkylating anticancer drugs, such as chlorambucil (4-[p-[bis [2-chloroethyl] amino] phenyl]-butanoic acid; CHB), is often limited by the emergence of drug resistant tumor cells. Increased glutathione (gamma-glutamylcysteinylglycine; GSH) conjugation (inactivation) of alkylating anticancer drugs due to overexpression of cytosolic glutathione S-transferase (GST) is believed to be an important mechanism in tumor cell resistance to alkylating agents. However, the potential involvement of microsomal GST in the establishment of acquired drug resistance (ADR) to CHB remains uncertain. In our experiments, a combination of lipid chromatography/electrospray ionization mass spectrometry (LC/ESI/MS) was employed for structural characterization of the resulting conjugates between CHB and GSH. The spontaneous reaction of 1mM CHB with 5 mM GSH at 37 degrees C in aqueous phosphate buffer for 1 h gave primarily the monoglutathionyl derivative, 4-[p-[N-2-chloroethyl, N-2-S-glutathionylethyl] amino]phenyl]-butanoic acid (CHBSG) and the diglutathionyl derivative, 4-[p-[2-S-glutathionylethyl] amino]phenyl]-butanoic acid (CHBSG2) with small amounts of the hydroxy-derivative, 4-[p-[N-2-S-glutathionylethyl, N-2-hydroxyethyl] amino]phenyl]-butanoic acid (CHBSGOH), 4-[p-[bis[2-hydroxyethyl] amino]phenyl]-butanoic acid (CHBOH2), 4-[p-[N-2-chloroethyl, N-2-S-hydroxyethyl]amino]phenyl]-butanoic acid (CHBOH). We demonstrated that rat liver microsomal GST presented a strong catalytic effect on these reactions as determined by the increase of CHBSG2, CHBSGOH and CHBSG and the decrease of CHB. We showed that microsomal GST was activated by CHB in a concentration and time dependent manner. Microsomal GST which was stimulated approximately two-fold with CHB had a stronger catalytic effect. Thus, microsomal GST may play a potential role in the metabolism of CHB in biological membranes, and in the development of ADR.  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