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.
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   

2.
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   

3.
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   

4.
1. The kinetics of oxidation of ethanol, propan-1-ol, butan-1-ol and propan-2-ol by NAD(+) and of reduction of acetaldehyde and butyraldehyde by NADH catalysed by yeast alcohol dehydrogenase were studied. 2. Results for the aldehyde-NADH reactions are consistent with a compulsory-order mechanism with the rate-limiting step being the dissociation of the product enzyme-NAD(+) complex. In contrast the results for the alcohol-NAD(+) reactions indicate that some dissociation of coenzyme from the active enzyme-NAD(+)-alcohol ternary complexes must occur and that the mechanism is not strictly compulsory-order. The rate-limiting step in ethanol oxidation is the dissociation of the product enzyme-NADH complex but with the other alcohols it is probably the catalytic interconversion of ternary complexes. 3. The rate constants describing the combination of NAD(+) and NADH with the enzyme and the dissociations of these coenzymes from binary complexes with the enzyme were measured.  相似文献   

5.
The kinetic properties of highly purified preparations of sheep liver cytoplasmic aldehyde dehydrogenase (preparations that had been shown to be free from contamination with the corresponding mitochondrial enzyme) were investigated with both propionaldehyde and butyraldehyde as substrates. At low aldehyde concentrations, double-reciprocal plots with aldehyde as the variable substrate are linear, and the mechanism appears to be ordered, with NAD+ as the first substrate to bind. Stopped-flow experiments following absorbance and fluorescence changes show bursts of NADH production in the pre-steady state, but the observed course of reaction depends on the pre-mixing conditions. Pre-mixing enzyme with NAD+ activates the enzyme in the pre-steady state and we suggest that the reaction mechanism may involve isomeric enzyme--NAD+ complexes. High concentrations of aldehyde in steady-state experiments produce significant activation (about 3-fold) at high concentrations of NAD+, but inhibition at low concentrations of NAD+. Such behaviour may be explained by postulating the participation of an abortive complex in product release. Stopped-flow measurements at high aldehyde concentrations indicate that the mechanism of reaction under these conditions is complex.  相似文献   

6.
The cDNA for the alpha-isoenzyme from rhesus monkey (Macaca mulatta) liver was cloned and expressed in yeast. The alpha-isoenzymes of human and monkey liver alcohol dehydrogenase differ from the other human and horse liver enzymes in having Met57, Ala93, and Val116 instead of Leu57, Phe93, and Leu116 in the substrate binding pocket and Gly47 instead of Arg47 near the pyrophosphate moiety of the coenzyme. The effects of these differences on the kinetic mechanism, substrate specificity, and coenzyme binding were studied with the purified, recombinant monkey alpha-isoenzyme (MmADH alpha) and mutated enzymes with Gly47 substituted with His or Arg. The mechanism appears to be random for the binding of NAD+ and ethanol and ordered for NADH and acetaldehyde, with formation of a dead-end enzyme-NADH-ethanol complex. MmADH alpha reacts 130-fold slower (V/K) with ethanol and 3-25-fold slower with 2-methyl alcohols but 20-fold faster with cyclohexanol, as compared with horse (Equus caballus) liver EE isoenzyme (EqADH). MmADH alpha is stereoselective for the R isomer of 2-butanol, whereas EqADH favors the S isomer. Both enzymes have comparable reactivity with larger primary alcohols. MmADH alpha is more reactive with secondary alcohols and has highest activity with cyclohexanol. However, it does not react with steroids such as 5 beta-androstane-17 beta-ol-3-one. Molecular modeling suggests that the differences between MmADH alpha and EqADH are a result of the substitution of Ala for Phe93 and Thr for Ser48. MmADH alpha binds NAD+ most rapidly when a group with a pK of 7.4 is unprotonated, implicating His51 in this reaction. The G47R substitution decreased the dissociation constants for NAD+ and NADH and turnover numbers only about 2-fold, whereas the G47H substitution increased dissociation constants 7-14-fold and turnover numbers 4-fold. A basic residue at position 47 is not crucial for activity, as multiple interactions determine coenzyme affinity.  相似文献   

7.
Alcohol dehydrogenase (E. C. 1.1.1.1) from Thermoanaerobium brockii at 25 degrees C and at 65 degrees C is more active with secondary than primary alcohols. The enzyme utilizes NADP and NADPH as cosubstrates better than NAD and NADH. The maximum velocities (V(m)) for secondary alcohols at 65 degrees C are 10 to 100 times higher than those at 25 degrees C, whereas the K(m) values are more comparable.At both 25 degrees C and 65 degrees C the substrate analogue 1,1,1,3,3,3-hexafluoro-2-propanol inhibited the oxidation of alcohol competitively with respect to cyclopentanol, and uncompetitively with respect to NADP. Dimethylsulfoxide inhibited the reduction of cyclopentanone competitively with respect to cyclopentanone, and uncompetitively with respect to NADPH. As a product inhibitor, NADP was competitive with respect to NADPH. These results demonstrate that the enzyme binds the nucleotide and then the alcohol or ketone to form a ternary complex which is converted to a product ternary complex that releases product and nucleotide in that order.At 25 degrees C, all aldehydes and ketones examined inhibited the enzyme at concentrations above their Michaelis constants. The substrate inhibition by cyclopentanone was incomplete, and it was uncompetitive with respect to NADPH. Furthermore, cyclopentanone as a product inhibitor showed intercept-linear, slope-parabolic inhibition with respect to cyclopentanol. These results indicate that cyclopentanone binds to the enzyme-NADP complex at high concentrations. The resulting ternary complex slowly dissociates NADP and cyclopentanone.At 65 degrees C, all of the secondary alcohols, with the exception of cyclohexanol, show substrate activation at high concentration. Experiments in which NADP was the variable substrate and cyclopentanol as the constant-variable substrate over a wide range of concentrations gave double reciprocal plots in which the intercepts showed substrate activation and the slopes showed substrate inhibition. These results indicate that the secondary alcohols bind to the enzyme-NADPH complex at high concentrations and that the resulting ternary complex dissociates NADPH faster than the enzyme-NADPH complex. (c) 1993 John Wiley & Sons, Inc.  相似文献   

8.
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   

9.
In the direction of reductive condensation of alpha-ketoglutarate and lysine, saccharopine dehydrogenase (N6-(glutar-2-yl)-L-lysine:NAD oxidoreductase (lysine-forming) is inhibited by high concentrations of alpha-ketoglutarate and lysine, but not by NADH. NAD+ and saccharopine show no substrate inhibition in the reverse direction. Substrate inhibition by alpha-ketoglutarate and lysine is linear uncompetitive versus NADH. However, when the inhibition is examined with alpha-ketoglutarate or lysine as the variable substrate, the double reciprocal plots show a family of curved lines concave up. The curvature is more pronounced with increasing concentrations of the inhibitory substrate, suggesting an interaction of variable substrate with the enzyme form carrying the inhibitory substrate. These inhibition patterns, the lack of interaction of structural analogs of lysine such as ornithine and norleucine with the E-NAD+ complex (Fujioka M., and Nakatani, Y. (1972) Eur. J. Biochem. 25, 301-307), the identity of values of inhibition constants of alpha-ketoglutarate and lysine obtained with either one as the substrate inhibitor, and the substrate inhibition data in the presence of a reaction product, NAD+, are consistent with the mechanism that substrate inhibition results from the formation of a dead-end E-NAD+-alpha-ketoglutarate complex followed by the addition of lysine to this abortive complex.  相似文献   

10.
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   

11.
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   

12.
Formaldehyde, a major industrial chemical, is classified as a carcinogen because of its high reactivity with DNA. It is inactivated by oxidative metabolism to formate in humans by glutathione-dependent formaldehyde dehydrogenase. This NAD(+)-dependent enzyme belongs to the family of zinc-dependent alcohol dehydrogenases with 40 kDa subunits and is also called ADH3 or chi-ADH. The first step in the reaction involves the nonenzymatic formation of the S-(hydroxymethyl)glutathione adduct from formaldehyde and glutathione. When formaldehyde concentrations exceed that of glutathione, nonoxidizable adducts can be formed in vitro. The S-(hydroxymethyl)glutathione adduct will be predominant in vivo, since circulating glutathione concentrations are reported to be 50 times that of formaldehyde in humans. Initial velocity, product inhibition, dead-end inhibition, and equilibrium binding studies indicate that the catalytic mechanism for oxidation of S-(hydroxymethyl)glutathione and 12-hydroxydodecanoic acid (12-HDDA) with NAD(+) is random bi-bi. Formation of an E.NADH.12-HDDA abortive complex was evident from equilibrium binding studies, but no substrate inhibition was seen with 12-HDDA. 12-Oxododecanoic acid (12-ODDA) exhibited substrate inhibition, which is consistent with a preferred pathway for substrate addition in the reductive reaction and formation of an abortive E.NAD(+).12-ODDA complex. The random mechanism is consistent with the published three-dimensional structure of the formaldehyde dehydrogenase.NAD(+) complex, which exhibits a unique semi-open coenzyme-catalytic domain conformation where substrates can bind or dissociate in any order.  相似文献   

13.
The effects of quercetin and resveratrol (substances found in red wine) on the activity of cytosolic aldehyde dehydrogenase in vitro are compared with those of the synthetic hormone diethylstilbestrol. It is proposed that quercetin inhibits the enzyme by binding competitively in both the aldehyde substrate binding-pocket and the NAD(+)-binding site, whereas resveratrol and diethylstilbestrol can only bind in the aldehyde site. When inhibition is overcome by high aldehyde and NAD(+) concentrations (1 mM of each), the modifiers enhance the activity of the enzyme; we hypothesise that this occurs through binding to the enzyme-NADH complex and consequent acceleration of the rate of dissociation of NADH. The proposed ability of quercetin to bind in both enzyme sites is supported by gel filtration experiments with and without NAD(+), by studies of the esterase activity of the enzyme, and by modelling the quercetin molecule into the known three-dimensional structure of the enzyme. The possibility that interaction between aldehyde dehydrogenase and quercetin may be of physiological significance is discussed.  相似文献   

14.
1. Initial rates of oxidative deamination of L-glutamate with NAD+ as coenzyme, and of reductive aminiation of 2-oxoglutarate with NADH as coenzyme, catalysed by bovine liver glutamate dehydrogenase were measured in 0.111 M-sodium phosphate buffer, pH 7, at 25 degrees C, in the absence and presence of product inhibitors. All 12 possible combinations of variable substrate and product inhibitor were used. 2. Strict competition was observed between NAD+ and NADH, and between glutamate and 2-oxoglutarate. All other inhibition patterns were clearly non-competitive, except for inhibition by NH4+ with NAD+ as variable substrate. Here the extrapolation did not permit a clear distinction between competitive and non-competitive inhibition. 3. Mutually non-competitive behaviour between glutamate and NH4+ indicates that these substrates can be bound at the active site simultaneously. 4. Primary Lineweaver-Burk plots and derived secondary plots of slopes and intercepts against inhibitor concentration were linear, with one exception: with 2-oxoglutarate as variable substrate, the replot of primary intercepts against inhibitory NAD+ concentration was curved. 5. Separate Ki values were evaluated for the effect of each product inhibitor on the individual terms in the reciprocal initial-rate equations. With this information it is possible to calculate rates for any combination of substrate concentrations within the experimental range with any concentration of a single product inhibitor. 6. The inhibition patterns are consistent with neither a simple compulsory-order mechanism nor a rapid-equilibrium random-order mechanism without modification. They can, however, be reconciled with either type of mechanism by postulating appropirate abortive complexes. Of the two compulsory sequences that have been proposed, one, that in which the order of binding is NADH, NH4+, 2-oxoglutarate, requires an implausible pattern of abortive complex-formation to account for the results. 7. On the basis of a rapid-equilibrium random-order mechanism, dissociation constants can be calculated from the Ki values. Where these can be compared with independent estimates from the kinetics of the uninhibited reaction or from direct measurements of substrate binding, the agreement is reasonable good. On balance, therefore, the results provide further support for the rapid-equilibrium random-order mechanism under these conditions.  相似文献   

15.
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   

16.
1. Pig heart lactate dehydrogenase is inhibited by addition of one equivalent of diethyl pyrocarbonate. The inhibition is due to the acylation of a unique histidine residue which is 10-fold more reactive than free histidine. No other amino acid side chains are modified. 2. The carbethoxyhistidine residue slowly decomposes and the enzyme activity reappears. 3. The essential histidine residue is only slightly protected by the presence of NADH but is completely protected when substrate and substrate analogues bind to the enzyme-NADH complex. The protection is interpreted in terms of a model in which substrates can only bind to the enzyme in which the histidine residue is protonated and is thus not available for reaction with the acylating agent. 4. The apparent pK(a) of the histidine residue in the apoenzyme is 6.8+/-0.2. In the enzyme-NADH complex it is 6.7+/-0.2. 5. Acylated enzyme binds NADH with unchanged affinity. The enzyme is inhibited because substrates and substrate analogues cannot bind at the acylated histidine residue in the enzyme-NADH complex.  相似文献   

17.
Intra- and intermolecular electron transfer processes in redox proteins   总被引:2,自引:0,他引:2  
Initial velocity and product inhibition experiments were performed to characterize the kinetic mechanism of branched chain ketoacid dehydrogenase (the branched chain complex) activity. The results were directly compared to predicted patterns for a three-site ping-pong mechanism. Product inhibition experiments confirmed that NADH is competitive versus NAD+ and isovaleryl CoA is competitive versus CoA. Furthermore, both NADH and isovaleryl CoA were uncompetitive versus ketoisovaleric acid. These results are consistent with a ping-pong mechanism and are similar to pyruvate dehydrogenase and alpha-ketoglutarate dehydrogenase. However, inhibition patterns for isovaleryl CoA versus NAD+ and NADH versus CoA are not consistent with a ping-pong mechanism. These patterns may result from a steric interaction between the flavoprotein and transacetylase subunits of the complex. To determine the kinetic mechanism of the substrates and feedback inhibitors (NADH and isovaleryl CoA) of the branched chain complex, it was necessary to define the interaction of the inhibitors at nonsaturating fixed substrate (CoA and NAD+) concentrations. While the competitive inhibition patterns were maintained, slope replots for NADH versus NAD+ at nonsaturating CoA concentrations were parabolic. This unexpected finding resembles a linear mixed type of inhibition where the inhibition is a combination of pure competitive and noncompetitive inhibition.  相似文献   

18.
The effects of adenine nucleotides on initial velocity and NADH binding have been studied with the malate dehydrogenase reaction. ATP, ADP, and AMP were inhibitors competitive with NADH and uncompetitive with oxaloacetate but caused only 50–60% inhibition at saturating concentrations. Direct fluorescence titrations indicated that saturating concentrations of the adenine nucleotides displaced 50–60% of the bound NADH from enzyme-NADH complex. Adenine and adenosine had no inhibitory effect but ADP-ribose caused complete inhibition and NADH dissociation. The possible mechanistic basis for these results and their physiological implications are discussed.  相似文献   

19.
The rate effects of imidazole on the EE isoenzyme of horse liver alcohol dehydrogenase have been analysed in terms of the elucidated kinetic mechanism of the enzyme. These imidazole effects on both directions of the reaction within nonexcess as well as excess ranges of substrate concentrations pointed to the competition between imidazole and ethanol for binding to the same three enzyme species in the kinetic mechanism, namely the free enzyme, the enzyme-NAD+ complex, and the enzyme-NADH complex. Moreover, both imidazole and ethanol brought about an enhancement in the rate of dissociation of NAD+ from its binding site on the enzyme.  相似文献   

20.
Dihydrolipoyl dehydrogenase from bovine kidney catalyzes NAD-linked redox reaction of lipoamide. Hates of the catalyzed reaction were studied in both directions. Saturation curves for NAD and lipoamide are nonhyperbolic, suggesting homotropic cooperative interactions of these substrates with the enzyme. The cooperative effect was analyzed by Hill plots according to the diagnostic procedure of Levitzki and Koshland. Dihydrolipoyl dehydrogenase is subject to homotropic regulation in which NAD acts as a negative cooperative effector, whereas lipoamide acts as a positive cooperative effector. At high concentrations, dihydrolipoamide normalizes the saturation curve of NAD, while NADH tends to enhance the cooperative interaction of lipoamide with the enzyme.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号