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 15 毫秒
1.
Aklavinone-11-hydroxylase (RdmE) is a FAD monooxygenase participating in the biosynthesis of daunorubicin, doxorubicin and rhodomycins. The rdmE gene encodes an enzyme of 535 amino acids. The sequence of the Streptomyces purpurascens enzyme is similar to other Streptomyces aromatic polyketide hydroxylases. We overexpressed the gene in Streptomyces lividans and purified aklavinone-11-hydroxylase to apparent homogeneity with four chromatographic steps utilizing a kinetic photometric enzyme assay. The enzyme is active as the monomer with a molecular mass of 60 kDa; it hydroxylates aklavinone and other anthracyclinones. Aklavinone-11-hydroxylase can use both NADH and NADPH as coenzyme but it is slowly inactivated in the presence of NADH. The apparent Km for NADPH is 2 mM and for aklavinone 10 microM. The enzyme is inactivated in the presence of phenylglyoxal and 2,3-butanedione. NADPH protects against inactivation of aklavinone-11-hydroxylase by phenylglyoxal.  相似文献   

2.
The assimilatory NADPH-nitrate reductase (NADPH:nitrate oxidoreductase, EC 1.6.6.3) from Neurospora crassa is competitively inhibited by 3-aminopyridine adenine dinucleotide (AAD) and 3-aminopyridine adenine dinucleotide phosphate (AADP) which are structural analogs of NAD and NADP, respectively. The amino group of the pyridine ring of AAD(P) can react with nitrous acid to yield the diazonium derivative which may covalently bind at the NAD(P) site. As a result of covalent attachment, diazotized AAD(P) causes time-dependent irreversible inactivation of nitrate reductase. However, only the NADPH-dependent activities of the nitrate reductase, i.e. the overall NADPH-nitrate reductase and the NADPH-cytochrome c reductase activities, are inactivated. The reduced methyl viologen- and reduced FAD-nitrate reductase activities which do not utilize NADPH are not inhibited. This inactivation by diazotized AADP is prevented by 1 mM NADP. The inclusion of 1 muM FAD can also prevent inactivation, but the FAD effect differs from the NADP protection in that even after removal of the exogenous FAD by extensive dialysis or Sephadex G-25 filtration chromatography, the enzyme is still protected against inactivation. The FAD-generated protected form of nitrate reductase could again be inactivated if the enzyme was treated with NADPH, dialyzed to remove the NADPH, and then exposed to diazotized AADP. When NADP was substituted for NADPH in this experiment, the enzyme remained in the FAD-protected state. Difference spectra of the inactivated nitrate reductase demonstrated the presence of bound AADP, and titration of the sulfhydryl groups of the inactivated enzyme revealed that a loss of accessible sulfhydryls had occurred. The hypothesis generated by these experiments is that diazotized AADP binds at the NADPH site on nitrate reductase and reacts with a functional sulfhydryl at the site. FAD protects the enzyme against inactivation by modifying the sulfhydryl. Since NADPH reverses this protection, it appears the modifications occurring are oxidation-reduction reactions. On the basis of these results, the physiological electron flow in the nitrate reductase is postulated to be from NADPH via sulfhydryls to FAD and then the remainder of the electron carriers as follows: NADPH leads to -SH leads to FAD leads to cytochrome b-557 leads to Mo leads to NO-3.  相似文献   

3.
The flavin-containing NADH peroxidase of Streptococcus faecalis 10C1, which catalyzes the reaction: NADH + H+ + H2O2----NAD+ + 2H2O, has been purified to homogeneity in our laboratory for analyses of both its structure and redox behavior. Our findings indicate that the enzyme is a tetramer of four apparently identical subunits (Mr = 46,000/subunit), each containing one FAD coenzyme and a second non-flavin, nonmetal redox center. There is no evidence of nonequivalence among the flavins. Dithionite reduction of the enzyme occurs in two steps, with end points of 0.96 and 2.05 eq/FAD. The first step generates a two-electron reduced form of the enzyme (EH2) which is spectrally identical with that generated by aerobic addition of NADH. Our studies suggest that the long-wavelength absorbance band (lambda max approximately 540 nm) exhibited by this form results from charge-transfer interaction between the reduced non-flavin redox center and the oxidized flavin. A second type of long-wavelength charge-transfer absorbance band (lambda max approximately 770 nm) is generated on anaerobic addition of 1 eq of NADH to EH2 and results from interaction between oxidized FAD and the reduced pyridine nucleotide. Either the EH2 X NAD+ or the EH2 X NAD+ X NADH forms may be involved in the catalytic mechanism of the enzyme, as both are reactive with hydrogen peroxide.  相似文献   

4.
Fatty acid synthase of chicken liver is inactivated rapidly and irreversibly by incubation with chloroacetyl-CoA or with bromopyruvate. Inactivation by both reagents follows saturation kinetics, indicating the formation of an E ... I complex (dissociation constants of 0.36 microM for chloroacetyl-CoA and 31 microM for bromopyruvate) prior to alkylation. The limiting rate constants are 0.15 s-1 for bromopyruvate and 0.041 s-1 for chloroacetyl-CoA. Inactivation by both reagents is protected by NADPH and 200 mM KCl, and by saturating amounts of thioester substrates which reduced the limiting rate constants 6.5-30-fold. Active-site-directed reaction of chloroacetyl-CoA is supported by the ability of this compound to form a kinetically viable complex with the enzyme as competitive inhibitor of acetyl-CoA. Chloroacetyl-CoA interacts initially at the CoA binding pocket, since the nucleotide afforded competitive protection of inactivation and caused a large decrease in its affinity. Subsequently, the phosphopantetheine prosthetic group is alkylated. Evidence is presented to show that bromopyruvate competes with chloroacetyl-CoA for the same target site.  相似文献   

5.
Crude extracts of Methanospirillum hungatei strain GP1 contained NADH and NADPH diaphorase activities. After a 483-fold purification of the NADH diaphorase the enzyme was further separated from contaminating proteins by polyacrylamide disc gel electrophoresis. Two distinct activity bands were extracted from the acrylamide, each one having oxygen, 2,6-dichlorophenolindophenol, and cytochrome c linked activities. In these preparations NADPH could not replace NADH as electron donor. During the initial purification steps all activity was lost due to the removal of a readily released cofactor. Enzyme activity was restored by either FAD or a FAD fraction isolated from M. hungatei. Oxidase activity exhibited a broad pH optimum from 7.0 to 8.5 and apparent Km values of 26 microM for NADH and 0.2 microM for FAD. Superoxide anion, formed in the presence of oxygen, accounted for all of the NADH consumed in the reaction. The molecular weight of the diaphorase was about 117 500 by sodium dodecyl sulfate gel electrophoresis. Sulfhydryl reagents and chelating agents were inhibitory. Inactivation, which occurred during storage in phosphate buffer at 4 degrees C, was delayed by dithiothreitol. The isolated NADH diaphorase lacked NADPH:NAD transhydrogenase and NAD reductase activities.  相似文献   

6.
It was found that when Escherichia coli is grown in the presence of 0.2-0.3 mM menadione (2-methyl-1,4-naphthoquinone), an FMN-dependent NADH-quinone reductase increases more than 20-fold in the cytoplasmic fraction. The menadione-induced quinone reductase was isolated from the cytoplasmic fraction of induced cells. The purified enzyme had an Mr of 24 kDa on SDS-polyacrylamide gel electrophoresis. The enzyme required flavin as a cofactor and a half-maximum activity was obtained with 0.54 microM FMN or 16.5 microM FAD. The enzyme had a broad pH optimum at pH 7.0-8.0 and reacted with NADH, but not with NADPH. The reaction followed a ping-pong mechanism and the intrinsic Km values for NADH and menadione were estimated to be 132 microM and 2.0 microM, respectively. Dicoumarol was a simple competitive inhibitor with respect to NADH with a Ki value of 0.22 microM. The electron acceptor specificity of this enzyme was very similar to that of NAD(P)H: (quinone acceptor) oxidoreductase (EC 1.6.99.2, DT-diaphorase) from rat liver. Since menadione is reduced by the two-electron reduction pathway to menadiol, the induction of this enzyme is likely to be an adaptive response of E. coli to partially alleviate the toxicity of menadione.  相似文献   

7.
Evidence is presented which suggests that the NAD(P)H-cytochrome c reductase component of nitrate reductase is the main site of action of the inactivating enzyme. When tested on the nitrate reductase (NADH) from the maize root and scutella, the NADH-cytochrome c reductase was inactivated at a greater rate than was the FADH2-nitrate reductase component. With the Neurospora nitrate reductase (NADPH) only the NADPH-cytochrome c reductase was inactivated. p-Chloromercuribenzoate at 50 muM, which gave almost complete inhibition of the NADH-cytochrome c reductase fraction of the maize nitrate reductase, had no marked effect on the action of the inactivating enzyme. A reversible inactivation of the maize nitrate reductase has been shown to occur during incubation with NAD(P)H. In contrast to the action of the inactivating enzyme, it is the FADH2-nitrate reductase alone which is inactivated. No inactivation of the Neurospora nitrate reductase was produced by NAD(P)H alone and also in the presence of FAD. The lack of effect of the inactivating enzyme and NAD(P)H on the FADH2-nitrate reductase of Neurospora suggests some differences in its structure or conformation from that of the maize enzyme. A low level of cyanide (0.4 mu M) markedly enhanced the action of NAD(P)H on the maize enzyme; Cyanide at a higher level (6 mu M) did give inactivation of the Neurospora nitrate reductase in the presence of NADPH and FAD. The maize nitrate reductase, when partially inactivated by NADH and cyanide, was not altered as a substrate for the inactivating enzyme. The maize root inactivating enzyme was also shown to inactivate the nitrate reductase (NADH) in the pea leaf. It had no effect on the nitrate reductase from either Pseudomonas denitrificans or Nitrobacter agilis.  相似文献   

8.
The reaction process of adrenodoxin reductase with NADPH and NADH were investigated. The appearance of new intermediate with a broad absorption band at around 520 nm has been detected by rapid-scan stopped-flow spectrophotometry. Although the formation of this intermediate is more rapid with NADPH than with NADH, the rates of the subsequent decay to the fully reduced state are almost identical (Kobs values were 20.5 and 16.0s-1). These results indicate that the new intermediate is the complex formed between the oxidized enzyme and reduced pyridine nucleotide (enzyme-substrate complex), and that subsequent decay of the intermidiate is caused by a two-electron transfer process from the reduced pyridine nucleotide to the enzyme flavin. On the other hand, spectral and kinetic properties in the steady state of the reoxidation reaction of the enzyme reduced with NADPH and NADH were somewhat different. The rate of reoxidation of the enzyme under aerobic conditions from the reduced state to the oxidized state was 6.5 times faster when a 10-fold molar excess of NADH was used than when NADPH of the same concentration was used. This result is consistent with the fact that the NADH-dependent oxidase activity was 6.4 times greater than that dependent on NADPH. During reoxidation of the reduced enzyme under aerobic conditions in the presence of an excess of NADPH or NADH, the EPR spectra indicated the formation of the flavin semiquinone radical species. Similarly, the formation of semiquinone was observed in the absorption spectrum with either NADPH or NADH under the same conditions as in the EPR measurement. The intensity of the semiquinone signal on EPR was considerably smaller with NADH than with NADPH. These results suggest that NADP+ complex with the enzyme semiquinone protects the radical from oxidation by oxygen to a greater extent than NAD+, and consequently the semiquinone is easier to detect with NADPH than with NADH.  相似文献   

9.
Inactivation of apo-glyceraldehyde-phosphate dehydrogenase (D-glyceraldehyde-3-phosphate: NAD+ oxidoreductase(phosphorylating) (EC 1.2.1.12) from rat skeletal muscle at 4 degrees C in 0.15 M NaC1, 5 mM EDTA, 4 mM 2-mercaptoethanol pH 7.2 is a first-order reaction. The rate constant of inactivation depends on protein concentration. With one molecule of NAD bound per tetrameric enzyme, a 50 per cent loss in activity is observed and the rate constant of inactivation becomes independent of the protein concentration over a 30-fold range. Two moles of NAD bound per mole of enzyme fully protect it against inactivation. NADH affords a cooperative effect on enzyme structure similar to that of NAD. Inactivation of 7.8 S apoenzyme is reflected in its dissociation into 4.8-S dimers. In the case of enzyme-NAD1 complex, no direct relationship between the extent of inactivation and dissociation is observed, suggesting that these two processes do not occur simultaneously; we may say that dissociation is slower than inactivation. A mechanism in which the rate-limiting step for inactivation is a conformational change in the tetramer occurring prior to dissociation and affecting only the structure of the non-liganded dimer, is consistent with the experimental observations. Inorganic phosphate protects apoenzyme against inactivation. Its effect is shown to be due to the anion binding at specific sites on the protein with a dissociation constant of 2.6 plus or minus 0.4 mM. The NaC1-induced cold inactivation of glyceraldehyde-phosphate dehydrogenase is fully reversible at 25 degrees C in the presence of 20 mM dithiothreitol and 50 mM inorganic phosphate. The rate of reactivation is independent of protein concentration. Inactivated enzyme retains the ability to bind specific antibodies produced in rabbits, but diminishes its precipitating capability.  相似文献   

10.
A novel aerobic mechanism of 2-aminobenzoate metabolism was proposed in a denitrifying Pseudomonas species. 2-Aminobenzoic acid is activated in a coenzyme-A-ligase reaction to 2-aminobenzoyl-CoA and this intermediate is dearomatized by a unique enzyme, tentatively named 2-aminobenzoyl-CoA monooxygenase/reductase. This paper describes the purification and some molecular, kinetic and spectral properties of this flavoenzyme which catalyzes the hydroxylation and reduction of 2-aminobenzoyl-CoA to an unknown non-aromatic compound. 2-Aminobenzoyl-CoA monooxygenase/reductase was purified 25-fold to a specific activity of 25 mumol.min-1.mg-1 protein using ammonium sulfate precipitation, DEAE-cellulose anion-exchange, hydroxylapatite and Mono Q FPLC anion-exchange chromatography. Superose 6 gel filtration for estimation of molecular mass resulted in one symmetrical protein peak corresponding to a molecular mass of 170 kDa. Several experimental data suggest that the protein is probably an alpha 2 dimer; however, it may exist in three dimeric forms, alpha alpha, alpha alpha' and alpha' alpha', where alpha' may be a subunit with a different conformation. Approximately 2 mol noncovalently bound FAD/mol enzyme was found, which in the absence of O2 was reduced by NADH. The enzyme was specific for the substrates 2-aminobenzoyl-CoA (Km less than or equal to 25 microM) and O2 (Km less than or equal to 5 microM), but less specific for the reduced pyridine nucleotides NADH (Km = 42 microM) or NADPH [Km = 500 microM; Vmax (NADH)/Vmax (NADPH) = 1.7:1]. The turnover number was 4250 min-1. The enzyme also reduced N-ethylmaleimide and maleimide with NAD(P)H. The substrate, the products and the reaction stoichiometry are described in two following papers.  相似文献   

11.
Monodehydroascorbate reductase (EC 1.6.5.4) was purified from cucumber fruit to a homogeneous state as judged by polyacrylamide gel electrophoresis. The cucumber monodehydroascorbate reductase was a monomer with a molecular weight of 47,000. It contained 1 mol of FAD/mol of enzyme which was reduced by NAD(P)H and reoxidized by monodehydroascorbate. The enzyme had an exposed thiol group whose blockage with thiol reagents inhibited the electron transfer from NAD(P)H to the enzyme FAD. Both NADH and NADPH served as electron donors with Km values of 4.6 and 23 microM, respectively, and Vmax of 200 mol of NADH and 150 mol of NADPH oxidized mol of enzyme-1 s-1. The Km for monodehydroascorbate was 1.4 microM. The amino acid composition of the enzyme is presented. In addition to monodehydroascorbate, the enzyme catalyzed the reduction of ferricyanide and 2,6-dichloroindophenol but showed little reactivity with calf liver cytochrome b5 and horse heart cytochrome c. The kinetic data suggested a ping-pong mechanism for the monodehydroascorbate reductase-catalyzed reaction. Cucumber monodehydroascorbate reductase occurs in soluble form and can be distinguished from NADPH dehydrogenase, NADH dehydrogenase, DT diaphorase, microsome-bound NADH-cytochrome b5 reductase, and NADPH-cytochrome c reductase by its molecular weight, amino acid composition, and specificity of electron acceptors and donors.  相似文献   

12.
R J Auchus  D F Covey 《Biochemistry》1986,25(23):7295-7300
14,15-Secoestra-1,3,5(10)-trien-15-yne-3,17 beta-diol (1) is a mechanism-based inactivator of human placental 17 beta,20 alpha-hydroxysteroid dehydrogenase (estradiol dehydrogenase, EC 1.1.1.62). Inactivation with alcohol 1 requires NAD-dependent enzymic oxidation and follows approximately pseudo-first-order kinetics with a limiting t1/2 of 82 min and a "Ki" of 2.0 microM at pH 9.2 and 25 degrees C. At saturating concentrations of NAD, the initial rate of inactivation is slower than in the presence of 5 microM NAD, suggesting that cofactor binding to free enzyme impedes the inactivation process. Glutathione completely protects the enzyme from inactivation at both cofactor concentrations. Inactivation with 45 microM tritiated alcohol 1 followed by dialysis and gel filtration demonstrates a covalent interaction and affords an estimated stoichiometry of 1.4 molecules of steroid per subunit (2.8 per dimer). Chemically prepared 3-hydroxy-14,15-secoestra-1,3,5(10)-trien-15-yn-17-one (2) rapidly inactivates estradiol dehydrogenase with biphasic kinetics. From the latter phase, a Ki of 2.8 microM and a limiting t1/2 of 12 min at pH 9.2 were determined. Estradiol, NADH, and NAD all retard this latter inactivation phase. We propose that enzymatically generated ketone 2 inactivates estradiol dehydrogenase after its release from and return to the active site of free enzyme.  相似文献   

13.
NAD(P)H:(quinone-acceptor) oxidoreductase (EC 1.6.99.2) is a widely distributed enzyme which promotes two-electron reductions of quinones and thereby protects cells against damage by reactive oxygen species generated during oxidative cycling of quinones and semiquinone radicals. Quinone reductase activity represents a minor component (about 0.006%) of mouse liver cytosolic proteins under basal (uninduced) conditions. Two isofunctional forms of this quinone reductase have been purified to homogeneity (1700-fold) in 30% yield from the liver cytosols of female CD-1 mice in which the enzymes were induced by administration of 2(3)-tert-butyl-4-hydroxyanisole. The purification involved ion exchange, hydrophobic, and affinity chromatographies. The two enzyme forms have been designated "hydrophilic" and "hydrophobic" based on the order of elution from phenyl-Sepharose. The more abundant hydrophilic form has been crystallized in the presence of FAD in the form of macroscopic tetragonal crystals. The two forms have similar isoelectric points (pI 9.2) and subunit molecular weights (Mr = 30,000) and probably exist as dimers in the native state. Purified preparations of the enzymes are equiactive with NADH and NADPH and show almost complete dependence on added FAD for catalytic activity. The Km values for FAD of the hydrophilic and hydrophobic forms are 2.72 and 1.72 nM, respectively. Their catalytic activities are the same and are remarkably high for nicotinamide nucleotide-linked dehydrogenases; maximum velocities (expressed per mg of pure enzyme) approach 4000 units/mg of protein under appropriate assay conditions. When menadione is the electron acceptor, the Km value for this quinone is very low (Km congruent to 2 microM). Both enzyme forms are potently inhibited by dicoumarol. Rabbit antisera against the hydrophilic quinone reductase precipitate quantitatively the entire quinone reductase activity of mouse liver cytosols obtained from animals maintained on a standard diet or those induced with 3-tert-butyl-4-hydroxyanisole. The quinone reductase activity of rat liver cytosols is also quantitatively precipitated by this antiserum.  相似文献   

14.
NAD(P)H:quinone oxidoreductase (NQO1; EC 1.6.99.2) catalyzes a two-electron transfer involved in the protection of cells from reactive oxygen species. These reactive oxygen species are often generated by the one-electron reduction of quinones or quinone analogs. We report here on the previously unreported Fe(III) reduction activity of human NQO1. Under steady state conditions with Fe(III) citrate, the apparent Michaelis-Menten constant (Km(app)) was approximately 0.3 nM and the apparent maximum velocity (Vmax(app)) was 16 U mg(-1). Substrate inhibition was observed above 5 nM. NADH was the electron donor, Km(app)= 340 microM and Vmax(app) = 46 Umg(-1). FAD was also a cofactor with a Km(app) of 3.1 microM and Vmax(app) of 89 U mg(-1). The turnover number for NADH oxidation was 25 s(-1). Possible physiological roles of the Fe(III) reduction by this enzyme are discussed.  相似文献   

15.
Arscott LD  Veine DM  Williams CH 《Biochemistry》2000,39(16):4711-4721
Glutathione reductase catalyzes the reduction of glutathione disulfide by NADPH. The FAD of the reductase is reduced by NADPH, and reducing equivalents are passed to a redox-active disulfide to complete the first half-reaction. The nascent dithiol of two-electron reduced enzyme (EH(2)) interchanges with glutathione disulfide forming two molecules of glutathione in the second half-reaction. It has long been assumed that a mixed disulfide (MDS) between one of the nascent thiols and glutathione is an intermediate in this reaction. In addition to the nascent dithiol composed of Cys(45) and Cys(50), the enzyme contains an acid catalyst, His(456), having a pK(a) of 9.2 that protonates the first glutathione (residue numbers refer to the yeast enzyme sequence). Reduction of yeast glutathione reductase by glutathione and reoxidation of EH(2) by glutathione disulfide indicate that the mixed disulfide accumulates, in particular, at low pH. The reaction of glutathione disulfide with EH(2) is stoichiometric in the absence of an excess of glutathione. The equilibrium position among E(ox), MDS, and EH(2) is determined by the glutathione concentration and is not markedly influenced by pH between 6.2 and 8.5. The mixed disulfide is the principal product in the reaction of glutathione with oxidized enzyme (E(ox)) at pH 6. 2. Its spectrum can be distinguished from that of EH(2) by a slightly lower thiolate (Cys(50))-FAD charge-transfer absorbance at 540 nm. The high GSH/GSSG ratio in the cytoplasm dictates that the mixed disulfide will be the major enzyme species.  相似文献   

16.
3 beta-Hydroxy-delta 5-steroid dehydrogenase and steroid delta-isomerase copurify from human placental microsomes as a single enzyme protein. The affinity-alkylating secosteroid, 5,10-secoestr-4-yne-3,10,17-trione, inactivates the dehydrogenase and isomerase reactions in a time-dependent manner, but which of the two activities is targeted depends on the concentration of secosteroid. At 2-5 microM secosteroid, the dehydrogenase activity is alkylated in a site-specific manner (pregnenolone slows inactivation) that follows first-order inactivation kinetics (KI = 4.2 microM, k3 = 1.31 x 10(-2) min-1). As the secosteroid level increases from 11 to 30 microM, dehydrogenase is paradoxically inactivated at progressively slower rates, and pregnenolone no longer protects against the alkylator. The inactivation of isomerase exhibits the expected first-order kinetics (KI = 31.3 microM, k3 = 6.42 x 10(-2) min-1) at 11-30 microM secosteroid. 5-Androstene-3,17-dione protects isomerase from inactivation by 15 microM secosteroid, but the substrate steroid unexpectedly fails to slow the inactivation of isomerase by a lower concentration of alkylator (5 microM). A shift from a dehydrogenase to an isomerase conformation in response to rising secosteroid levels explains these results. Analysis of the ligand-induced conformational change along with cofactor protection data suggests that the enzyme expresses both activities at a bifunctional catalytic site. According to this model, the protein begins the reaction sequence as 3 beta-hydroxysteroid dehydrogenase. The products of the first step (principally NADH) promote a change in protein conformation that triggers the isomerase reaction.  相似文献   

17.
Summary The glutathione reductase from E. coli was rapidly inactivated following aerobic incubation of the pure and cell-free extract enzymes with NADPH, NADH and other reductants. The inactivation of the pure enzyme depended on the time and temperature of incubation (t1/2 = 2 min at 37°C), and was proportional to the |INADPH|/|enzyme| ratio, reaching 50% in the presence of 0.3 M NADPH and 45 M NADH respectively, at a subunit concentration of 20 nM. Higher pyridine nucleotide concentrations were required to inactivate the enzyme from cell-free extracts. Two apparent pKa, corresponding to pH 5.8 and 7.3, were determined for the redox inactivation. The enzyme remained inactive even after eliminating the excess NADPH by gel chromatography. E. coli glutathione reductase was protected by oxidized and reduced glutathione against redox inactivation with both pure and cell-free extract enzymes. Ferricyanide and dithiothreitol protected only the pure enzyme, while NADP+ exclusively protected the cell-free extract enzyme. The inactive glutathione reductase was reactivated by treatment with oxidized and reduced glutathione, ferricyanide, and dithiothreitol in a time-and temperature-dependent process. The oxidized form of glutathione was more efficient and specific than the reduced form in the protection and reactivation of the pure enzyme.The molecular weight of the redox-inactivated E. coli glutathione reductase was similar to that of the dimeric native enzyme, ruling out aggregation as a possible cause of inactivation. A tentative model is discussed for the redox inactivation, involving the formation of an erroneous disulfide bridge at the glutathione-binding site.  相似文献   

18.
Mercuric reductase, with FAD and a reducible disulfide at the active site, catalyzes the two-electron reduction of Hg(II) by NADPH. Addition of reducing equivalents rapidly produces a spectrally distinct EH2 form of the enzyme containing oxidized FAD and reduced active site thiols. Formation of EH2 has previously been reported to require only 2 electrons for reduction of the active site disulfide. We present results of anaerobic titrations of mercuric reductase with NADPH and dithionite showing that the equilibrium conversion of oxidized enzyme to EH2 actually requires 2 equiv of reducing agent or 4 electrons. Kinetic studies conducted both at 4 degrees C and at 25 degrees C indicate that reduction of the active site occurs rapidly, as previously reported [Sahlman, L., & Lindskog, S. (1983) Biochem. Biophys. Res. Commun. 117, 231-237]; this is followed by a slower reduction of another redox group via reaction with the active site. Thiol titrations of denatured Eox and EH2 enzyme forms show that an additional disulfide is the group in communication with the active site. [14C]Iodoacetamide labeling experiments demonstrate that the C-terminal residues, Cys558 and Cys559, are involved in this disulfide. The fluorescence, but not the absorbance, of the enzyme-bound FAD was found to be highly dependent on the redox state of the C-terminal thiols. Thus, Eox with Cys558 and Cys559 as thiols exhibits less than 50% of the fluorescence of Eox where these residues are present as a disulfide, indicating that the thiols remain intimately associated with the active site.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

19.
The reduction of yeast glutathione reductase by reduced nicotinamide hypoxanthine dinucleotide phosphate (NHxDPH) has been examined by stopped-flow kinetic methods. Like reduced glutathione or NADPH, this pyridine nucleotide generates the catalytically active two-electron reduced form of the enzyme. This reductive half-reaction with NHxDPH has only one detectable kinetic step which shows saturation kinetics (Kd = 76 microM), and has a limiting rate constant of 56 s-1. Comparison of stopped-flow and steady-state turnover data indicates that the reductive half-reaction is rate-limiting in the overall catalytic reaction. No evidence was found for a preequilibrium charge-transfer complex between NHxDPH and the active site FAD, like that seen when NADPH is the electron donor.  相似文献   

20.
Luba J  Charrier V  Claiborne A 《Biochemistry》1999,38(9):2725-2737
An unusual flavoprotein disulfide reductase, which catalyzes the NADPH-dependent reduction of CoASSCoA, has recently been purified from the human pathogen Staphylococcus aureus [delCardayré, S. B., Stock, K. P., Newton, G. L., Fahey, R. C., and Davies, J. E. (1998) J. Biol. Chem. 273, 5744-5751]. Coenzyme A-disulfide reductase (CoADR) lacks the redox-active protein disulfide characteristic of the disulfide reductases; instead, NADPH reduction yields 1 protein-SH and 1 CoASH. Furthermore, the CoADR sequence reveals the presence of a single putative active-site Cys (Cys43) within an SFXXC motif also seen in the Enterococcus faecalis NADH oxidase and NADH peroxidase, which use a single redox-active cysteine-sulfenic acid in catalysis. In this report, we provide a detailed examination of the equilibrium properties of both wild-type and C43S CoADRs, focusing on the role of Cys43 in the catalytic redox cycle, the behavior of both enzyme forms on reduction with dithionite and NADPH, and the interaction of NADP+ with the corresponding reduced enzyme species. The results of these analyses, combined with electrospray mass spectrometric data for the two oxidized enzyme forms, fully support the catalytic redox role proposed for Cys43 and confirm that this is the attachment site for bound CoASH. In addition, we provide evidence indicating dramatic thermodynamic inequivalence between the two active sites per dimer, similar to that documented for the related enzymes mercuric reductase and NADH oxidase; only 1 FAD is reduced with NADPH in wild-type CoADR. The EH2.NADPH/EH4.NADP+ complex which results is reoxidized quantitatively in titrations with CoASSCoA, supporting a possible role for the asymmetric reduced dimer in catalysis.  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