首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
GSH is rapidly oxidized by HOCl (hypochlorous acid), which is produced physiologically by the neutrophil enzyme myeloperoxidase. It is converted into, mainly, oxidized glutathione. Glutathione sulfonamide is an additional product that is proposed to be covalently bonded between the cysteinyl thiol and amino group of the gamma-glutamyl residue of GSH. We have developed a sensitive liquid chromatography-tandem MS assay for the detection and quantification of glutathione sulfonamide as well as GSH and GSSG. The assay was used to determine whether glutathione sulfonamide is a major product of the reaction between GSH and HOCl, and whether it is formed by other two-electron oxidants. At sub-stoichiometric ratios of HOCl to GSH, glutathione sulfonamide accounted for up to 32% of the GSH that was oxidized. It was also formed when HOCl was generated by myeloperoxidase and its yield increased with the flux of oxidant. Of the other oxidants tested, only hypobromous acid and peroxynitrite produced substantial amounts of glutathione sulfonamide, but much less than with HOCl. Chloramines were able to generate detectable levels only when at a stoichiometric excess over GSH. We conclude that glutathione sulfonamide is sufficiently selective for HOCl to be useful as a biomarker for myeloperoxidase activity in biological systems. We have also identified a novel oxidation product of GSH with a molecular weight two mass units less than GSH, which we have consequently named dehydroglutathione. Dehydroglutathione represented a few percent of the total products and was formed with all of the oxidants except H2O2.  相似文献   

2.
In vivo 13C NMR has been used to detect the transient formation of S-(hydroxymethyl)glutathione (GSCH2OH) from glutathione and [13C]formaldehyde in Escherichia coli. Two-dimensional 1H-13C shift correlation was used to locate the chemical shift of the formaldehyde-derived protons of the adduct. The adduct GSCH2OH is formed by chemical reaction in the first few minutes after cells are challenged with formaldehyde and remains within the cell until consumed by metabolism.  相似文献   

3.
Uptake of the thioether S-(2,4-dinitrophenyl)glutathione (DNPSG) in canalicular plasma membrane vesicles from rat liver is enhanced in the presence of ATP and exhibits an overshoot with a transient 5.5-fold accumulation of DNPSG. Stimulation by ATP is not caused by the generation of a membrane potential, based on responses of the indicator dye oxonol V. ATP-dependent uptake has an apparent Km of 71 microM for DNPSG and a Vmax of 0.34 nmol.min-1.mg of vesicle protein-1. Protein thiol groups are essential for transport activity as indicated by the sensitivity of DNPSG transport to sulfhydryl reagents. There is competitive inhibition with other thioethers, S-hexylglutathione (Ki = 66 microM), the photoaffinity label S-(4-azidophenacyl)glutathione (Ki = 56 microM), as well as with glutathione disulfide (Ki = 0.44 mM) and with the bile acid taurocholate (Ki = 0.61 mM). GSH (2 mM) or cholate (0.4 mM) does not inhibit. Both glutathione disulfide and taurocholate show ATP-dependent transport in the canalicular membrane vesicles which is inhibited by DNPSG. No ATP-dependent transport is found for GSH. Transport of DNPSG is also inhibited competitively by alpha-naphthyl-beta-D-glucuronide (Ki = 0.42 mM) but not by alpha-naphthylsulfate (2 mM), and there is substantial inhibition with the glucuronides from ebselen and p-nitrophenol. The results indicate that the canalicular transport system for DNPSG is directly driven by ATP and that the biliary transport of other classes of compounds may also proceed via this system.  相似文献   

4.
The steady-state kinetics of glutathione S-transferase I1 (GST I1) from housefly Musca domestica expressed in Escherichia coli were investigated with glutathione (GSH) and 1-chloro-2,4-dinitrobenzene (CDNB). Concentrations of the varied substrates were from 0.03 to 1 mM for GSH and 0.05 to 1 mM for CDNB. Within this range, Michaelis-Menten behaviour was observed and convergent straight lines in double reciprocal plots excluded a ping-pong kinetic mechanism. Instead, data were consistent either with rapid-equilibrium random or with steady-state ordered sequential mechanisms because of abscissa convergence. Discrimination was achieved by studying the reaction with another electrophilic partner, p-nitrophenyl-acetate (PNPA). Concentrations of PNPA and GSH varied within the ranges 0.5 to 10 mM and 0.03 to 0.6 mM, respectively. The complete set of data supports the proposal of a rapid-equilibrium random-sequential model with strictly independent sites for GSH and CDNB or PNPA. Kinetic parameters are thus true dissociation equilibrium constants with values of 0.15 mM for GSH, 0.15 mM for CDNB, and 7 mM for PNPA. Analysis of the inhibition by the product (S-(2,4-dinitrophenyl)-glutathione, 10 to 100 microM), on the coupling reaction between GSH and CDNB with either GSH (0.05 to 0.5 mM, CDNB 0.2 mM) or CDNB (0.05 to 0.5 mM, GSH 0.2 mM) varied, consistent with the proposed mechanism. Binding of product to the free enzyme excludes GSH (competitive inhibition pattern with Kp = 12 microM) but only slightly hinders binding of CDNB. Binding free energies, together with the inhibition pattern, suggest that the non-peptidic moiety of product interacts with an alternative sub-site within the large open pocket accommodating the various electrophilic substrates. These results lead us to propose a model for intra-pocket shifting of the non-peptidic moiety upon product formation which contributes to the product release.  相似文献   

5.
Levels of reduced glutathione (GSH) in C3H/10T1/2 cells were selectively altered to determine what quantitative role GSH transferase-catalyzed conjugation plays in regulating the cytotoxic effects of benzo(a)pyrene anti-7,8-dihydrodiol 9,10-epoxide (r-7,t-8-dihydroxy-t-9,10-oxy-7,8,9,10-tetrahydrobenzo(a)pyrene, anti-diol epoxide). A 65% decrease in 10T1/2 cell GSH content from 0.16 mM (control cell GSH concentration) to 0.06 mM was accompanied by a 46% decrease in the anti-diol epoxide LD80; a 98% increase in GSH content resulted in a 44% increase in anti-diol epoxide LD80. This nonlinear relationship between changes in cellular GSH concentration and anti-diol epoxide LD80 was directly relatable to the nonlinear change in the rate of anti-diol epoxide conjugation which was catalyzed by 10T1/2 cell GSH transferases. Purified 10T1/2 cell cytosol catalyzed the GSH conjugation of anti-diol epoxide to yield a GSH conjugation product with a distinct UV absorbance spectrum; the apparent GSH Km for this cell cytosol-catalyzed reaction was 0.20 mM. Variations in the cellular GSH concentration around the GSH Km resulted in a nonlinear change in the amount of anti-diol epoxide-GSH conjugate formed, and a reciprocal change in the amount of free anti-diol epoxide available for cytotoxic alkylation events. These results clarify in quantitative, biochemical terms how GSH transferase-catalyzed conjugation can regulate the level of an electrophilic carcinogen metabolite in a biological system.  相似文献   

6.
Maize, wheat and other grasses synthesise large quantities of benzoxazinones and their glucosides, which act as antifeedant and allelopathic agents. These activities are probably due to the electrophilic nature of the aglycones, however, the mechanism of their action is unclear. In biological systems, glutathione (GSH) is the major electrophile-reactive compound so the reaction of the major maize benzoxazinone DIMBOA with GSH was studied. GSH reacts with DIMBOA to form eight isomeric mono-conjugates and eight isomeric di-conjugates. Through NMR studies with the model thiol 2-mercaptoethanol, these were structurally elucidated as unusual spirocycles. Similar reactivity was observed with proteins, with cysteinyl thiols being modified by DIMBOA. The thioether bonds formed were stable and not easily reduced to the parent thiol. DIMBOA can therefore readily deplete GSH levels and irreversibly inactivate enzymes with active-site cysteine residues, with clear implications for potentially toxic effects when young grasses are ingested, whether by insect pests or humans.  相似文献   

7.
Glycation (nonenzymatic glycosylation of proteins) is known to be increased as a result of hyperglycaemia in diabetes. Moreover, cell glutathione concentration has been found to be lower in diabetics and such depletion may impair the cell defence against toxic radical species. Ribose being a potent reducing sugar expected to be increased in cells of diabetics where the pentose phosphate pathway is enhanced, its putative condensation with glutathione was investigated. Reduced glutathione (GSH) was incubated with ribose and the structure of the resultant product was assessed by mass spectrometry, as well as the measurement of its remaining thiol group. A covalent reaction clearly occurred between the reducing sugar and GSH, to give an adduct named N-ribosyl-1-glutathione. This adduct appears to be the Amadori product resulting from the condensation of the primary amine group of GSH with the aldehyde group of ribose. Interestingly, the adduct could not be used as a proper substrate by glutathione peroxidase although it keeps its thiol group. We conclude that the coupling of GSH with a monosaccharide such as ribose might contribute to the decreased cell GSH and glutathione peroxidase activity observed in diabetics.  相似文献   

8.
Refolding of dimeric porcine cytosolic or mitochondrial malate dehydrogenases and of tetrameric pig heart and skeletal muscle lactate dehydrogenases (containing 5-7 cysteine residues), as well as reformation of the four cystine cross-bridges of bovine pancreatic ribonuclease, were studied in the presence of reduced and oxidized glutathione (GSH and GSSG). At the intracellular GSH level (5 mM) reduced ribonuclease can be reoxidized by 0.01-0.5 mM GSSG (pH 7.4) both at 20 degrees C and 37 degrees C. In this physiological range of GSSG concentrations and pH, the dehydrogenases show at least partial reactivation. With GSSG concentrations greater than 5 mM, reactivation is found to be completely inhibited for all the enzymes given. The results show that at the intracellular level of GSH and GSSG, thiol groups in reduced, unfolded ribonuclease are oxidized to form intramolecular cystine cross-bridges, while thiol groups of typical cysteine enzymes, such as lactate and malate dehydrogenase, remain in their reduced state during refolding. The rate of reactivation of lactate dehydrogenase (porcine muscle) is not affected by GSSG. In the case of ribonuclease, increasing concentrations of GSSG increase the rate of reactivation: At 20 degrees C, the halftime of the correct disulfide bond formation varies from approximately equal to 80 h in the presence of 0.01 mM GSSG to approximately equal to 10 h in the presence of 0.25 mM GSSG. A further increase in the rate of reactivation at higher GSSG concentrations is accompanied by a decrease in yield. Reactivation of ribonuclease is also observed at the low glutathione level found in blood plasma (5-25 microM GSH).  相似文献   

9.
4-Hydroxy-2,3-trans-nonenal, a lipid peroxidation product, inhibits glutathione peroxidase in a concentration-dependent manner. The concentration providing 50% inhibition is 0.12 mM. This inhibition can be almost completely (89%) prevented by 1 mM glutathione added to the incubation mixture 30 min before 4-hydroxy-2,3-trans-nonenal or 2,3-trans-nonenal, but not by other thiol-containing antioxidants such as 0.5 mM dithiothreitol or beta-mercaptoethanol. Again the addition of 1 mM glutathione, and not of 0.5 mM dithiothreitol or beta-mercaptoethanol, to the enzyme 30 min after incubation with 4-hydroxy-2,3-trans-nonenal restores activity to the same extent as does the preincubation with GSH. In view of the known reactivity of 4-hydroxy-2,3-trans-nonenal with lysine residues and the reversibility of the inhibition, the involvement of a lysine residue in GSH binding to glutathione peroxidase is proposed. The potential relevance of the inhibition of glutathione peroxidase by 4-hydroxy-nonenal to oxidative tissue damage is discussed with particular emphasis on neurological disorders.  相似文献   

10.
Double-reciprocal plots of initial-rate data for the conjugation of 1-chloro-2,4-dinitrobenzene (CDNB) and GSH by human placental GSH S-transferase pi were linear for both substrates. Computer modelling of the initial-rate data using nonlinear least-squares regression analysis favoured a rapid equilibrium random sequential bi-bi mechanism, over a steady-state random sequential mechanism or a steady-state or rapid equilibrium ordered mechanism. KGSH was calculated as 0.125 +/- 0.006 mM, KCDNB was 0.87 +/- 0.07 mM and alpha was 2.1 +/- 0.3 for the rapid equilibrium random model. The product, S-(2,4-dinitrophenyl)glutathione, was a competitive inhibitor with respect to GSH, and a mixed-type inhibitor toward CDNB (KP = 18 +/- 3 microM). The observed pattern of inhibition is consistent with a rapid equilibrium random mechanism, with a dead-end enzyme.CDNB.product complex, but inconsistent with the inhibition patterns of other bireactant mechanisms. Since rat liver GSH S-transferase 3-3 acts via a steady-state random sequential mechanism [1], while human placental GSH S-transferase and perhaps also rat liver GSH S-transferase 1-1 [2] exhibit rapid equilibrium random mechanisms, we conclude that the kinetic mechanism of the GSH S-transferases is isoenzyme-dependent.  相似文献   

11.
Curcumin (diferuloylmethane), a yellow pigment of turmeric with antioxidant properties has been shown to be a cancer preventative in animal studies. It contains two electrophilic alpha, beta-unsaturated carbonyl groups, which can react with nucleophilic compounds such as glutathione (GSH), but formation of the GSH-curcumin conjugates has not previously been demonstrated. In the present studies, we investigated the reactions of curcumin with GSH and the effect of recombinant human glutathione S-transferase(GST)P1-1 on reaction kinetics. Glutathionylated products of curcumin identified by FAB-MS and MALDI-MS included mono- and di-glutathionyl-adducts of curcumin as well as cyclic rearrangement products of GSH adducts of feruloylmethylketone (FMK) and feruloylaldehyde (FAL). The presence of GSTP1-1 significantly accelerated the initial rate of GSH-mediated consumption of curcumin in 10 mM potassium phosphate, pH 7.0, and 1 mM GSH. GSTP1-1 kinetics determined using HPLC indicated substrate inhibition (apparent K(m) for curcumin of 25+/-11 microM, and apparent K(i) for curcumin of 8+/-3 microM). GSTP1-1 was also shown to catalyze the reverse reaction leading to the formation of curcumin from GSH adducts of FMK and FAL.  相似文献   

12.
Selenite, selenate and selenocystine catalyzed the reduction of methemoglobin (metHb) by glutathione (GSH), while selenomethionine did not. Maximal reduction of metHb was observed with 10?5 M selenite and 2 mM GSH, at pH 7.4. Selenite also catalyzed the reduction of metHb with cysteine or 2-mercaptoethylamine in place of GSH. Heavy metals and arsenite completely prevented the effect of selenite. These findings suggest that certain seleno-compounds catalyze the reduction of metHb by thiol compounds.  相似文献   

13.
Nutritional biochemistry of cellular glutathione   总被引:12,自引:0,他引:12  
Glutathione (GSH) has emerged to be one of the most fascinating endogenous molecules virtually present in all animal cells often in quite high (mM) concentrations. In addition to the detoxicant, antioxidant, and cysteine-reservoir functions of cellular glutathione, the potential of this ubiquitous thiol to modulate cellular signal transduction processes has been recently evident. Lowered tissue GSH levels have been observed in several disease conditions. Restoration of cell GSH levels in a number of these conditions have proven to be beneficial. Thus, strategies to boost cell glutathione level are of marked therapeutic significance. Availability of cysteine, a precursor for glutathione synthesis, inside the cell is a critical determinant of cellular glutathione level. N-acetylcysteine and -lipoic acid are two pro-glutathione agents that have remarkable clinical potential. The ability of these two clinical drugs to enhance cellular glutathione level, coupled with their favorable effect on the molecular biology of HIV infection may make them useful tools for AIDS treatment.  相似文献   

14.
In the normal rat given a single dose of one mg N,N-dimethyl-4-aminoazobenzene (DAB) via the hepatic portal vein the following biliary metabolites reached their maximal rates of excretion in the sequence: 4'-sulphonyloxy-DAB, N-(glutathione-S-methylene)-4-aminoazobenzene (GSCH2AB), 4'-sulphonyloxy-N-methyl-4-aminoazobenzene (4'-sulphonyloxy-MAB) 4'-sulphonyloxy-GSCH2AB and MAB-4'-beta-glucuronide. The unusual and relatively unstable N-methylene glutathione conjugates were major metabolites accounting for up to 70% of the whole. It was shown that all the 4-aminoazobenzene (AB) and perhaps all of the 4'-sulphonyloxy-AB, which may be observed in bile, are artefacts due to decomposition of GSCH2AB and 4'-sulphonyloxy-GSCH2AB respectively and that biliary excretion of N-methyl oxidised products of MAB and 4'-hydroxy-MAB is dependent on their conversion to the GSH conjugates, GSCH2AB and 4'-hydroxy-GSCH2AB respectively. Sulphotransferase inhibition by pentachlorophenol caused a reduction in the excretion of all sulphate conjugates, but biliary excretion as a whole was not reduced significantly due to a compensatory increase in the excretion of MAB-4'-beta-glucuronide and the appearance of 4'-OH-GSCH2AB. Glutathione (GSH) depletion by diethylmaleate caused a reduction in biliary metabolites of DAB by lowering the levels of GSH conjugates. This was because the amount of N-methyl oxidation of MAB and 4'-hydroxy-MAB were proportional to the amount of GSH present. The fall in N-methyl oxidation was not compensated for by an increase in 4'-hydroxylation and was accompanied by a delay in the appearance of 4'-hydroxylated metabolites. The administration of potential precursors of 4'-sulphonyloxy-GSCH2AB establishes the sequence of reactions resulting in its formation to be 4'-hydroxylation, N-methyl oxidation, GSH conjugation and O-sulphation.  相似文献   

15.
Glutathione (γ-glutamylcysteinylglycine, GSH and oxidized glutathione, GSSG), may function as a neuromodulator at the glutamate receptors and as a neurotransmitter at its own receptors. We studied now the effects of GSH, GSSG, glutathione derivatives and thiol redox agents on the spontaneous, K+- and glutamate-agonist-evoked releases of [3H]dopamine from mouse striatal slices. The release evoked by 25 mM K+ was inhibited by GSH, S-ethyl-, -propyl-, -butyl- and pentylglutathione and glutathione sulfonate. 5,5′-Dithio-bis-2-nitrobenzoate (DTNB) and l-cystine were also inhibitory, while dithiothreitol (DTT) and l-cysteine enhanced the K+-evoked release. Ten min preperfusion with 50 μM ZnCl2 enhanced the basal unstimulated release but prevented the activation of K+-evoked release by DTT. Kainate and 2-amino-3-hydroxy-5-methyl-4-isoxazolepropionate (AMPA) evoked dopamine release but the other glutamate receptor agonists N-methyl-d-aspartate (NMDA), glycine (1 mM) and trans-1-aminocyclopentane-1,3-dicarboxylate (t-ACPD, 0.5 mM), and the modulators GSH, GSSG, glutathione sulfonate, S-alkyl-derivatives of glutathione, DTNB, cystine, cysteine and DTT (all 1 mM) were without effect. The release evoked by 1 mM glutamate was enhanced by 1 mM GSH, while GSSG, glutathionesulfonate and S-alkyl derivatives of glutathione were generally without effect or inhibitory. NMDA (1 mM) evoked release only in the presence of 1 mM GSH but not with GSSG, other peptides or thiol modulators. l-Cysteine (1 mM) enhanced the glutamate-evoked release similarly to GSH. The activation by 1 mM kainate was inhibited by S-ethyl-, -propyl-, and -butylglutathione and the activation by 0.5 mM AMPA was inhibited by S-ethylglutathione but enhanced by GSSG. Glutathione alone does not directly evoke dopamine release but may inhibit the depolarization-evoked release by preventing the toxic effects of high glutamate, and by modulating the cysteine–cystine redox state in Ca2+ channels. GSH also seems to enhance the glutamate-agonist-evoked release via both non-NMDA and NMDA receptors. In this action, the γ-glutamyl and cysteinyl moieties of glutathione are involved.  相似文献   

16.
The physiological consequences of glutathione variations.   总被引:11,自引:0,他引:11  
S Uhlig  A Wendel 《Life sciences》1992,51(14):1083-1094
The major low molecular weight thiol inside cells, the tripeptide glutathione (GSH), is of importance for protection of the cell against oxidative challenge, for thiol homeostasis required to guarantee basic functions, and for defence mechanisms against xenobiotics. Since the pathophysiological significance of a perturbed GSH status in human disease is less clear, this review evaluates the consequences of in vivo variations of GSH. Owing to intracellular GSH concentrations above 2 mM depletion of GSH as such has little metabolic consequences unless an additional stress is superimposed. The kinetic properties of GSH-dependent enzymes imply that loss of up to 90% of intracellular GSH may still be compatible with cellular integrity. Mitochondrial GSH, which accounts for about 10% of total cellular GSH, may define the threshold beyond that toxicity commences. Thus, in cases of severe GSH-depletion a substitution of GSH as a therapeutic measure seems justified. Such a severe depletion of GSH has been described for some diseases such as liver dysfunction, AIDS or pulmonary fibrosis.  相似文献   

17.
An important aspect of the catalytic mechanism of microsomal glutathione transferase (MGST1) is the activation of the thiol of bound glutathione (GSH). GSH binding to MGST1 as measured by thiolate anion formation, proton release, and Meisenheimer complex formation is a slow process that can be described by a rapid binding step (K(GSH)d = 47 +/- 7 mM) of the peptide followed by slow deprotonation (k2 = 0.42 +/- 0.03 s(-1). Release of the GSH thiolate anion is very slow (apparent first-order rate k(-2) = 0.0006 +/- 0.00002 s(-)(1)) and thus explains the overall tight binding of GSH. It has been known for some time that the turnover (kcat) of MGST1 does not correlate well with the chemical reactivity of the electrophilic substrate. The steady-state kinetic parameters determined for GSH and 1-chloro-2,4-dinitrobenzene (CDNB) are consistent with thiolate anion formation (k2) being largely rate-determining in enzyme turnover (kcat = 0.26 +/- 0.07 s(-1). Thus, the chemical step of thiolate addition is not rate-limiting and can be studied as a burst of product formation on reaction of halo-nitroarene electrophiles with the E.GS- complex. The saturation behavior of the concentration dependence of the product burst with CDNB indicates that the reaction occurs in a two-step process that is characterized by rapid equilibrium binding ( = 0.53 +/- 0.08 mM) to the E.GS- complex and a relatively fast chemical reaction with the thiolate (k3 = 500 +/- 40 s(-1). In a series of substrate analogues, it is observed that log k3 is linearly related (rho value 3.5 +/- 0.3) to second substrate reactivity as described by Hammett sigma- values demonstrating a strong dependence on chemical reactivity that is similar to the nonenzymatic reaction (rho = 3.4). Microsomal glutathione transferase 1 displays the unusual property of being activated by sulfhydryl reagents. When the enzyme is activated by N-ethylmaleimide, the rate of thiolate anion formation is greatly enhanced, demonstrating for the first time the specific step that is activated. This result explains earlier observations that the enzyme is activated only with more reactive substrates. Taken together, the observations show that the kinetic mechanism of MGST1 can be described by slow GSH binding/thiolate formation followed by a chemical step that depends on the reactivity of the electrophilic substrate. As the chemical reactivity of the electrophile becomes lower the rate-determining step shifts from thiolate formation to the chemical reaction.  相似文献   

18.
A complex of haemoglobin and GSH was prepared by incubating haemoglobin with GSH and acetylphenylhydrazine. GSH could be released from the crude preparation by incubation with NADPH. However, when the haemoglobin preparation was separated from glutathione reductase by DEAE-Sephadex chromatography, NADPH no longer released GSH. Rather, the addition of a combination of either partially purified human erythrocyte or crystalline glutathione reductase and NADPH was required to release GSH from the haemoglobin-GSH complex. This complex is commonly believed to represent a mixed disulphide of GSH and the cysteine-beta-93 thiol group. This interpretation was supported by the finding that prior alkylation of available haemoglobin thiol groups prevented the formation of the complex. By using haemoglobin-[(35)S]GSH complex as a substrate, it was shown that GSH itself released the radioactivity from the complex only very slowly. In contrast, the release of [(35)S]GSH was very rapid in the presence of NADPH and glutathione reductase. This suggests that the cleavage of the haemoglobin-GSH complex is not mediated by GSH with cyclic reduction of GSSG formed, but rather proceeds enzymically through glutathione reductase.  相似文献   

19.
A glutathione S-transferase (GST) with activity toward 1, 2-epoxy-2-methyl-3-butene (isoprene monoxide) and cis-1, 2-dichloroepoxyethane was purified from the isoprene-utilizing bacterium Rhodococcus sp. strain AD45. The homodimeric enzyme (two subunits of 27 kDa each) catalyzed the glutathione (GSH)-dependent ring opening of various epoxides. At 5 mM GSH, the enzyme followed Michaelis-Menten kinetics for isoprene monoxide and cis-1, 2-dichloroepoxyethane, with Vmax values of 66 and 2.4 micromol min-1 mg of protein-1 and Km values of 0.3 and 0.1 mM for isoprene monoxide and cis-1,2-dichloroepoxyethane, respectively. Activities increased linearly with the GSH concentration up to 25 mM. 1H nuclear magnetic resonance spectroscopy showed that the product of GSH conjugation to isoprene monoxide was 1-hydroxy-2-glutathionyl-2-methyl-3-butene (HGMB). Thus, nucleophilic attack of GSH occurred on the tertiary carbon atom of the epoxide ring. HGMB was further converted by an NAD+-dependent dehydrogenase, and this enzyme was also purified from isoprene-grown cells. The homodimeric enzyme (two subunits of 25 kDa each) showed a high activity for HGMB, whereas simple primary and secondary alcohols were not oxidized. The enzyme catalyzed the sequential oxidation of the alcohol function to the corresponding aldehyde and carboxylic acid and followed Michaelis-Menten kinetics with respect to NAD+ and HGMB. The results suggest that the initial steps in isoprene metabolism are a monooxygenase-catalyzed conversion to isoprene monoxide, a GST-catalyzed conjugation to HGMB, and a dehydrogenase-catalyzed two-step oxidation to 2-glutathionyl-2-methyl-3-butenoic acid.  相似文献   

20.
A reduction of cellular glutathione (GSH) content was observed when isolated rat hepatocytes were incubated with a stereoisomer of a uricosuric diuretic (S-8666) at a high concentration. Subsequent studies have revealed it was due to conjugation of GSH and S-8666 (-)-enantiomer in the liver cytosol. The (+)-enantiomer strongly inhibited the conjugation reaction, therefore, GSH depletion did not take place when a racemic form of S-8666 was incubated with the liver cells. A possible chemical structure of the GSH-conjugate is tentatively proposed.  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