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 109 毫秒
1.
The reduction of duroquinone (DQ) and 2,3-dimethoxy-5-methyl-6-decyl-1,4-benzoquinone (DB) by NADH and ethanol was investigated in intact yeast mitochondria with good respiratory control ratios. In these mitochondria, exogenous NADH is oxidized by the NADH dehydrogenase localized on the outer surface of the inner membrane, whereas the NADH produced by ethanol oxidation in the mitochondrial matrix is oxidized by the NADH dehydrogenase localized on the inner surface of the inner membrane. The reduction of DQ by ethanol was inhibited 86% by myxothiazol; however, the reduction of DQ by NADH was inhibited 18% by myxothiazol, suggesting that protein-protein interactions between the internal (but not the external) NADH: ubiquinone oxidoreductase and ubiquinol:cytochrome c oxidoreductase (the cytochrome bc1 complex) are involved in the reduction of DQ by NADH. The reduction of DQ and DB by NADH and ethanol was also investigated in mutants of yeast lacking cytochrome b, the iron-sulfur protein, and ubiquinone. The reduction of both quinone analogues by exogenous NADH was reduced to levels that were 10 to 20% of those observed in wild-type mitochondria; however, the rate of their reduction by ethanol in the mutants was equal to or greater than that observed in the wild-type mitochondria. Furthermore, the reduction of DQ in the cytochrome b and iron-sulfur protein lacking mitochondria was myxothiazol sensitive, suggesting that neither of these proteins is an essential binding site for myxothiazol. The mitochondria from the three mutants also contained significant amounts of antimycin- and myxothiazol-insensitive NADH:cytochrome c reductase activity, but had no detectable succinate:cytochrome c reductase activity. These results suggest that the mutants lacking a functional cytochrome bc1 complex have adapted to oxidize NADH.  相似文献   

2.
The oxidation of matrix NADH in the presence and absence of rotenone was investigated in submitochondrial particles prepared from purified beetroot ( Beta vulgaris L.) mitochondria. The submitochondrial particles oxidised NADH using oxygen and artificial electron acceptors such as ferricyanide (FeCN) and short-chain analogues of ubiquinone(UQ)-10, although the NADH-FeCN reductase activity was not inhibited by rotenone. NADH-oxygen reductase activity in the presence and absence of rotenone displayed different affinities for NADH (145 ± 37 and 24 ± 9 μ M , respectively). However, in the presence of 0.15 m M UQ-1 where any contribution from non-specific sites of UQ-reduction was minimal, the rotenone-insensitive oxygen uptake was stimulated dramatically and the Km(NADH) decreased from 167 ± 55 μ M to 11 ± 1 μ M ; a value close to that determined for the total oxygen uptake which itself was virtually unaffected by the addition of UO-1 [Km(NADH) of 13 ± 3 μ M ).
The similar affinity of NADH-oxygen reductase for NADH when UQ-1 was present in both the presence and absence of rotenone, suggested that there may be only one NADH binding site involved in the two activities. A quantitative two-stage model for Complex I is postulated with one NADH binding site and two sites of UQ-reduction (one of which is insensitive to rotenone) with a common intermediate 'P' whose level of reduction can influence the NADH binding site. The poor affinity that rotenone-insensitive NADH-oxygen reductase activity displayed for NADH results from a limitation on the interaction of its UQ-reduction site with UQ-10 in the membrane; possibly from a low concentration of UQ-10 around this site or from steric hindrance restricting the access of UQ-10 to this reduction site.  相似文献   

3.
Guard cell protoplasts of Commelina communis L. reduced exogenousferricyanide at pH values lower than 5?0; upon addition of NADH,reduction of ferricyanide by guard cell protoplasts was stimulatedover the pH range 4?0 to 9?0 with two peaks of activity at pH5?0 and between pH 8?0 and pH 9?0. Calcium chloride (1?0 molm–3) and MgCl2 (1?0 mol m–3) increased the NADH-stimulatedreduction of ferricyanide. Superoxide dismutase and cyanidehad little effect on the NADH-stimulated reduction of ferricyanideby guard cell protoplasts, but, salicylhydroxamic acid completelyinhibited this activity. The NADH-stimulated reduction of ferricyanidealso occurred in the cell-free supernatant. Horseradish peroxidasedid not reduce ferricyanide in the absence of NADH over a broadrange of pH (4?0 to 9?0). However, in the presence of NADH,horseradish peroxidase reduced ferricyanide over the pH range5?0 to 9?0 with maximal activity at pH 8?0. The NADH-stimulatedreduction of ferricyanide by horseradish peroxidase showed similarproperties to those observed with guard cell protoplasts. Mannitol,superoxide dismutase, and cyanide did not inhibit the NADH-stimulatedreduction of ferricyanide by horseradish peroxidase; SHAM, however,completely inhibited the reduction of ferricyanide by horseradishperoxidase. Catalase inhibited the NADH-stimulated reductionof ferricyanide by horseradish peroxidase by 20%, while absenceof oxygen in the assay medium stimulated this activity over60%. We propose that the reduction of ferricyanide in the presenceof NADH by guard cell protoplasts, can be explained in termsof peroxidase activity associated with the plasma membrane andsecreted to the extracellular medium. However, the capacityof guard cell protoplasts to reduce ferricyanide at acid pHvalues where little peroxidase activity occurs may indicatethe presence of a plasma membrane redox system in guard cellsof C. communis. Key words: Commelina, guard cell protoplasts, ferricyanide reduction, peroxidase, redox system  相似文献   

4.
The processes of NADH oxidation by p-NTF violet and ubiquinone catalyzed by isolated yeast alcohol dehydrogenase in aqueous and water-alcohol buffer solutions were studied. In the presence of p-NTF in aqueous solution at a pH of 6–7, NADH oxidation was extremely slow due to inhibition of the enzyme by the remaining enzyme-bound hydrophobic product, formazan, which forms during the reduction of p-Nitrotetrazolium. However, when the medium was alkalinized to a pH of 8–9 or when alcohol (ethanol or isopropanol) was added, formazan was desorbed from the enzyme, leading to an increase in the NADH oxidation rate. It was assumed that this redox reaction can be used as the basis for colorimetric measurement of the activity of different alcohol dehydrogenases. Tetrazolium reduction by alcohol was not observed at any value within the entire pH range. NADH oxidation in the presence of the enzyme and ubiquinone was also slow, even with the addition of alcohol, but its rate increased when the medium was acidified to a pH of 5.5–6. When a Tris-phosphate buffer was replaced with HEPES, a quasi-vibrational process was observed: NADH oxidization with ubiquinone to NAD+ and its subsequent reverse recovery with alcohol to NADH.  相似文献   

5.
Three rotenone-insensitive NADH dehydrogenases are present in the mitochondria of yeast Saccharomyces cerevisiae, which lack complex I. To elucidate the functions of these enzymes, superoxide production was determined in yeast mitochondria. The low levels of hydrogen peroxide (0.10 to 0.18 nmol/min/mg) produced in mitochondria incubated with succinate, malate, or NADH were stimulated 9-fold by antimycin A. Myxothiazol and stigmatellin blocked completely hydrogen peroxide formation with succinate or malate, indicating that the cytochrome bc(1) complex is the source of superoxide; however, these inhibitors only inhibited 46% hydrogen peroxide formation with NADH as substrate. Diphenyliodonium inhibited hydrogen peroxide formation (with NADH as substrate) by 64%. Superoxide formation, determined by EPR and acetylated cytochrome c reduction in mitochondria was stimulated by antimycin A, and partially inhibited by myxothiazol and stigmatellin. Proteinase K digestion of mitoplasts reduced 95% NADH dehydrogenase activity with a similar inhibition of superoxide production. Mild detergent treatment of the proteinase-treated mitoplasts resulted in an increase in NADH dehydrogenase activity due to the oxidation of exogenous NADH by the internal NADH dehydrogenase; however, little increase in superoxide production was observed. These results suggest that the external NADH dehydrogenase is a potential source of superoxide in S. cerevisiae mitochondria.  相似文献   

6.
The S9 fraction of MCF-7 human breast carcinoma cells has NAD(P)H (quinone-acceptor) oxidoreductase activity as measured by the reduction of dichlorophenol-indophenol (DCPIP). This reduction is dependent on the activators Tween-20 and bovine serum albumin and it is inhibitable by dicumarol. The S9 fraction also has cytochrome c reductase activity which is approximately 29 times less than the two-electron reduction activity of NAD(P)H (quinone-acceptor) oxidoreductase. Diaziquone (AZQ) is a substrate for this NAD(P)H oxidoreductase active S9 fraction as judged by its enzymatic reduction detected spectrophotometrically and by electron spin resonance spectroscopy. Two-electron mediated enzymatic reduction of AZQ was evidenced by the formation of the colorless dihydroquinone (AZQH2) which could be followed at 340 nm. The production of the dihydroquinone was inhibitable by dicumarol implicating NAD(P)H oxidoreductase in its formation. Under aerobic conditions, electron spin resonance spectroscopy showed evidence for the production of AZQ semiquinone (AZQH) and oxygen radicals. Under anaerobic conditions no oxygen radicals were observed, but the semiquinone was stable for hours. These results are also inhibitable by dicumarol and suggest a two-step one-electron oxidation process of the dihydroquinone. The production of semiquinone and oxygen radicals as detected by electron spin resonance spectroscopy was more sensitive to dicumarol when NADPH was used as cofactor (68% inhibition of OH and 65% inhibition of AZQH) than when NADH was used (28% inhibition of OH and 5% inhibition of AZQH). This suggests that NADH flavin reductases play a more important role in the one-electron reduction pathway of AZQ in MCF-7 S9 fraction than NADPH reductases. The reduction of AZQ by NAD(P)H (quinone-acceptor) oxidoreductase may play an important role in the bioreductive alkylating properties of AZQ.  相似文献   

7.
The effect of extracellular NADH on the rate of reduction of nitrite-induced methaemoglobin in erythrocytes from man, cattle, dog, horse, grey kangaroo, pig and sheep was investigated. Extracellular NADH was found to enhance the rate of methaemoglobin reduction in man, dog, pig and kangaroo erythrocytes, but had essentially no effect on the rate of methaemoglobin reduction in erythrocytes from cattle, horse and sheep. In erythrocytes of those animals affected by extracellular NADH the rate of reduction of metHb in the presence of NADH was the same or greater than that observed in the presence of nutrients such as glucose and inosine. The combination of nutrient and NADH produced a more profound increase in the rate of methaemoglobin reduction. The rate of methaemoglobin reduction in all cases was significantly less than that observed with methylene blue, the standard treatment of methaemoglobinaemia. Extracellular NADH was found to indirectly increase the intracellular NADH concentration through displacement of the pseudo-equilibrium of the intracellular LDH reaction and relied upon the presence of sufficient LDH activity released into the extracellular medium through haemolysis. The lack of response of cattle, horse and sheep RBCs to extracellular NADH was found to derive mainly from their low extracellular LDH activity, but also correlated with their lower NADH-methaemoglobin reductase activity compared to the other species.  相似文献   

8.
H2S reduced growth of spinach at levels higher than 0.075 µl/literand inhibited NADH oxidation by shoot extracts. Inhibition ofNADH oxidation was maximum after a 48-h exposure. Oxidationof NADH was equally sensitive to sulfide and cyanide. NADH oxidationshowed promise as an early indicator for growth reduction byH2S. 1Present address: Centre for Agrobiological Research, P.O. Box14, 6700 AA Wageningen, The Netherlands. (Received May 18, 1987; Accepted February 9, 1988)  相似文献   

9.
Ammonium-dependent induction of a GDH isozyme in oat leaveswas proportional to light intensity and inhibited by DCMU. Thestimulation of GDH synthesis in response to ammonia was partiallyrepressed by benzimidazole. The inducible (no. 1) and noninducible(no. 2) GDH isozymes wereseparated and purified by approximately54 and 24 fold respectively. The two isozymes were highly specificfor NAD and the rate of NADH oxidation was 7 to 9 times higherthan NAD reduction. Both isozymes showed similar Km values forsubstrates of the reductive amination reaction and pH optimafor NADH oxidation. The pH optima for NAD reduction were 9 and8.2 for isozymes1 and 2 respectively. The two isozymes had asimilar molecular weight, 2.2–2.4 x 105 but differed intheir isoelectric point and temperature sensitivity. Resultssuggest that the GDH isozymes in oat leaves are two differententities but might possess a similar metabolic function. (Received January 6, 1976; )  相似文献   

10.
Preparations with a selectively decreased (by 85-90%) content of NADH dehydrogenase were isolated by means of heating treatment of M. lysodeikticus isolated membranes. The degree of the reduction of the NADH dehydrogenase nearest neighbour in the respiration chain of cytochrome b556 in heated membranes is similar to that in intact membranes. It is concluded that cytochrome b556 and (or) NADH dehydrogenase are capable to lateral migration in the membrane of M. lysodeikticus, resulting in the inter-chain electrone transport. A coefficient of their lateral diffusion is calculated (D equals 8-10(-10)-2-10(-9) CM2SEC-1 At 30 degrees C) on the basis of kinetics of cytochrome reduction by NADH dehydrogenase. The electron transport, due to a diffusion of respiration carriers from one assambly to another, proceeds 100 times as slow as the electrone transport in the respiratory chain. The data obtained allow to consider the aggregation of respiration enzymes as a dynamic formation.  相似文献   

11.
The membrane fraction of Bacillus subtilis catalyzes the reduction of fumarate to succinate by NADH. The activity is inhibited by low concentrations of 2-(heptyl)-4-hydroxyquinoline-N-oxide (HOQNO), an inhibitor of succinate: quinone reductase. In sdh or aro mutant strains, which lack succinate dehydrogenase or menaquinone, respectively, the activity of fumarate reduction by NADH was missing. In resting cells fumarate reduction required glycerol or glucose as the electron donor, which presumably supply NADH for fumarate reduction. Thus in the bacteria, fumarate reduction by NADH is catalyzed by an electron transport chain consisting of NADH dehydrogenase (NADH:menaquinone reductase), menaquinone, and succinate dehydrogenase operating in the reverse direction (menaquinol:fumarate reductase). Poor anaerobic growth of B. subtilis was observed when fumarate was present. The fumarate reduction catalyzed by the bacteria in the presence of glycerol or glucose was not inhibited by the protonophore carbonyl cyanide m-chlorophenyl hydrazone (CCCP) or by membrane disruption, in contrast to succinate oxidation by O2. Fumarate reduction caused the uptake by the bacteria of the tetraphenyphosphonium cation (TPP+) which was released after fumarate had been consumed. TPP+ uptake was prevented by the presence of CCCP or HOQNO, but not by N,N'-dicyclohexylcarbodiimide, an inhibitor of ATP synthase. From the TPP+ uptake the electrochemical potential generated by fumarate reduction was calculated (Deltapsi = -132 mV) which was comparable to that generated by glucose oxidation with O2 (Deltapsi = -120 mV). The Deltapsi generated by fumarate reduction is suggested to stem from menaquinol:fumarate reductase functioning in a redox half-loop.  相似文献   

12.
The coupling of the quinoprotein glucose dehydrogenase to the electron transport chain has been investigated in Acinetobacter calcoaceticus. No evidence was obtained to support a previous suggestion that the soluble form of the dehydrogenase and the soluble cytochrome b associated with it are involved in the oxidation of glucose. Analysis of cytochrome content, and of reduction of cytochromes in membranes by substrates, and of sensitivity to cyanide indicated that glucose, succinate and NADH are all oxidized by way of the same b-type cytochrome(s) and cytochrome oxidases (cytochrome o and cytochrome d). Mixed inhibition studies [with KCN and hydroxyquinoline N-oxide (HQNO)] showed that the b-type cytochrome(s) formed a binary complex with the o-type oxidase and that there was thus no communication between the electron transport chains at the cytochrome level. Measurements of the reduction of ubiquinone-9 by glucose and NADH, and inhibitor studies using HQNO, indicated that the ubiquinone mediates electron transport from both the glucose and NADH dehydrogenases. In some conditions the quinone pool facilitated communication between the 'glucose oxidase' and 'NADH oxidase' electron transport chains, but in normal conditions these chains were kinetically distinct.  相似文献   

13.
Nidetzky B  Klimacek M  Mayr P 《Biochemistry》2001,40(34):10371-10381
Microbial xylose reductase, a representative aldo-keto reductase of primary sugar metabolism, catalyzes the NAD(P)H-dependent reduction of D-xylose with a turnover number approximately 100 times that of human aldose reductase for the same reaction. To determine the mechanistic basis for that physiologically relevant difference and pinpoint features that are unique to the microbial enzyme among other aldo/keto reductases, we carried out stopped-flow studies with wild-type xylose reductase from the yeast Candida tenuis. Analysis of transient kinetic data for binding of NAD(+) and NADH, and reduction of D-xylose and oxidation of xylitol at pH 7.0 and 25 degrees C provided estimates of rate constants for the following mechanism: E + NADH right arrow over left arrow E.NADH right arrow over left arrow E.NADH + D-xylose right arrow over left arrow E.NADH.D-xylose right arrow over left arrow E.NAD(+).xylitol right arrow over left arrow E.NAD(+) right arrow over left arrow E.NAD(+) right arrow over left arrow E + NAD(+). The net rate constant of dissociation of NAD(+) is approximately 90% rate limiting for k(cat) of D-xylose reduction. It is controlled by the conformational change which precedes nucleotide release and whose rate constant of 40 s(-)(1) is 200 times that of completely rate-limiting E.NADP(+) --> E.NADP(+) step in aldehyde reduction catalyzed by human aldose reductase [Grimshaw, C. E., et al. (1995) Biochemistry 34, 14356-14365]. Hydride transfer from NADH occurs with a rate constant of approximately 170 s(-1). In reverse reaction, the E.NADH --> E.NADH step takes place with a rate constant of 15 s(-1), and the rate constant of ternary-complex interconversion (3.8 s(-1)) largely determines xylitol turnover (0.9 s(-1)). The bound-state equilibrium constant for C. tenuis xylose reductase is estimated to be approximately 45 (=170/3.8), thus greatly favoring aldehyde reduction. Formation of productive complexes, E.NAD(+) and E.NADH, leads to a 7- and 9-fold decrease of dissociation constants of initial binary complexes, respectively, demonstrating that 12-fold differential binding of NADH (K(i) = 16 microM) vs NAD(+) (K(i) = 195 microM) chiefly reflects difference in stabilities of E.NADH and E.NAD(+). Primary deuterium isotope effects on k(cat) and k(cat)/K(xylose) were, respectively, 1.55 +/- 0.09 and 2.09 +/- 0.31 in H(2)O, and 1.26 +/- 0.06 and 1.58 +/- 0.17 in D(2)O. No deuterium solvent isotope effect on k(cat)/K(xylose) was observed. When deuteration of coenzyme selectively slowed the hydride transfer step, (D)()2(O)(k(cat)/K(xylose)) was inverse (0.89 +/- 0.14). The isotope effect data suggest a chemical mechanism of carbonyl reduction by xylose reductase in which transfer of hydride ion is a partially rate-limiting step and precedes the proton-transfer step.  相似文献   

14.
The mechanism of disulphide reduction by mitochondria   总被引:1,自引:0,他引:1       下载免费PDF全文
1. Cystamine was reduced to the corresponding thiol by rat liver mitochondria, even in the presence of rotenone or antimycin A. 2. The reduction of disulphides was stimulated by the accumulation of NADH or by the addition of NADH to osmotically ;shocked' mitochondria. 3. Energy made available by oxidative phosphorylation was not essential for the reduction of disulphides. 4. Cystamine was not reduced during the oxidation of NADH by ultrasonically treated particles, which had lost their capacity for oxidation of alpha-oxo acids. 5. In intact mitochondria, arsenite and other inhibitors of vicinal dithiols caused a decrease in the capacity for reduction of disulphides concomitantly with an inhibition of the oxidation of alpha-oxo acids. 6. Isolated lipoamide dehydrogenase reduced cystamine at the expense of NADH, provided that lipoic acid was also present. 7. It is concluded that in mitochondria the reduction of cystamine and related disulphides is probably brought about by interaction with reduced lipoic acid, generated by the alpha-oxo acid dehydrogenase complexes during the oxidation of alpha-oxo acids or by reaction of lipoamide dehydrogenase with NADH.  相似文献   

15.
The degree of myocardial oxygen delivery (Do2) that is necessary to reestablish functional contractile activity after short-term global ischemia in heart is not known. To determine the relationship between Do2 and recovery of contractile and metabolic functions, we used tissue NADH fluorometric changes to characterize adequacy of reperfusion flow. Isolated perfused rat hearts were subjected to global ischemia and were reperfused at variable flow rates that ranged from 1 to 100% of baseline flow. Myocardial function and tissue NADH changes were continuously measured. NADH fluorescence rapidly increased and plateaued during ischemia. A strong inverse logarithmic correlation between NADH fluorescence and reperfusion Do2 was demonstrated (r = -0.952). Left ventricular function (rate-pressure product) was inversely related to NADH fluorescence at reperfusion flows from 25 to 100% of baseline (r = -0.922) but not at lower reperfusion flow levels. An apparent reperfusion threshold of 25% of baseline Do2 was necessary to resume contractile function. At very low reperfusion flows (1% of baseline), another threshold flow was identified at which NADH levels increased beyond that observed during global ischemia (3.4 +/- 3.0%, means +/- SE, n = 9), which suggests further reduction of the cellular redox state. This NADH increase at 1% of baseline reperfusion flow was blocked by removing glucose from the perfusate. NADH fluorescence is a sensitive indicator of myocardial cellular oxygen utilization over a wide range of reperfusion Do2 values. Although oxygen is utilized at very low flow rates, as indicated by changes in NADH, a critical threshold of approximately 25% of baseline Do2 is necessary to restore contractile function after short-term global ischemia.  相似文献   

16.
Deoxyhypusine is a modified lysine residue. It is formed posttranslationally in the precursor of eukaryotic initiation factor 5A (eIF5A) by deoxyhypusine synthase, employing spermidine as a butylamine donor. In the initial step of this reaction, deoxyhypusine synthase catalyzes the production of NADH through dehydrogenation of spermidine. Fluorescence measurements of this reaction revealed a -22-nm blue shift in the emission peak of NADH and a approximately 15-fold increase in peak intensity, characteristics of tightly bound NADH that were not seen by simply mixing NADH and enzyme. The fluorescent properties of the bound NADH can be ascribed to a hydrophobic environment and a rigidly held, open conformation of NADH, features in accord with the known crystal structure of the enzyme. Considerable fluorescence resonance energy transfer from tryptophan 327 in the active site to the dihydronicotinamide ring of NADH was seen. Upon addition of the eIF5A precursor, utilization of the enzyme-bound NADH for reduction of the eIF5A-imine intermediate to deoxyhypusine was reflected by a rapid decrease in the NADH fluorescence, indicating a transient hydride transfer mechanism as an integral part of the reaction. The number of NADH molecules bound approached four/enzyme tetramer; not all of the bound NADH was available for reduction of the eIF5A-imine intermediate.  相似文献   

17.
C Meyer  I Cherel  T Moureaux  J Hoarau  J Gabard  P Rouze 《Biochimie》1987,69(6-7):735-742
NADH: nitrate reductase (EC 1.6.6.1) was purified from Nicotiana plumbaginifolia leaves. As recently observed with nitrate reductase from other sources, this enzyme is able to reduce nitrate using reduced bromphenol blue (rBPB) as the electron donor. In contrast to the physiological NADH-dependent activity, the rBPB-dependent activity is stable in vitro. The latter activity is non-competitively inhibited by NADH. The monoclonal antibody ZM.96(9)25, which inhibits the NADH: nitrate reductase total activity as well as the NADH: cytochrome c reductase and reduced methyl viologen (rMV): nitrate reductase partial activities, has no inhibitory effect on the rBPB: nitrate reductase activity. Conversely, the monoclonal antibody NP.17-7(6) inhibits nitrate reduction with all three electron donors: NADH, MV or BPB. Among various nitrate reductase-deficient mutants, an apoprotein gene mutant (nia. E56) shows reduced terminal activities but a highly increased rBPB:nitrate reductase activity. rBPB:nitrate reductase thus appears to be a new terminal activity of higher plant nitrate reductase and involves specific sites which are not shared by the other activities.  相似文献   

18.
3-Hydroxybenzoate 6-hydroxylase (3HB6H) from Rhodococcus jostii RHA1 is a nicotinamide adenine dinucleotide (NADH)-specific flavoprotein monooxygenase involved in microbial aromatic degradation. The enzyme catalyzes the para hydroxylation of 3-hydroxybenzoate (3-HB) to 2,5-dihydroxybenzoate (2,5-DHB), the ring-fission fuel of the gentisate pathway. In this study, the kinetics of reduction of the enzyme-bound flavin by NADH was investigated at pH 8.0 using a stopped-flow spectrophotometer, and the data were analyzed comprehensively according to kinetic derivations and simulations. Observed rate constants for reduction of the free enzyme by NADH under anaerobic conditions were linearly dependent on NADH concentrations, consistent with a one-step irreversible reduction model with a bimolecular rate constant of 43 ± 2 M(-1) s(-1). In the presence of 3-HB, observed rate constants for flavin reduction were hyperbolically dependent on NADH concentrations and approached a limiting value of 48 ± 2 s(-1). At saturating concentrations of NADH (10 mM) and 3-HB (10 mM), the reduction rate constant is ~51 s(-1), whereas without 3-HB, the rate constant is 0.43 s(-1) at a similar NADH concentration. A similar stimulation of flavin reduction was found for the enzyme-product (2,5-DHB) complex, with a rate constant of 45 ± 2 s(-1). The rate enhancement induced by aromatic ligands is not due to a thermodynamic driving force because Em 0 for the enzyme-substrate complex is -179 ± 1 mV compared to an E(m)(0) of -175 ± 2 mV for the free enzyme. It is proposed that the reduction mechanism of 3HB6H involves an isomerization of the initial enzyme-ligand complex to a fully activated form before flavin reduction takes place.  相似文献   

19.
The respiratory chain-linked NADH-quinone reductase (NQR) and NADH-ferricyanide dehydrogenase (NFD) were extracted from membranes of Escherichia coli by n-dodecyl octaethyleneglycol monoether detergent and purified by DEAE-Sephacel, DEAE-5PW and Bio-Gel HTP column chromatography. The purified NQR contained FAD as a cofactor, catalyzed the reduction of ubiquinone-1 (Q1) and reacted with NADH, but not with deamino-NADH (d-NADH), with an apparent Km of 48 microM. On the other hand, the purified NFD contained FMN as a cofactor, reacted with both NADH and d-NADH, and catalyzed the reduction of ferricyanide but not Q1. NFD showed a high affinity for both NADH and d-NADH with a Km of 7-9 microM. NFD was inactivated, whereas NQR was rather activated, by preincubation with an electron donor in the absence of electron acceptor. These properties were compared with those of activities observed with inverted membrane vesicles with special reference to the generation of inside-positive membrane potential (delta psi). It was found that d-NADH-reactive FMN-containing NFD is a dehydrogenase part of energy-generating NADH-quinone reductase complex. The FAD-containing NQR was very similar to that purified by Jaworowski et al. (Biochemistry (1981) 20, 2041-2047), and reduced Q1 without generating delta psi.  相似文献   

20.
1. The mechanisms of the reduction of oxaloacetate and of 3-fluoro-oxaloacetate by NADH catalysed by cytoplasmic pig heart malate dehydrogenase (MDH) were investigated. 2. One mol of dimeric enzyme produces 1.7+/-0.4 mol of enzyme-bound NADH when mixed with saturating NAD+ and L-malate at a rate much higher than the subsequent turnover at pH 7.5. 3. Transient measurements of protein and nucleotide fluorescence show that the steady-state complex in the forward direction is MDH-NADH and in the reverse direction MDH-NADH-oxaloacetate. 4. The rate of dissociation of MDH-NADH was measured and is the same as Vmax. in the forward direction at pH 7.5. Both NADH-binding sites are kinetically equivalent. The rate of dissociation varies with pH, as does the equilibrium binding constant for NADH. 5. 3-Fluoro-oxaloacetate is composed of three forms (F1, F2 and S) of which F1 and F2 are immediately substrates for the enzyme. The third form, S, is not a substrate, but when the F forms are used up form S slowly and non-enzymically equilibrates to yield the active substrate forms. S is 2,2-dihydroxy-3-fluorosuccinate. 6. The steady-state compound during the reduction of form F1 is an enzyme form that does not contain NADH, probably MDH-NAD+-fluoromalate. The steady-state compound for form F2 is an enzyme form containing NADH, probably MDH-NADH-fluoro-oxaloacetate. 7. The rate-limiting reaction in the reduction of form F2 shows a deuterium isotope rate ratio of 4 when NADH is replaced by its deuterium analogue, and the rate-limiting reaction is concluded to be hydride transfer. 8. A novel titration was used to show that dimeric cytoplasmic malate dehydrogenase contains two sites that can rapidly reduce the F1 form of 3-fluoro-oxaloacetate. The enzyme shows 'all-of-the-sites' behaviour. 9. Partial mechanisms are proposed to explain the enzyme-catalysed transformations of the natural and the fluoro substrates. These mechanisms are similar to the mechanism of pig heart lactate dehydrogenase and this, and the structural results of others, can be explained if the two enzymes are a product of divergent evolution.  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