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.
R M Gould  B V Plapp 《Biochemistry》1990,29(23):5463-5468
Molecular modeling of alcohol dehydrogenase suggests that His-47 in the yeast enzyme (His-44 in the protein sequence, corresponding to Arg-47 in the horse liver enzyme) binds the pyrophosphate of the NAD coenzyme. His-47 in the Saccharomyces cerevisiae isoenzyme I was substituted with an arginine by a directed mutation. Steady-state kinetic results at pH 7.3 and 30 degrees C of the mutant and wild-type enzymes were consistent with an ordered Bi-Bi mechanism. The substitution decreased dissociation constants by 4-fold for NAD+ and 2-fold for NADH while turnover numbers were decreased by 4-fold for ethanol oxidation and 6-fold for acetaldehyde reduction. The magnitudes of these effects are smaller than those found for the same mutation in the human liver beta enzyme, suggesting that other amino acid residues in the active site modulate the effects of the substitution. The pH dependencies of dissociation constants and other kinetic constants were similar in the two yeast enzymes. Thus, it appears that His-47 is not solely responsible for a pK value near 7 that controls activity and coenzyme binding rates in the wild-type enzyme. The small substrate deuterium isotope effect above pH 7 and the single exponential phase of NADH production during the transient oxidation of ethanol by the Arg-47 enzyme suggest that the mutation makes an isomerization of the enzyme-NAD+ complex limiting for turnover with ethanol.  相似文献   

2.
Arg-47 of human beta 1 beta 1 alcohol dehydrogenase has been replaced with Lys, His, Gln, and Gly by site-directed mutagenesis. The mutated enzymes were expressed in Escherichia coli and purified to homogeneity. The recombinant enzymes with Arg and His at position 47 exhibit kinetic constants and stability which are similar to beta 1 beta 1 and beta 2 beta 2, respectively. The substitution of Lys, His, or Gln for Arg-47 resulted in active enzymes with lower affinity for coenzyme and higher Vmax values than beta 1 beta 1. The substitution of Gln at position 47 resulted in an enzyme with the highest Vmax for ethanol oxidation of any mammalian alcohol dehydrogenase. In this series of enzymes, the affinity for coenzyme decreases with decreasing pKa of the substituted amino acid side chains. The substitution of Gly at position 47 resulted in an enzyme with a Vmax that was one-half that of the low activity beta 1 beta 1 and coenzyme affinities that are lower than beta 1 beta 1, but are equal to or greater than the affinities exhibited by the His-47 or Gln-47 enzymes. Product inhibition studies indicated a change in mechanism from ordered Bi Bi for beta 1 beta 1 to rapid equilibrium random Bi Bi for the Gly-47 enzyme. The kinetic properties of the Gly-47 enzyme are substantially different from human liver alpha alpha which also has Gly at position 47.  相似文献   

3.
J C Burnell  T K Li  W F Bosron 《Biochemistry》1989,28(17):6810-6815
Human liver alcohol dehydrogenase catalyzes the NAD+-dependent oxidation of alcohols. Isoenzymes are produced in liver by five different genes, two of which are polymorphic. We have studied the three beta beta isoenzymes produced at ADH2 because they exhibit very different kinetic properties and they appear with different frequencies in different racial populations. The beta 3 beta 3 isoenzyme which appears in 25% of black Americans was purified to homogeneity, and conditions were found to stabilize this labile isoenzyme. The comparison of substrate specificity among beta beta isoenzymes for primary straight-chain alcohols indicates that there is a positive correlation between Vmax/KM and the log octanol/water partition coefficient for alcohols with beta 2 beta 2 and beta 3 beta 3 but not with beta 1 beta 1. Methyl substitutions at C1 or C2 of these alcohols reduce the catalytic efficiency with all three isoenzymes. The KM and Ki values of beta 3 beta 3 for NAD+ and NADH are substantially higher than values for beta 1 beta 1 or beta 2 beta 2. The Vmax of beta 3 beta 3 ethanol oxidation is 90 times that of beta 1 beta 1. Sequencing of the beta 3 subunit and gene indicates that the polymorphism results from a single amino acid exchange of Cys-369 in beta 3 for Arg-369 in beta 1 and beta 2 [Burnell et al. (1987) Biochem. Biophys. Res. Commun. 146, 1227-1233]. In horse alcohol dehydrogenase and beta 1 beta 1, the guanidino group of Arg-369 is thought to stabilize the NAD(H)-enzyme complex by bonding to one of the pyrophosphate oxygens.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

4.
1. The transient-state kinetics of ligand-displacement reactions have been analyzed. Methods based on this analysis have been used to obtain reliable estimates of on-velocity and off-velocity constants for coenzyme binding to liver alcohol dehydrogenase at different pH values between 6 and 10. 2. The rate of NADH dissociation from the enzyme shows no pronounced dependence on pH. The rate of NAD+ dissociation is controlled by a group with a pKa of 7.6, agreeing with the pKa reported to regulate the binding of certain inhibitory substrate analogues to the enzyme . NAD+ complex. 3. Critical experiments have been performed to test a recent proposal that on-velocity constants for the binding of NADH and NAD+ are controlled by proton equilibria exhibiting different pKa values. The results show that association rates for NADH and NAD+ exhibit the same pH dependence corresponding to a pKa of 9.2. Titrimetric evidence is presented indicating that the latter effect of pH derives from ionization of a group which affects the anion-binding capacity of the coenzyme-binding site.  相似文献   

5.
Soybean nodule xanthine dehydrogenase: a kinetic study   总被引:1,自引:0,他引:1  
Xanthine dehydrogenase was purified from soybean nodules and the kinetic properties were studied at pH 7.5. Km values of 5.0 +/- 0.6 and 12.5 +/- 2.5 microM were obtained for xanthine and NAD+, respectively. The pattern of substrate dependence suggested a Ping-Pong mechanism. Reaction with hypoxanthine gave Km's of 52 +/- 3 and 20 +/- 2.5 microM for hypoxanthine and NAD+, respectively. The Vmax for this reaction was twice that for the xanthine-dependent reaction. The pH dependence of Vmax gave a pKa of 7.6 +/- 0.1 for either xanthine or hypoxanthine oxidation. In addition the Km for xanthine had a pKa of 7.5 consistent with the protonated form of xanthine being the true substrate. Km for hypoxanthine varied only 2.5-fold between pH 6 and 10.7. Product inhibition studies were carried out with urate and NADH. Both products gave mixed inhibition with respect to both substrates. Xanthine dehydrogenase was able to use APAD+ as an electron acceptor for xanthine oxidation, with a Km at pH 7.5 of 21.2 +/- 2.5 microM and Vmax the same as that obtained with NAD+. Reduction of APAD+ by NADH was also catalyzed by xanthine dehydrogenase with a Km of 102 +/- 15 microM; Vmax was approximately 2.5 times that for the xanthine-dependent reaction, and was independent of pH between 6 and 9. Reaction with group-specific reagents indicated the possibility of an essential histidyl group. A thiol-modifying reagent did not cause inactivation of the enzyme. A role for the histidyl side chain in catalysis is proposed.  相似文献   

6.
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   

7.
LeBrun LA  Park DH  Ramaswamy S  Plapp BV 《Biochemistry》2004,43(11):3014-3026
Histidine-51 in horse liver alcohol dehydrogenase (ADH) is part of a hydrogen-bonded system that appears to facilitate deprotonation of the hydroxyl group of water or alcohol ligated to the catalytic zinc. The contribution of His-51 to catalysis was studied by characterizing ADH with His-51 substituted with Gln (H51Q). The steady-state kinetic constants for ethanol oxidation and acetaldehyde reduction at pH 8 are similar for wild-type and H51Q enzymes. In contrast, the H51Q substitution significantly shifts the pH dependencies for steady-state and transient reactions and decreases by 11-fold the rate constant for the transient oxidation of ethanol at pH 8. Modest substrate deuterium isotope effects indicate that hydride transfer only partially limits the transient oxidation and turnover. Transient data show that the H51Q substitution significantly decreases the rate of isomerization of the enzyme-NAD(+) complex and becomes a limiting step for ethanol oxidation. Isomerization of the enzyme-NAD(+) complex is rate limiting for acetaldehyde reduction catalyzed by the wild-type enzyme, but release of alcohol is limiting for the H51Q enzyme. X-ray crystallography of doubly substituted His51Gln:Lys228Arg ADH complexed with NAD(+) and 2,3- or 2,4-difluorobenzyl alcohol shows that Gln-51 isosterically replaces histidine in interactions with the nicotinamide ribose of the coenzyme and that Arg-228 interacts with the adenosine monophosphate of the coenzyme without affecting the protein conformation. The difluorobenzyl alcohols bind in one conformation. His-51 participates in, but is not essential for, proton transfers in the mechanism.  相似文献   

8.
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   

9.
Coenzyme analogues with the adenosine ribose replaced with n-propyl, n-butyl, and n-pentyl groups; coenzyme analogues with the adenosine replaced with 3-(4-acetylanilino)propyl and 6-(4-acetylanilino)hexyl moieties; and nicotinamide mononucleotide, nicotinamide hypoxanthine dinucleotide, and 3-acetylpyridine adenine dinucleotide were used in steady-state kinetic studies with native and activated, amidinated enzymes. The Michaelis and inhibition constants increased up to 100-fold upon modification of coenzyme or enzyme. Turnover numbers with NAD+ and ethanol increased in some cases up to 10-fold due to increased rates of dissociation of enzyme-reduced coenzyme complexes. Rates of dissociation of oxidized coenzyme appeared to be mostly unaffected, but the values calculated (10-60 s-1) were significantly less than the turnover numbers with acetaldehyde and reduced coenzyme (20-900 s-1, at pH 8, 25 degrees C). Rates of association of coenzyme analogues also decreased up to 100-fold. When Lys-228 in the adenosine binding site was picolinimidylated, turnover numbers increased about 10-fold with NAD(H). Furthermore, the pH dependencies for association and dissociation of NAD+ and turnover number with NAD+ and ethanol showed the fastest rates above a pK value of 8.0. Turnover with NADH and acetaldehyde was fastest below a pK value of 8.1. These results can be explained by a mechanism in which isomerization of the enzyme-NAD+ complex (110 s-1) is partially rate limiting in turnover with NAD+ and ethanol (60 s-1) and is controlled by ionization of the hydrogen-bonded system that includes the water ligated to the catalytic zinc and the imidazole group of His-51.  相似文献   

10.
S Ramaswamy  D H Park  B V Plapp 《Biochemistry》1999,38(42):13951-13959
When horse liver alcohol dehydrogenase binds coenzyme, a rotation of about 10 degrees brings the catalytic domain closer to the coenzyme binding domain and closes the active site cleft. The conformational change requires that a flexible loop containing residues 293-298 in the coenzyme binding domain rearranges so that the coenzyme and some amino acid residues from the catalytic domain can be accommodated. The change appears to control the rate of dissociation of the coenzyme and to be necessary for installation of the proton relay system. In this study, directed mutagenesis produced the activated Gly293Ala/Pro295Thr enzyme. X-ray crystallography shows that the conformations of both free and complexed forms of the mutated enzyme and wild-type apoenzyme are very similar. Binding of NAD(+) and 2,2, 2-trifluoroethanol do not cause the conformational change, but the nicotinamide ribose moiety and alcohol are not in a fixed position. Although the Gly293Ala and Pro295Thr substitutions do not disturb the apoenzyme structure, molecular modeling shows that the new side chains cannot be accommodated in the closed native holoenzyme complex without steric alterations. The mutated enzyme may be active in the "open" conformation. The turnover numbers with ethanol and acetaldehyde increase 1.5- and 5.5-fold, respectively, and dissociation constants for coenzymes and other kinetic constants increase 40-2,000-fold compared to those of the native enzyme. Substrate deuterium isotope effects on the steady state V or V/K(m) parameters of 4-6 with ethanol or benzyl alcohol indicate that hydrogen transfer is a major rate-limiting step in catalysis. Steady state oxidation of benzyl alcohol is most rapid above a pK of about 9 for V and V/K(m) and is 2-fold faster in D(2)O than in H(2)O. The results are consistent with hydride transfer from a ground state zinc alkoxide that forms a low-barrier hydrogen bond with the hydroxyl group of Ser48.  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.
The velocity of acetaldehyde metabolism in rat liver may be governed either by the rate of regeneration of NAD from NADH through the electron transport system or by the activity of aldehyde dehydrogenase (ALDH). Measurements of oxygen consumption revealed that the electron transport system was capable of reoxidizing ALDH-generated NADH much faster than it was produced and hence was not rate-limiting for aldehyde metabolism. To confirm that ALDH activity was the rate-limiting factor, low-Km ALDH in slices or intact mitochondria was partially inhibited by treatment with cyanamide and the rate of acetaldehyde metabolism measured. Any inhibition of low-Km ALDH resulted in a decreased rate of acetaldehyde metabolism, indicating that no excess of low-Km ALDH existed. Approximately 40% of the metabolism of 200 microM acetaldehyde in slices was not catalyzed by low-Km ALDH. Fifteen of this 40% was catalyzed by high-Km ALDH. A possible contribution by aldehyde oxidase was ruled out through the use of a competitive inhibitor, quinacrine. Acetaldehyde binding to cytosolic proteins may account for the remainder. By measuring acetaldehyde accumulation during ethanol metabolism, it was also established that low-Km ALDH activity was rate-limiting for acetaldehyde oxidation during concomitant ethanol oxidation.  相似文献   

13.
The structure of mouse class II alcohol dehydrogenase (ADH2) has been determined in a binary complex with the coenzyme NADH and in a ternary complex with both NADH and the inhibitor N-cyclohexylformamide to 2.2 A and 2.1 A resolution, respectively. The ADH2 dimer is asymmetric in the crystal with different orientations of the catalytic domains relative to the coenzyme-binding domains in the two subunits, resulting in a slightly different closure of the active-site cleft. Both conformations are about half way between the open apo structure and the closed holo structure of horse ADH1, thus resembling that of ADH3. The semi-open conformation and structural differences around the active-site cleft contribute to a substantially different substrate-binding pocket architecture as compared to other classes of alcohol dehydrogenase, and provide the structural basis for recognition and selectivity of alcohols and quinones. The active-site cleft is more voluminous than that of ADH1 but not as open and funnel-shaped as that of ADH3. The loop with residues 296-301 from the coenzyme-binding domain is short, thus opening up the pocket towards the coenzyme. On the opposite side, the loop with residues 114-121 stretches out over the inter-domain cleft. A cavity is formed below this loop and adds an appendix to the substrate-binding pocket. Asp301 is positioned at the entrance of the pocket and may control the binding of omega-hydroxy fatty acids, which act as inhibitors rather than substrates. Mouse ADH2 is known as an inefficient ADH with a slow hydrogen-transfer step. By replacing Pro47 with His, the alcohol dehydrogenase activity is restored. Here, the structure of this P47H mutant was determined in complex with NADH to 2.5 A resolution. His47 is suitably positioned to act as a catalytic base in the deprotonation of the substrate. Moreover, in the more closed subunit, the coenzyme is allowed a position closer to the catalytic zinc. This is consistent with hydrogen transfer from an alcoholate intermediate where the Pro/His replacement focuses on the function of the enzyme.  相似文献   

14.
The effect of pH on steady state kinetic parameters for the yeast alcohol dehydrogenase-catalyzed reduction of aldehydes and oxidation of alcohols has been studied. The oxidation of p-CH3 benzyl alcohol-1,1-h2 and -1,1-d2 by NAD+ was found to be characterized by large deuterium isotope effects (kH/kD = 4.1 plus or minus 0.1) between pH 7.5 and 9.5, indicating a rate-limiting hydride trahsfer step in this pH range; a plot of kCAT versus pH could be fit to a theoretical titration curve, pK = 8.25, where kCAT increases with increasing pH. The Michaelis constnat for p-CH3 benzyl alcohol was independent of pH. The reduction of p-CH3 benzaldehyde by NADH and reduced nicotinamide adenine dinucleotide with deuterium in the 4-A position (NADD) cound not be studied below pH 8.5 due to substrate inhibition; however, between pH 8.5 and 9.5, kCAT was found to decrease with increasing pH and to be characterized by significant isotope effects (kH/kD = 3.3 plus or minus 0.3). In the case of acetaldehyde reduction by NADH and NADD, isotope effects were found to be small and exxentially invariant (kH/kD = 2.O plus or minus 0.4) between pH 7.2 and 9.5, suggesting a partially rate-limiting hydride transger step for this substrate; a plot of kCAT/K'b (where K'b is the Michaelis constant for acetaldehyde) versus pH could be fit to a titration curve, pK = 8.25. The titration curve for acetaldehyde reduction has the same pK but is opposite in direction to that observed for p-CH3 benzyl alcohol oxidation. The data presented in this paper indicate a dependence on different enzyme forms for aldehyde reduction and alcohol oxidation and are consistent with a single active site side chain, pK = 8.25, which functions in acid-base catalysis of the hydride transfer step.  相似文献   

15.
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   

16.
Formation and hydrolysis rate constants as well as equilibrium constants of the Schiff base derived from pyridoxal 5'-phosphate and n-hexylamine were determined between pH 3.5 and 7.5 in ethanol/water mixtures (3:17, v/v, and 49:1, v/v). The results indicate that solvent polarity scarcely alters the values of these constants but that they are dependent on the pH. Spectrophotometric titration of this Schiff base was also carried out. We found that a pKa value of 6.1, attributed in high-polarity media to protonation of the pyridine nitrogen atom, is independent of solvent polarity, whereas the pKa of the monoprotonated form of the imine falls from 12.5 in ethanol/water (3:17) to 11.3 in ethanol/water (49:1). Fitting of the experimental results for the hydrolysis to a theoretical model indicates the existence of a group with a pKa value of 6.1 that is crucial in the variation of kinetic constant of hydrolysis with pH. Studies of the reactivity of the coenzyme (pyridoxal 5'-phosphate) of glycogen phosphorylase b with hydroxylamine show that this reaction only occurs when the pH value of solution is below 6.5 and the hydrolysis of imine bond has started. We propose that the decrease in activity of phosphorylase b when the pH value is less than 6.2 must be caused by the cleavage of enzyme-coenzyme binding and that this may be related with protonation of the pyridine nitrogen atom of pyridoxal 5'-phosphate.  相似文献   

17.
M J Rao  A S Acharya 《Biochemistry》1992,31(32):7231-7236
Glu-43(beta) of hemoglobin A exhibits a high degree of chemical reactivity around neutral pH for amidation with nucleophiles in the presence of carbodiimide. Such a reactivity is unusual for the side-chain carboxyl groups of proteins. In addition, the reactivity of Glu-43(beta) is also sensitive to the ligation state of the protein [Rao, M. J., & Acharya, A. S. (1991) J. Protein Chem. 10, 129-138]. The influence of deoxygenation of hemoglobin A on the chemical reactivity of the gamma-carboxyl group of Glu-43(beta) has now been investigated as a function of pH (from 5.5 to 7.5). The chemical reactivity of Glu-43(beta) for amidation increases upon deoxygenation only when the modification reaction is carried out above pH 6.0. The pH-chemical reactivity profile of the amidation of hemoglobin A in the deoxy conformation reflects an apparent pKa of 7.0 for the gamma-carboxyl group of Glu-43(beta). This pKa is considerably higher than the pKa of 6.35 for the oxy conformation. The deoxy conformational transition mediated increase in the pKa of the gamma-carboxyl group of Glu-43(beta) implicates this carboxyl group as an alkaline Bohr group. The amidated derivative of hemoglobin A with 2 mol of glycine ethyl ester covalently bound to the protein was isolated by CM-cellulose chromatography.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

18.
The transient state kinetics of the oxidation of reduced nicotinamide adenine dinucleotide (NADH) by horseradish peroxidase compound I and II (HRP-I and HRP-II) was investigated as a function of pH at 25.0 degrees C in aqueous solutions of ionic strength 0.11 using both a stopped-flow apparatus and a conventional spectrophotometer. In agreement with studies using many other substrates, the pH dependence of the HRP-I-NADH reaction can be explained in terms of a single ionization of pKa = 4.7 +/- 0.5 at the active site of HRP-I. Contrary to studies with other substrates, the pH dependence of the HRP-II-NADH reaction can be interpreted in terms of a single ionization with pKa of 4.2 +/- 1.4 at the active site of HRP-II. An apparent reversibility of the HRP-II-NADH reaction was observed. Over the pH range of 4-10 the rate constant for the reaction of HRP-I with NADH varied from 2.6 X 10(5) to 5.6 X 10(2) M-1 s-1 and of HRP-II with NADH varied from 4.4 X 10(4) to 4.1 M-1 s-1. These rate constants must be taken into consideration to explain quantitatively the oxidase reaction of horseradish peroxidase with NADH.  相似文献   

19.
On the basis of the three-dimensional structure of horse liver alcohol dehydrogenase determined by X-ray crystallography, His 51 has been proposed to act as a general base during catalysis by abstracting a proton from the alcohol substrate. A hydrogen-bonding system (proton relay system) connecting the alcohol substrate and His 51 has been proposed to mediate proton transfer. We have mutated His 51 to Gln in the homologous human liver beta 1 beta 1 alcohol dehydrogenase isoenzyme which is expected to have a similar proton relay system. The mutation resulted in an about 6-fold drop in V/Kb (Vmax for ethanol oxidation divided by Km for ethanol) at pH 7.0 and a 12-fold drop at pH 6.5. V/Kb could be restored completely or partially by the presence of high concentrations of glycylglycine, glycine, and phosphate buffers. A Br?nsted plot of the effect on V/Kb versus the pKa of these bases plus H2O and OH- was linear. Only secondary or tertiary amine buffers differed from linearity, presumably due to steric hindrance. These results suggest that His 51 acts as a general base catalyst during alcohol oxidation in the wild-type enzyme and can be functionally replaced in the mutant enzyme by general base catalysts present in the solvent. Steady-state kinetic constants for NAD+ and the trifluoroethanol inhibition patterns were similar between the wild-type and the mutant enzyme. Differences in the inhibition constants (Ki) of caprate and trifluoroethanol below pH 7.8 and in the pH dependence of Ki can be explained by the substitution of neutral Gln for positively charged His.  相似文献   

20.
In an attempt to trace the source of phosphate activation of the enzyme-catalysed pyruvate-lactate interconversion by rabbit muscle lactate dehydrogenase, equilibrium constants were measured to examine the effects of phosphate on interactions pertinent to the enzymic process. Frontal gel-chromatographic studies of the binding of NADH to the enzyme established that the intrinsic association constant is doubled in the presence of 50 mM-phosphate in the buffer (pH 7.4, I0.15). From kinetic studies of the competition between NAD+ and NADH for the coenzyme-binding sites of the enzyme it is concluded that the binding of oxidized nicotinamide nucleotide is also doubled in the presence of 50 mM-phosphate. Competitive-inhibition studies and fluorescence-quenching measurements indicated the lack of a phosphate effect on ternary-complex formation between enzyme-NADH complex and oxamate, a substrate analogue of pyruvate. The equilibrium constant for the interaction between enzyme-NAD+ complex and oxalate, an analogue of lactate, was also shown, by difference spectroscopy, to be insensitive to phosphate concentration. Provided that the effects observed with the substrate analogues mimic those operative in the kinetic situation, the equilibrium constant governing the isomerization of ternary complex is also independent of phosphate concentration. It is concluded that enhanced coenzyme binding is the source of phosphate activation of the rabbit muscle lactate dehydrogenase system.  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