首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Superoxide production by inside-out coupled bovine heart submitochondrial particles, respiring with succinate or NADH, was measured. The succinate-supported production was inhibited by rotenone and uncouplers, showing that most part of superoxide produced during succinate oxidation is originated from univalent oxygen reduction by Complex I. The rate of the superoxide (O2*-)) production during respiration at a high concentration of NADH (1 mM) was significantly lower than that with succinate. Moreover, the succinate-supported O2*- production was significantly decreased in the presence of 1 mM NADH. The titration curves, i.e., initial rates of superoxide production versus NADH concentration, were bell-shaped with the maximal rate (at 50 microM NADH) approaching that seen with succinate. Both NAD+ and acetyl-NAD+ inhibited the succinate-supported reaction with apparent Ki's close to their Km's in the Complex I-catalyzed succinate-dependent energy-linked NAD+ reduction (reverse electron transfer) and NADH:acetyl-NAD+ transhydrogenase reaction, respectively. We conclude that: (i) under the artificial experimental conditions the major part of superoxide produced by the respiratory chain is formed by some redox component of Complex I (most likely FMN in its reduced or free radical form); (ii) two different binding sites for NADH (F-site) and NAD+ (R-site) in Complex I provide accessibility of the substrates-nucleotides to the enzyme red-ox component(s); F-site operates as an entry for NADH oxidation, whereas R-site operates in the reverse electron transfer and univalent oxygen reduction; (iii) it is unlikely that under the physiological conditions (high concentrations of NADH and NAD+) Complex I is responsible for the mitochondrial superoxide generation. We propose that the specific NAD(P)H:oxygen superoxide (hydrogen peroxide) producing oxidoreductase(s) poised in equilibrium with NAD(P)H/NAD(P)+ couple should exist in the mitochondrial matrix, if mitochondria are, indeed, participate in ROS-controlled processes under physiologically relevant conditions.  相似文献   

2.
The kinetics of the NAD: artificial acceptor-oxidoreductase and delta mu H(+)-dependent succinate: NAD(+)-oxidoreductase reactions (reverse electron transfer) reactions catalyzed by the membrane-bound complex I was studied. The values of apparent rate constants of dissociation of complexes of the oxidized and reduced enzyme with NAD+ and NADH were determined. It was shown that the apparent affinity of NADH for the oxidized complex I is by nearly three orders of magnitude as high as that of the reduced one; a reverse correlation is found for NAD+. A kinetic scheme of complex I functioning in the forward and reverse reactions, according to which the free reduced enzyme is not an intermediate of the forward (NADH-oxidase) reaction and the free oxidized enzyme is not an intermediate of the reverse (NAD(+)-reductase) reaction, is proposed.  相似文献   

3.
Transhydrogenase (TH) is a dimeric integral membrane enzyme in mitochondria and prokaryotes that couples proton translocation across a membrane with hydride transfer between NAD(H) and NADP(H) in soluble domains. Crystal structures of the NAD(H) binding alpha1 subunit (domain I) of Rhodospirillum rubrum TH have been determined at 1.8 A resolution in the absence of dinucleotide and at 1.9 A resolution with NADH bound. Each structure contains two domain I dimers in the asymmetric unit (AB and CD); the dimers are intimately associated and related by noncrystallographic 2-fold axes. NADH binds to subunits A and D, consistent with the half-of-the-sites reactivity of the enzyme. The conformation of NADH in subunits A and D is very similar; the nicotinamide is in the anti conformation, the A-face is exposed to solvent, and both N7 and O7 participate in hydrogen bonds. Comparison of subunits A and D to six independent copies of the subunit without bound NADH reveals multiple conformations for residues and loops surrounding the NADH site, indicating flexibility for binding and release of the substrate (product). The NADH-bound structure is also compared to the structures of R. rubrum domain I with NAD bound (PDB code 1F8G) and with NAD bound in complex with domain III of TH (PDB code 1HZZ). The NADH- vs NAD-bound domain I structures reveal conformational differences in conserved residues in the NAD(H) binding site and in dinucleotide conformation that are correlated with the net charge, i.e., oxidation state, of the nicotinamides. The comparisons illustrate how nicotinamide oxidation state can affect the domain I conformation, which is relevant to the hydride transfer step of the overall reaction.  相似文献   

4.
C T Grubmeyer  K W Chu  S Insinga 《Biochemistry》1987,26(12):3369-3373
Salmonella typhimurium histidinol dehydrogenase produces histidine from the amino alcohol histidinol by two sequential NAD-linked oxidations which form and oxidize a stable enzyme-bound histidinaldehyde intermediate. The enzyme was found to catalyze the exchange of 3H between histidinol and [4(R)-3H]NADH and between NAD and [4(S)-3H]NADH. The latter reaction proceeded at rates greater than kcat for the net reaction and was about 3-fold faster than the former. Histidine did not support an NAD/NADH exchange, demonstrating kinetic irreversibility in the second half-reaction. Specific activity measurements on [3H]histidinol produced during the histidinol/NADH exchange reaction showed that only a single hydrogen was exchanged between the two reactants, demonstrating that under the conditions employed this exchange reaction arises only from the reversal of the alcohol dehydrogenase step and not the aldehyde dehydrogenase reaction. The kinetics of the NAD/NADH exchange reaction demonstrated a hyperbolic dependence on the concentration of NAD and NADH when the two were present in a 1:2 molar ratio. The histidinol/NADH exchange showed severe inhibition by high NAD and NADH under the same conditions, indicating that histidinol cannot dissociate directly from the ternary enzyme-NAD-histidinol complex; in other words, the binding of substrate is ordered with histidinol leading. Binding studies indicated that [3H]histidinol bound to 1.7 sites on the dimeric enzyme (0.85 site/monomer) with a KD of 10 microM. No binding of [3H]NAD or [3H]NADH was detected. The nucleotides could, however, displace histidinol dehydrogenase from Cibacron Blue-agarose.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

5.
A simple in situ model of alamethicin-permeabilized isolated rat liver mitochondria was used to investigate the channeling of NADH between mitochondrial malate dehydrogenase (MDH) and NADH:ubiquinone oxidoreductase (complex I). Alamethicin-induced pores in the mitochondrial inner membrane allow effective transport of low molecular mass components such as NAD+/NADH but not soluble proteins. Permeabilized mitochondria demonstrate high rates of respiration in the presence of malate/glutamate and NAD+ due to coupled reaction between MDH and complex I. In the presence of pyruvate and lactate dehydrogenase, an extramitochondrial competitive NADH utilizing system, respiration of permeabilized mitochondria with malate/glutamate and NAD+ was completely abolished. These data are in agreement with the free diffusion of NADH and do not support the suggestion of direct channeling of NADH from MDH to complex I.  相似文献   

6.
Initial velocity and product inhibition studies were carried out on UDP-glucose dehydrogenase (UDPglucose: NAD+ 6-oxidoreductase, EC 1.1.1.22) from beef liver to determine if the kinetics of the reaction are compatible with the established mechanism. An intersecting initial velocity pattern was observed with NAD+ as the variable substrate and UDPG as the changing fixed substrate. UDPglucuronic acid gave competitive inhibition of UDPG and non-competitive inhibition of NAD+. Inhibition by NADH gave complex patterns.Lineweaver-Burk plots of 1/upsilon versus 1/NAD+ at varied levels of NADH gave highly non-linear curves. At levels of NAD+ below 0.05 mM, non-competitive inhibition patterns were observed giving parabolic curves. Extrapolation to saturation with NAD+ showed NADH gave linear uncompetitive inhibition of UDPG if NAD+ was saturating. However, at levels of NAD+ above 0.10 mM, NADH became a competitive inhibitor of NAD+ (parabolic curves) and when NAD+ was saturating NADH gave no inhibition of UDPG. NADH was non-competitive versus UDPG when NAD+ was not saturating. These results are compatible with a mechanism in which UDPG binds first, followed by NAD+, which is reduced and released. A second mol of NAD+ is then bound, reduced, and released. The irreversible step in the reaction must occur after the release of the second mol of NADH but before the release of UDPglucuronic acid. This is apparently caused by the hydrolysis of a thiol ester between UDPglucoronic acid and the essential thiol group of the enzyme. Examination of rate equations indicated that this hydrolysis is the rate-limiting step in the overall reaction. The discontinuity in the velocities observed at high NAD+ concentrations is apparently caused by the binding of NAD+ in the active site after the release of the second mol of NADH, eliminating the NADH inhibition when NAD+ becomes saturating.  相似文献   

7.
Xu H  West AH  Cook PF 《Biochemistry》2006,45(39):12156-12166
Kinetic data have been measured for the histidine-tagged saccharopine dehydrogenase from Saccharomyces cerevisiae, suggesting the ordered addition of nicotinamide adenine dinucleotide (NAD) followed by saccharopine in the physiologic reaction direction. In the opposite direction, the reduced nicotinamide adenine dinucleotide (NADH) adds to the enzyme first, while there is no preference for the order of binding of alpha-ketoglutarate (alpha-Kg) and lysine. In the direction of saccharopine formation, data also suggest that, at high concentrations, lysine inhibits the reaction by binding to free enzyme. In addition, uncompetitive substrate inhibition by alpha-Kg and double inhibition by NAD and alpha-Kg suggest the existence of an abortive E:NAD:alpha-Kg complex. Product inhibition by saccharopine is uncompetitive versus NADH, suggesting a practical irreversibility of the reaction at pH 7.0 in agreement with the overall K(eq). Saccharopine is noncompetitive versus lysine or alpha-Kg, suggesting the existence of both E:NADH:saccharopine and E:NAD:saccharopine complexes. NAD is competitive versus NADH, and noncompetitive versus lysine and alpha-Kg, indicating the combination of the dinucleotides with free enzyme. Dead-end inhibition studies are also consistent with the random addition of alpha-Kg and lysine. Leucine and oxalylglycine serve as lysine and alpha-Kg dead-end analogues, respectively, and are uncompetitive against NADH and noncompetitive against alpha-Kg and lysine, respectively. Oxaloacetate (OAA), pyruvate, and glutarate behave as dead-end analogues of lysine, which suggests that the lysine-binding site has a higher affinity for keto acid analogues than does the alpha-Kg site or that dicarboxylic acids have more than one binding mode on the enzyme. In addition, OAA and glutarate also bind to free enzyme as does lysine at high concentrations. Glutarate gives S-parabolic noncompetitive inhibition versus NADH, indicating the formation of a E:(glutarate)2 complex as a result of occupying both the lysine- and alpha-Kg-binding sites. Pyruvate, a slow alternative keto acid substrate, exhibits competitive inhibition versus both lysine and alpha-Kg, suggesting the combination to the E:NADH:alpha-Kg and E:NADH:lysine enzyme forms. The equilibrium constant for the reaction has been measured at pH 7.0 as 3.9 x 10(-7) M by monitoring the change in NADH upon the addition of the enzyme. The Haldane relationship is in very good agreement with the directly measured value.  相似文献   

8.
Treatment of submitochondrial particles (ETP) with trypsin at 0 degrees destroyed NADPH leads to NAD (or 3-acetylpyridine adenine dinucleotide, AcPyAD) transhydrogenase activity. NADH oxidase activity was unaffected; NADPH oxidase and NADH leads to AcPyAD transhydrogenase activities were diminished by less than 10%. When ETP was incubated with trypsin at 30 degrees, NADPH leads to NAD transhydrogenase activity was rapidly lost, NADPH oxidase activity was slowly destroyed, but NADH oxidase activity remained intact. The reduction pattern by NADPH, NADPH + NAD, and NADH of chromophores absorbing at 475 minus 510 nm (flavin and iron-sulfur centers) in complex I (NADH-ubiquinone reductase) or ETP treated with trypsin at 0 degrees also indicated specific destruction of transhydrogenase activity. The sensitivity of the NADPH leads to NAD transhydrogenase reaction to trypsin suggested the involvement of susceptible arginyl residues in the enzyme. Arginyl residues are considered to be positively charged binding sites for anionic substrates and ligands in many enzymes. Treatment of ETP with the specific arginine-binding reagent, butanedione, inhibited transhydrogenation from NADPH leads to NAD (or AcPyAD). It had no effect on NADH oxidation, and inhibited NADPH oxidation and NADH leads to AcPyAD transhydrogenation by only 10 to 15% even after 30 to 60 min incubation of ETP with butanedione. The inhibition of NADPH leads to NAD transhydrogenation was diminished considerably when butanedione was added to ETP in the presence of NAD or NADP. When both NAD and NADP were present, the butanedione effect was completely abolished, thus suggesting the possible presence of arginyl residues at the nucleotide binding site of the NADPH leads to NAD transhydrogenase enzyme. Under conditions that transhydrogenation from NADPH to NAD was completely inhibited by trypsin or butanedione, NADPH oxidation rate was larger than or equal to 220 nmol min-1 mg-1 ETP protein at pH 6.0 and 30 degrees. The above results establish that in the respiratory chain of beef-heart mitochondria NADH oxidation, NADPH oxidation, and NADPH leads to NAD transhydrogenation are independent reactions.  相似文献   

9.
The rotenone-insensitive reduction of quinones and aromatic nitrocompounds by mitochondrial NADH: ubiquinone reductase (complex I, EC 1.6.99.3) has been studied. It was found that these reactions proceed via a mixed one- and two-electron transfer. The logarithms of the bimolecular rate constants of oxidation (TN/Km) are proportional to the one-electron-reduction potentials of oxidizers. The reactivities of nitrocompounds are close to those of quinones. Unlike the reduction of ferricyanide, these reactions are not inhibited by NADH. However, they are inhibited by NAD+ and ADP-ribose, which also act as the mixed-type inhibitors for ferricyanide. TN/Km of quinones and nitrocompounds depend on the NAD+/NADH ratio, but not on NAD+ concentration. They are diminished by the limiting factors of 2.5-3.5 at NAD+/NADH greater than 200. It seems that rotenone-insensitive reduction of quinones and nitrocompounds takes place near the NAD+/NADH and ferricyanide binding site, and the inhibition is caused by induced conformational changes after the binding of NAD+ or ADP-ribose.  相似文献   

10.
The respiratory complex I couples the electron transfer from NADH to ubiquinone with a translocation of protons across the membrane. Its nucleotide-binding site is made up of a unique Rossmann fold to accommodate the binding of the substrate NADH and of the primary electron acceptor flavin mononucleotide. Binding of NADH includes interactions of the hydroxyl groups of the adenosine ribose with a conserved glutamic acid residue. Structural analysis revealed that due to steric hindrance and electrostatic repulsion, this residue most likely prevents the binding of NADPH, which is a poor substrate of the complex. We produced several variants with mutations at this position exhibiting up to 200-fold enhanced catalytic efficiency with NADPH. The reaction of the variants with NAD(P)H is coupled with proton translocation in an inhibitor-sensitive manner. Thus, we have created an energy-converting NADPH:ubiquinone oxidoreductase, an activity so far not found in nature. Remarkably, the oxidation of NAD(P)H by the variants leads to an enhanced production of reactive oxygen species.  相似文献   

11.
The haem-distal pocket of nitric oxide reductase cytochrome P450 contains many Arg and Lys residues that are clustered to form a putative access channel for NADH. Asp88 is the sole negatively charged amino acid in this positive charge cluster, and thus it would be interesting to know its functional role. Here we found the intriguing phenomenon that mutation at this site of P450nor (D88A or D88V) considerably decreased the overall nitric oxide reductase activity without blocking the reducing half reaction in which the ferric enzyme-NO complex is reduced with NADH to yield a specific intermediate (I). The results indicate that the catalytic turnover subsequent to the I formation was blocked by such mutation. This property of the mutants made it possible to perform kinetic analysis of the reduction step, which is impossible with the wild-type P450nor. These results are the first kinetic evidence for direct complex formation between P450nor and an electron donor (NADH or NADPH). The kinetic analysis also showed that the inhibition by chloride ions (Cl(-)) is competitive with respect to NAD(P)H, which highlights the importance of the binding site for Cl(-) (the anion hole) in the interaction with NAD(P)H. We also characterized another mutant (D393A) of P450nor. The results demonstrated that both Asp residues play important roles in the interaction with NADH, whereas the role of Asp88 is unique in that it must be essential for the release of NAD(+) rather than binding to NADH.  相似文献   

12.
Low concentrations of NAD+ inhibit the NADH: acceptor reductase reactions catalyzed by soluble NADH dehydrogenase from bovine heart mitochondria. The degree of incomplete inhibition of the enzyme depends on the nature and concentration of artificial electron acceptors and is manifested only at low concentrations of the latter. Marked inhibition was demonstrated for the 2.6-dichlorophenolindophenol-, ferricyanide- and O2-reductase reactions, being weakly pronounced during the measurement of the NADH: cytochrome c reductase activity. The inhibition of the above reactions by oxidized NAD+ isn't competitive towards NADH. A kinetic scheme is proposed, which postulates NADH: acceptor reductase reactions occurrence via two mechanisms, namely, a ping-pong mechanism and oxidation of the product-enzyme complex by the acceptor. It was shown that low concentrations of NAD+ also inhibit the NADH oxidase reaction catalyzed by complex I.  相似文献   

13.
V B Lawlis  T E Roche 《Biochemistry》1981,20(9):2519-2524
Micromolar Ca2+ markedly reduces NADH inhibition of bovine kidney alpha-ketoglutarate dehydrogenase complex [Lawlis, V. B., & Roche, T. E. (1980) Mol. Cell. Biochem. 32, 147-152]. Product inhibition patterns from initial velocity studies conducted at less than 10(-9) M or at 1.5 X 10(-5) M Ca2+ with NAD+, CoA, or alpha-ketoglutarate as the variable substrate showed that NADH was a noncompetitive inhibitor with respect to each of these substrates, except at high NAD+ concentrations, where reciprocal plots were nonlinear and the inhibition pattern for NADH vs. NAD+ changed from a noncompetitive to a competitive pattern. From slope and intercept replots, 2-fold to 12-fold higher inhibition constants were estimated for inhibition by NADH vs. the various substrates in the presence of 1.5 X 10(-5) M Ca2+ than for inhibition at less than 10(-9) M Ca2+. These inhibition patterns and the lack of an effect of Ca2+ on the inhibition of the dihydrolipoyl dehydrogenase component suggested that Ca2+-modulated NADH inhibition occurs at an allosteric site with competitive binding at the site by high levels of NAD+. Decarboxylation of alpha-keto[1-14C]glutarate by the resolved alpha-ketoglutarate dehydrogenase component was investigated in the presence of 5.0 mM glyoxylate which served as an efficient acceptor. NADH (0.2 mM) or 1.0 mM ATP inhibited the partial reaction whereas 15 muM Ca2+, 1.0 mM ADP, or 10 mM NAD+ stimulated the partial reaction and reduced NADH inhibition of this reaction. Thus these effectors alter the activity of the alpha-ketoglutarate dehydrogenase complex by binding at allosteric sites on the alpha-ketoglutarate dehydrogenase component. Inhibition by NADH over a wide range of NADH/NAD+ ratios was measured under conditions in which the level of alpha-ketoglutarate was adjusted to give matching control activities at less than 10(-9) M Ca2+ or 1.5 X 10(-5) M Ca2+ in either the presence or the absence of 1.6 mM ADP. These studies establish that both Ca2+ and ADP decreased NADH inhibition under conditions compensating for the effects of Ca2+ and ADP on S0.5 for alpha-ketoglutarate. ADP was particularly effective in reducing NADH inhibition; further studies are required to determine whether this occurs through binding of NADH and ADP at the same, overlapping, or interacting sites.  相似文献   

14.
The binding of porcine heart mitochondrial malate dehydrogenase and beta-hydroxyacyl-CoA dehydrogenase to bovine heart NADH:ubiquinone oxidoreductase (complex I), but not that of bovine heart alpha-ketoglutarate dehydrogenase complex, is virtually abolished by 0.1 mM NADH. The malate dehydrogenase and beta-hydroxyacyl-CoA enzymes compete in part for the same binding site(s) on complex I as do the malate dehydrogenase and alpha-ketoglutarate dehydrogenase complex enzymes. Associations between mitochondrial malate dehydrogenase and bovine serum albumin were observed. Subtle convection artifacts in short-time centrifugation tests of enzyme association with the Beckman Airfuge are described. Substrate channeling of NADH from both the mitochondrial and cytoplasmic malate dehydrogenase isozymes to complex I and reduction of ubiquinone-1 were shown to occur in vitro by transient enzyme-enzyme complex formation. Excess apoenzyme causes little inhibition of the substrate channeling reaction with both malate dehydrogenase isozymes in spite of tighter equilibrium binding than the holoenzyme to complex I. This substrate channeling could, in principle, provide a dynamic microcompartmentation of mitochondrial NADH.  相似文献   

15.
Preincubation of submitochondrial particles (SMP) from beef heart in a reaction mixture containing low concentrations of Mg2+ induces a time lag in the NADH:oxidase activity. Preconditioning of the SMP by NADH, but not by NAD+, prevents the Mg2+-related time lag. The data obtained show that there exists a tight binding site for Mg2+ regulating the rate of electron transfer from NADH to the natural acceptor. The ability of Mg2+ to form a catalytically inactive complex with the enzyme is regulated by NADH.  相似文献   

16.
Mitochondria are considered major regulators of longevity, although their exact role in aging is not fully understood. Data from different laboratories show a negative correlation between reactive oxygen species (ROS) generated by complex I and lifespan. This suggests that complex I has a central role in the regulation of longevity. Here, we review data that both support and refute the role of complex I as a pacemaker of aging. We include data from our laboratory, where we have manipulated ROS production by the electron transport chain (ETC) in Drosophila melanogaster. The bypass of complex I increases the lifespan of the fruit fly, but it is not clear if this is caused by a reduction in ROS or by a change in the NAD+ to NADH ratio. We propose that complex I regulates aging through at least two mechanisms: (1) an ROS-dependent mechanism that leads to mitochondrial DNA damage and (2) an ROS-independent mechanism through the control of the NAD+ to NADH ratio. Control of the relative levels of NAD+ and NADH would allow the regulation of (1) glyco- and (2) lipoxidative-damage and (3) the activation of sirtuins.  相似文献   

17.
Yakovlev G  Hirst J 《Biochemistry》2007,46(49):14250-14258
NADH-ubiquinone oxidoreductase (complex I) is the first enzyme of the respiratory electron transport chain in mitochondria. It conserves the energy from NADH oxidation, coupled to ubiquinone reduction, as a proton motive force across the inner membrane. Complex I catalyzes NADPH oxidation, NAD+ reduction, and hydride transfers from reduced to oxidized nicotinamide nucleotides also. Here, we investigate the transhydrogenation reactions of complex I, using four different nucleotide pairs to encompass a range of reaction rates. Our experimental data are described accurately by a ping-pong mechanism with double substrate inhibition. Thus, we contend that complex I contains only one functional nucleotide binding site, in agreement with recent structural information, but in disagreement with previous mechanistic models which have suggested that two different binding sites are employed to catalyze the two half reactions. We apply the Michaelis-Menten equation to describe the productive states formed when the nucleotide and the active-site flavin mononucleotide have complementary oxidation states, and dissociation constants to describe the nonproductive states formed when they have the same oxidation state. Consequently, we derive kinetic and thermodynamic information about nucleotide binding and interconversion in complex I, relevant to understanding the mechanisms of coupled NADH oxidation and NAD+ reduction, and to understanding how superoxide formation by the reduced flavin is controlled. Finally, we discuss whether NADPH oxidation and/or transhydrogenation by complex I are physiologically relevant processes.  相似文献   

18.
Complex I (NADH:ubiquinone oxidoreductase) is the largest protein complex of bacterial and mitochondrial respiratory chains. The first three-dimensional structure of bacterial complex I in vitrified ice was determined by electron cryo-microscopy and single particle analysis. The structure of the Escherichia coli enzyme incubated with either NAD(+) (as a reference) or NADH was calculated to 35 and 39 A resolution, respectively. The X-ray structure of the peripheral arm of Thermus thermophilus complex I was docked into the reference EM structure. The model obtained indicates that Fe-S cluster N2 is close to the membrane domain interface, allowing for effective electron transfer to membrane-embedded quinone. At the current resolution, the structures in the presence of NAD(+) or NADH are similar. Additionally, side-view class averages were calculated for the negatively stained bovine enzyme. The structures of bovine complex I in the presence of either NAD(+) or NADH also appeared to be similar. These observations indicate that conformational changes upon reduction with NADH, suggested to occur by a range of studies, are smaller than had been thought previously. The model of the entire bacterial complex I could be built from the crystal structures of subcomplexes using the EM envelope described here.  相似文献   

19.
The role of phospholipid in the binding of coenzyme, NAD(H), to 3-hydroxybutyrate dehydrogenase, a lipid-requiring membrane enzyme, has been studied with the ultrafiltration binding method, which we optimized to quantitate weak ligand binding (KD in the range 10-100 microM). 3-Hydroxybutyrate dehydrogenase has a specific requirement of phosphatidylcholine (PC) for optimal function and is a tetramer quantitated both for the apodehydrogenase, which is devoid of phospholipid, and for the enzyme reconstituted into phospholipid vesicles in either the presence or absence of PC. We find that (i) the stoichiometry for NADH and NAD binding is 0.5 mol/mol of enzyme monomer (2 mol/mol of tetramer); (ii) the dissociation constant for NADH binding is essentially the same for the enzyme reconstituted into the mixture of mitochondrial phospholipids (MPL) (KD = 15 +/- 3 microM) or into dioleoyl-PC (KD = 12 +/- 3 microM); (iii) the binding of NAD+ to the enzyme-MPL complex is more than an order of magnitude weaker than NADH binding (KD approximately 200 microM versus 15 microM) but can be enhanced by formation of a ternary complex with either 2-methylmalonate (apparent KD = 1.1 +/- 0.2 microM) or sulfite to form the NAD-SO3- adduct (KD = 0.5 +/- 0.1 microM); (iv) the binding stoichiometry for NADH is the same (0.5 mol/mol) for binary (NADH alone) and ternary complexes (NADH plus monomethyl malonate); (v) binding of NAD+ and NADH together totals 0.5 mol of NAD(H)/mol of enzyme monomer, i.e., two nucleotide binding sites per enzyme tetramer; and (vi) the binding of nucleotide to the enzyme reconstituted with phospholipid devoid of PC is weak, being detected only for the NAD+ plus 2-methylmalonate ternary complex (apparent KD approximately 50 microM or approximately 50-fold weaker binding than that for the same complex in the presence of PC). The binding of NADH by equilibrium dialysis or of spin-labeled analogues of NAD+ by EPR spectroscopy gave complementary results, indicating that the ultrafiltration studies approximated equilibrium conditions. In addition to specific binding of NAD(H) to 3-hydroxybutyrate dehydrogenase, we find significant binding of NAD(H) to phospholipid vesicles. An important new finding is that the nucleotide binding site is present in 3-hydroxybutyrate dehydrogenase in the absence of activating phospholipid since (a) NAD+, as the ternary complex with 2-methylmalonate, binds to the enzyme reconstituted with phospholipid devoid of PC and (b) the apodehydrogenase, devoid of phospholipid, binds NADH or NAD-SO3- weakly (half-maximal binding at approximately 75 microM NAD-SO3- and somewhat weaker binding for NADH).(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

20.
The reduction of both NO2- and hydroxylamine by the NADH-dependent nitrite reductase of Escherichia coli K 12 (EC 1.6.6.4) appears to follow Michaelis-Menten kinetics over a wide range of NADH concentrations. Substrate inhibition can, however, be detected at low concentrations of the product NAD+. In addition, NAD+ displays mixed product inhibition with respect to NADH and mixed or uncompetitive inhibition with respect to hydroxylamine. These inhibition characteristics are consistent with a mechanism in which hydroxylamine binds during catalysis to a different enzyme form from that generated when NAD+ is released. The apparent maximum velocity with NADH as varied substrate increases as the NAD+ concentration increases from 0.05 to 0.7 mM with 1 mM-NO2- or 100 mM-hydroxylamine as oxidized substrate. This increase is more marked for hydroxylamine reduction than for NO2- reduction. Models incorporating only one binding site for NAD can account for the variation in the Michaelis-Menten parameters for both NADH and hydroxylamine with [NAD+] for hydroxylamine reduction. According to these models, activation of the reaction occurs by reversal of an over-reduction of the enzyme by NADH. If the observed activation of the enzyme by NAD+ derives both from activation of the generation of the enzyme-hydroxylamine complex from the enzyme-NO2- complex during NO2- reduction and from activation of the reduction of the enzyme-hydroxylamine complex to form NH4+, then the variation of Vapp. for NO2- or hydroxylamine with [NAD+] is consistent with the occurrence of the same enzyme-hydroxylamine complex as an intermediate in both reactions.  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