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 234 毫秒
1.
Thioredoxin reductase, lipoamide dehydrogenase, and glutathione reductase are members of the pyridine nucleotide-disulfide oxidoreductase family of dimeric flavoenzymes. The mechanisms and structures of lipoamide dehydrogenase and glutathione reductase are alike irrespective of the source (subunit M(r) approximately 55,000). Although the mechanism and structure of thioredoxin reductase from Escherichia coli are distinct (M(r) approximately 35,000), this enzyme must be placed in the same family because there are significant amino acid sequence similarities with the other two enzymes, the presence of a redox-active disulfide, and the substrate specificities. Thioredoxin reductase from higher eukaryotes on the other hand has a M(r) of approximately 55,000 [Luthman, M. & Holmgren, A. (1982) Biochemistry 21, 6628-6633; Gasdaska, P. Y., Gasdaska, J. R., Cochran, S. & Powis, G. (1995) FEBS Lett 373, 5-9; Gladyshev, V. N., Jeang, K. T. & Stadtman, T.C. (1996) Proc. Natl. Acad. Sci. USA 93, 6146-6151]. Thus, the evolution of this family is highly unusual. The mechanism of thioredoxin reductase from higher eukaryotes is not known. As reported here, thioredoxin reductase from human placenta reacts with only a single molecule of NADPH, which leads to a stable intermediate similar to that observed in titrations of lipoamide dehydrogenase or glutathione reductase. Titration of thioredoxin reductase from human placenta with dithionite takes place in two spectral phases: formation of a thiolate-flavin charge transfer complex followed by reduction of the flavin, just as with lipoamide dehydrogenase or glutathione reductase. The first phase requires more than one equivalent of dithionite. This suggests that the penultimate selenocysteine [Tamura, T. & Stadtman, T.C. (1996) Proc. Natl. Acad. Sci. USA 93, 1006-1011] is in redox communication with the active site disulfide/dithiol. Nitrosoureas of the carmustine type inhibit only the NADPH reduced form of human thioredoxin reductase. These compounds are widely used as cytostatic agents, so this enzyme should be studied as a target in cancer chemotherapy. In conclusion, three lines of evidence indicate that the mechanism of human thioredoxin reductase is like the mechanisms of lipoamide dehydrogenase and glutathione reductase and differs fundamentally from the mechanism of E. coli thioredoxin reductase.  相似文献   

2.
Glutathione reductase (EC 1.6.4.2) is a pivotal enzyme of the glutathione antioxidant system in a cell. The kinetic studies of the interaction of glutathione reductase with unfractionated and low molecular weight heparin and dextran sulfate can contribute to explanation of polyanions effect on the conformation changes of glutathione reductase. The tested polyanions inhibit this enzyme and the inhibition effect depends on the ionic strengths and pH value. The most potent inhibitor is dextran sulfate (ID50 is 4.1 micrograms/ml, pH = 6.8, without NaCl). The ionic strength (> 100 mM) allows the reactivating of GR if the concentration of DS is not higher than 80 micrograms/ml. The inhibition effect of tested polyanions is caused by electrostatic interactions with enzyme; the kinetic analyses indicate that it is a mixed inhibition with respect to oxidized glutathione or NADPH.  相似文献   

3.
Flavin reductase catalyzes the reduction of free flavins by NAD(P)H. As isolated, Escherichia coli flavin reductase does not contain any flavin prosthetic group but accommodates both the reduced pyridine nucleotide and the flavin substrate in a ternary complex prior to oxidoreduction. The reduction of riboflavin by NADPH catalyzed by flavin reductase has been studied by static and rapid kinetics absorption spectroscopies. Static absorption spectroscopy experiments revealed that, in the presence of riboflavin and reduced pyridine nucleotide, flavin reductase stabilizes, although to a small extent, a charge-transfer complex of NADP+ and reduced riboflavin. In addition, reduction of riboflavin was found to be essentially irreversible. Rapid kinetics absorption spectroscopy studies demonstrated the occurrence of two intermediates with long-wavelength absorption during the catalytic cycle. Such intermediate species exhibit spectroscopic properties similar to those of charge-transfer complexes of oxidized flavin and NAD(P)H, and reduced flavin and NAD(P)+, respectively, which have been identified as intermediates during the reaction of flavoenzymes of the ferredoxin-NADP+ reductase family. Thus, a minimal kinetic scheme for the reaction of flavin reductase with NADPH and riboflavin can be proposed. After formation of the Michaelis complex of flavin reductase with NADPH and riboflavin, a first intermediate, identified as a charge-transfer complex of NADPH and riboflavin, is formed. It is followed by a second charge-transfer intermediate of enzyme-bound NADP+ and reduced riboflavin. The latter decays, yielding the Michaelis complex of flavin reductase with NADP+ and reduced riboflavin, which then dissociates to complete the reaction. These results support the initial hypothesis of a structural similarity between flavin reductase and the enzymes of the ferredoxin-NADP+ reductase family and extend it at a functional level.  相似文献   

4.
The flavoenzyme glutathione reductase catalyzes the NADPH-dependent reduction of glutathione disulfide, yielding two molecules of glutathione. The oxidation-reduction potentials, Eox/EH2 (two-electron reduced enzyme), for yeast, Escherichia coli, and human glutathione reductase have been determined between pH 6.0 and 9.8 relative to the nonphysiological substrate couple NAD+/NADH and were found to be -237, -243, and -227 mV (+/-5 mV) at pH 7.0 and 20 degreesC, respectively. The potential as a function of pH demonstrated slopes of -51, -45, and -42 mV/pH unit, respectively, at low pH and -37, -31, and -34 mV/pH unit, respectively, at high pH. The change in slope indicated pKa values of 7.4, 8.5, and 7.6, respectively. The slopes indicate that two protons are associated with the two-electron reduction of Eox at low pH and that only one proton is involved with the two-electron reduction of Eox at high pH, provided that the effects of nearby titratable residues are considered in the data analysis. The influence of four such groups, Cys50, Cys45, His456', and either Tyr107 or the flavin-(N3), has been included (residue numbering refers to the yeast sequence). The enzyme loses activity upon deprotonation of the acid-base catalyst at high pH. Since the pKa ascribed to the EH2-to-EH- ionization is lower than the pKa of the acid-base catalyst, both the EH2 and EH- forms of glutathione reductase must be catalytically active, in contrast to the closely related enzyme lipoamide dehydrogenase, for which only EH2 is active.  相似文献   

5.
Saccharomyces, human and two Arabidopsis (ATR1 and ATR2) NADPH-P-450 reductases were expressed in yeast, purified to homogeneity and used to raise antibodies. Among the P-450-reductases, ATR2 contrasted by its very low FMN affinity and required a thiol-reducing agent for efficient cofactor binding to the FMN-depleted enzyme. Analysis of reductase kinetic properties using artificial acceptors and different salt conditions suggested marked differences between reductases in their FAD and FMN environments and confirmed the unusual properties of the ATR2 FMN-binding domain. Courses of flavin reductions by NADPH were analysed by rapid kinetic studies. The human enzyme was characterized by a FAD reduction rate sixfold to tenfold slower than values for the three other reductases. Following the fast phase of reduction, expected accumulation of flavin semiquinone was observed for the human and ATR1 but not for ATR2 and the yeast reductases. Consistently, redox potential for the FMN semiquinone/reduced couple in the yeast enzyme was found to be more positive than the value for the FMN oxidized/semiquinone couple. This situation was reminiscent of similar inversion observed in bacterial P-450 BM3 reductase. Affinities of reductases for rabbit P-450 2B4 and supported monooxygenase activities in reconstituted systems highly depended on the reductase source. The human enzyme exhibited the highest affinity but supported the lowest kcat whereas the yeast reductase gave the best kcat but with the lowest affinity. ATR1 exhibited both high affinity and efficiency. No simple relation was found between reductase activities with artificial and natural (P-450) acceptors. Thus marked differences in kinetic and redox parameters between reductases dramatically affect their respective abilities to to support P-450 functions.  相似文献   

6.
A glutathione reductase null mutant of Saccharomyces cerevisiae was isolated in a synthetic lethal genetic screen for mutations which confer a requirement for thioredoxin. Yeast mutants that lack glutathione reductase (glr1 delta) accumulate high levels of oxidized glutathione and have a twofold increase in total glutathione. The disulfide form of glutathione increases 200-fold and represents 63% of the total glutathione in a glr1 delta mutant compared with only 6% in wild type. High levels of oxidized glutathione are also observed in a trx1 delta, trx2 delta double mutant (22% of total), in a glr1 delta, trx1 delta double mutant (71% of total), and in a glr1 delta, trx2 delta double mutant (69% of total). Despite the exceptionally high ratio of oxidized/reduced glutathione, the glr1 delta mutant grows with a normal cell cycle. However, either one of the two thioredoxins is essential for growth. Cells lacking both thioredoxins and glutathione reductase are not viable under aerobic conditions and grow poorly anaerobically. In addition, the glr1 delta mutant shows increased sensitivity to the thiol oxidant diamide. The sensitivity to diamide was suppressed by deletion of the TRX2 gene. The genetic analysis of thioredoxin and glutathione reductase in yeast runs counter to previous studies in Escherichia coli and for the first time links thioredoxin with the redox state of glutathione in vivo.  相似文献   

7.
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   

8.
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   

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.
A K-12 strain of Escherichia coli that overproduces methylenetetrahydrofolate reductase (MetF) has been constructed, and the enzyme has been purified to apparent homogeneity. A plasmid specifying MetF with six histidine residues added to the C terminus has been used to purify histidine-tagged MetF to homogeneity in a single step by affinity chromatography on nickel-agarose, yielding a preparation with specific activity comparable to that of the unmodified enzyme. The native protein comprises four identical 33-kDa subunits, each of which contains a molecule of noncovalently bound flavin adenine dinucleotide (FAD). No additional cofactors or metals have been detected. The purified enzyme catalyzes the reduction of methylenetetrahydrofolate to methyltetrahydrofolate, using NADH as the reductant. Kinetic parameters have been determined at 15 degreesC and pH 7.2 in a stopped-flow spectrophotometer; the Km for NADH is 13 microM, the Km for CH2-H4folate is 0.8 microM, and the turnover number under Vmax conditions estimated for the reaction is 1,800 mol of NADH oxidized min-1 (mol of enzyme-bound FAD)-1. NADPH also serves as a reductant, but exhibits a much higher Km. MetF also catalyzes the oxidation of methyltetrahydrofolate to methylenetetrahydrofolate in the presence of menadione, which serves as an electron acceptor. The properties of MetF from E. coli differ from those of the ferredoxin-dependent methylenetetrahydrofolate reductase isolated from the homoacetogen Clostridium formicoaceticum and more closely resemble those of the NADH-dependent enzyme from Peptostreptococcus productus and the NADPH-dependent enzymes from eukaryotes.  相似文献   

11.
Nitrosothiols, the product of the reaction of nitric oxide-derived species (NOx) with thiols, participate in both cell signaling and cytotoxic events. Glutathione has recently been shown to modulate nitrosothiol-mediated signal transduction and to protect against NOx-mediated cytotoxicity. We have investigated the role of protein S-thiolation/dethiolation as a potential mechanism by which glutathione regulates nitrosothiol signaling and toxicity. Our data show that exogenous sources of NOx decreased both free protein thiol and total glutathione levels in endothelial cells. The decrease in glutathione levels could not be accounted for by formation of S-nitrosoglutathione (GSNO) since borohydride treatment of the nonprotein fraction of cell extracts did not restore glutathione levels, whereas borohydride treatment of protein-containing cell extracts led to recovery of glutathione levels. The NOx-mediated decrease in glutathione and protein thiol content was correlated with an increase in protein mixed disulfide formation, as measured by the incorporation of [35S]glutathione into cellular proteins. [35S]glutathione was incorporated into proteins via a covalent disulfide bond since dithiothreitol removed the radiolabel from cellular proteins. The withdrawal of the exogenous NOx source led to recovery of free protein thiol and cellular glutathione levels, which correlated with the dethiolation of proteins. Dethiolation required the action of the glutathione redox system since 1, 3-bis(2-chloroethyl)-1-nitrosourea, an inhibitor of glutathione reductase, blocked both the recovery of glutathione levels and the dethiolation of proteins. These results suggest that exposure of cells to NOx does not lead to accumulation of GSNO but rather stimulates protein S-thiolation, a mechanism which may have important implications with respect to nitrosothiol signaling and toxicity.  相似文献   

12.
Glutaredoxins (Grxs) catalyze reversible oxidation/reduction of protein disulfide groups and glutathione-containing mixed disulfide groups via an active site Grx-glutathione mixed disulfide (Grx-SG) intermediate. The NMR solution structure of the Escherichia coli Grx3 mixed disulfide with glutathione (Grx3-SG) was determined using a C14S mutant which traps this intermediate in the redox reaction. The structure contains a thioredoxin fold, with a well-defined binding site for glutathione which involves two intermolecular backbone-backbone hydrogen bonds forming an antiparallel intermolecular beta-bridge between the protein and glutathione. The solution structure of E. coli Grx3-SG also suggests a binding site for a second glutathione in the reduction of the Grx3-SG intermediate, which is consistent with the specificity of reduction observed in Grxs. Molecular details of the structure in relation to the stability of the intermediate and the activity of Grx3 as a reductant of glutathione mixed disulfide groups are discussed. A comparison of glutathione binding in Grx3-SG and ligand binding in other members of the thioredoxin superfamily is presented, which illustrates the highly conserved intermolecular interactions in this protein family.  相似文献   

13.
Nitrate reductase of Neurospora crassa is a complex multi-redox protein composed of two identical subunits, each of which contains three distinct domains, an amino-terminal domain that contains a molybdopterin cofactor, a central heme-containing domain, and a carboxy-terminal domain which binds a flavin and a pyridine nucleotide cofactor. The flavin domain of nitrate reductase appears to have structural and functional similarity to ferredoxin NADPH reductase (FNR). Using the crystal structure of FNR and amino acid identities in numerous nitrate reductases as guides, site-directed mutagenesis was used to replace specific amino acids suspected to be involved in the binding of the flavin or pyridine nucleotide cofactors and thus important for the catalytic function of the flavin domain. Each mutant flavin domain protein was expressed in Escherichia coli and analyzed for NADPH: ferricyanide reductase activity. The effect of each amino acid substitution upon the activity of the complete nitrate reductase reaction was also examined by transforming each manipulated gene into a nit-3- null mutant of N. crassa. Our results identify amino acid residues which are critical for function of the flavin domain of nitrate reductase and appear to be important for the binding of the flavin or the pyridine nucleotide cofactors.  相似文献   

14.
The thermal dependence of kinetic parameters has been determined in purified or partially purified preparations of cold-hardiness-specific glutathione reductase isozymes from red spruce (Picea rubens Sarg.) needles to investigate a possible functional adaptation of these isozymes to environmental temperature. We have previously purified glutathione reductase isozymes specific for nonhardened (GR-1NH) or hardened (GR-1H) needles. Isozymes that were distinct from GR-1NH and GR-1H, but appeared to be very similar to each other, were also purified from nonhardened (GR-2NH) or hardened (GR-2H) needles (A. Hausladen, R.G. Alscher [1994] Plant Physiol 105: 205-213). GR-1NH had 2-fold higher Km values for NADPH and 2- to 4-fold lower Km values for oxidized glutathione (GSSG) than GR-2NH, and a similar difference was found between GR-1H and GR-2H. However, no differences in Km values were found between the hardiness-specific isozymes GR-1NH and GR-1H. There was only a small effect of temperature on the Km(GSSG) of GR-1H and GR-2H, and no significant temperature effect on Km(NADPH) or Km(GSSG) could be found for the other isozymes. These results are discussed with respect to "thermal kinetic windows," and it is proposed that the relative independence of Km values to temperature ensures adequate enzyme function in a species that is exposed to extreme temperature differences in its natural habitat. A variety of substrates has been tested to characterize any further differences among the isozymes, but all isozymes are highly specific for their substrates, NADPH and GSSG. The reversible reductive inactivation by NADPH (redox interconversion) is more pronounced in GR-1H than in GR-2H. Reduced, partially inactive GR-1H is further deactivated by H2O2, whereas GR-2H is fully reactivated by the same treatment. Both isozymes are reactivated by GSSG or reduced glutathione. It is proposed that this property of GR-2H ensures enzyme function under oxidative conditions, and that in vivo the enzyme may exist in its partially inactive form and be activated in the presence of increased levels of GSSG or oxidants.  相似文献   

15.
Extracellular manipulation of protein disulfide bonds has been implied in diverse biological processes, including penetration of viruses and endotoxin into cells and activation of certain cytokine receptors. We now demonstrate reduction of one or more disulfide bonds in the serine proteinase, plasmin, by a reductase secreted by Chinese hamster ovary or HT1080 cells. Reduction of plasmin disulfide bond(s) triggered proteolysis of the enzyme, generating fragments with the domain structure of the angiogenesis inhibitor, angiostatin. Two of the known reductases secreted by cultured cells are protein disulfide isomerase and thioredoxin, and incubation of plasmin with these purified reductases resulted in angiostatin fragments comparable with those generated from plasmin in cell culture. Thioredoxin-derived angiostatin inhibited proliferation of human dermal microvascular endothelial cells with half-maximal effect at approximately 0.2 microg/ml. Angiostatin made by cells and by purified reductases contained free sulfhydryl group(s), and S-carbamidomethylation of these thiol group(s) ablated biological activity. Neither protein disulfide isomerase nor thioredoxin were the reductases used by cultured cells, because immunodepletion of conditioned medium of these proteins did not affect angiostatin generating activity. The plasmin reductase secreted by HT1080 cells required a small cofactor for activity, and physiologically relevant concentrations of reduced glutathione fulfilled this role. These results have consequences for plasmin activity and angiogenesis, particularly in the context of tumor growth and metastasis. Moreover, this is the first demonstration of extracellular reduction of a protein disulfide bond, which has general implications for cell biology.  相似文献   

16.
The purpose of this study was to determine whether the major thiol-disulfide oxidoreductase activities of the rat liver were altered as a consequence of aging, and whether the alterations had any consequences in terms of hepatic thiol concentrations. Liver fractions were prepared from male and female Fischer 344 rats at ages representing young adulthood (5 months), middle age (15 months) and old age (24-29 months), and the activities of the major thiol-disulfide exchange enzymes, together with protein and nonprotein sulfhydryl contents, were measured using spectrophotometric procedures. Thioltransferase, protein disulfide isomerase and thioredoxin reductase activities in livers of male and female rats were unchanged with aging, while glutathione disulfide (GSSG) reductase activity remained the same (in male livers) or increased (in female livers) as a consequence of aging. Both protein and nonprotein sulfhydryl concentrations were well maintained in old age. The absence of age-dependent alterations in the thiol-protein disulfide exchange enzymes and the lack of compromise in the glutathione GSSG reductase system suggest that aged livers retain their capacity to regulate their thiol-disulfide redox balance under normal physiological conditions.  相似文献   

17.
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   

18.
Aquacobalamin reductase (NADPH), which catalyzes the reduction of aquacobalamin to cob(II)alamin in the synthesis of cobalamin coenzymes, has already been purified from mitochondria of Euglena gracilis and partly characterized. Here, the enzyme was further characterized to clarify its enzymatic properties. The enzyme reduced 2 mol of aquacobalamin per mole of NADPH and had NADPH diaphorase-like activity. The 16 amino acid residues at the NH2-terminal of the enzyme were identical with those of the NADPH diaphorase domain of pyruvate: NADP+ oxidoreductase, which is involved in Euglena wax ester fermentation. Peptide mapping of the aquacobalamin reductase showed that elution during C-18 reversed-phase high-performance liquid chromatography was identical to that of the NADPH diaphorase domain. Immunoblotting indicated that the Euglena aquacobalamin reductase had a higher molecular weight (166,000) in the intact mitochondria than the purified enzyme (65,000), and that the molecular weights of the native and purified enzyme were identical with those of the subunit and the NADPH diaphorase domain, respectively. These results showed that the aquacobalamin reductase isolated earlier was the NADPH diaphorase domain, cleaved by trypsin during preparation of the mitochondrial homogenate from the native enzyme. Purified pyruvate:NADP+ oxidoreductase also had the activity of aquacobalamin reductase, which suggests that the enzyme in Euglena mitochondria has more than one function in the synthesis of cobalamin co-enzymes.  相似文献   

19.
5'-Adenylylsulfate (APS) reductase (EC 1.8.99.-) catalyzes the reduction of activated sulfate to sulfite in plants. The evidence presented here shows that a domain of the enzyme is a glutathione (GSH)-dependent reductase that functions similarly to the redox cofactor glutaredoxin. The APR1 cDNA encoding APS reductase from Arabidopsis thaliana is able to complement the cysteine auxotrophy of an Escherichia coli cysH [3'-phosphoadenosine-5'-phosphosulfate (PAPS) reductase] mutant, only if the E. coli strain produces glutathione. The purified recombinant enzyme (APR1p) can use GSH efficiently as a hydrogen donor in vitro, showing aKm[GSH] approximately of 0.6 mM. Gene dissection was used to express separately the regions of APR1p from amino acids 73-327 (the R domain), homologous with microbial PAPS reductase, and from amino acids 328-465 (the C domain), homologous with thioredoxin. The R and C domains alone are inactive in APS reduction, but the activity is partially restored by mixing the two domains. The C domain shows a number of activities that are typical of E. coli glutaredoxin rather than thioredoxin. Both the C domain and APR1p are highly active in GSH-dependent reduction of hydroxyethyldisulfide, cystine, and dehydroascorbate, showing a Km[GSH] in these assays of approximately 1 mM. The R domain does not show these activities. The C domain is active in GSH-dependent reduction of insulin disulfides and ribonucleotide reductase, whereas APR1p and R domain are inactive. The C domain can substitute for glutaredoxin in vivo as demonstrated by complementation of an E. coli mutant, underscoring the functional similarity between the two enzymes.  相似文献   

20.
BACKGROUND: Disulfide exchange reactions are catalyzed by thiol/disulfide oxidoreductases. These enzymes possess a thioredoxin fold and contain a catalytic disulfide with the sequence Cys-X-X-Cys at the N terminus of an alpha helix. Despite these similarities, the various members differ strongly in their redox potentials (-122 mV to -270 mV). Using the strong oxidant DsbA from Escherichia coli as a model system, we investigated whether the redox properties of these enzymes can be modulated rationally by exchange of the X-X dipeptide. RESULTS: The X-X dipeptide of DsbA (Cys30-Pro31-His32-Cys33) was exchanged by the dipeptides of eukaryotic protein disulfide isomerase (PDI; Gly-His), glutaredoxin (Pro-Tyr), and thioredoxin (Gly-Pro) from E. coli. All variants were less oxidizing than wild-type DsbA and their redox potentials were in the order of the related natural enzymes (DsbA > PDI > glutaredoxin > thioredoxin). The equilibrium constant between glutathione and the thioredoxin-like variant increased 1200-fold compared with wild-type DsbA. The variants also showed a strong increase in the pKa of the nucleophilic cysteine (Cys30). As for glutaredoxin and thioredoxin, the catalytic disulfide stabilized the corresponding variants while destabilizing wild-type DsbA and the PDI-like variant. CONCLUSIONS: The X-X dipeptide in the active site of thiol/disulfide oxidoreductases appears to be the main determinant of the redox properties of these enzymes. This empirical finding should be very useful for the design of new thiol/disulfide oxidoreductases with altered redox potentials and for studying the function of these enzymes in vivo.  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