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.
Glutathione reductase from S. cerevisiae (EC 1.6.4.2) catalyzes the NADPH oxidation by glutathione in accordance with a "ping-pong" scheme. The catalytic constant kcat) is 240 s-1 (pH 7.0, 25 degrees C); kcat for the diaphorase reaction is 4-5 s-1. The enzyme activity does not change markedly at pH 5.5-8.0. At pH less than or equal to 7.0, NADP+ acts as a competitive inhibitor towards NADPH and as a noncompetitive inhibitor towards glutathione. NADP+ increases the diaphorase activity of the enzyme. The maximal activity is observed, when the NADP+/NADPH ratio exceeds 100. At pH 8.0, NADP+ acts as a mixed type inhibitor during the reduction of glutathione. High concentrations of NADP+ also inhibit the diaphorase activity due to the reoxidation of the reduced enzyme by NADP+ at pH 8.0. The redox potential of glutathione reductase calculated from the inhibition data is--306 mV (pH 8.0). Glutathione reductase reduces quinoidal compounds in an one-electron way. The hyperbolic dependence of the logarithm of the oxidation constant on the one electron reduction potential of quinone is observed. It is assumed that quinones oxidize the equilibtium fraction of the two-electron reduced enzyme containing reduced FAD.  相似文献   

2.
The reduced glutathione-linked NADP+ reduction, catalyzed by yeast glutathione reductase, follows a 'sequential' or 'ping-pong' mechanism at high or low NADP+ concentrations, respectively. The pattern of the NADPH and NADP+ cross-inhibition reflects not only the competition for the binding site, but the shift of the reaction equilibrium as well. A 'branched' scheme of the glutathione reductase reaction is presented. The enzyme standard potential (-255 mV, pH 7.0) was estimated from the ratio of the NADPH and NADP+ rate constants corresponding to the ping-pong mechanism.  相似文献   

3.
One- and two-electron reduction of quinones by glutathione reductase   总被引:1,自引:0,他引:1  
Yeast glutathione reductase (E.C. 1.6.4.2) catalyzes the oxidation of NADPH by p-quinones and ferricyanide with a maximal turnover number (TNmax) of 4-5 s-1.NADP+ stimulates the reaction and the TNmax/Km value of acceptors is reached at NADP+/NADPH greater than or equal to 100. TNmax is increased up to 30-33 s-1. The stimulatory effect of NADP+ may be associated with its complexation with the NADPH-binding site in the reduced enzyme (Kd = 40-60 microM). It is suggested that NADP+ shifts the electron density towards FAD in the two-electron-reduced enzyme and, evidently, changes its one-electron-reduction potentials, while quinones oxidize an equilibrium form of glutathione reductase containing reduced FAD. In the absence of NADP+ the reduction of quinones by glutathione reductase proceeds mainly in a two-electron manner. At NADP+/NADPH = 100 a one-electron reduction makes up 44% of the total process. At pH 6.0-7.0 the reduced forms of naphthoquinones undergo cyclic redox conversions. A hyperbolic dependence exists of the log TN/Km of quinones on their one-electron-reduction potentials.  相似文献   

4.
Hyperglycemia is associated with metabolic disturbances affecting cell redox potential, particularly the NADPH/NADP+ ratio and reduced glutathione levels. Under oxidative stress, the NADPH supply for reduced glutathione regeneration is dependent on glucose-6-phosphate dehydrogenase. We assessed the effect of different hyperglycemic conditions on enzymatic activities involved in glutathione regeneration (glucose-6-phosphate dehydrogenase and glutathione reductase), NADP(H) and reduced glutathione concentrations in order to analyze the relative role of these enzymes in the control of glutathione restoration. Male Sprague-Dawley rats with mild, moderate and severe hyperglycemia were obtained using different regimens of streptozotocin and nicotinamide. Fifteen days after treatment, rats were killed and enzymatic activities, NADP(H) and reduced glutathione were measured in liver and pancreas. Severe hyperglycemia was associated with decreased body weight, plasma insulin, glucose-6-phosphate dehydrogenase activity, NADPH/NADP+ ratio and glutathione levels in the liver and pancreas, and enhanced NADP+ and glutathione reductase activity in the liver. Moderate hyperglycemia caused similar changes, although body weight and liver NADP+ concentration were not affected and pancreatic glutathione reductase activity decreased. Mild hyperglycemia was associated with a reduction in pancreatic glucose-6-phosphate dehydrogenase activity. Glucose-6-phosphate dehydrogenase, NADPH/NADP+ ratio and glutathione level, vary inversely in relation to blood glucose concentrations, whereas liver glutathione reductase was enhanced during severe hyperglycemia. We conclude that glucose-6-phosphate dehydrogenase and NADPH/NADP+ were highly sensitive to low levels of hyperglycemia. NADPH/NADP+ is regulated by glucose-6-phosphate dehydrogenase in the liver and pancreas, whereas levels of reduced glutathione are mainly dependent on the NADPH supply.  相似文献   

5.
Nitrofurans inhibit the oxidation of NADPH by glutathione, catalyzed by yeast glutathione reductase (EC 1.6.4.2). acting as uncompetitive incomplete inhibitors for NADPH and glutathione. The quinoline-substituted nitrofurans were the most effective inhibitors. These compounds increased the turnover numbers of enzyme at fixed concentrations of reduced glutathione, in the reverse reaction of glutathione reductase, but in most cases diminished the affinity of the enzyme for NAD+. Nitrofurans are weak one-electron oxidants of glutathione reductase. Their reactivity is close to that of p-quinones possessing the analoguous one-electron reduction potential (Cénas, N.K., Rakauskiené, G.A. and Kulys, J.J. (1989) Biochim. Biophys. Acta 973, 399-404), and reaction is stimulated by NADP+. It is assumed, that nitrofurans bind to the 'regulative' site of glutathione reductase (Karplus, P.A., Pai, E.F. and Schulz, G.E. (1989) Eur. J. Biochem. 178, 693-703).  相似文献   

6.
Previous studies have shown that the interaction of P450 reductase with bound NADP(H) is essential to ensure fast electron transfer through the two flavin cofactors. In this study we investigated in detail the interaction of the house fly flavoprotein with NADP(H) and a number of nucleotide analogues. 1,4,5,6-Tetrahydro-NADP, an analogue of NADPH, was used to characterize the interaction of P450 reductase with the reduced nucleotide. This analogue is inactive as electron donor, but its binding affinity and rate constant of release are very close to those for NADPH. The 2'-phosphate contributes about 5 kcal/mol of the binding energy of NADP(H). Oxidized nicotinamide does not interact with the oxidized flavoprotein, while reduced nicotinamide contributes 1.3 kcal/mol of the binding energy. Oxidized P450 reductase binds NADPH with a K(d) of 0.3 microM, while the affinity of the reduced enzyme is considerably lower, K(d) = 1.9 microM. P450 reductase catalyzes a transhydrogenase reaction between NADPH and oxidized nucleotides, such as thionicotinamide-NADP(+), acetylpyridine-NADP(+), or [(3)H]NADP(+). The reverse reaction, reduction of [(3)H]NADP(+) by the reduced analogues, is also catalyzed by P450 reductase. We define the mechanism of the transhydrogenase reaction as follows: NADPH binding, hydride ion transfer, and release of the NADP(+) formed. An NADP(+) or its analogue binds to the two-electron-reduced flavoprotein, and the electron-transfer steps reverse to transfer hydride ion to the oxidized nucleotide, which is released. Measurements of the flavin semiquinone content, rate constant for NADPH release, and transhydrogenase turnover rates allowed us to estimate the steady-state distribution of P450 reductase species during catalysis, and to calculate equilibrium constants for the interconversion of catalytic intermediates. Our results demonstrate that equilibrium redox potentials of the flavin cofactors are not the sole factor governing rapid electron transfer during catalysis, but conformational changes must be considered to understand P450 reductase catalysis.  相似文献   

7.
Biochemical and clinical evidence indicates that monomethylated selenium compounds are crucial for the tumor preventive effects of the trace element selenium and that methylselenol (CH(3)SeH) is a key metabolite. As suggested by Ganther (Ganther, H. E. (1999) Carcinogenesis 20, 1657-1666), methylselenol and its precursor methylseleninate might exert their effects by inhibition of the selenoenzyme thioredoxin reductase via the irreversible formation of a diselenide bridge. Here we report that methylseleninate does not act as an inhibitor of mammalian thioredoxin reductase but is in fact an excellent substrate (K(m) of 18 microm, k(cat) of 23 s(-1)), which is reduced by the enzyme according to the equation 2 NADPH + 2 H(+) + CH(3)SeO(2)H --> 2 NADP(+) + 2 H(2)O + CH(3)SeH. The selenium-containing product of this reaction was identified by mass spectrometry. Nascent methylselenol was found to efficiently reduce both H(2)O(2) and glutathione disulfide. The implications of these findings for the antitumor activity of selenium are discussed. Methylseleninate was a poor substrate not only for human glutathione reductase but also for the non-selenium thioredoxin reductases enzymes from Drosophila melanogaster and Plasmodium falciparum. This suggests that the catalytic selenocysteine residue of mammalian thioredoxin reductase is essential for methylseleninate reduction.  相似文献   

8.
Rat liver 3-hydroxy-3-methylglutaryl-CoA (HMG-CoA) reductase catalyzes, in addition to its normal biosynthetic or forward reaction (HMG-CoA + 2 NADPH + 2H+----mevalonate + 2 NAD+ + CoASH), the reverse reaction (mevalonate + CoASH + 2 NADP+----HMG-CoA + 2 NADPH + 2H+) and two "half-reactions" that involve the presumed intermediate mevaldate (mevaldate + CoASH + NADP+----HMG-CoA + NADPH + H+ and mevaldate + NADPH + H+----mevalonate + NADP+). These reactions were studied using both enzyme solubilized by the traditional freeze-thaw method and enzyme solubilized with a nonionic detergent in the presence of inhibitors of proteolysis. All four reactions were inhibited by mevinolin, a known inhibitor of the forward (biosynthetic) reaction catalyzed by HMG-CoA reductase. When the enzyme was inactivated by ATP and a cytosolic, ADP-dependent HMG-CoA reductase kinase, the rates of both the forward reaction and the half-reactions decreased to comparable extents. Although coenzyme A is not a stoichiometric participant in the second half-reaction (mevaldate + NADPH + H+----mevalonate + NADP+), it was required as an activator of this reaction. This observation implies that coenzyme A may remain bound to the enzyme throughout the normal catalytic cycle of HMG-CoA reductase.  相似文献   

9.
The binding site of NADPH in NADPH-adrenodoxin reductase was examined using crystalline enzyme from bovine adrenocortical mitochondria by studies on the effects of photooxidation and chemical modifications of amino acid residues in the reductase. (1) Photoxication decreased the enzymatic activity of NADPH-adrenodoxin reductase. Photooxidation of the reductase was prevented by NADP+, adrenodoxin, or reduced glutathione, but not NAD+. Photoinactivation caused loss of a histidyl residue, but not of tyrosyl, tryptophanyl, cysteinyl, or methionyl residues of the reductase. It did not affect the circular dichroism spectrum of the reductase appreciably. (2) NADPH-adrenodoxin reductase activity was inhibited by diethyl pyrocarbonate and the inhibition was partially reversed by addition of hydroxylamine. The inhibition was prevented by NADP+, but not NAD+. (3) NADPH-adrenodoxin reductase activity was inhibited by 5,5'-dithiobis(2-nitrobenzoate) and the inhibition was reversed by reduced glutathione. It was also protected by NADP+, but not NAD+. The results indicate that a histidyl residue and a cysteinyl residue of NADPH-adrenodoxin reductase are essential for the binding of NADPH by the reductase.  相似文献   

10.
Glutathione reductase has been found to catalyze an NAD(P)H-dependent electron transfer to 2,4,6-trinitrobenzenesulfonate (TNBS). In the presence of oxygen TNBS is not consumed in the reaction, but is rapidly reoxidized with concomitant production of hydrogen peroxide. Cytochrome c can replace oxygen as the final electron acceptor, indicating that a one-electron transfer takes place. The rate is slightly higher in the absence than in the presence of oxygen, ruling out superoxide anion as an obligatory intermediate in cytochrome c reduction. In the absence of oxygen (or cytochrome c), TNBS limits the reaction and accepts a total of four electrons. The TNBS-dependent NADPH (or NADH) oxidation is markedly stimulated by NADP+, and to a smaller extent also by NAD+. The TNBS-dependent reactions are inhibited by excess of NADPH but not by NADH. The kinetics of these reactions are consistent with a branching reaction mechanism in which a pathway including a ternary complex between the two-electron reduced enzyme and NADP+ has the highest turnover. NADPH-dependent reductions of ferricyanide or 2,6-dichloroindophenol catalyzed by glutathione reductase are also markedly influenced by NADP+. Evidently NADP+ facilitates a shift of the catalyzed reaction from the normal two-electron reduction of glutathione disulfide to a more unspecific one-electron reduction of other acceptors. Spectral as well as kinetic data suggest that the rate of radical formation limits the reactions with the artificial electron acceptors and that NADP+ promotes this rate-limiting step.  相似文献   

11.
Glutathione reductase has been purified to homogeneity by a method which is an improvement of an earlier procedure (Carlberg, I. and Mannervik, B. (1975) J. Biol. Chem. 250, 5475-5480). The new steps in the purification scheme include affinity chromatography on 2',5' ADP-Sepharose 4B. Antibodies to glutathione reductase from rat liver were raised in rabbits and used for analysis of the enzyme by quantitative 'rocket' immunoelectrophoresis. Glutathione reductase from human erythrocytes, porcine erythrocytes, and calf-liver gave precipitin lines showing partial identity with the rat liver enzyme in Ouchterlony double diffusion experiments. Enzyme from spinach, yeast (Saccharomyces cerevisiae), and the photosynthetic bacterium Rhodospirillum rubrum did not give precipitates with the antibodies to the enzyme from rat liver. Titration of glutathione reductase from the different sources with antibodies confirmed the cross-reactivity of the mammalian enzymes; the human enzyme giving the strongest heterologous reaction. No reaction was observed with the enzyme from spinach, yeast, and Rhodospirillum rubrum. NADPH, NADP+, and 2',5' ADP were found to inhibit the interaction between antibodies and glutathione reductase from rat liver and human erythrocytes. NADH, glutathione, or glutathione disulfide did not protect the enzyme from reacting with the antibodies. It is concluded that glutathione reductase has an antigenic binding site for the antibodies at the pyridine nucleotide-binding site of the enzyme molecule.  相似文献   

12.
Dismutation of dihydrofolate by dihydrofolate reductase   总被引:1,自引:0,他引:1  
R L Blakley  L Cocco 《Biochemistry》1984,23(11):2377-2383
Degradation of 7,8-dihydrofolate (H2folate) in the presence of dihydrofolate reductase (DHFR) has been shown due not to an oxygenase activity of the reductase as previously reported but to dismutation of H2folate to folate and 5,6,7,8-tetrahydrofolate (H4folate). The reaction can be followed spectrophotometrically or by analysis of the reaction mixture by high-performance liquid chromatography (HPLC). The products have also been isolated and characterized. Oxygen uptake during the reaction is much less than stoichiometric with H2folate disappearance and is attributed to autoxidation of the H4folate formed. The dismutation activity is a property of highly purified Streptococcus faecium DHFR isoenzyme 2 (but not isoenzyme 1) and of Lactobacillus casei DHFR, but not of bovine liver DHFR. The activity is dependent on tightly bound NADP+ and/or NADPH. Removal of the nucleotide results in loss of dismutation activity, which is restored by adding NADP+ or NADPH. Maximum activity is obtained when approximately 1 mol equiv of nucleotide is added per mol of DHFR. It is proposed that in the dismutation reaction bound NADP(H) is alternately reduced and oxidized by incoming molecules of H2folate with release of folate and H4folate, respectively. The relatively slow rate of folate formation presumably limits the rate of the overall reaction. The equilibrium constant for the dismutation reaction is 19.4 +/- 7.4 at 22 degrees C and pH 7.0. Calculation of standard oxidation-reduction potentials at pH 7 gave values of -0.230 V for the H2folate/H4 folate pair and -0.268 V for the folate/H2folate pair. The mechanism by which NADP+ is retained by the enzyme from some sources during purification procedures is unclear.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

13.
Anaerobic reduction of the flavoprotein adrenodoxin reductase with NADPH yields a spectrum with long wavelength absorbance, 750 nm and higher. No EPR signal is observed. This spectrum is produced by titration of oxidized adrenodoxin reductase with NADPH, or of dithionite-reduced adrenodoxin reductase with NADP+. Both titrations yield a sharp endpoint at 1 NADP(H) added per flavin. Reduction with other reductants, including dithionite, excess NADH, and catalytic NADP+ with an NADPH generating system, yields a typical fully reduced flavin spectrum, without long wavelength absorbance. The species formed on NADPH reduction appears to be a two-electron-containing complex, with a low dissociation constant, between reduced adrenodoxin reductase and NADP+, designated ARH2-NADP+. Titration of dithionite-reduced adrenodoxin reductase with NADPH also produces a distinctive spectrum, with a sharp endpoint at 1 NADPH added per reduced flavin, indicating formation of a four-electron-containing complex between reduced adrenodoxin reductase and NADPH. Titration of adrenodoxin reductase with NADH, instead of NADPH, provides a curved titration plot rather than the sharp break seen with NADPH, and permits calculation of a potential for the AR/ARH2 couple of -0.291 V, close to that of NAD(P)H (-0.316 V). Oxidized adrenodoxin reductase binds NADP+ much more weakly (Kdiss=1.4 X 10(-5) M) than does reduced adrenodoxin reductase, with a single binding site. The preferential binding of NADP+ to reduced enzyme permits prediction of a more positive oxidation-reduction potential of the flavoprotein in the presence of NADP+; a change of about + 0.1 V has been demonstrated by titration with safranine T. From this alteration in potential, a Kdiss of 1.0 X 10(-8) M for binding of NADP+ to reduced adrenodoxin reductase is calculated. It is concluded that the strong binding of NADP+ to reduced adrenodoxin reductase provides the thermodynamic driving force for formation of a fully reduced flavoprotein form under conditions wherein incomplete reduction would otherwise be expected. Stopped flow studies demonstrate that reduction of adrenodoxin reductase by equimolar NADPH to form the ARH2-NADP+ complex is first order (k=28 s-1). When a large excess of NADPH is used, a second apparently first order process is observed (k=4.25 s-1), which is interpreted as replacement of NADPH for NADP+ in the ARH2-NADP+ complex. Comparison of these rate constants to catalytic flavin turnover numbers for reduction of various oxidants by NADPH, suggests an ordered sequential mechanism in which reduction of oxidant is accomplished by the ARH2-NADP+ complex, followed by dissociation of NADP+. The absolute dependence of NADPH-cytochrome c reduction on both adrenodoxin reductase and adrenodoxin is confirmed...  相似文献   

14.
Yeast glutathione reductase exists in a single molecular form which exhibits preferred NADPH and weak NADH linked multifunctional activities. Kinetic parameters for the NADPH and NADH linked reductase, transhydrogenase, electron transferase and diaphorase reactions have been determined. The functional preference for the NADPH linked reductase reaction is kinetically related to the high catalytic efficiency and low dissociation constants for substrates. NADP+ and NAD+ may interact with two different sites or different kinetic forms of the enzyme. The active site disulfide and histidine are required for the reductase activity but are not essential to the transhydrogenase, electron transferase and diaphorase activities. Amidation of carboxyl groups and Co(II) chelation of glutathione reductase facilitate the electron transferase reaction presumably by encouraging the formation of an anionic flavosemiquinone.  相似文献   

15.
The methylenetetrahydrofolate reductase from the carbon-monoxide-utilizing homoacetogen Peptostreptococcus productus (strain Marburg) has been purified to apparent homogeneity. The purified enzyme catalyzed the oxidation of NADH with methylenetetrahydrofolate as the electron acceptor at a specific activity of 380 mumols.min-1 mg protein-1 (37 degrees C; pH 5.5). The apparent Km for NADH was near 10 microM. The apparent molecular mass of the enzyme was determined by gel filtration to be approximately 250.0 kDa. The enzyme consists of eight identical subunits with a molecular mass of 32 kDa. It contains 4 FAD/mol octamer which were reduced by the enzyme with NADH as the electron donor; iron could not be detected. Oxygen had no effect on the enzyme. Ultracentrifugation of cell extracts revealed that about 40% of the enzyme activity was recovered in the particulate fraction, suggesting that the enzyme is associated with the membrane. The enzyme also catalyzed the methylenetetrahydrofolate reduction with methylene blue as an artificial electron donor. The oxidation of methyltetrahydrofolate was mediated with methylene blue as the electron acceptor; neither NAD+ nor viologen dyes could replace methylene blue in this reaction. NADP(H) or FAD(H2) were not used to substrates for the reaction in either direction. The activity of the purified enzyme, which was proposed to be involved in sodium translocation across the cytoplasmic membrane, was not affected by the absence or presence of added sodium. The properties of the enzyme differ from those of the ferredoxin-dependent methylenetetrahydrofolate reductase of the homoacetogen Clostridium formicoaceticum and of the NADP(+)-dependent reductase of eucaryotes investigated so far.  相似文献   

16.
Ferredoxin:NADP+ oxidoreductase (ferredoxin: NADP+ reductase, EC 1.18.1.2) was shown to form a ternary complex with its substrates ferredoxin (Fd) and NADP(H), but the ternary complex was less stable than the separate binary complexes. Kd for oxidized binary Fd-ferredoxin NADP+ reductase complex was less than 50 nM; Kd(Fd) increased with NADP+ concentration, approaching 0.5-0.6 microM when the flavoprotein was saturated with NADP+ K(NADP+) also increased from about 14 microM to about 310 microM, on addition of excess Fd. The changes in Kd were consistent with negative cooperativity between the associations of Fd and NADP+ and with our unpublished observations which suggest that product dissociation is rate-limiting in the reaction mechanism. Similar interference in binding was observed in more reduced states; NADPH released much ferredoxin:NADP+ reductase from Fd-Sepharose whether the proteins were initially oxidized or reduced. Complexation between Fd and ferredoxin: NADP+ reductase was found to shield each center from paramagnetic probes; charge specificity suggested that the active sites of Fd and ferredoxin:NADP+ reductase were, respectively, negatively and positively charged.  相似文献   

17.
Pure glutathione reductase from Saccharomyces cerevisiae catalyzed under anaerobic conditions the enzymatic reduction of GSSG using electrochemically reduced methyl viologen as electron donor. The new assay was completely dependent on the amount of active enzyme present, and involved the formation of 1 mol GSH per mole of reduced methyl viologen consumed. The enzyme followed a standard Michaelis-Menten kinetics; a Km = 230 microM for reduced methyl viologen and a turnover number of 969 mumol GSSG reduced per minute per micromole enzyme were determined. The enzymatic activity seemed to depend on the redox potential, showing half-maximal activity at -0.407 V. The enzyme was quite specific: the activity using reduced benzyl viologen as electron donor was just 1.5% of that obtained with reduced methyl viologen at the same concentration and potential. Glutathione reductase was totally inactivated after a brief anaerobic exposure with reduced methyl viologen in the absence of GSSG; a partial reactivation was observed following addition of glutathione disulfide. No inhibition of the methyl viologen-dependent activity was observed in the presence of 2',5'-ADP or 2'-P-5'-ADP-ribose, two NADP(H) analogs, at concentrations which drastically inhibited the NADPH-dependent activity, thus suggesting that the reduced viologen does not interact with the pyridine nucleotide-binding site.  相似文献   

18.
J A Navarro  M Roncel  G Tollin 《Biochemistry》1990,29(25):6102-6107
Steady-state and laser flash photolysis techniques have been used to examine the photoreduction of yeast glutathione reductase by the one-electron reduction products of 5-deazariboflavin and the viologen analogue 1,1'-propylene-2,2'-bipyridyl. Steady-state photoreduction of the enzyme with the viologen generates the two-electron-reduced form, whereas photoreduction with deazaflavin generates the anion semiquinone. Flash photolysis indicates that the product of viologen radical reduction is also a semiquinone, suggesting that this species is rapidly further reduced by viologen in the steady-state experiment to form the EH2 enzyme. This reduction is apparently inhibited when deazaflavin is the photoreductant, perhaps due to complexation of the anion semiquinone with deazaflavin. Steady-state experiments demonstrate that complexation of the anion semiquinone with NADP+ also inhibits further reduction. Both one-electron reduction reactions of oxidized glutathione reductase proceed at close to diffusion-controlled rates (second-order rate constants = 10(8)-10(9) M-1 s-1), despite the relatively buried nature of the FAD cofactor. Addition of NADP+ and oxidized glutathione produced no effects on the kinetics of the initial entry of the electron into the enzyme. No kinetic evidence of intramolecular electron transfer involving the FAD and the protein disulfide was obtained during or subsequent to the initial one-electron reduction process. Thus, if this reaction occurs in the semiquinone, it must be quite rapid (k greater than 8000 s-1).  相似文献   

19.
Chloroplast ferredoxin-NADP(+) reductase has a 32,000-fold preference for NADPH over NADH, consistent with its main physiological role of NADP(+) photoreduction for de novo carbohydrate biosynthesis. Although it is distant from the 2'-phosphoryl group of NADP(+), replacement of the C-terminal tyrosine (Tyr(308) in the pea enzyme) by Trp, Phe, Gly, and Ser produced enzyme forms in which the preference for NADPH over NADH was decreased about 2-, 10-, 300-, and 400-fold, respectively. Remarkably, in the case of the Y308S mutant, the k(cat) value for the NADH-dependent activity approached that of the NADPH-dependent activity of the wild-type enzyme. Furthermore, difference spectra of the NAD(+) complexes revealed that the nicotinamide ring of NAD(+) binds at nearly full occupancy in the active site of both the Y308G and Y308S mutants. These results correlate well with the k(cat) values obtained with these mutants in the NADH-ferricyanide reaction. The data presented support the hypothesis that specific recognition of the 2'-phosphate group of NADP(H) is required but not sufficient to ensure a high degree of discrimination against NAD(H) in ferredoxin-NADP(+) reductase. Thus, the C-terminal tyrosine enhances the specificity of the reductase for NADP(H) by destabilizing the interaction of a moiety common to both coenzymes, i.e. the nicotinamide.  相似文献   

20.
A thioredoxin reductase (TrxR) has been identified in the hyperthermophilic archaeon Sulfolobus solfataricus (Ss). This enzyme is a homodimeric flavoprotein that was previously identified as NADH oxidase in the same micro-organism ('Biotechnol. Appl. Biochem. 23 (1996) 47'). The primary structure of SsTrxR is made of 323 amino acid residues and contains two putative betaalphabeta regions for the binding of FAD, and a NADP(H) binding consensus sequence in the proximity of a CXXC motif. These findings indicate that SsTrxR is structurally related to the class II of the pyridine nucleotide-disulphide oxidoreductases family. Moreover, the enzyme exhibits a NADP(H) dependent thioredoxin reductase activity requiring the presence of FAD. Surprisingly, the reductase activity of SsTrxR is reduced in the presence of a specific inhibitor of mammalian TrxR. This finding demonstrates that the archaeal enzyme, although structurally related to eubacterial TrxR, is functionally closer to eukaryal enzymes. Experimental evidences indicate that a disulphide bridge is required for the reductase but also for the NADH oxidase activity of the enzyme. These results are further supported by the significantly reduced activities exerted by the C147A mutant. The integrity of the CXXC motif is also involved in the stability of the enzyme.  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