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 62 毫秒
1.
Trichloroethylene (Tri) caused modest cytotoxicity in freshly isolated human proximal tubular (hPT) cells, as assessed by significant decreases in lactate dehydrogenase (LDH) activity after 1 h of exposure to 500 microM Tri. Oxidative metabolism of Tri by cytochrome P-450 to form chloral hydrate (CH) was only detectable in kidney microsomes from one patient out of four tested and was not detected in hPT cells. In contrast, GSH conjugation of Tri was detected in cells from every patient tested. The kinetics of Tri metabolism to its GSH conjugate S-(1,2-dichlorovinyl)glutathione (DCVG) followed biphasic kinetics, with apparent Km and Vmax values of 0.51 and 24.9 mM and 0.10 and 1.0 nmol/min per mg protein, respectively. S-(1,2-dichlorovinyl)-L-cysteine (DCVC), the cysteine conjugate metabolite of Tri that is considered the penultimate nephrotoxic species, caused both time- and concentration-dependent increases in LDH release in freshly isolated hPT cells. Preincubation of hPT cells with 0.1 mM aminooxyacetic acid did not protect hPT cells from DCVC-induced cellular injury, suggesting that another enzyme besides the cysteine conjugate beta-lyase may be important in DCVC bioactivation. This study is the first to measure the cytotoxicity and metabolism of Tri and DCVC in freshly isolated cells from the human kidney. These data indicate that the pathway involved in the cytotoxicity and metabolism of Tri in hPT cells is the GSH conjugation pathway and that the cytochrome P-450-dependent pathway has little direct role in renal Tri metabolism in humans.  相似文献   

2.
1. The metabolism of 14C-hexachloro-1,3-butadiene (HCBD) was studied in mice and in subcellular fractions from mouse liver and kidney. 2. In the presence of glutathione (GSH), liver microsomes and cytosol transformed HCBD to S-(pentachlorobutadienyl)glutathione (PCBG). PCBG formation in subcellular fractions from mouse kidney was very limited. Oxidative metabolism of HCBD by cytochrome P-450 could not be demonstrated. 3. Cysteine conjugate beta-lyase was present in mitochondria and cytosol from mouse liver and kidney. 4. After an oral dose of 30 mg/kg 14C-HCBD, mice eliminated 67.5-76.7% of dose in faeces; urinary elimination accounted for 6.6-7.6%. 5. Metabolites of HCBD identified are: S-(pentachlorobutadienyl)glutathione in faeces; S-(pentachlorobutadienyl)-L-cysteine, N-acetyl-S-(pentachlorobutadienyl)-L-cysteine and 1,1,2,3-tetrachlorobutenoic acid in urine. 6. The results suggest that conjugation of HCBD with GSH in liver, followed by renal processing of the glutathione S-conjugates and beta-lyase-catalysed formation of reactive intermediates, accounts for the organ specific toxicity of HCBD in mice.  相似文献   

3.
Mechanisms of 1,1-dichloroethylene-induced cytotoxicity in lung and liver   总被引:2,自引:0,他引:2  
Exposure to 1,1-dichloroethylene (DCE) causes lung and liver toxicities in mice. The lesions are characterized by damage preferentially to bronchiolar Clara cells in the lung and necrosis of centrilobular hepatocytes in the liver. The primary metabolites formed from DCE in lung and liver microsomal incubations are the epoxide, 2,2-dichloroacetaldehyde and 2-chloroacetyl chloride, which undergo hydrolysis and/or conjugation with glutathione (GSH). The major products formed are the epoxide-derived GSH conjugates 2-(S-glutathionyl) acetyl glutathione [B] and 2-S-glutathionyl acetate [C]. The hydrate of 2,2-dichloroacetaldehyde (acetal) is also detected. These metabolites are detected in vivo in murine lung and liver cytosol and in bile, and importantly, also in human lung and liver microsomal incubations. Formation of the epoxide is mediated mainly by CYP2E1. Immunohistochemical studies localized the epoxide-derived GSH conjugate [C] and cysteine-containing proteins in Clara cells and centrilobular hepatocytes. These findings are consistent with the premise that the lung and liver cytotoxicities induced by DCE are associated with in situ formation of the epoxide within the target cells.  相似文献   

4.
Glutathione (GSH) conjugation of the separate alpha-bromoisovalerylurea (BIU) enantiomers was studied in the rat. Administration of (R)-BIU resulted in excretion of a single glutathione conjugate in bile (IU-S-G/I) and a single mercapturate in urine (IU-S-MA/B). The other enantiomer, (S)-BIU, was exclusively metabolized to the other diastereomeric conjugates, IU-S-G/II and IU-S-MA/A. Thus, the conjugation of BIU with glutathione was completely stereospecific. Both the GSH conjugate and mercapturate derived from (R)-BIU were excreted two to three times more rapidly than their diastereomeric (S)-BIU counterparts. The enantiomers did not influence each others metabolism as reflected by identical metabolite excretion rates when the BIU enantiomers were administered either separately or as the racemic mixture. A similar rate difference for GSH conjugation of the separate BIU enantiomers was observed in incubations with rat liver cytosol as source of GSH transferases, suggesting that the stereoselectivity in vivo was due to glutathione conjugation properly. Similar results were obtained with a rat liver microsomal fraction, indicating that microsomal GSH transferases are active towards BIU and have a similar stereoselectivity as the cytosolic enzymes. Comparison of the GSH conjugation of BIU with that of its analogue alpha-bromoisovaleric acid (BI, which lacks the amide-linked urea group) revealed an opposite stereoselectivity: while (R)-BIU was conjugated faster than (S)-BIU, the (R) enantiomer of the acid was conjugated more slowly than (S)-BI. The alpha-bromocarbonyl compounds BI and BIU present a new type of substrate for the GSH transferases and allow studies of these enzymes in intact organisms as well as investigations on the stereoselectivity of GSH conjugation.  相似文献   

5.
It is not clear whether glutathione (GSH) conjugation to 1,2-dibromo-3-chloropropane (DBCP) results in genotoxic activation. Therefore S9, cytosolic, and microsomal fractions from uninduced rat liver were evaluated for their relative ability to activate DBCP in a modified Ames system. The S9 enzymes, either alone or in combination with exogenous GSH, did not enhance the mutagenicity of DBCP; identical results were obtained with cytosolic enzymes. Significant mutagenic activation of DBCP was produced by either S9 or microsomal fractions in the presence of NADPH. Activation was proportional to cytochrome P-450 concentrations, and was diminished by exogenous GSH. The protection against genotoxicity exerted by GSH did not require cytosolic glutathione S-transferases (GST). Thus, mutagenic activation of DBCP as obtained with S9 fractions is primarily due to biotransformation by microsomal rather than by cytosolic enzymes. Kinetic studies of cytosol-catalyzed conjugation of GSH to DBCP revealed tissue-specific differences in apparent Km and Vmax. Renal and testicular GSTs were associated with 28-46% smaller Vmax values when compared to hepatic GSTs (31.2 +/- 1.9 nmol/min X mg protein). However, renal and testicular GSTs had relatively higher affinities for DBCP. Thus, extrahepatic tissues possess significant capacity to conjugate GSH to DBCP. DBCP-GSH conjugates may undergo enzymatic modification by extrahepatic peptidase and beta-lyase to yield other sulfur-containing moieties that perhaps mediate DBCP's extrahepatic toxicity.  相似文献   

6.
1. The metabolism of 14C-hexachloro-1,3-butadiene (HCBD) was studied in mice and in subcellular fractions from mouse liver and kidney.

2. In the presence of glutathione (GSH), liver microsomes and cytosol transformed HCBD to S-(pentachlorobutadienyl)glutathione (PCBG). PCBG formation in sub-cellular fractions from mouse kidney was very limited. Oxidative metabolism of HCBD by cytochrome P-450 could not be demonstrated.

3. Cysteine conjugate β-lyase was present in mitochondria and cytosol from mouse liver and kidney.

4. After an oral dose of 30 mg/kg 14C-HCBD, mice eliminated 67˙5-76˙7% of dose in faeces; urinary elimination accounted for 6˙6-7˙6%.

5. Metabolites of HCBD identified are: S-(pentachlorobutadienyl)glutathione in faeces; S-(pentachlorobutadienyl)-L-cysteine, N-acetyl-S-(pentachlorobutadienyl)-L-cysteine and 1,1,2,3-tetrachlorobutenoic acid in urine.

6. The results suggest that conjugation of HCBD with GSH in liver, followed by renal processing of the glutathione S-conjugates and β-lyase-catalysed formation of reactive intermediates, accounts for the organ specific toxicity of HCBD in mice.  相似文献   

7.
Previous lab studies implicated the sulfoxidation pathway of molinate metabolism to induce testicular toxicity. Once molinate is metabolized to molinate sulfoxide, it undergoes further phase II metabolism either spontaneously, enzyme catalyzed, or both to form glutathione-conjugated molinate. This study compared the metabolic capability of rat and human liver cytosol to form a glutathione (GSH)-conjugated metabolite of molinate. The GSH conjugation of molinate sulfoxide in rat cytosol was described by the constants Km of 305 microM and Vmax of 4.21 nmol/min/mg cytosol whereas the human values were 91 microM and 0.32 nmol/min/mg protein for Km and Vmax, respectively. At the same 1 mM GSH concentration, the in vitro bimolecular nonenzymatic rate constant of 3.02 x 10(-6) microM(-1) min(-1) was calculated for GSH conjugation of molinate sulfoxide. Specific activity for rat and human glutathione transferase was calculated to equal 1.202 +/- 0.25 and 0.809 +/- 0.45 micromol/min/mg protein, respectively by 1-chloro-2,4-dinitrobenzene (CDNB) assay. Compared to a conventional GSH depletion model (BSO + DEM combination), molinate alone was nearly as effective in reducing GSH levels by approximately 90 and 25% in liver and testes, respectively. The impact of molinate sulfoxide's ability to adduct glutathione transferase and inhibit the production of the glutathione conjugated metabolite was examined and found to be negligible.  相似文献   

8.
Hexachloro-1,3-butadiene (HCBD) is a substrate for the hepatic microsomal glutathione transferases and is metabolised at higher rates by these enzymes than their cytosolic counterparts. Conjugation reactions catalysed by the microsomal and cytosolic transferases have been studied and characterized using this substrate and 1-chloro-2,4-dinitrobenzene (CDNB). In rat liver microsomes the Km values for HCBD and CDNB were 0.91 and 0.012 mM and in cytosol 0.51 and 0.10 mM respectively. Vmax values for HCBD were 1.39 and 0.35 nmol conjugate formed/min/mg protein for microsomes and cytosol respectively. In microsomal systems HCBD was a potent competitive inhibitor of the metabolism of CDNB with a Ki value of approximately 10 microM. However, CDNB did not inhibit HCBD metabolism significantly. These data suggest that more than one microsomal enzyme is involved in HCBD metabolism. The microsomal membrane could be solubilized without significant inhibition of HCBD activity; however, some detergents did inhibit the conjugation reaction. Activity was also lost on treatment of microsomal membranes with trypsin indicating the enzyme is localized on the cytoplasmic surface of the endoplasmic reticulum. Pretreatment of the rats with Aroclor 1254, 3-methylcholanthrene or phenobarbital did not change the microsomal conjugation of HCBD or CDNB with glutathione. Of seven species investigated, a human liver sample showed the highest ratio of microsomal to cytosolic glutathione transferase activity for HCBD (in microsomes 40-fold higher specific activity than in cytosol). Glutathione conjugation appears to play a critical role in the toxicity and carcinogenicity of some halogenated hydrocarbons. These data substantiate the potentially important role for the microsomal glutathione transferase in catalysing these reactions.  相似文献   

9.
《Toxicology letters》1995,76(2):145-153
We compared the effects of 3 dipyridyl isomers, 2,2′-dipyridyl, 2,4′-dipyridyl and 4,4′-dipyridyl, on hepatic microsomal heme oxygenase and drug-metabolizing enzyme activities in male rats. 2,2′-Dipyridyl increased cytochrome P450 (P450) content at lower doses, but decreased with increasing dose levels. Immunoblot analysis revealed that 2,2′-dipyridyl did not induce both P450 1A1/2 and P450 2B1/2, in contrast to 2,4′- and 4,4′-dipyridyls, both of which were inducers of either P450 1A1/2 and/or P450 2B1/2. Some drug-metabolizing enzyme activities gradually declined with the increasing dose level of 2,2′-dipyridyl. 2,2′-Dipyridyl was able to induce hepatic microsomal heme oxygenase in a dose-dependent manner, but 2,4′- and 4,4′-dipyridyls did not, even at the highest dose (0.80 mmol/kg) examined. Treatment of rats with 2,2′-dipyridyl resulted in the increase of glutathione (GSH) content in a dose-dependent manner, but not 4-substituted isomers. A time course study with 2,2′-dipyridyl revealed that it produced a significant decrease in hepatic GSH content at early time periods (2–6 h) after its administration with an inverse increase in heme oxygenase activity. The present investigation has revealed that in contrast to the induction of P450 by 4-substituted dipyridyl compounds, 2,2′-dipyridyl is a novel inducer of hepatic microsomal heme oxygenase, together with the change in hepatic GSH content. This study would provide information on the differential effects of simple dipyridyl isomers on hepatic enzymes involved in heme and drug metabolism.  相似文献   

10.
The enzyme kinetics of the glutathione (GSH) conjugation of cryptophycin 52 (C52, R-stereoisomer) and cryptophycin 53 (C53, S-stereoisomer) by cytosolic glutathione S-transferases (cGSTs) from human, rat, mouse, dog and monkey liver were studied. Vmax, Km, and CLint values for glutathione conjugation of C52 (R-stereoisomer) were 0.10 +/- 0.01 nmol min-1 mg-1, 3.24 +/- 0.23 microM, and (3.15 +/- 0.09) x 10(-2) ml min-1 mg-1, respectively, in human cytosol. Due to limited solubility relative to the Km, only CLint values were determined in rat ((7.76 +/- 0.10) x 10-2 ml min-1 mg-1) and mouse ((7.61 +/- 0.50) x 10(-2) ml min-1 mg-1) cytosol. Enzyme kinetic parameters could not be determined for C53 (S-stereoisomer). Microsomal GSH conjugation in human, rat, and mouse was attributed to cytosolic contamination. No GSH conjugation was seen in any biological matrix from dog or monkey. There was little GSH conjugation of C53 by cytosol or microsomes from any species. The metabolism of C52 and C53 by epoxide hydrolase was also investigated. No diol product was observed in any biological matrix from any species. Thus, cGSTs are primarily responsible for C52 metabolism.  相似文献   

11.
The drinking water disinfection byproduct bromodichloromethane (CHBrCl(2)) was previously shown to be mutagenic in Salmonella typhimurium that overexpress rat glutathione transferase theta 1-1 (GSTT1-1). Several experimental approaches were undertaken in this study to investigate the DNA covalent binding potential of reactive intermediates generated by GSTT1-1-mediated metabolism of CHBrCl(2). First, rodent hepatic cytosol incubations containing [(14)C]CHBrCl(2), supplemented glutathione (GSH), and calf thymus DNA resulted in approximately 3-fold (rat liver cytosol) and 7-fold (mouse liver cytosol) greater amounts of total radioactivity (RAD) associated with the purified DNA as compared to a control (absence of rodent cytosol) following liquid scintillation counting (LSC) of isolated DNA. The relative increase in DNA labeling is consistent with the conjugation activity of these rodent cytosols toward CHBrCl(2). Second, exposure of GSTT1-1-expressing S. typhimurium to [(14)C]CHBrCl(2) resulted in a concentration-dependent increase of bacterial DNA-associated total radioactivity. Characterization of DNA-associated radioactivity could not be assigned to a specific deoxynucleoside adduct(s) following enzymatic hydrolysis of DNA and subsequent HPLC analysis. A possible explanation for this observation was formation of a 'transient' adduct that was unstable in the DNA isolation and hydrolysis procedures employed. To circumvent problems of adduct instability, reactions of [(14)C]CHBrCl(2) with GSH catalyzed by recombinant rat GSTT1-1 were performed in the presence of calf thymus DNA or, alternatively, the model nucleophile deoxyguanosine. Hydroxyapatite chromatography of [(14)C]-labeled DNA or HPLC chromatography of [(14)C]-labeled deoxyguanosine derivatives demonstrated the covalent binding of [(14)C]CHBrCl(2)-derived metabolites to DNA and deoxyguanosine in low yield (approximately 0.02% of [(14)C]CHBrCl(2) biotransformed by GSTT1-1 resulted in DNA adducts). Cytochrome P450 (CYP)- and GST-catalyzed biotransformation of CHBrCl(2) in rat tissues (kidney and large intestine) that develop tumors following chronic CHBrCl(2) exposure were compared with rat liver (a nontarget tissue). Rat liver had a significant capacity to detoxify CHBrCl(2) (to carbon dioxide) compared with kidney and large intestine as a result of CYP-catalyzed oxidation, liver was approximately 16-fold more efficient than kidney and large intestine when intrinsic clearance values (V(max)/K(m)) were compared. In contrast, the efficiency of GST-mediated GSH conjugation of CHBrCl(2) in kidney and large intestine was only slightly lower than liver (approximately 2- to 4-fold lower), thus, the relative amounts of reactive intermediates that are produced with the capacity to covalently modify DNA may be enhanced in these extrahepatic tissues. The significance of these findings is that conjugation of CHBrCl(2) with GSH can result in the covalent modification of DNA and that cancer target tissues in rats have a much reduced detoxification capacity, but only a modest decrease in bioactivation capacity, as compared to the liver (a nontarget tissue in rats).  相似文献   

12.
The toxicity and metabolism of trichloroethylene (TRI) were studied in renal proximal tubular (PT) and distal tubular (DT) cells from male Fischer 344 rats. TRI was slightly toxic to both PT and DT cells, and inhibition of cytochrome P450 (P450; substrate, reduced-flavoprotein:oxygen oxidoreductase [RH-hydroxylating or -epoxidizing]; EC 1.14.14.1) increased TRI toxicity only in DT cells. In untreated cells, glutathione (GSH) conjugation of TRI to form S-(1,2-dichlorovinyl)glutathione (DCVG) was detected only in PT cells. Inhibition of P450 transiently increased DCVG formation in PT cells and resulted in detection of DCVG formation in DT cells. Formation of DCVG in PT cells was described by a two-component model (apparent Vmax values of 0.65 and 0.47 nmol/min per mg protein and Km values of 2.91 and 0.46 mM). Cytosol isolated from rat renal cortical, PT, and DT cells expressed high levels of GSH S-transferase (GST; RX:glutathione R-transferase; EC 2.5.1.18) alpha (GSTalpha) but not GSTpi. Low levels of GSTmu were detected in cortical and DT cells. Purified rat GSTalpha2-2 exhibited markedly higher affinity for TRI than did GSTalpha1-1 or GSTalpha1-2, but each isoform exhibited similar VmaX values. Triethyltinbromide (TETB) (9 microM) inhibited DCVG formation by purified GSTalpha-1 and GSTalpha2-2, but not GSTalpha1-2. Bromosulfophthalein (BSP) (4 microM) only inhibited DCVG formation by GSTalpha2-2. TETB and BSP inhibited approximately 90% of DCVG formation in PT cytosol but had no effect in DT cytosol. This suggests that GSTalpha1-1 is the primary isoform in rat renal PT cells responsible for GSH conjugation of TRI. These data, for the first time, describe the metabolism of TRI by individual GST isoforms and suggest that DCVG feedback inhibits TRI metabolism by GSTs.  相似文献   

13.
At the present time, comprehensive metabolism studies of 2,3-dichloro-1-propene (2,3-DCP) have not yet been reported. We have investigated the biotransformation of 2,3-DCP using female Wistar rats in order to elucidate the bioactivation mechanisms. 175 mg/kg, 1,3-14C-2,3-DCP in corn oil was administered to a rat. The animal was killed 20 hr later. Approximately 56.7% of the radioactivity was excreted in the urine, 1.6% in the feces, 5.3% was exhaled as unchanged 2,3-DCP, and 0.3% as CO2. 31.3% remained in the organs and the carcass. Three metabolic pathways were established. 1) Conjugation with GSH leading to S-(2-chloro-2-propenyl)mercapturic acid. 2) The P450 induced epoxidation with subsequent rearrangement to highly mutagenic 1,3-dichloroacetone. 1,3-Dichloroacetone was further converted to the dimercapturic acid, 1,3-(2-propanone)-bis-S-(N-acetylcysteine). 3) The hydrolysis to 2-chloroallyl alcohol followed by alcohol dehydrogenase catalyzed formation of highly mutagenic 2-chloroacrolein. The 2-chloroallyl alcohol is excreted directly in the urine and as the glucuronide. 2-Chloroacrolein is further oxidized to 2-chloroacrylic acid which is also excreted in the urine.  相似文献   

14.
S-(2-Hydroxy-3,4-epoxybutyl)glutathione (DEB-GSH conjugate) is formed from the reaction of 1,2:3,4-diepoxybutane (DEB) with glutathione (GSH), and the conjugate is considerably more mutagenic than several other butadiene-derived epoxides-including DEB-in Salmonella typhimurium TA1535 [Cho, S.-H., (2010) Chem. Res. Toxicol. 23, 1544-1546]. We previously identified six DNA adducts in the reaction of the DEB-GSH conjugate with nucleosides and calf thymus DNA and two DNA adducts in livers of mice and rats treated with DEB [Cho, S.-H. and Guengerich, F. P. (2012) Chem. Res. Toxicol. 25, 706-712]. To define the role of GSH conjugation in 1,3-butadiene (BD) metabolism and characterize the mechanism of GSH transferase (GST)-enhanced mutagenicity of DEB, mutation spectra of BD and its metabolites in the absence and presence of GST/GSH and mouse liver microsomes were compared in the rpoB gene of Escherichia coli TRG8. The presence of GST considerably enhanced mutations. The mutation spectra derived from the DEB-GSH conjugate, the DEB/GST/GSH system, and the BD/mouse liver microsomes/GST/GSH system matched each other and were different from those derived from the other systems devoid of GSH. The major adducts in E. coli TRG8 cells treated with the DEB/GST/GSH system, the BD/mouse liver microsomes/GST/GSH system, or the DEB-GSH conjugate were S-[4-(N(7)-guanyl)-2,3-dihydroxybutyl]GSH, S-[4-(N(3)-adenyl)-2,3-dihydroxybutyl]GSH, and S-[4-(N(6)-deoxyadenosinyl)-2,3-dihydroxybutyl]GSH, indicating the presence of the GSH-containing DNA adducts in the systems. These results, along with the strong enhancement of mutagenicity by GST in this system, indicate the relevance of these GSH-containing DNA adducts.  相似文献   

15.
The accepted pathway for MC biotransformation is GSH conjugation, occurring either spontaneously or catalyzed by GST. In the present work, the already available information on human MC metabolism have been expanded and the capacity of human GST to conjugate MC-LR has been confirmed in human liver cytosol. At physiological GSH content the spontaneous reaction predominated on the enzymatic one; the prevalence of the enzymatic reaction occurred following GSH depletion, and the shift was detectable at higher GSH levels, the lower was MC concentration. However, at low MC-LR concentrations (≤10 μM), representative of repeated oral exposure, the relevance of the enzymatic reaction became predominant at GSH concentration between 1 and 2 mM. MC-LR conjugate was detectable at ≥0.5 mM GSH, whereas, with 10 μM MC-RR detectable levels of conjugate were observed at 0.05 mM GSH, a 10-fold lower concentration. Overall, our data indicate that MC-RR is more efficiently conjugated than MC-LR, especially at low concentrations. Cytosol samples from rat and mouse were used to characterize GSH conjugation of MC-LR and MC-RR, and to check for possible species differences. At physiological GSH content, in both rodent species the enzymatic reaction accounted for half of the total conjugate formation, reducing the impact of spontaneous reaction with respect to human. Rat and mouse GST showed similar MC-LR and-RR GSH conjugation, but a two-fold higher catalytic efficiency than human sample. This is mainly due to higher affinity for the substrate, with Kmapp values being an order of magnitude lower in the animal models than in human liver cytosol. More pronounced differences in the metabolism of the two variants were evidenced in rodents than in humans.  相似文献   

16.
1. Three mercapturic acid metabolites were identified in the urine of male and female Fischer 344 rats given 1,2-dichloropropane (DCP) orally (100 mg/kg) or by inhalation exposure (100 ppm, 6 h). 2. These compounds (N-acetyl-S-(2-hydroxypropyl)-L-cysteine, N-acetyl-S-(2-oxopropyl)-L-cysteine and N-acetyl-S-(1-carboxyethyl)-L-cysteine) were isolated from the urine following acidification and extraction with ethyl acetate. The extracts were derivatized with diazomethane and N,O-bis(trimethylsilyl)trifluoroacetamide and analysed by chemical ionization g.l.c.-mass spectrometry. 3. Further mechanistic studies were carried out with the stable isotope-labelled analogue, D6-DCP (105 mg/kg, orally). Analysis of the resulting mass spectra indicated retention of primarily three deuterium atoms in the 2-hydroxypropyl-mercapturic acid formed from D6-DCP. Similar isotope retention was observed for the 2-oxopropyl-mercapturic acid metabolite. 4. These results do not support a sulphonium ion intermediate in the formation of the 2-hydroxypropyl-mercapturic acid metabolite of DCP. Instead, this metabolite is thought to arise via direct oxidation of DCP, either prior to or following conjugation with glutathione.  相似文献   

17.
The metabolism of [14C]tetrachloroethylene (Tetra) and its metabolite S-(1,2,2-trichlorovinyl)-L-cysteine (TCVC) was investigated with in vitro systems to substantiate metabolic pathways of Tetra deduced from in vivo experiments. In the presence of NADPH, rat hepatic microsomal fractions metabolized Tetra to soluble metabolites, which were identified as trichloroacetic acid and oxalic acid by gas chromatography/mass spectroscopy and a metabolite largely bound to microsomal macromolecules. The majority of the alkylated macromolecules were identified as N-trichloroacetylated phospholipids by high performance liquid chromatography and GC/MS. When Tetra was incubated with hepatic microsomes and cytosol in the presence of 10 mM glutathione, but in the absence of NADPH, the formation of a polar metabolite other than trichloroacetic acid and oxalic acid was observed. This metabolite was identified, after hydrolysis to the corresponding cysteine conjugate, as S-(1,2,2-trichlorovinyl)-glutathione (TCVG). Microsomal GSH S-transferases catalyzed TCVG formation more efficiently than cytosolic GSH S-transferases; the competitive substrate 1-chloro-2,4-dinitrobenzene inhibited TCVG formation. In the presence of both NADPH and GSH, TCVG formation in microsomes was decreased, indicating that oxidative metabolism and GSH conjugation of Tetra are competitive reactions. The Tetra metabolite TCVC was cleaved by bacterial cysteine conjugate b-lyase to dichloroacetic acid and pyruvate. The obtained results substantiate the postulated pathways of Tetra biotransformation and demonstrate that both oxidative and conjugative reactions occur in hepatic Tetra metabolism. Phospholipid alkylation, which occurs during oxidative metabolism, may be a deactivation reaction, whereas TCVG formation, renal metabolism to TCVC, and cleavage of TCVC by b-lyase under formation of mutagenic intermediates may contribute to the nephrocarcinogenic effect of Tetra.  相似文献   

18.
[1,2-14C]1,2-Dichloroethane was metabolized by rat liver enzyme systems to nonvolatile products and to products irreversibly bound to protein and calf thymus DNA. Cytosolic metabolism to all three types of metabolites was dependent upon the presence of reduced glutathione (GSH), suggesting the role of GSH transferases. Microsomal metabolism to all three types of products occurred via mixed function oxidation; microsomal GSH transferase(s) catalyzed the formation of metabolites irreversibly bound to DNA. GSH blocked microsomal mixed function oxidase (MFO)-catalyzed binding to protein but stimulated binding to DNA in a synergistic manner. 2-Chloroacetaldehyde, S-(2-chloroethyl)-GSH, and 1-chloroso-2-chloroethane are proposed as major species involved in irreversible binding but vinyl chloride, 2-chloroethanol, and chloroethyl radicals are not. The microsomal MFO, synergistic microsomal MFO-GSH transferase, and cytosolic GSH transferase systems differed in their preferences for irreversible binding of label from dichloroethane to various homopolyribonucleotides and only the latter system produced metabolites mutagenic to Salmonella typhimurium TA 1535. We postulate that several 1,2-dichloroethane activation pathways are operative which produce different adducts; the relative contribution of each pathway to total nonvolatile metabolites, mutagenic metabolites, and DNA and protein adducts under these in vitro assay conditions was estimated.  相似文献   

19.
Chlorothalonil (2,4,5,6-tetrachloroisophthalonitrile) is a broad spectrum fungicide that is a potent acute toxicant to fish. Therefore, the metabolism of chlorothalonil was investigated in liver and gill cytosolic and microsomal fractions from channel catfish (Ictalurus punctatus) using HPLC. All fractions catalyzed the metabolism of chlorothalonil to polar metabolites. Chlorothalonil metabolism by cytosolic fractions was reduced markedly when glutathione (GSH) was omitted from the reaction mixtures. The lack of microsomal metabolism in the presence of either NADPH or an NADPH-regenerating system indicated direct glutathione S-transferase (GST)-catalyzed conjugation with GSH without prior oxidation by cytochrome P450. Cytosolic and microsomal GSTs from both tissues were also active toward 1-chloro-2,4-dinitrobenzene (CDNB), a commonly employed reference substrate. In summary, channel catfish detoxified chlorothalonil in vitro by GST-catalyzed GSH conjugation in the liver and gill. The present report is the first to confirm microsomal GST activity toward CDNB in gill and toward chlorothalonil in liver, and also of gill cytosolic GST activity towards chlorothalonil, in an aquatic species.  相似文献   

20.
Bromodichloromethane (CHBrCl(2)), a prevalent drinking water disinfection byproduct, was previously shown to be mutagenic in Salmonella that express rat GSH transferase (GST) theta 1-1 (GST T1-1). In the present study, in vitro experiments were performed to study the kinetics of CHBrCl(2) reactions mediated by GST in different species as well as the isoform specificity and reaction products of the GST pathway. Conjugation activity of CHBrCl(2) with GSH in mouse liver cytosol was time- and protein-dependent, was not inhibited by the GST alpha, mu and pi inhibitor S-hexyl-GSH, and correlated with GST T1-1 activity toward the substrate 1,2-epoxy-3-(4'-nitrophenoxy)propane. Conjugation activities in hepatic cytosols of different species toward CHBrCl(2) followed the order mouse > rat > human. As compared with CH(2)Cl(2), the catalytic efficiency (k(cat)/K(m)) of conjugation of CHBrCl(2) with GSH by pure recombinant rat GST T1-1 was approximately 3-6-fold less. Taken together, this suggests that GST T1-1 is the primary catalyst for conjugation of CHBrCl(2) with GSH and that flux through this pathway is less than for CH(2)Cl(2). The initial GSCHCl(2) conjugate formed was unstable and degraded to several metabolites, including GSCH(2)OH, S-formyl-GSH, and HCOOH. Addition of NAD(+) to cytosol did not alter the rate of conjugation of CHBrCl(2) with GSH; however, it did increase the amount of [(14)C]HCOOH produced ( approximately 10-fold). A similar result was seen in a reaction containing pure rat GST T1-1 and GSH-dependent formaldehyde dehydrogenase, indicating that GSCH(2)OH was formed as a precursor to S-formyl-GSH. The half-life of synthetic S-formyl-GSH in pH 7.4 buffer was approximately 1 h at ambient temperature and decreased to approximately 7 min in pH 9.0 buffer, and it does not react with deoxyguanosine. In conclusion, GST T1-1 conjugation of CHBrCl(2) has been definitively demonstrated and the kinetics of conjugation of CHBrCl(2) with GSH characterized in mouse, rat, and human hepatic cytosols. The significance of this GST pathway is that reactive GSH conjugates are produced resulting in possible formation of DNA adducts. Comparisons with CH(2)Cl(2) suggest that the reactive intermediates specific to GSH conjugation of CHBrCl(2) are more mutagenic/genotoxic than those derived from CH(2)Cl(2).  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