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.
The variation with pH of the kinetic parameters for the alcohol and acetaldehyde reactions were studied for the alleloenzyme AdhS from Drosophila melanogaster. The variation of Ki (KEO,I) with pH for two ethanol-competitive inhibitors, pyrazole and 2,2,2-trifluoroethanol, was also studied. Both alcohol oxidation and acetaldehyde reduction follow a compulsory ordered pathway, with coenzyme binding first. The rate-limiting step for ethanol oxidation is complex and involves at least hydride transfer and dissociation of the enzyme-NADH complex (ER). In contrast with this, the rate-limiting step for the back reaction, i.e. the reduction of acetaldehyde, is dissociation of the enzyme-NAD+ complex (EO). A rate-limiting ER dissociation appears in the oxidation of the secondary alcohol propan-2-ol, whereas for the back reaction, i.e. acetone reduction, hydride transfer in the ternary complexes is rate-limiting. There is one group in the free enzyme, with a pK of approx. 8.0, that regulates the kon velocity for NADH, whereas for NAD+ several groups seem to be involved. A group in the enzyme is drastically perturbed by the formation of the binary EO complex. Protonation of this group with a pK of 7.6 in the EO complex resulted in weakened alcohol and inhibitor binding, in addition to an increased dissociation rate of NAD+ from the binary EO complex. Neither the binding of acetaldehyde nor the dissociation rate of NADH from the binary ER complex varied within the pH region studied.  相似文献   

2.
1. The reaction of pig heart lactate dehydrogenase (EC 1.1.1.27) with NAD(+) and lactate to form pyruvate and NADH was followed by rapid spectrophotometric methods. The distinct spectrum of enzyme-bound NADH permits the measurement of the rate of dissociation of this compound. 2. The reduction of the first mole equivalent of NAD(+) per mole of enzyme sites can also be observed, and is much more rapid than the steady-state rate of NADH production. 3. At pH8 the dissociation of the enzyme-NADH complex is rate-determining for the steady-state oxidation of lactate. At lower pH some other step after the interconversion of the ternary complex and before the dissociation of NADH is rate-determining. Other evidence for a compulsory-order mechanism is provided.  相似文献   

3.
1. The kinetics of oxidation of butan-1-ol and propan-2-ol by NAD+, catalysed by yeast alcohol dehydrogenase, were studied at 25 degrees C from pH 5.5 to 10, and at pH 7.05 from 14 degrees to 44 degrees C, 2. Under all conditions studied the results are consistent with a mechanism whereby some dissociation of coenzyme from the active enzyme-NAD+-alcohol ternary complexes occurs, and the mechanism is therefore not strictly compulsory order. 3. A primary 2H isotopic effect on the maximum rates of oxidation of [1-2H2]butan-1-ol and [2H7]propan-2-ol was found at 25 degrees C over the pH range 5.5-10. Further, in stopped-flow experiments at pH 7.05 and 25 degrees C, there was no transient formation of NADH in the oxidation of butan-1-ol and propan-2-ol. The principal rate-limiting step in the oxidation of dependence on pH of the maximum rates of oxidation of butan-1-ol and propan-2-ol is consisten with the possibility that histidine and cysteine residues may affect or control catalysis.  相似文献   

4.
Stopped-flow studies of oxidation of butan-1-ol and propan-2-ol by NAD(+) in the presence of Phenol Red and large concentrations of yeast alcohol dehydrogenase give no evidence for the participation of a group of pK(a) approx. 7.6 in alcohol binding. Such a group has been implicated in ethanol binding to horse liver alcohol dehydrogenase [Shore, Gutfreund, Brooks, Santiago & Santiago (1974) Biochemistry13, 4185-4190]. The present result supports previous findings based on steady-state kinetic studies with the yeast enzyme. Stopped-flow studies of the yeast alcohol dehydrogenase-catalysed reduction of acetaldehyde by NADH in the presence of ethanol as product inhibitor indicate that the rate-limiting step is NAD(+) release from the enzyme-NAD(+)-ethanol product complex. This finding permits calculation of K(3), the dissociation constant for ethanol from the enzyme-NAD(+)-ethanol complex, by using the product-inhibition data of Dickenson & Dickinson (1978) (Biochem. J.171, 613-627). The calculations show that K(3) varies very little with pH in the range 5.95-8.9, and this agrees with the findings of the stopped-flow experiments described above. Absorption and fluorescence measurements on mixtures of substrates and coenzymes in the presence of high concentrations of alcohol dehydrogenase have been used to estimate values for the ratio [enzyme-NADH-acetaldehyde]/ [enzyme-NAD(+)-ethanol] at equilibrium. The values obtained were in the range 0.11+/-0.04, and this value together with estimates of K(3) was used to provide estimates of values for rate constants and dissociation constants for steps within the catalytic mechanism.  相似文献   

5.
1. The activity of liver alcohol dehydrogenase with cyclohexanol and cyclohexanone as substrates was studied, and the initial-rate parameters were determined from measurements at low substrate concentrations. In contrast with aliphatic ketones, cyclohexanone is a fairly good substrate, although less active than aliphatic aldehydes. The Michaelis constant for cyclohexanol is of the same order as that for ethanol, and the maximum rate and Michaelis constant for NAD(+) obtained with cyclohexanol are very similar to those obtained with primary aliphatic alcohols. The data for this substrate at low concentrations are therefore consistent with a compulsory-order mechanism in which ternary complexes are not rate-limiting. 2. With large concentrations of NAD(+), substrate activation is observed with increasing concentrations of cyclohexanol, whereas with small NAD(+) concentrations substrate inhibition is observed. This complex behaviour is explained by a mechanism previously proposed for this enzyme, which also satisfactorily described the kinetics of oxidation of primary and secondary aliphatic alcohols and aldehydes, including the substrate inhibition exhibited by primary alcohols, and the reduction of aldehydes. The activation with large concentrations of both NAD(+) and cyclohexanol is attributed to the formation of an abortive complex, E.NADH.ROH, from which NADH dissociates more rapidly than from the normal product complex E.NADH. Substrate inhibition in the presence of small NAD(+) concentrations is attributed to the formation of an active complex E.ROH, with which NAD(+) reacts more slowly than with the free enzyme. 3. Some support for these mechanisms of substrate activation and inhibition is obtained by approximate theoretical calculations, and their applicability to other two-substrate reactions that exhibit complex initial-rate behaviour, as a more likely alternative to the postulate of a second binding site for the substrate, is suggested.  相似文献   

6.
An NAD+-dependent alcohol dehydrogenase (ADH) was purified to homogeneity from an aerobic strain of Bacillus stearothermophilus, DSM 2334 (ADH 2334), and compared with the ADH from B. stearothermophilus NCA 1503 (ADH 1503). When an antibody raised against ADH 2334 was used, no cross-reactivity with ADH 1503 was observed on Western blots; by means of an enzyme-linked-immunoabsorbent-assay ('e.l.i.s.a.') procedure, it was found that ADH 1503 had less than 6% of the antigenic activity of ADH 2334. Amino acid analyses detected very small differences in composition, equivalent to about 40 sequence changes, between the two enzymes. The new enzyme has the same six-amino-acid N-terminal sequence as ADH 1503. ADH 2334, but not ADH 1503, is reactive towards methanol; both enzymes can oxidize ethanol, propan-1-ol, butan-1-ol and butan-2-ol. The new enzyme has a distinctive pH optimum at pH 5.5-6 and has significantly lower KEthanolm and kEthanolcat. values than those of ADH 1503. From steady-state kinetic parameters of the reaction with ethanol, propan-1-ol and butan-1-ol, it was shown that ADH 2334 has an ordered mechanism in both directions, with NAD+ being the compulsory first substrate in alcohol oxidation and NADH release being the rate-limiting step. ADH 1503 has an ordered addition of NAD+ and alcohol, but NADH release is not rate-limiting.  相似文献   

7.
1. The activity of liver alcohol dehydrogenase with propan-2-ol and butan-2-ol has been confirmed. The activity with the corresponding ketones is small. Initial-rate parameters are reported for the oxidation of these secondary alcohols, and of propan-1-ol and 2-methylpropan-1-ol, and for the reduction of propionaldehyde and 2-methylpropionaldehyde. Substrate inhibition with primary alcohols is also described. 2. The requirements of the Theorell-Chance mechanism are satisfied by the data for all the primary alcohols and aldehydes, but not by the data for the secondary alcohols. A mechanism that provides for dissociation of either coenzyme or substrate from the reactive ternary complex is described, and shown to account for the initial-rate data for both primary and secondary alcohols, and for isotope-exchange results for the former. With primary alcohols, the rapid rate of reaction of the ternary complex, and its small steady-state concentration, result in conformity of initial-rate data to the requirements of the Theorell-Chance mechanisms. With secondary alcohols, the ternary complex reacts more slowly, its steady-state concentration is greater, and therefore dissociation of coenzyme from it is rate-limiting with non-saturating coenzyme concentrations. 3. Substrate inhibition with large concentrations of primary alcohols is attributed to the formation of an abortive complex of enzyme, NADH and alcohol from which NADH dissociates more slowly than from the enzyme-NADH complex. The initial-rate equation is derived for the complete mechanism, which includes a binary enzyme-alcohol complex and alternative pathways for formation of the reactive ternary complex. This mechanism would also provide, under suitable conditions, for substrate activation or substrate inhibition in a two-substrate reaction, according to the relative rates of reaction through the two pathways.  相似文献   

8.
The rate-limiting step of ethanol oxidation by alcohol dehydrogenase (E) at substrate inhibitory conditions (greater than 500 mM ethanol) is shown to be the dissociation rate of NADH from the abortive E-ethanol-NADH complex. The dissociation rate constant of NADH decreased hyperbolically from 5.2 to 1.4 s-1 in the presence of ethanol causing a decrease in the Kd of NADH binding from 0.3 microM for the binary complex to 0.1 microM for the abortive complex. Correspondingly, ethanol binding to E-NADH (Kd = 37 mM) was tighter than to enzyme (Kd = 109 mM). The binding rate of NAD+ (7 X 10(5) M-1s-1) to enzyme was not affected by the presence of ethanol, further substantiating that substrate inhibition is totally due to a decrease in the dissociation rate constant of NADH from the abortive complex. Substrate inhibition was also observed with the coenzyme analog, APAD+, but a single transient was not found to be rate limiting. Nevertheless, the presence of substrate inhibition with APAD+ is ascribed to a decrease in the dissociation rate of APADH from 120 to 22 s-1 for the abortive complex. Studies to discern the additional limiting transient(s) in turnover with APAD+ and NAD+ were unsuccessful but showed that any isomerization of the enzyme-reduced coenzyme-aldehyde complex is not rate limiting. Chloride increases the rate of ethanol oxidation by hyperbolically increasing the dissociation rate constant of NADH from enzyme and the abortive complex to 12 and 2.8 s-1, respectively. The chloride effect is attributed to the binding of chloride to these complexes, destabilizing the binding of NADH while not affecting the binding of ethanol.  相似文献   

9.
The alcohol dehydrogenase (ADH) from Drosophila lebanonensis shows 82% positional identity to the alcohol dehydrogenases from Drosophila melanogaster. These insect ADHs belong to the short-chain dehydrogenase/reductase family which lack metal ions in their active site. In this family, it appears that the function of zinc in medium chain dehydrogenases has been replaced by three amino acids, Ser138, Tyr151 and Lys155. The present work on D. lebanonensis ADH has been performed in order to obtain information about reaction mechanism, and possible differences in topology and electrostatic properties in the vicinity of the catalytic residues in ADHs from various species of Drosophila. Thus the pH dependence of various kinetic coefficients has been studied. Both in the oxidation of alcohols and in the reduction of aldehydes, the reaction mechanism of D. lebanonensis ADH in the pH 6-10 region was consistent with a compulsory ordered pathway, with the coenzymes as the outer substrates. Over the entire pH region, the rate limiting step for the oxidation of secondary alcohols such as propan-2-ol was the release of the coenzyme product from the enzyme-NADH complex. In the oxidation of ethanol at least two steps were rate limiting, the hydride transfer step and the dissociation of NADH from the binary enzyme-NADH product complex. In the reduction of acetaldehyde, the rate limiting step was the dissociation of NAD+ from the binary enzyme-NAD+ product complex. The pH dependences of the kon velocity curves for the two coenzymes were the opposite of each other, i.e. kon increased for NAD+ and decreased for NADH with increasing pH. The two curves appeared complex and the kon velocity for the two coenzymes seemed to be regulated by several groups in the free enzyme. The kon velocity for ethanol and the ethanol competitive inhibitor pyrazole increased with pH and was regulated through the ionization of a single group in the binary enzyme-NAD+ complex, with a pKa value of 7.1. The kon velocity for acetaldehyde was pH independent and showed that in the enzyme-NADH complex, the pKa value of the catalytic residue must be above 10. The koff velocity of NAD+ appeared to be partly regulated by the catalytic residue, and protonation resulted in an increased dissociation rate. The koff velocity for NADH and the hydride transfer step was pH independent. In D. lebanonensis ADH, the pKa value of the catalytic residue was 0.5 pH units lower than in the ADHS alleloenzyme from D. melanogaster. Thus it can be concluded that while most of the topology of the active site is mainly conserved in these two distantly related enzymes, the microenvironment and electrostatic properties around the catalytic residues differ.  相似文献   

10.
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   

11.
Kinetic studies of formate dehydrogenase   总被引:4,自引:1,他引:3       下载免费PDF全文
1. The kinetic mechanism of formate dehydrogenase is a sequential pathway. 2. The binding of the substrates proceeds in an obligatory order, NAD(+) binding first, followed by formate. 3. It seems most likely that the interconversion of the central ternary complex is extremely rapid, and that the rate-limiting step is the formation or possible isomerization of the enzyme-coenzyme complexes. 4. The secondary plots of the inhibitions with HCO(3) (-) and NO(3) (-) are non-linear, which suggests that more than one molecule of each species is able to bind to the same enzyme form. 5. The rate of the reverse reaction with carbon dioxide at pH6.0 is 20 times that with bicarbonate at pH8.0, although no product inhibition could be detected with carbon dioxide. The low rate of the reverse reaction precluded any steady-state analysis as the enzyme concentrations needed to obtain a measurable rate are of the same order as the K(m) values for NAD(+) and NADH.  相似文献   

12.
Horse-liver alcohol dehydrogenase was carboxymethylated with iodoacetate, which is known to selectively alkylate cysteine-46 in the polypeptide sequence. Carboxymethyl and native enzyme had the same electrophoretic mobility on starch or polyacrylamide gel, but some separation was achieved when isobutyramide and a low concentration of NADH were present (under these conditions NADH was bound by native enzyme but not by Carboxymethyl enzyme).The Carboxymethyl enzyme formed ternary complexes with NAD+ and pyrazole or decanoate. The fluorescence emission of NADH was enhanced 7- to 8-fold (at 410 nm), and a dissociation-constant of 1.7 μM was calculated at pH 7.4; but, in contrast to native enzyme, neither the affinity nor fluorescence were increased by amides (acetamide or isobutyramide).Carboxymethyl alcohol dehydrogenase possesses catalytic activity. Higher alcohols gave maximum velocities up to 7-fold higher than ethanol (reaching nearly 20% of the activity of native enzyme) while [2H]ethanol showed an isotope-rate effect of 3.3. Although the affinity for aldehydes was considerably increased, the maximum velocity of aldehyde-reduction was always at least 20% of that shown by native enzyme, and at pH 9.9 it was almost 2-fold greater than with native enzyme. The rate-limiting step in alcohol-oxidation is likely to be the interconversion of ternary complexes (possibly the hydride-transfer step), while in aldehyde-reduction it could still be the dissociation of the enzyme/NAD+ complex. This is also indicated by inhibition experiments with decanoate, pyrazole, and isobutyramide.These results suggest that a major effect of carboxymethylation is upon ternary complexes of enzyme and NADH, which become much more reluctant to form, either by combination of NADH and ligand with the modified enzyme, or by catalytic conversion of the enzyme/NAD +/alcohol complex.  相似文献   

13.
The kinetic mechanism of betaine aldehyde dehydrogenase from leaves of the plant Amaranthus hypochondriacus is ordered with NAD(+) adding first. NADH is a noncompetitive inhibitor against NAD(+), which was interpreted before as evidence of an iso mechanism, in which NAD(+) and NADH binds to different forms of free enzyme. With the aim of testing the proposed kinetic mechanism, we have now investigated the ability of NADH to form different complexes with the enzyme. By initial velocity and equilibrium binding studies, we found that the steady-state levels of E.glycine betaine are negligible, ruling out binding of NADH to this complex. However, NADH readily bind to E.betaine aldehyde, whose levels most likely are kinetically significant given its low dissociation constant. Also, NADH combined with E.NADH and E.NAD(+). Finally, NADH was not able to revert the hydride transfer step, what suggest that there is no acyl-enzyme intermediate, i.e. the release of the reduced dinucleotide takes place after the deacylation step. Although formation of the complex E.NAD(+).NADH would produce an uncompetitive effect in the inhibition of NADH against NAD(+), the iso mechanism cannot be conclusively discarded.  相似文献   

14.
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   

15.
Pig heart lactate dehydrogenase was studied in the direction of pyruvate and NADH formation by recording rapid changes in extinction, proton concentration, nucleotide fluorescence and protein fluorescence. Experiments measuring extinction changes show that there is a very rapid formation of NADH within the first millisecond and that the amplitude of this phase (phase 1) increases threefold over the pH range 6-8. A second transient rate (phase 2) can also be distinguished (whose rate is pH-dependent), followed by a steady-state rate (phase 3) of NADH production. The sum of the amplitudes of the first two phases corresponds to 1mol of NADH produced/mol of active sites of lactate dehydrogenase. Experiments that measured the liberation of protons by using Phenol Red as an indicator show that no proton release occurs during the initial very rapid formation of NADH (phase 1), but protons are released during subsequent phases of NADH production. Fluorescence experiments help to characterize these phases, and show that the very rapid phase 1 corresponds to the establishment of an equilibrium between E(NAD) (Lactate) right harpoon over left harpoon H(+)E(NADH) (Pyruvate). This equilibrium can be altered by changing lactate concentration or pH, and the H(+)E(NADH) (Pyruvate) species formed has very low nucleotide fluorescence and quenched protein fluorescence. Phase 2 corresponds to the dissociation of pyruvate and a proton from the complex with a rate constant of 1150s(-1). The observed rate constant is slower than this and is proportional to the position of the preceding equilibrium. The E(NADH) formed has high nucleotide fluorescence and quenched protein fluorescence. The reaction, which is rate-limiting during steady-state turnover, must then follow this step and be involved with dissociation of NADH from the enzyme or some conformational change immediately preceding dissociation. Several inhibitory complexes have also been studied including E(NAD+) (Oxamate) and E(NADH) (Oxamate') and the abortive ternary complex E(NADH) (Lactate). The rate of NADH dissociation from the enzyme was measured and found to be the same whether measured by ligand displacement or by relaxation experiments. These results are discussed in relation to the overall mechanism of lactate dehydrogenase turnover and the independence of the four binding sites in the active tetramer.  相似文献   

16.
Previous studies showed that natural human liver alcohol dehydrogenase gamma exhibits negative cooperativity (substrate activation) with ethanol. Studies with the recombinant gamma(2) isoenzyme now confirm that observation and show that the saturation kinetics with other alcohols are also nonhyperbolic, whereas the kinetics for reactions with NAD(+), NADH, and acetaldehyde are hyperbolic. The substrate activation with ethanol and 1-butanol are explained by an ordered mechanism with an abortive enzyme-NADH-alcohol complex that releases NADH more rapidly than does the enzyme-NADH complex. In contrast, high concentrations of cyclohexanol produce noncompetitive substrate inhibition against varied concentrations of NAD(+) and decrease the maximum velocity to 25% of the value that is observed at optimal concentrations of cyclohexanol. Transient kinetics experiments show that cyclohexanol inhibition is due to a slower rate of dissociation of NADH from the abortive enzyme-NADH-cyclohexanol complex than from the enzyme-NADH complex. Fluorescence quenching experiments confirm that the alcohols bind to the enzyme-NADH complex. The nonhyperbolic saturation kinetics for oxidation of ethanol, cyclohexanol, and 1-butanol are quantitatively explained with the abortive complex mechanism. Physiologically relevant concentrations of ethanol would be oxidized predominantly by the abortive complex pathway.  相似文献   

17.
Site-directed mutagenesis of Thr(66) in porcine liver NADH-cytochrome b(5) reductase demonstrated that this residue modulates the semiquinone form of FAD and the rate-limiting step in the catalytic sequence of electron transfer. The absorption spectrum of the T66V mutant showed a typical neutral blue semiquinone intermediate during turnover in the electron transfer from NADH to ferricyanide but showed an anionic red semiquinone form during anaerobic photoreduction. The apparent k(cat) values of this mutant were approximately 10% of that of the wild type enzyme (WT). These data suggest that the T66V mutation stabilizes the neutral blue semiquinone and that the conversion of the neutral blue to the anionic red semiquinone form is the rate-limiting step. In the WT, the value of the rate constant of FAD reduction (k(red)) was consistent with the k(cat) values, and the oxidized enzyme-NADH complex was observed during the turnover with ferricyanide. This indicates that the reduction of FAD by NADH in the WT-NADH complex is the rate-limiting step. In the T66A mutant, the k(red) value was larger than the k(cat) values, but the k(red) value in the presence of NAD(+) was consistent with the k(cat) values. The spectral shape of this mutant observed during turnover was similar to that during the reduction with NADH in the presence of NAD(+). These data suggest that the oxidized T66A-NADH-NAD(+) ternary complex is a major intermediate in the turnover and that the release of NAD(+) from this complex is the rate-limiting step. These results substantiate the important role of Thr(66) in the one-electron transfer reaction catalyzed by this enzyme. On the basis of these data, we present a new kinetic scheme to explain the mechanism of electron transfer from NADH to one-electron acceptors including cytochrome b(5).  相似文献   

18.
Transient-kinetic studies of pig muscle lactate dehydrogenase   总被引:9,自引:8,他引:1  
1. The very fast pre-steady-state formation of NADH catalysed by pig M(4) lactate dehydrogenase was equivalent to the enzyme-site concentration at pH values greater than 8.0 and to one-half the site concentration at pH6.8. 2. The rate of dissociation of NADH from the enzyme at pH8.0 (450s(-1)) in the absence of other substrates is faster than the steady-state oxidation of lactate (80s(-1)). The latter process is therefore controlled by a step before NADH dissociation but subsequent to the hydride transfer. 3. The oxidation of enzyme-NADH by excess of pyruvate was studied as a first-order process at pH9.0. There was no effect of NADD on this reaction and it was concluded that the ternary complex undergoes a rate-limiting change before the hydride-transfer step. 4. Some conclusions about the reactions catalysed by the M(4) isoenzyme were drawn from a comparison of these results with those obtained with the H(4) isoenzyme and liver alcohol dehydrogenase.  相似文献   

19.
l-3-Hydroxyacyl-CoA dehydrogenase reversibly catalyzes the conversion of l-3-hydroxyacyl-CoA to 3-ketoacyl-CoA concomitant with the reduction of NAD(+) to NADH as part of the beta-oxidation spiral. In this report, crystal structures have been solved for the apoenzyme, binary complexes of the enzyme with reduced cofactor or 3-hydroxybutyryl-CoA substrate, and an abortive ternary complex of the enzyme with NAD(+) and acetoacetyl-CoA. The models illustrate positioning of cofactor and substrate within the active site of the enzyme. Comparison of these structures with the previous model of the enzyme-NAD(+) complex reveals that although significant shifting of the NAD(+)-binding domain relative to the C-terminal domain occurs in the ternary and substrate-bound complexes, there are few differences between the apoenzyme and cofactor-bound complexes. Analysis of these models clarifies the role of key amino acids implicated in catalysis and highlights additional critical residues. Furthermore, a novel charge transfer complex has been identified in the course of abortive ternary complex formation, and its characterization provides additional insight into aspects of the catalytic mechanism of l-3-hydroxyacyl-CoA dehydrogenase.  相似文献   

20.
Purified Drosophila lebanonensis alcohol dehydrogenase (Adh) revealed one enzymically active zone in starch gel electrophoresis at pH 8.5. This zone was located on the cathode side of the origin. Incubation of D. lebanonensis Adh with NAD+ and acetone altered the electrophoretic pattern to more anodal migrating zones. D. lebanonensis Adh has an Mr of 56,000, a subunit of Mr of 28 000 and is a dimer with two active sites per enzyme molecule. This agrees with a polypeptide chain of 247 residues. Metal analysis by plasma emission spectroscopy indicated that this insect alcohol dehydrogenase is not a metalloenzyme. In studies of the substrate specificity and stereospecificity, D. lebanonensis Adh was more active with secondary than with primary alcohols. Both alkyl groups in the secondary alcohols interacted hydrophobically with the alcohol binding region of the active site. The catalytic centre activity for propan-2-ol was 7.4 s-1 and the maximum velocity of most secondary alcohols was approximately the same and indicative of rate-limiting enzyme-coenzyme dissociation. For primary alcohols the maximum velocity varied and was much lower than for secondary alcohols. The catalytic centre activity for ethanol was 2.4 s-1. With [2H6]ethanol a primary kinetic 2H isotope effect of 2.8 indicated that the interconversion of the ternary complexes was rate-limiting. Pyrazole was an ethanol-competitive inhibitor of the enzyme. The difference spectra of the enzyme-NAD+-pyrazole complex gave an absorption peak at 305 nm with epsilon 305 14.5 X 10(3) M-1 X cm-1. Concentrations and amounts of active enzyme can thus be determined. A kinetic rate assay to determine the concentration of enzyme active sites is also presented. This has been developed from active site concentrations established by titration at 305 nm of the enzyme and pyrazole with NAD+. In contrast with the amino acid composition, which indicated that D. lebanonensis Adh and the D. melanogaster alleloenzymes were not closely related, the enzymological studies showed that their active sites were similar although differing markedly from those of zinc alcohol dehydrogenases.  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