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 29 毫秒
1.
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   

2.
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   

3.
The presence of the DNA repair protein O(6)-alkylguanine-DNA alkyltransferase (AGT) paradoxically increases the mutagenicity and cytotoxicity of 1,2-dibromoethane (DBE) in Escherichia coli. This enhancement of genotoxicity did not occur when the inactive C145A mutant of human AGT (hAGT) was used. Also, hAGT did not enhance the genotoxicity of S-(2-haloethyl)glutathiones that mimic the reactive product of the reaction of DBE with glutathione, which is catalyzed by glutathione S-transferase. These experiments support a mechanism by which hAGT activates DBE. Studies in vitro showed a direct reaction between purified recombinant hAGT and DBE resulting in a loss of AGT repair activity and a formation of an hAGT-DBE conjugate at Cys(145). A 2-hydroxyethyl adduct was found by mass spectrometry to be present in the Gly(136)-Arg(147) peptide from tryptic digests of AGT reacted with DBE. Incubation of AGT with DBE and oligodeoxyribonucleotides led to the formation of covalent AGT-oligonucleotide complexes. These results indicate that DBE reacts at the active site of AGT to generate an S-(2-bromoethyl) intermediate, which forms a highly reactive half-mustard at Cys(145). In the presence of DNA, the DNA-binding function of AGT facilitates formation of DNA adducts. In the absence of DNA, the intermediate undergoes hydrolytic decomposition to form AGT-Cys(145)-SCH(2)CH(2)OH.  相似文献   

4.
Adult male Sprague-Dawley rats were exposed by inhalation to various concentrations of styrene vapors (25, 50, 100, or 200 ppm) 6 h/day, 5 days/week, for 4 consecutive weeks. The concentrations were varied from day to day according to a random pattern allowing treated animals to be exposed five times to each concentration of styrene. Each day, the following urinary metabolites were analysed from samples collected during exposure (0-6 h) and after exposure (6-24 h): mandelic acid; phenylglyoxylic acid; and two mercapturic acids, N-acetyl-S-(1-phenyl-2-hydroxyethyl)-L-cysteine (M1) and N-acetyl-S-(2-phenyl-2-hydroxyethyl)-L-cysteine (M2). Various parameters of renal toxicity and hepatic microsomal and cytosolic enzyme activities were also measured. The results show that there is a very good relationship between the excretion of all four styrene metabolites and the degree of daily exposure to styrene over the entire period of urine collection, with correlation coefficients ranging from 0.82 to 0.98. The correlation was poor for mandelic acid during the 0-6 h period. There was no evidence that repeated exposure to styrene caused renal toxicity, nor induced hepatic microsomal enzyme activities; cytosolic glutathione S-transferase activity was increased moderately by 1.5 times. Thus, under conditions of exposure to styrene likely to be found in the workplace, all four metabolites measured were good indicators of styrene exposure throughout the length of the experiment. Since mercapturic acids result from the conjugation of styrene oxide with glutathione, the data suggest that measurement of these metabolites offers the possibility to monitor internal exposure to a toxic electrophilic compound more directly.  相似文献   

5.
The cysteine S conjugate of 1,2-dichloroethane, S-(2-chloroethyl)-DL-cysteine (CEC), is hepatotoxic, nephrotoxic, and mutagenic. To determine the cellular and chemical mechanisms involved in CEC-induced toxicity and to assess the role of an episulfonium ion, the effect of CEC on the viability of isolated rat hepatocytes was studied. CEC addition resulted in both a time- and concentration-dependent loss of cell viability. Depletion of intracellular glutathione concentrations (greater than 70%) and inhibition of microsomal Ca2+ transport and Ca2+-ATPase activity preceded the loss of cell viability, and initiation of lipid peroxidation paralleled the loss of viability. The depletion of glutathione concentrations was partially attributable to a reaction between glutathione and CEC to form S-[2-(DL-cysteinyl)ethyl]glutathione, which was identified by NMR and mass spectrometry. N-Acetyl-L-cysteine, vitamin E, and N,N'-diphenyl-p-phenylenediamine protected against the loss of cell viability. N,N'-Diphenyl-p-phenylenediamine inhibited CEC-initiated lipid peroxidation but did not protect against cell death at 4 h, indicating that lipid peroxidation was not the cause of cell death. The analogues S-ethyl-L-cysteine, S-(3-chloropropyl)-DL-cysteine, and S-(2-hydroxyethyl)-L-cysteine, which cannot form an episulfonium ion, were not cytotoxic, thus demonstrating a role for an episulfonium ion in the cytotoxicity associated with exposure to CEC and, possibly, 1,2-dichloroethane. These results show that an alteration in Ca2+ homeostasis and the generation of an electrophilic intermediate may be involved in the mechanism of cell death.  相似文献   

6.
Kubec R  Kim S  Musah RA 《Phytochemistry》2002,61(6):675-680
Three cysteine derivatives, (R)-S-(2-hydroxyethyl)cysteine, together with (R(S)R(C))- and (S(S)R(C))-S-(2-hydroxyethyl)cysteine sulfoxides, have been isolated from the roots of Petiveria alliacea. Furthermore, three additional amino acids, S-methyl-, S-ethyl-, and S-propylcysteine derivatives, were detected. They were present only in trace amounts (<3 microg g(-1) fr. wt), precluding determination of their absolute configurations and oxidation states. In addition, four thiosulfinates, S-(2-hydroxyethyl) (2-hydroxyethane)-, S-(2-hydroxyethyl) phenylmethane-, S-benzyl (2-hydroxyethane)- and S-benzyl phenylmethanethiosulfinates, have been found in a homogenate of the roots. The formation pathways of various benzyl/phenyl-containing compounds previously found in the plant were also discussed.  相似文献   

7.
Zn2+ in native glyoxalase I from human erythrocytes can be replaced by Cu2+, giving an inactive enzyme. Cu2+ was demonstrated to compete with the activating metals Zn2+ and Mn2+, indicating a common binding site on the enzyme for these metal ions. The electron paramagnetic resonance (EPR) spectra of 63Cu(II) glyoxalase I at 77 K and of its complexes with glutathione and some glutathione derivatives are characteristic of Cu2+ in an elongated octahedral coordination (g parallel = 2.34, g perpendicular = 2.09, and A parallel = 14.2 mT). The low-field bands of the free enzyme are asymmetric and become symmetrical upon addition of glutathione or S-(p-bromobenzyl)glutathione but not S-(D-lactoyl)glutathione. The results indicate the existence of two conformations of Cu(II) glyoxalase I, in agreement with the effects caused by these compounds on the protein fluorescence. The copper hyperfine line at low field in the EPR spectrum of the S-(p-bromobenzyl)glutathione complex of 63Cu(II) glyoxalase I shows a triplet structure, indicative of coupling to one nitrogen ligand in the equatorial plane. Similar results were obtained with the glutathione complex. By addition of the spectrum of the S-(p-bromobenzyl)glutathione complex and a spectrum corresponding to two nitrogen ligands with two different coupling constants, a good fit was obtained for the low-field region of the asymmetric spectrum of free 63Cu(II) glyoxalase I. The first two spectra are assumed to correspond to two separate conformational states of the enzyme. The results demonstrate that at least one nitrogen ligand is involved in the binding of Cu2+.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

8.
Chlorotrifluoroethene, a potent nephrotoxin, is a substrate for the glutathione S-transferases present in the cytosolic and microsomal fractions of rat liver. The glutathione conjugate formed by both subcellular fractions has been identified as S-(2-chloro-1,1,2-trifluoroethyl)glutathione by 1H and 19F NMR and by secondary ion mass spectrometry. The conjugate formed by the cytosolic fraction is an equimolar mixture of two diastereomers, whereas the conjugate formed by the microsomal fraction is predominantly one diastereomer, as judged by the 19F NMR spectra. No evidence for the formation of S-(trihalovinyl)glutathione derivatives by an addition/elimination reaction was found. High-performance liquid chromatography was employed to measure the rates of glutathione conjugate formation in vitro. The rates of S-(2-chloro-1,1,2-trifluoroethyl)glutathione formation were 75-107 nmol min-1 (mg of protein)-1 and 151-200 nmol min-1 (mg of protein)-1 catalyzed by the cytosolic and microsomal fractions, respectively (measured at pH 7.4, 37 degrees C, with 5 mM glutathione). These results suggest that glutathione conjugation occurs at high rates in vivo to produce the highly nephrotoxic S-(2-chloro-1,1,2-trifluoroethyl)glutathione.  相似文献   

9.
S-1358 was rapidly absorbed, metabolized and readily excreted via urine and feces from orally dosed rats. Excretion of radioactivity was almost complete within 4 days. The radioactivity was distributed mainly in stomach, intestines, liver and kidneys. It seems that S-1358 and its metabolites do not persist in organs and tissues following a single oral dosing.

Major urinary metabolites of the benzyl-labeled S-1358 were p-(1,1-dimethyl-2-hydroxyethyl)benzyl methyl sulfide [B], p-(1,1-dimethyl-2-hydroxyethyl)benzyl methyl sulfone [A], p-(1-methyl-1-carboxylethyl)benzyl methyl sulfide [D], p-(1-methyl-1-carboxylethyl)benzyl methyl sulfone [C] and their glucuronide conjugates. Fecal metabolites were S-n-butyl S′-(1, 1-dimethyl-2-hydroxyethyl)benzyl N-3-pyridyldithiocarbonimidate [MR], A, B, C and D. These metabolites were also found in the bile. The pyridine-labeled S-1358 gave rise to 2-(3′-pyridylimino)-4-carboxylthiazolidine [HM] and 3-aminopyridine [AP] in the urine, and MR and AP in the feces. Intact S-1358 was a major component of the fecal radioactivity.  相似文献   

10.
G L Foureman  D J Reed 《Biochemistry》1987,26(7):2028-2033
The formation of S-[2-(N7-guanyl)ethyl]glutathione (GEG) from dihaloethanes is postulated to occur through two intermediates: the S-(2-haloethyl)glutathione conjugate and the corresponding episulfonium ion. We report the formation of GEG when deoxyguanosine (dG) was incubated with chemically synthesized S-(2-chloroethyl)glutathione (CEG). The depurination of GEG was shown to be first order with a half-life of 7.4 +/- 0.4 h at 27 degrees C. Evidence is also presented for the formation of S-[2-(N7-guanyl)ethyl]-L-cysteine (GEC) in incubation mixtures containing dG and S-(2-chloroethyl)-L-cysteine (CEC), the corresponding cysteine conjugate of CEG. This finding demonstrates that this (haloethyl)cysteine conjugate does not require activation by enzymatic action of cysteine conjugate beta-lyase but, instead, can directly alkylate DNA. The half-life of the depurination of GEC was 6.5 +/- 0.9 h, which is no different from that of GEG. Of the two conjugates, CEC is a somewhat more active alkylating agent toward dG than CEG as N7-guanylic adduct was detected in reaction mixtures with lower concentrations of CEC than with CEG.  相似文献   

11.
Conversion of the toxic vinyl halide 1,1-dichloroethylene (DCE) to S-(2-S-glutathionyl-acetyl)glutathione (GSCH2COSG) involves sequential acylation and alkylation of two glutathione (GSH) molecules by the microsomal DCE metabolite ClCH2COCl. To examine its possible role in DCE biotransformation, we synthesized the putative intermediate S-(2-chloroacetyl)glutathione (ClCH2COSG). In aqueous buffer, ClCH2COSG did not hydrolyze to release GSH, but instead underwent a two-step rearrangement to yield a cyclic product. Product analyses by liquid secondary ion mass spectrometry and 1H-13C heteronuclear correlation nuclear magnetic resonance spectroscopy indicated that rearrangement involved initial transfer of the chloroacetyl moiety from the cysteinyl thiol to the gamma-glutamyl alpha-amine. The cysteinyl thiol then displaced chloride from the 2-chloroacetyl methylene carbon to yield the cyclic product. Incubation of 2 mM ClCH2COSG with 20 mM GSH yielded approximately 4.5-fold more cyclic product than GSCH2COSG. ClCH2COSG alkylated oxytocindithiol and N-acetyl-L-cysteine to yield S-[2-(alkylthio)acetyl]glutathione adducts analogous to GSCH2COSG. S-2-Chloroacetylation products were absent. In reacting with thiols by alkylation and in decomposing by rearrangement, ClCH2COSG displayed properties strikingly different from those of ClCH2COCl. Although much less reactive than its acyl halide precursor, ClCH2COSG may display greater selectivity in covalent modification of cellular targets in DCE intoxication.  相似文献   

12.
The stereochemistry of S-(2-chloro-1,1,2-trifluoroethyl)glutathione formation was studied in rat liver cytosol, microsomes, N-ethylmaleimide-treated microsomes, 9000g supernatant fractions, purified rat liver microsomal glutathione S-transferase, and isolated rat hepatocytes. The absolute configuration of the chiral center generated by the addition of glutathione to chlorotrifluoroethene was determined by degradation of S-(2-chloro-1,1,2-trifluoroethyl)glutathione to chlorofluoroacetic acid, followed by derivatization to form the diastereomeric amides N-(S)-alpha-methylbenzyl-(S)-chlorofluoacetamide and N-(S)-alpha-methylbenzyl-(R)-chlorofluoroacetamide, which were separated by gas chromatography. Native and N-ethylmaleimide-treated rat liver microsomes, purified rat liver microsomal glutathione S-transferase, rat liver 9000g supernatant, and isolated rat hepatocytes catalyzed the formation of 75-81% (2S)-S-(2-chloro-1,1,2-trifluoroethyl)glutathione; rat liver cytosol catalyzed the formation of equal amounts of (2R)- and (2S)-S-(2-chloro-1,1,2-trifluoroethyl)glutathione. In rat hepatocytes, microsomal glutathione S-transferase catalyzed the formation of 83% of the total S-(2-chloro-1,1,2-trifluoroethyl)glutathione formed. These observations show that the microsomal glutathione S-transferase catalyzes the first step in the intracellular, glutathione-dependent bioactivation of the nephrotoxin chlorotrifluoroethene.  相似文献   

13.
The nephrotoxic and nephrocarcinogenic potential of the haloalkenes is associated with the conjugation of the chemicals to L-glutathione. Subsequent processing of the haloalkene glutathione S-conjugates via the cysteine conjugate beta-lyase pathway in the mammalian kidney yields nephrotoxic and mutagenic species. To investigate whether S-conjugates of the model chlorofluoroalkenes 1,1,2-trichloro-3,3,3-trifluoro-1-propene (CAS # 431-52-7) and trichlorofluoroethene (CAS # 359-29-5) show comparable effects, we have synthesised the respective cysteine and glutathione S-conjugates and subjected them to the Ames test. The cysteine and glutathione S-conjugates of tetrachloroethene (CAS # 127-18-4), S-(1,2,2-trichlorovinyl)-L-cysteine (TCVC) and S-(1,2,2-trichlorovinyl)glutathione (TCVG) were used as positive controls and reference substances. S-(1,2-dichloro-3,3,3-trifluoro-1-propenyl)-L-cysteine (DCTFPC) and S-(2,2-dichloro-1-fluorovinyl)-L-cysteine (DCFVC) showed clear dose-dependent mutagenic effects with the Salmonella typhimurium tester strains TA100 and TA98. Using TCVC as a reference substance the following ranking in mutagenic response was established: TCVC>DCTFPC>DCFVC. S-(1,2-dichloro-3,3,3-trifluoro-1-propenyl)glutathione (DCTFPG) and S-(2,2-dichloro-1-fluorovinyl)glutathione (DCFVG) showed potent dose-dependent mutagenic effects with the S. typhimurium tester strain TA100 in the presence of a rat kidney S9-protein fraction; tests carried out in the absence of the bioactivation system resulted only in background rates of revertants. Using TCVG as a reference substance the following ranking in mutagenic response was established: TCVG=DCTFPG>DCFVG.The data obtained provide a basis for further studies on the mutagenic and presumable carcinogenic potential of the substances.  相似文献   

14.
Incubation of isolated hepatocytes in the presence of either the parkinsonian-inducing compound 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP) or its putative toxic metabolite 1-methyl-4-phenylpyridinium ion (MPP+) led to a depletion of intracellular reduced glutathione (GSH), which was mostly recovered as glutathione disulfide (GSSG). However, both MPTP- and MPP+-induced glutathione perturbances were relatively unaffected by the prior inhibition of glutathione reductase with 1,3-bis(2-chloroethyl)-1-nitrosourea (BCNU), suggesting that intracellular oxidation was not the major mechanism involved in the GSH loss. Inclusion of cystine in the incubation mixtures revealed a time-dependent formation of cysteinyl glutathione (CySSG), indicating that an increased efflux was mostly responsible for the MPTP- and MPP+-induced GSH depletion. Therefore, the measurement of GSSG, which is apparently formed extracellularly, was not associated with oxidative stress.  相似文献   

15.
Mutagenicity of amino acid and glutathione S-conjugates in the Ames test   总被引:1,自引:0,他引:1  
The mutagenicity of the glutathione S-conjugate S-(1,2-dichlorovinyl)glutathione (DCVG), the cysteine conjugates S-(1,2-dichlorovinyl)-L-cysteine (DCVC) and S-(1,2-dichlorovinyl)-DL-alpha-methylcysteine (DCVMC), and the homocysteine conjugates S-(1,2-dichlorovinyl)-L-homocysteine (DCVHC) and S-(1,2-dichlorovinyl)-DL-alpha-methylhomocysteine (DCVMHC) was investigated in Salmonella typhimurium strain TA2638 with the preincubation assay. DCVC was a strong, direct-acting mutagen; the cysteine conjugate beta-lyase inhibitor aminooxyacetic acid decreased significantly the number of revertants induced by DCVC; rat renal mitochondria (11,000 X g pellet) and cytosol (105,000 X g supernatant) with high beta-lyase activity increased DCVC mutagenicity at high DCVC concentrations. DCVG was also mutagenic without the addition of mammalian activating enzymes; the presence of low gamma-glutamyltransferase activity in bacteria, the reduction of DCVG mutagenicity by aminooxyacetic acid, and the potentiation of DCVG mutagenicity by rat kidney mitochondria and microsomes (105,000 X g pellet) with high gamma-glutamyltransferase activity indicate that gamma-glutamyltransferase and beta-lyase participate in the metabolism of DCVG to mutagenic intermediates. The homocysteine conjugate DCVHC was only weakly mutagenic in the presence of rat renal cytosol, which exhibits considerable gamma-lyase activity, this mutagenic effect was also inhibited by aminooxyacetic acid. The conjugates DCVMC and DCVMHC, which are not metabolized to reactive intermediates, were not mutagenic at concentrations up to 1 mumole/plate. The results demonstrate that gamma-glutamyltransferase and beta-lyase are the key enzymes in the biotransformation of cysteine and glutathione conjugates to reactive intermediates that interact with DNA and thereby cause mutagenicity.  相似文献   

16.
The glutathione S-transferases are a family of dimeric enzymes that catalyze the reaction between GSH and a variety of electrophiles. Two closely related isozymes, referred to as YaYa and YcYc, were purified from rat liver. A radiolabeled azido derivative of glutathione (S-(p-azidophenacyl)[3H]glutathione) was prepared and used to label covalently the active site of the above two glutathione S-transferases. The noncovalently bound affinity label was a competitive inhibitor of glutathione S-transferase YaYa toward both 1-chloro-2,4-dinitrobenzene and GSH. The covalently labeled enzymes no longer bound to a GSH-affinity column, and covalent labeling was reduced in the presence of GSH and S-(dinitrophenyl)glutathione. These results suggest that the affinity label was binding at the active site. The covalently labeled enzymes were digested with trypsin, and the labeled peptides were purified by HPLC and then sequenced. A single-labeled peptide was identified in the tryptic digest of the YaYa isozyme, whereas two labeled peptides were present in the tryptic digest of YcYc. The Ya peptide sequence was identical with the published deduced sequence of amino acids between residues 212 and 218 and the sequences of the two peptides purified from Yc were identical with the deduced sequence of amino acids between 91 and 110 and 206 and 218. Hence, the Ya peptide and the smaller peptide purified from Yc came from the same region of the Ya and Yc subunits. This common region and a second region of the Yc subunit appear to form a portion of the active site of these two forms of glutathione S-transferase.  相似文献   

17.
Excretion of glutathione conjugates by primary cultured rat hepatocytes   总被引:2,自引:0,他引:2  
Conjugation of xenobiotics with glutathione occurs commonly within the liver, and these glutathione conjugates are then preferentially excreted into bile. We have characterized this excretory process using primary cultured hepatocytes (24 h). 1-Chloro-2,4-dinitrobenzene rapidly entered the cells and formed a glutathione conjugate, S-(dinitrophenyl)glutathione, irrespective of the temperature of incubation. In contrast, the efflux of the glutathione conjugate was essentially absent in the cold but recovered rapidly upon rewarming of the cells. Therefore, initial rates of efflux of the conjugate at 37 degrees C were measured from cells preloaded biosynthetically at 10 degrees C. Efflux was a saturable process with respect to intracellular S-(dinitrophenyl)glutathione with an apparent Km of 0.58 +/- 0.12 mM and Vmax of 0.15 +/- 0.05 nmol/min/mg of protein. The excretion of S-(dinitrophenyl)glutathione had an energy of activation of 15.3 kcal/mol. The glutathione conjugate of p-nitrobenzylchloride when formed within the hepatocytes acted as a competitive inhibitor of S-(dinitrophenyl)glutathione efflux. Cultured hepatocytes, therefore, appeared to have a specific transport process for the excretion of glutathione conjugates. The addition of S-(dinitrophenyl)glutathione, but not GSH, GSSG, or methionine, to the medium caused a decrease in the rate of efflux of radiolabeled S-(dinitrophenyl)glutathione. The hepatocytes were able, however, to excrete the glutathione conjugate against an excess of extracellular S-(dinitrophenyl)glutathione. This observation suggested that extracellular S-(dinitrophenyl)glutathione, although capable of binding to the carrier, entered the hepatocytes quite slowly relative to rates of efflux. This carrier may function in a manner that would minimize the reuptake by hepatocytes of conjugates that have been excreted into the bile.  相似文献   

18.
Monobromobimane (mBBr), functions as a substrate of porcine glutathione S-transferase pi (GST pi): The enzyme catalyzes the reaction of mBBr with glutathione. S-(Hydroxyethyl)bimane, a nonreactive analog of monobromobimane, acts as a competitive inhibitor with respect to mBBr as substrate but does not affect the reaction of GST pi with another substrate, 1-chloro-2,4-dinitrobenzene (CDNB). In the absence of glutathione, monobromobimane inactivates GST pi at pH 7.0 and 25 degrees C as assayed using mBBr as substrate, with a lesser effect on the enzyme's use of CDNB as substrate. These results indicate that the sites occupied by CDNB and mBBr are not identical. Inactivation is proportional to the incorporation of 2 moles of bimane/mole of subunit. Modification of GST pi with mBBr does not interfere with its binding of 8-anilino-1-naphthalene sulfonate, indicating that this hydrophobic site is not the target of monobromobimane. S-Methylglutathione and S-(hydroxyethyl)bimane each yield partial protection against inactivation and decrease reagent incorporation, while glutathionyl-bimane protects completely against inactivation. Peptide analysis after trypsin digestion indicates that mBBr modifies Cys45 and Cys99 equally. Modification of Cys45 is reduced in the presence of S-methylglutathione, indicating that this residue is at or near the glutathione binding region. In contrast, modification of Cys99 is reduced in the presence of S-(hydroxyethyl)bimane, suggesting that this residue is at or near the mBBr xenobiotic substrate binding site. Modification of Cys99 can best be understood by reaction with monobromobimane while it is bound to its xenobiotic substrate site in an alternate orientation. These results support the concept that glutathione S-transferase accomplishes its ability to react with a diversity of substrates in part by harboring distinct xenobiotic substrate sites.  相似文献   

19.
The relative ease of formation of five-, six-, seven-, and eight-membered cyclic benzeneboronates is discussed. Pentane-1,3,5-triol forms, by interaction with benzene-boronic anhydride, exclusively DL-4-(2-hydroxyethyl)-2-phenyl-1,3,2-dioxaborinane. It is suggested that, in this compound, the oxygen atom of the hydroxyl group is intramolecularly co-ordinated with the boron atom.  相似文献   

20.
Liposome encapsulation of saffron effectively enhanced its antitumor activity towards Sarcoma-180 (S-180) and Ehrlich ascites carcinoma solid tumors in mice. Significant inhibition (P < 0.001) in the growth of these tumors was observed as compared with vehicle (control) mice. In the presence of phytohemagglutinin (PHA), a T cell mitogen, saffron stimulated non-specific proliferation of lymphocytes in vitro. The intracellular reduced glutathione and related enzymes, i.e. glutathione reductase and glutathione-S-transferase, of S-180 tumor cells were significantly elevated when incubated with saffron, possibly acting to maintain functional levels of other antioxidants. Our studies indicate the antioxidant activity of saffron.  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