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.
Tyr114 and Tyr197 are highly conserved residues in the active site of human glutathione reductase, Tyr114 in the glutathione disulfide (GSSG) binding site and Tyr197 in the NADPH site. Mutation of either residue has profound effects on catalysis. Y197S and Y114L have 17% and 14% the activity of the wild-type enzyme, respectively. Mutation of Tyr197, in the NADPH site, leads to a decrease in Km for GSSG, and mutation of Tyr114, in the GSSG site, leads to a decrease in Km for NADPH. This behavior is predicted for enzymes operating by a ping-pong mechanism where both half-reactions partially limit turnover. Titration of the wild-type enzyme or Y114L with NADPH proceeds in two phases, Eox to EH2 and EH2 to EH2-NADPH. In contrast, Y197S reacts monophasically, showing that excess NADPH fails to enhance the absorbance of the thiolate-FAD charge-transfer complex, the predominant EH2 form of glutathione reductase. The reductive half-reactions of the wild-type enzyme and of Y114L are similar; FAD reduction is fast (approximately 500 s-1 at 4 degreesC) and thiolate-FAD charge-transfer complex formation has a rate of 100 s-1. In Y197S, these rates are only 78 and 5 s-1, respectively. The oxidative half-reaction, the rate of reoxidation of EH2 by GSSG, of the wild-type enzyme is approximately 4-fold faster than that of Y114L. These results are consistent with Tyr197 serving as a gate in the binding of NADPH, and they indicate that Tyr114 assists the acid catalyst His467'.  相似文献   

2.
1. Purification of horse-liver glutathione reductase was obtained by affinity chromatography on N6-(6-aminohexyl)-adenosine-1'5'-bisphosphate Sepharose (N6-2'5'-ADP-Sepharose) and Reactive Red-120-Agarose, and chromatography on DEAE-Sephadex and Sephacryl S-300. 2. The final preparation had 248 U/mg specific activity after 11,174-fold purification with 47% final recovery, and was homogeneous by SDS-electrophoresis. It showed charge heterogeneity in non-denaturing electrophoresis and chromatofocusing, with several peaks of pI between 5.7 and 6.7. 3. The enzyme was homodimeric (107,000 native MW), with S20w = 6.31 S, and 41.22 A of hydrodynamic radius. It showed absorption peaks at 270, 370 and 462 nm, a characteristic of flavoproteins. 4. When NADPH was substituted by deamino-NADPH or NADH the enzyme showed 69 and 8.5% activity, respectively, while with glutathione-CoA mixed disulfide the enzyme had 23% of the activity shown with GSSG. Apparent Km values of 8.8, 680, 59, and 560 microM were measured for NADPH, NADH, GSSG and ferricyanide, respectively.  相似文献   

3.
The nitric-oxide synthase (NOS; EC 1.14.13.39) reaction is formulated as a partially tetrahydrobiopterin (H4Bip)-dependent 5-electron oxidation of a terminal guanidino nitrogen of L-arginine (Arg) associated with stoichiometric consumption of dioxygen (O2) and 1.5 mol of NADPH to form L-citrulline (Cit) and nitric oxide (.NO). Analysis of NOS activity has relied largely on indirect methods such as quantification of nitrite/nitrate or the coproduct Cit; we therefore sought to directly quantify .NO formation from purified NOS. However, by two independent methods, NOS did not yield detectable .NO unless superoxide dismutase (SOD; EC 1.15.1.1) was present. In the presence of H4Bip, internal .NO standards were only partially recovered and the dismutation of superoxide (O2-.), which otherwise scavenges. .NO to yield ONOO-, was a plausible mechanism of action of SOD. Under these conditions, a reaction between NADPH and ONOO- resulted in considerable overestimation of enzymatic NADPH consumption. SOD lowered the NADPH:Cit stoichiometry to 0.8-1.1, suggesting either that additional reducing equivalents besides NADPH are required to explain Arg oxidation to .NO or that .NO was not primarily formed. The latter was supported by an additional set of experiments in the absence of H4Bip. Here, recovery of internal .NO standards was unaffected. Thus, a second activity of SOD, the conversion of nitroxyl (NO-) to .NO, was a more likely mechanism of action of SOD. Detection of NOS-derived nitrous oxide (N2O) and hydroxylamine (NH2OH), which cannot arise from .NO decomposition, was consistent with formation of an .NO precursor molecule such as NO-. When, in the presence of SOD, glutathione was added, S-nitrosoglutathione was detected. Our results indicate that .NO is not the primary reaction product of NOS-catalyzed Arg turnover and an alternative reaction mechanism and stoichiometry have to be taken into account.  相似文献   

4.
The effects of reduced glutathione (GSH) and glutathione disulfide (GSSG) on lipid peroxidation were investigated in rat liver microsomes containing deficient or adequate amounts of alpha-tocopherol (alpha-TH). Rates of formation of thiobarbituric acid reactive substances (TBARS) as well as rates of consumption of alpha-TH and O2 were decreased by GSH and were more pronounced in the NADPH-dependent assay system than in the ascorbate-dependent system. The GSH-dependent inhibition of lipid peroxidation was potentiated by GSSG in the NADPH-dependent assay system, but it had no effect in the nonenzymatic system. Diphenyliodonium chloride, an inhibitor of NADPH cytochrome P-450 reductase, completely prevented lipid peroxidation in the NADPH-dependent assay system whereas it had no effect on the ascorbate-dependent system. This is further evidenced by the fact that purified rat liver microsomal NADPH cytochrome P-450 reductase (EC 1.6.2.4) was inhibited approximately 24% and 52% by 5 mM GSH and 5 mM GSH + 2.5 mM GSSG, respectively. Glutathione disulfide alone had no effect on reductase activity. Similarly, other disulfides such as cystine, cystamine and lipoic acid were without effect on reductase activity. These results clearly delineate different mechanisms underlying the combined effects of GSH and GSSG on microsomal lipid peroxidation in rat liver. One mechanism involves recycling of microsomal alpha-TH by GSH during oxidative stress via a labile protein, ostensibly associated with "free radical reductase" activity. A second glutathione-dependent mechanism appears to be mediated through the inhibition of NADPH cytochrome P-450 reductase. The enhanced inhibition by GSH + GSSG of microsomal lipid peroxidation in the NADPH-dependent assay system suggests suppression of the initiation phase at the level of NADPH cytochrome P-450 reductase which is independent of microsomal alpha-TH.  相似文献   

5.
Previous studies [Kondo, T., Dale, G. L. and Beutler, E. (1981) Biochim. Biophys. Acta, 645, 132-136] have shown evidence for the existence of two different active-transport processes for glutathione disulphide (GSSG) in human erythrocytes (the high-Km and low-Km processes). In the present investigation adenosine-triphosphate-dependent transport of glutathione S-conjugate was characterized in comparison with active glutathione transport using inside-out vesicles from human erythrocytes. Incubation of the vesicles with glutathione S-conjugate (S-2,4-dinitrophenylglutathione) was found to inhibit competitively the high-Km process of GSSG transport but not significantly affect the low-Km process. The glutathione S-conjugate transport required ATP. A lineweaver-Burk plot of the transport rate as a function of the conjugate concentration gave an apparent Km value of 0.94 mM. The Km value of ATP-Mg was 0.76 mM. The transport of glutathione S-conjugate was dependent on temperature. Preincubation of vesicles with dithiothreitol resulted in an increase of the transport rate while thiol reagents, such as iodoacetamide, N-ethylmaleimide and p-chloromercuribenzoate inhibited the transport. Addition of nucleotides, such as CTP, UTP or GTP had no effect on the transport. These findings suggest that glutathione S-conjugate formed by the catalytic reaction of glutathione S-transferase in erythrocytes under the exposure to electrophilic compounds, is eliminated via the same transport process for GSSG elevated under oxidative stress.  相似文献   

6.
The tripeptide glutathione (GSH) is the predominant low molecular weight thiol reductant in mammalian cells. In this report, we show that at concentrations at which GSH is typically present in the intracellular milieu, GSH and the oxidized GSH derivatives GSH disulfide (GSSG) and glutathione sulfonate each irreversibly inactivate up to 100% of the activity of purified Ca2+- and phosphatidylserine (PS)-dependent protein kinase C (PKC) isozymes in a concentration-dependent manner by a novel nonredox mechanism that requires neither glutathiolation of PKC nor the reduction, formation, or isomerization of disulfide bridges within PKC. Our evidence for a nonredox mechanism of PKC inactivation can be summarized as follows. GSSG antagonized the Ca2+- and PS-dependent activity of purified rat brain PKC with the same efficacy (IC50 = 3 mM) whether or not the reductant dithiothreitol was present. Glutathione sulfonate, which is distinguished from GSSG and GSH by its inability to undergo disulfide/thiol exchange reactions, was as effective as GSSG in antagonizing Ca2+- and PS-dependent PKC catalysis. The irreversibility of the inactivation mechanism was indicated by the stability of the inactivated form of PKC to dilution and extensive dialysis. The inactivation mechanism did not involve the nonspecific phenomena of denaturation and aggregation of PKC because it obeyed pseudo-first order kinetics and because the hinge region of PKC-alpha remained a preferential target of tryptic attack following GSH inactivation. The selectivity of GSH in the inactivation of PKC was also indicated by the lack of effect of the tripeptides Tyr-Gly-Gly and Gly-Ala-Gly on the activity of PKC. Furthermore, GSH antagonism of the Ser/Thr kinase casein kinase 2 was by comparison weak (<25%). Inactivation of PKC-alpha was not accompanied by covalent modification of the isozyme by GSH or other irreversible binding interactions between PKC-alpha and the tripeptide, but it was associated with an increase in the susceptibility of PKC-alpha to trypsinolysis. Treatment of cultured rat fibroblast and human breast cancer cell lines with N-acetylcysteine resulted in a substantial loss of Ca2+- and PS- dependent PKC activity in the cells within 30 min. These results suggest that GSH exerts negative regulation over cellular PKC isozymes that may be lost when oxidative stress depletes the cellular GSH pool.  相似文献   

7.
Diabetic patients undergo a chronic oxidative stress. This phenomenon is demonstrated by low levels of reduced glutathione (GSH) levels. The NADPH used by glutathione reductase for the reduction of oxidized glutathione (GSSG) to GSH is also used by aldose reductase for the reduction of glucose to sorbitol through the polyol pathway. The competition for NADPH could be responsible for the decreased glutathione levels found in non-insulin-dependent diabetic patients. For this purpose, we investigated the effect of polyol pathway inhibition on the glutathione redox status in these patients. We measured GSH and GSSG levels in erythrocytes of non-insulin-dependent diabetic patients (n = 15) before and after 1 week of treatment with placebo, followed by 1 week of treatment with an aldose reductase inhibitor (tolrestat 200 mg/dl). We found lower GSH levels (7.7 +/- 1.4 mumol/g hemoglobin [Hb]), higher GSSG levels (0.35 +/- 0.09 mumol/g Hb), and lower GSH/GSSG ratios (23.9 +/- 7.7) in diabetics compared with controls (n = 15; 9.8 +/- 0.8 mumol/g Hb, P < .001; 0.17 +/- 0.02, P < .001; and 58.3 +/- 9.1, P < .001, respectively). We did not demonstrate any statistical difference after 1 week of treatment with placebo. In contrast, the treatment with tolrestat induced a significant increase in GSH (8.9 +/- 0.7 mumol/g Hb, P < .01), a decrease in GSSG (0.25 +/- 0.06 mumol/g Hb, P < .02), and an increase in the GSH/GSSG ratio (37.3 +/- 8.4, P < .01). These data strongly support the hypothesis that the polyol pathway plays an important role in the impairment of the glutathione redox status in diabetic patients.  相似文献   

8.
The diterpenoic compound steviol (ent-kaur-16-en-13-ol-19-oic acid) is the aglycone of sweet glycosides accumulated in Stevia rebaudiana Bertoni. This compound is the hydroxylated form of ent-kaurenoic acid (ent-kaur-16-en-19-oic acid; ent-KA). The hydroxylation of ent-KA to form steviol requiring NADPH and molecular oxygen was detected in stroma prepared from S. rebaudiana Bertoni. The enzyme was purified from leaf extract to apparent homogeneity with a molecular mass of 39 kDa. Taken together with the value of 160 kDa estimated for native enzyme, this suggested that the hydroxylating enzyme is a homotetramer. The N-terminal sequence was determined through 20 residues. The pH optimum was 7.5-7.8. Apparent Km values were 11.1 microM for ent-KA and 20.6 microM for NADPH. Its visible absorption spectrum suggested that the enzyme was flavoprotein. The stoichiometric relationship between the formation of steviol and the utilization of ent-KA and cofactors confirmed the equation ent-KA + NADPH + H(+) + O2-->steviol + NADPH(+) + H2O.  相似文献   

9.
Tri(gamma-glutamylcysteinylglycinyl)trithioarsenite (AsIII(GS)3) is formed in cells and is a more potent mixed-type inhibitor of the reduction of glutathione disulfide (GSSG) by yeast glutathione (GSH) reductase than either arsenite (AsIII) or GSH. The present work examines the effects of valence and complexation of arsenicals with GSH or L-cysteine (Cys) upon potency as competitive inhibitors of the reduction of GSH disulfide (GSSG) by yeast GSH reductase. Trivalent arsenicals were more potent inhibitors than their pentavalent analogs, and methylated trivalent arsenicals were more potent inhibitors than was inorganic trivalent As. Complexation of either inorganic trivalent As or methylarsonous diiodide (CH3As(III)I2) with Cys or GSH produced inhibitors of GSH reductase that were severalfold more potent than the parent arsenicals. In contrast, dimethylarsinous iodide ((CH3)2As(III)I) was a more potent inhibitor than its complexes with either GSH or Cys. Complexes of CH3AsIII with GSH (CH3-AsIII(GS)2) or with Cys (CH3AsIII(Cys)2) were the most potent inhibitors, with Ki's of 0.009 and 0.018 mM, respectively. Inhibition of GSH reductase by arsenicals or arsenothiols was prevented by addition of meso-2,3-dimercaptosuccinic acid (DMSA) to a mixture of enzyme, GSSG, and inhibitor before addition of NADPH. DMSA added to the reaction mixture after NADPH reversed inhibition by (CH3)2As(III)I but had little effect on inhibition by CH3As(III)I2, Ch3AsIII(GS)2, CH3AsIII(Cys)2, or AsIII(GS)3. Partial redox inactivation of the enzyme with NADPH increased the inhibitory potency of CH3As(III)I2 and (CH3)2As(III)I and changed the mode of inhibition for CH3As(III)I2 from competitive to noncompetitive. The greater potency of methylated trivalent arsenicals and arsenothiols than of inorganic trivalent As suggests that biomethylation of As could yield species that inhibit reduction of GSSG and alter the redox status of cells.  相似文献   

10.
PURPOSE: Methylene blue (MB) can be used as an intracellular electron acceptor. The purpose of this study was to demonstrate the usefulness of MB for the determination of total bioreductive capacity of cell suspensions. METHODS AND MATERIALS: We measured oxygen consumption by Clark electrode and pentose cycle activity by release of 14CO2 from 1-14C-glucose. RESULTS: Methylene blue catalyzes the reaction of intracellular reductants NADPH, NADH, and reduced glutathione (GSH) with oxygen, causing the production of hydrogen peroxide. The reaction rate correlates with the negative charge of molecule (NADPH(-4) > NADH(-2) > GSH(-1)), suggesting that reaction with positively charged oxidized MB is the limiting step of the reaction. In a cellular system MB causes the electron flow from cellular endogenous substrates to oxygen. It is activated by the disruption of the NADP+/NADPH ratio due to several processes. These are direct oxidation of NADPH and GSH, the GSH peroxidase catalyzed reaction of GSH with H2O2, followed by NADPH oxidation by oxidized glutathione (GSSG). This results in increased cellular oxygen consumption and stimulation of the oxidative limb of pentose cycle (PC) in the presence of MB. The cellular effect of MB differs from other electron accepting drugs. Diamide and tert-butylhydroperoxide act as direct oxidants, while MB is an electron carrier to oxygen. Accordingly, MB shows the highest effect on PC activation and oxygen consumption. CONCLUSIONS: Our results indicate that MB may be used for the determination of the total bioreductive capacity of the cells, measured by oxygen consumption and PC activation.  相似文献   

11.
Soybean lipoxygenase-mediated cooxidation of reduced glutathione (GSH) and concomitant superoxide generation was examined. The oxidation of GSH was dependent on the concentration of linoleic acid (LA), GSH, and the enzyme. The optimal conditions to observe maximal enzyme velocity included the presence of 0.42 mM LA, 2 mM GSH, and 50 pmole of enzyme/mL. The GSH oxidation was linear up to 10 minutes and exhibited a pH optimum of 9.0. The reaction displayed a Km of 1.49 mM for GSH and Vmax of 1.35 +/- 0.02 mumoles/min/nmole of enzyme. Besides LA, arachidonic and gamma-linolenic acids also supported the lipoxygenase-mediated GSH oxidation. Hydrogen peroxide and 13-hydroperoxylinoleic acid supported GSH cooxidation, but to a very limited extent. Oxidized glutathione (GSSG) was identified as the major product of the reaction based on the depletion of nicotinamide-adenine dinucleotide 3'-phosphate (NADPH) in the presence of glutathione reductase. The GSH oxidation was accompanied by the reduction of ferricytochrome c, which can be completely abolished by superoxide dismutase (SOD), suggesting the generation of superoxide anion radicals. Under optimal conditions, the rate of superoxide generation (measured as the SOD-inhibitable reduction of ferricytochrome c) was 10 +/- 1.0 nmole/min/nmole of enzyme. These results clearly suggest that lipoxygenase is capable of oxidizing GSH to GSSG and simultaneously generating superoxide anion radicals, which may contribute to oxidative stress in cells under certain conditions.  相似文献   

12.
Pyrococcus furiosus is an anaerobic archaeon that grows optimally at 100 degrees C by the fermentation of carbohydrates yielding acetate, CO2, and H2 as the primary products. If elemental sulfur (S0) or polysulfide is added to the growth medium, H2S is also produced. The cytoplasmic hydrogenase of P. furiosus, which is responsible for H2 production with ferredoxin as the electron donor, has been shown to also catalyze the reduction of polysulfide to H2S (K. Ma, R. N. Schicho, R. M. Kelly, and M. W. W. Adams, Proc. Natl. Acad. Sci. USA 90:5341-5344, 1993). From the cytoplasm of this organism, we have now purified an enzyme, sulfide dehydrogenase (SuDH), which catalyzes the reduction of polysulfide to H2S with NADPH as the electron donor. SuDH is a heterodimer with subunits of 52,000 and 29,000 Da. SuDH contains flavin and approximately 11 iron and 6 acid-labile sulfide atoms per mol, but no other metals were detected. Analysis of the enzyme by electron paramagnetic resonance spectroscopy indicated the presence of four iron-sulfur centers, one of which was specifically reduced by NADPH. SuDH has a half-life at 95 degrees C of about 12 h and shows a 50% increase in activity after 12 h at 82 degrees C. The pure enzyme has a specific activity of 7 mumol of H2S produced.min-1.mg of protein-1 at 80 degrees C with polysulfide (1.2 mM) and NADPH (0.4 mM) as substrates. The apparent Km values were 1.25 mM and 11 microM, respectively. NADH was not utilized as an electron donor for polysulfide reduction. P. furiosus rubredoxin (K(m) = 1.6 microM) also functioned as an electron acceptor for SuDH, and SuDH catalyzed the reduction of NADP with reduced P. furiosus ferredoxin (K(m) = 0.7 microM) as an electron donor. The multiple activities of SuDH and its proposed role in the metabolism of S(o) and polysulfide are discussed.  相似文献   

13.
An inducible membrane-bound L-4-hydroxymandelate oxidase (decarboxylating) from Pseudomonas convexa has been solubilized and partially purified. It catalyzes the conversion of L-4-hydroxymandelic acid to 4-hydroxybenzaldehyde in a single step with the stoichiometric consumption of O2 and liberation of CO2. The enzyme is optimally active at pH 6.6 and at 55 degrees C. It requires FAD and Mn2+ for its activity. The membrane-bound enzyme is more stable than the solubilized and purified enzyme. After solubilization it gradually loses its activity when kept at 5 degrees C which can be fully reactivated by freezing and thawing. The Km values for DL-4-hydroxymandelate and FAD are 0.44 mM and 0.038 mM respectively. The enzyme is highly specific for DL-4-hydroxymandelic acid. DL-3,4-Dihydroxymandelic acid competitively inhibited the enzyme reaction. From the Dixon plot the Ki for DL-3,4-dihydroxymandelic acid was calculated to be 1.8 X 10(-4) M. The enzyme is completely inactivated by thiol compounds and not affected by thiol inhibitors. The enzyme is also inhibited by denaturing agents, heavy metal ions and by chelating agents.  相似文献   

14.
2-Ketoaldonate reductase, which is involved in ketogluconate catabolism, was purified to homogeneity from Brevibacterium ketosoreductum ATCC21914. The enzyme was found to catalyze the reduction of 2,5-diketo-D-gluconate to 5-keto-D-gluconate, and to a lesser extent, 2-keto-D-gluconate to D-gluconate, and 2-keto-L-gluconate to L-idonate. The molecular mass of the reductase was 35 kDa by sodium dodecyl sulfate-polyacrylamide gel electrophoresis and 72 kDa by gel filtration, indicating that the native enzyme may exist as a dimer. The reductase was optimally active at pH 6.0 with NADPH as a preferred electron donor. The pI of 4.7 was measured for the enzyme. The apparent Km for 2,5-diketo-D-gluconate and NADPH were 5 microM and 10 microM, respectively. The amino-terminal amino acid sequence was NH2-Ala-Ser-Ile-Ser-Val-Ser-Val-Pro-Ser-Ala- Arg-Leu-Ala-Glu-Asp-Leu-Ser-Asp-Ile-Glu.  相似文献   

15.
Kinetic and structural changes in recombinant human aldose reductase (AR) due to modification by S-nitrosoglutathione (GSNO) were investigated. Incubation of the enzyme with 10-50 microM GSNO led to a time- and concentration-dependent inactivation of the enzyme, with a second-order rate constant of 0.087 +/- 0.009 M-1 min-1. However, upon exhaustive modification, 30-40% of the enzyme activity was retained. The non-inactivated enzyme displayed a 2-3-fold change in Km for NADPH and Km fordl-glyceraldehyde, whereas the Km for the lipid peroxidation product, 4-hydroxy-2-trans nonenal (HNE), was comparable to that of the untreated enzyme. The residual activity of the enzyme after GSNO treatment was less sensitive to inhibition by the active site inhibitor sorbinil or to activation by sulfate. Significantly higher catalytic activity was retained when the enzyme was modified in the presence of NADPH, suggesting relatively low reactivity of the E-NADPH complex with GSNO. The modification site was identified using site-directed mutants in which each of the solvent-exposed cysteines of the enzyme was replaced individually by serine. The mutant C298S was insensitive to GSNO, whereas the sensitivity of the mutants C303S and C80S was comparable to that of the wild-type enzyme. Electrospray ionization mass spectroscopy of the GSNO-modified enzyme revealed a major modified species (70% of the protein) with a molecular mass that was 306 Da higher than that of the untreated enzyme, which is consistent with the addition of a single glutathione molecule to the enzyme. The remaining 30% of the protein displayed a molecular mass that was not significantly different from that of the native enzyme. No nitrosated forms of the enzyme were observed. These results suggest that inactivation of AR by GSNO is due to the selective formation of a single mixed disulfide between glutathione and Cys-298 located at the NADP(H)-binding site of the enzyme.  相似文献   

16.
The eubacterial 3-hydroxy-3-methylglutaryl coenzyme A (HMG-CoA) reductase (EC 1.1.1.34) was purified 3,000-fold from Streptomyces sp. strain CL190 to apparent homogeneity with an overall yield of 2.1%. The purification procedure consisted of (NH4)2SO4 precipitation, heat treatment and anion exchange, hydrophobic interaction, and affinity chromatographies. The molecular mass of the enzyme was estimated to be 41 kDa by sodium dodecyl sulfate-polyacrylamide gel electrophoresis and 100 to 105 kDa by gel filtration chromatography, suggesting that the enzyme is most likely to be a dimer. The enzyme showed a pH optimum of around 7.2, with apparent Km values of 62 microM for NADPH and 7.7 microM for HMG-CoA. A gene from CL190 responsible for HMG-CoA reductase was cloned by the colony hybridization method with an oligonucleotide probe synthesized on the basis of the N-terminal sequence of the purified enzyme. The amino acid sequence of the CL190 HMG-CoA reductase revealed several limited motifs which were highly conserved and common to the eucaryotic and archaebacterial enzymes. These sequence conservations suggest a strong evolutionary pressure to maintain amino acid residues at specific positions, indicating that the conserved motifs might play important roles in the structural conformation and/or catalytic properties of the enzyme.  相似文献   

17.
Polyclonal antibodies against cysteine synthase (CSase; EC 4.2.99.8) isozymes 1, 2, and 3 were used for the detection of complexes of these isozymes with serine acetyltransferase (SATase; EC 2.3.1.30). SATase was partially purified and found to complex with these isozymes by western blotting and immunotitration. When the complexes were treated with a high concentration of O-acetyl-L-serine, they did not dissociate. However, some complexes with CSase 1 or 3 dissociated when left for 24 h at 4 degrees C. Results of western blotting on SDS-PAGE showed that CSase 2 complexed with SATase. CSases 1, 2, and 3 all could complex with SATase, but the tightness of the bond differed.  相似文献   

18.
A rapid decrease of intracellular glutathione (GSH) was observed when exponentially growing cells of Saccharomyces cerevisiae were treated with sublethal concentrations of either dimethyldithiocarbamic acid or thiram [bis(dimethylthiocarbamoyl) disulfide]. The underlying mechanism of this effect possibly involves the intracellular oxidation of dimethyldithiocarbamate anions to thiram, which in turn oxidizes GSH. Overall, a linear relationship was found between thiram concentrations up to 21 microM and production of oxidized GSH (GSSG). Cytochrome c can serve as the final electron acceptor for dimethyldithiocarbamate reoxidation, and it was demonstrated in vitro that NADPH handles the final electron transfer from GSSG to the fungicide by glutathione reductase. These cycling reactions induce transient alterations in the intracellular redox state of several electron carriers and interfere with the respiration of the yeast. Thiram and dimethyldithiocarbamic acid also inactivate yeast glutathione reductase when the fungicide is present within the cells as the disulfide. Hence, whenever the GSH regeneration rate falls below its oxidation rate, the GSH:GSSG molar ratio drops from 45 to 1. Inhibition of glutathione reductase may be responsible for the saturation kinetics observed in rates of thiram elimination and uptake by the yeast. The data suggest also a leading role for the GSH redox cycle in the control of thiram and dimethyldithiocarbamic acid fungitoxicity. Possible pathways for the handling of thiram and dimethyldithiocarbamic acid by yeast are considered with respect to the physiological status, the GSH content, and the activity of glutathione reductase of the cells.  相似文献   

19.
The human pathogen Staphylococcus aureus does not utilize the glutathione thiol/disulfide redox system employed by eukaryotes and many bacteria. Instead, this organism produces CoA as its major low molecular weight thiol. We report the identification and purification of the disulfide reductase component of this thiol/disulfide redox system. Coenzyme A disulfide reductase (CoADR) catalyzes the specific reduction of CoA disulfide by NADPH. CoADR has a pH optimum of 7.5-8.0 and is a dimer of identical subunits of Mr 49,000 each. The visible absorbance spectrum is indicative of a flavoprotein with a lambdamax = 452 nm. The liberated flavin from thermally denatured enzyme was identified as flavin adenine dinucleotide. Steady-state kinetic analysis revealed that CoADR catalyzes the reduction of CoA disulfide by NADPH at pH 7.8 with a Km for NADPH of 2 muM and for CoA disulfide of 11 muM. In addition to CoA disulfide CoADR reduces 4,4'-diphosphopantethine but has no measurable ability to reduce oxidized glutathione, cystine, pantethine, or H2O2. CoADR demonstrates a sequential kinetic mechanism and employs a single active site cysteine residue that forms a stable mixed disulfide with CoA during catalysis. These data suggest that S. aureus employs a thiol/disulfide redox system based on CoA/CoA-disulfide and CoADR, an unorthodox new member of the pyridine nucleotide-disulfide reductase superfamily.  相似文献   

20.
The cytochrome P450 isozymes in rat liver microsomes that catalyze the demethylenation of methylenedioxymethamphetamine enantiomers to the corresponding dihydroxymethamphetamine were characterized. Dihydroxymethamphetamine formation in liver microsomes from male Sprague-Dawley rats exhibited multienzyme kinetics, with Km values in the micromolar/millimolar range. The stereoselectivity [(+)-isomer versus (-)-isomer] varied from 0.78 to 1.94 after pretreatment of the rats with phenobarbital, 3-methylcholanthrene, pregnenolone-16 alpha-carbonitrile, or pyrazole, suggesting that different isozymes participate in the reaction. The low-Km demethylenation was not induced by these compounds and was not inhibited by antibodies raised against CYP2C11. Liver microsomes from female Dark-Agouti rats, a strain genetically deficient in CYP2D1, exhibited demethylenation activities that were 9% of those in microsomes from male Sprague-Dawley rats. The low-Km demethylenation was also inhibited by CYP2D substrates such as sparteine, bufuralol, or desipramine and was almost completely inhibited by antibodies against P450 BTL, which belongs to the CYP2D family. The higg-Km demethylation activity was induced by phenobarbital and pregnenolone-16 alpha-carbonitrile and the activity in both untreated and phenobarbital-induced microsomes was suppressed by anti-CYP2B1 IgG. Experiments with IgG raised against cytochrome b5 suggested that the hemoprotein contributed to the low-Km activity but not the high-Km activity. These results indicate that cytochrome P450 isozymes belonging to the CYP2D subfamily catalyze demethylenation with low Km values and that the reaction occurring with high Km values is likely to be mediated by members of the CYP2B family, but with the possible participation of other phenobarbital-inducible isoforms.  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