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.
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   

2.
The turnover numbers and other kinetic constants for human alcohol dehydrogenase (ADH) 4 ("stomach" isoenzyme) are substantially larger (10-100-fold) than those for human class I and horse liver alcohol dehydrogenases. Comparison of the primary amino acid sequences (69% identity) and tertiary structures of these enzymes led to the suggestion that residue 317, which makes a hydrogen bond with the nicotinamide amide nitrogen of the coenzyme, may account for these differences. Ala-317 in the class I enzymes is substituted with Cys in human ADH4, and locally different conformations of the peptide backbones could affect coenzyme binding. This hypothesis was tested by making the A317C substitution in horse liver ADH1E and comparisons to the wild-type ADH1E. The steady-state kinetic constants for the oxidation of benzyl alcohol and the reduction of benzaldehyde catalyzed by the A317C enzyme were very similar (up to about 2-fold differences) to those for the wild-type enzyme. Transient kinetics showed that the rate constants for binding of NAD(+) and NADH were also similar. Transient reaction data were fitted to the full Ordered Bi Bi mechanism and showed that the rate constants for hydride transfer decreased by about 2.8-fold with the A317C substitution. The structure of A317C ADH1E complexed with NAD(+) and 2,3,4,5,6-pentafluorobenzyl alcohol at 1.2 ? resolution is essentially identical to the structure of the wild-type enzyme, except near residue 317 where the additional sulfhydryl group displaces a water molecule that is present in the wild-type enzyme. ADH is adaptable and can tolerate internal substitutions, but the protein dynamics apparently are affected, as reflected in rates of hydride transfer. The A317C substitution is not solely responsible for the larger kinetic constants in human ADH4; thus, the differences in catalytic activity must arise from one or more of the other hundred substitutions in the enzyme.  相似文献   

3.
The EE and SS isozymes of horse liver alcohol dehydrogenase have been immobilized separately to weakly CNBr-activated Sepharose 4B. The resulting immobilized dimeric preparations lost practically all of their activity after treatment with 6 M urea. However, enzyme activity was regenerated by allowing the urea-treated Sepharose-bound alcohol dehydrogenase to interact specifically with either soluble subunits of dissociated horse liver alcohol dehydrogenase or soluble dimeric enzyme. The regeneration of steroid activity in the immobilized preparations after treatment of the bound S subunits with soluble E subunits seems to show that true reassociation of the enzyme had taken place on the solid phase, since only isozymes with an S-polypeptide chain are active when using 5 beta-dihydrotestosterone as substrate. The results presented in this paper indicate that immobilized single subunits of horse liver alcohol dehydrogenase are inactive and that dimer formation is a prerequisite for the enzymic activity.  相似文献   

4.
Various omega-halogenated carboxy acids and amides were evaluated as potential active-site-directed reagents for alcohol dehydrogenase. 2-Bromoacetamide and bromoacetic and 3-bromopropionic acids inactivated the enzyme; AMP, NAD+, and NADH markedly decreased the rate of inactivation. Some omega-halogenated carboxyamides, X(CH2)nCONH2, increased the activity of the enzyme with the rate and extent of activation depending on the number of methylene units (n) in the order 3 greater than 4 greater than 2 and on X in the order Br greater than Cl. 4-Chlorobutyramide (0.1 M) activated the horse liver enzyme 20-fold in 24 hr at pH 8.0 and 25 degrees. The activation was not prevented by AMP or 2,2-bipyridine, but was by NADH. The kinetic constants and turnover numbers for human and horse liver alcohol dehydrogenases treated with chlorobutyramide were increased markedly compared to those for native enzymes. Alcohol dehydrogenase treated with chlorobutyramide was not further activated by methyl picolinimidate, an imidoester which activates native enzyme by modifying amino groups in the active sites. Chlorobutyramide does not appear to react directly with the enzyme but cyclizes in the reaction medium to form an intermediate imidoester, 2-iminotetrahydrofuran, which reacts with most of the amino groups of the enzyme.  相似文献   

5.
Alcohol dehydrogenase from horse liver was reductively alkylated with aldehydes having varied alkyl substituents. Kinetic studies of alkylated liver alcohol dehydrogenases which were modified in the absence and in the presence of NADH indicate that the alkylation of the specific lysine residues generally activates the enzyme by increasing Michaelis and inhibition constants for substrates and maximum velocities for the reactions. These kinetic parameters were analyzed in terms of electronic, steric, and hydrophobic effects of alkyl substituents. The hydrophilic character of the lysine residues is the most important factor which affects all kinetic parameters, particularly Kia and V2. In addition, the nucleophilic character of the lysine residues enhances the enzyme activity by increasing the maximum velocity of ethanol oxidation and the affinity of alcohol dehydrogenase for NADH and acetaldehyde. The steric interaction at the lysine residues favors the affinity of the enzyme for NADH and ethanol.  相似文献   

6.
The structural framework of cod liver alcohol dehydrogenase is similar to that of horse and human alcohol dehydrogenases. In contrast, the substrate pocket differs significantly, and main differences are located in three loops. Nevertheless, the substrate pocket is hydrophobic like that of the mammalian class I enzymes and has a similar topography in spite of many main-chain and side-chain differences. The structural framework of alcohol dehydrogenase is also present in a number of related enzymes like glucose dehydrogenase and quinone oxidoreductase. These enzymes have completely different substrate specificity, but also for these enzymes, the corresponding loops of the substrate pocket have significantly different structures. The domains of the two subunits in the crystals of the cod enzyme further differ by a rotation of the catalytic domains by about 6 degrees. In one subunit, they close around the coenzyme similarly as in coenzyme complexes of the horse enzyme, but form a more open cleft in the other subunit, similar to the situation in coenzyme-free structures of the horse enzyme. The proton relay system differs from the mammalian class I alcohol dehydrogenases. His 51, which has been implicated in mammalian enzymes to be important for proton transfer from the buried active site to the surface is not present in the cod enzyme. A tyrosine in the corresponding position is turned into the substrate pocket and a water molecule occupies the same position in space as the His side chain, forming a shorter proton relay system.  相似文献   

7.
The kinetic mechanisms of the beta-hydroxybutyrate dehydrogenase from rat heart and liver mitochondria were investigated. Both enzymes, show an Ordered Bi Bi mechanism and there are no major differences in the kinetic constants. In both cases, the solubilized enzyme, re-activated with phosphatidylcholine, shows kinetic properties very similar to those of the enzyme bound to the mitochondrial membrane.  相似文献   

8.
Formaldehyde dehydrogenase (EC 1.2.1.1) is a widely occurring enzyme which catalyzes the oxidation of S-hydroxymethylglutathione, formed from formaldehyde and glutathione, into S-formyglutathione in the presence of NAD. We determined the amino acid sequences for 5 tryptic peptides (containing altogether 57 amino acids) from electrophoretically homogeneous rat liver formaldehyde dehydrogenase and found that they all were exactly homologous to the sequence of rat liver class III alcohol dehydrogenase (ADH-2). Formaldehyde dehydrogenase was found to be able at high pH values to catalyze the NAD-dependent oxidation of long-chain aliphatic alcohols like n-octanol and 12-hydroxydodecanoate but ethanol was used only at very high substrate concentrations and pyrazole was not inhibitory. The amino acid sequence homology and identical structural and kinetic properties indicate that formaldehyde dehydrogenase and the mammalian class III alcohol dehydrogenases are identical enzymes.  相似文献   

9.
Class III alcohol dehydrogenase, whose activity toward ethanol is negligible, has defined, specific properties and is not just a "variant" of the class I protein, the traditional liver enzyme. The primary structure of the horse class III protein has now been determined, and this allows the comparison of alcohol dehydrogenases from human, horse, and rat for both classes III and I, providing identical triads for both these enzyme types. Many consistent differences between the classes separate the two forms as distinct enzymes with characteristic properties. The mammalian class III enzymes are much less variable in structure than the corresponding typical liver enzymes of class I: there are 35 versus 84 positional differences in these identical three-species sets. The class III and class I subunits contain four versus two tryptophan residues, respectively. This makes the differences in absorbance at 280 nm a characteristic property. There are also 4-6 fewer positive charges in the class III enzymes accounting for their electrophoretic differences. The substrate binding site of class III differs from that of class I by replacements at positions that form the hydrophobic barrel typical for this site. In class III, two to four of these positions contain residues with polar or even charged side chains (positions 57 and 93 in all species, plus positions 116 in the horse and 140 in the human and the horse), while corresponding intraclass variation is small. All these structural features correlate with functional characteristics and suggest that the enzyme classes serve different roles. In addition, the replacements between these triad sets illustrate further general properties of the two mammalian alcohol dehydrogenase classes.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

10.
Pyridoxal compounds can either activate or inactivate horse liver alcohol dehydrogenase in differential labeling experiments. Amino groups outside of the active sites were modified with ethyl acetimidate, while the amino groups in the active sites were protected by the formation of the complex with NAD-plus and pyrazole. After removal of the NAD-plus and pyranzole, the partially acetimidylated enzyme was reductively alkylated with pyridoxal and NaBH4, with the incorporation of one pyridoxal group per subunit of the enzyme. The turnover numbers for the reaction of NAD-plus and ethanol increased by 15-fold, and for NADH and acetaldehyde by 32-fold. The Michaelis and inhibition constants increased 80-fold or more. Pyridoxal phosphate and NaBH4 also modified one group per subunit, but the turnover numbers decreased by 10-fold and the kinetic constants were intermediate between those obtained for pyridoxyl alcohol dehydrogenase and the partially acetimidylated enzyme. With native enzyme, the rates of dissociation of the enzyme-coenzyme complexes are rate-limiting in the catalytic reactions. The pyridoxyl enzyme is activated because the rates of dissociation of the enzyme-coenzyme complexes are increased. The rates of binding of coenzyme to phosphopyridoxyl enzyme have decreased due to the introduction of the negatively charged phosphate. The size of the group is not responsible for this decrease since these rates are not greatly decreased by the incorporation of pyridoxal. For both pyrodoxal and phosphopyridoxyl alcohol dehydrogenases, the interconversion of the ternary complex is at least partially rate-limiting. Chymotryptic-tryptic digestion of pryidoxyl enzyme produced a major peptide corresponding to residues 219 to 229, in which Lys 228 had reacted with pyridoxal. The same lysine residue reacted with pyridoxal phosphate.  相似文献   

11.
The mechanism of vesication from sulfur mustard remains unknown in spite of 80 years of investigation. We recently reported sulfur mustard–related inhibition of one or more protein (serine/threonine) phosphatases in tissue cytosol in vitro, suggesting a mechanism common to other vesicants such as cantharidin and Lewisite. Our investigation showed that this inhibition was related to the concentration of 2,2′-thiobis-ethanol (thiodiglycol), the hydrolysis product of sulfur mustard, rather than to the concentration of mustard itself. Related work showed an increase in the rate of NAD (but not NADP) reduction upon the addition of thiodiglycol to mouse liver cytosol. This result provided evidence that metabolism beyond thiodiglycol may be contributing to protein phosphatase inhibition. This observation indicated that metabolism involving one or more dehydrogenases may be necessary to produce the ultimate inhibitor of the protein phosphatases. We report here that thiodiglycol is a substrate for horse liver alcohol dehydrogenase (Km = 3.68 ± 0.45 mM and Vmax = 0.22 ± 0.01 μmol min−1 mg protein−1) and for pyridine nucleotide-linked enzymes in mouse liver and human skin cytosol. The alcohol dehydrogenase-specific inhibitor 4-methylpyrazole inhibited the oxidation of thiodiglycol by the pure horse liver enzyme as well as by the enzymes in human skin and mouse liver cytosol, indicating that the activity in the tissue preparations is also alcohol dehydrogenase. © 1998 John Wiley & Sons, Inc. J Biochem Mol Toxicol 12: 361–369, 1998  相似文献   

12.
Horse liver alcohol dehydrogenase was reacted with glyoxal at different pH values ranging from 6.0 to 9.0. At pH 9.0 the enzyme undergoes a rapid activation over the first minutes of reaction, followed by a decline of activity, which reaches 10% of that of the native enzyme. Chemical analysis of the inactivated enzyme after sodium borohydride reduction shows that 11 argi-ine and 11 lysine residues per mole are modified. At pH 7.7 the enzyme activity increases during the first hour of the reaction with glyoxal and then decreases slowly. Chemical analysis shows that 4 arginine and 3 lysine residues per mole are modified in the enzyme at the maximum of activation. At pH 7.0 the enzyme undergoes a 4-fold activation. Chemical analysis shows that in this activated enzyme 3 lysine and no arginine residues per mole have been modified. Steady-state kinetic analysis suggests that the activated enzyme is not subjected to substrate inhibition and that its Michaelis constant for ethanol is three times larger than that of the native enzyme. The possible role of arginine and lysine residues in the catalytic function of liver alcohol dehydrogenase is discussed.  相似文献   

13.
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   

14.
The transient kinetics of aldehyde reduction by NADH catalyzed by liver alcohol dehydrogenase consist of two kinetic processes. This biphasic rate behavior is consistent with a model in which one of the two identical subunits in the enzyme is inactive during the reaction at the adjacent protomer. Alternatively, enzyme heterogeneity could result in such biphasic behavior. We have prepared liver alcohol dehydrogenase containing a single major isozyme; and the transient kinetics of this purified enzyme are biphasic.Addition of two [14C]carboxymethyl groups per dimer to the two “reactive” sulfhydryl groups (Cys46) yields enzyme which is catalytically inactive toward alcohol oxidation. Alkylated enzyme, as initially isolated by gel filtration chromatography at pH 7·5, forms an NAD+-pyrazole complex. However, the ability to bind NAD+-pyrazole is rapidly lost in pH 8·75 buffer; therefore, our alkylated preparations, as isolated by chromatography at pH 8·75, are inactive toward NAD+-pyrazole complex formation. We have prepared partially inactivated enzyme by allowing iodoacetic acid to react with liver alcohol dehydrogenase until 50% of the NAD+-pyrazole binding capacity remains; under these reaction conditions one [14C]carboxymethyl group is added per dimer. This partially alkylated enzyme preparation is isolated by gel filtration and has been aged sufficiently to lose NAD+-pyrazole binding ability at alkylated subunits. When solutions of native liver alcohol dehydrogenase and partially alkylated liver alcohol dehydrogenase containing the same number of unmodified active sites are allowed to react with substrate under single turnover conditions, partially alkylated enzyme is only half as reactive as native enzyme. This indicates that some molecular species in partially alkylated liver alcohol dehydrogenase that react with pyrazole and NAD+ during the active site titration do not react with substrate. These data are consistent with a model in which a subunit adjacent to an alkylated protomer in the dimeric enzyme is inactive toward substrate. In addition, NAD+-pyrazole binding at the protomers adjacent to alkylated subunits is slowly lost so that 75% of the enzyme-NAD+-pyrazole binding capacity is lost in 50% alkylated enzyme. These data supply strong evidence for subunit interactions in liver alcohol dehydrogenase.Binding experiments performed on partially alkylated liver alcohol dehydrogenase indicate that coenzyme binding is normal at a subunit adjacent to an alkylated protomer even though active ternary complexes cannot be formed. One hypothesis consistent with these results is the unavailability of zinc for substrate binding at the active site in subunits adjacent to alkylated protomers in monoalkylated dimer.  相似文献   

15.
The amino acid sequence of alcohol dehydrogenase of class III from rat liver (the enzyme ADH-2) has been determined. This type of structure is quite different from those of both the class I and the class II alcohol dehydrogenases. The rat class III structure differs from the rat and human class I structures by 133-138 residues (exact value depending on species and isozyme type); and from that of human class II by 132 residues. In contrast, the rat/human species difference within the class III enzymes is only 21 residues. The protein was carboxymethylated with iodo[2(14)C]acetate, and cleaved with CNBr and proteolytic enzymes. Peptides purified by exclusion chromatography and reverse-phase high-performance liquid chromatography were analyzed by degradation with a gas-phase sequencer and with the manual 4-N,N-dimethylaminoazobenzene-4'-isothiocyanate double-coupling method. The protein chain has 373 residues with a blocked N terminus. No evidence was obtained for heterogeneity. The rat ADH-2 enzyme of class III contains an insertion of Cys at position 60 in relation to the class I enzymes, while the latter alcohol dehydrogenase in rat (ADH-3) has another Cys insertion (at position 111) relative to ADH-2. The structure deduced explains the characteristic differences of the class III alcohol dehydrogenase in relation to the other classes of alcohol dehydrogenase, including a high absorbance, an anodic electrophoretic mobility and special kinetic properties. The main amino acid substitutions are found in the catalytic domain and in the subunit interacting segments of the coenzyme-binding domain, the latter explaining the lack of hybrid dimers between subunits of different classes. Several substitutions provide an enlarged and more hydrophilic substrate-binding pocket, which appears compatible with a higher water content in the pocket and hence could possibly explain the higher Km for all substrates as compared with the corresponding values for the class I enzymes. Finally the class III structure supports evolutionary relationships suggesting that the three classes constitute clearly separate enzymes within the group of mammalian zinc-containing alcohol dehydrogenases.  相似文献   

16.
Starting from 6-chloropurine riboside and NAD+, different reactive analogues of NAD+ have been obtained by introducing diazoniumaryl or aromatic imidoester groups via flexible spacers into the nonfunctional adenine moiety of the coenzyme. The analogues react with different amino-acid residues of dehydrogenases and form stable amidine or azobridges, respectively. After the formation of a ternary complex by the coenzyme, the enzyme and a pseudosubstrate, the reactive spacer is anchored in the vicinity of the active site. Thus, the coenzyme remains covalently attached to the protein even after decomposition of the complex. On addition of substrates the covalently bound coenzyme is converted to the dihydro-form. In enzymatic tests the modified dehydrogenases show 80-90% of the specific activity of the native enzymes, but they need remarkably higher concentrations of free NAD+ to achieve these values. The dihydro-coenzymes can be reoxidized by oxidizing agents like phenazine methosulfate or by a second enzyme system. Various systems for coenzyme regeneration were investigated; the modified enzymes were lactate dehydrogenase from pig heart and alcohol dehydrogenase from horse liver; the auxiliary enzymes were alcohol dehydrogenase from yeast and liver, lactate dehydrogenase from pig heart, glutamate dehydrogenase and alanine dehydrogenase. Lactate dehydrogenase from heart muscle is inhibited by pyruvate. With alanine dehydrogenase as the auxiliary enzyme, the coenzyme is regenerated and the reaction product, pyruvate, is removed. This system succeeds to convert lactate quantitatively to L-alanine. The thermostability of the binary enzyme systems indicates an interaction of covalently bound coenzymes with both dehydrogenases; both binding sites seem to compete for the coenzyme. The comparison of dehydrogenases with different degrees of modifications shows that product formation mainly depends on the amount of incorporated coenzyme.  相似文献   

17.
Three-dimensional models of human alcohol dehydrogenase subunits have been constructed, based on the homologous horse enzyme, with computer graphics. All types of class I subunits (alpha, beta, and gamma) and the major allelic variants (beta 1/beta 2 and gamma 1/gamma 2) have been studied. Residue differences between the E-type subunit of the horse enzyme and any of the subunits of the human isozymes occur at 64 positions, about half of which are isozyme-specific. About two thirds of the substitutions are at the surface and all differences can be accommodated in highly conserved three-dimensional structures. The model of the gamma isozyme is most similar to the crystallographically analyzed horse liver E-type alcohol dehydrogenase, and has all the functional residues identical to those of the E subunit except for one which is slightly smaller: Val-141 in the substrate pocket. The residues involved in coenzyme binding are generally conserved between the horse enzyme and the alpha, beta, and gamma types of the human enzyme. In contrast, single exchanges of these residues are the ones involved in the major allelic differences (beta 1 versus beta 2 and gamma 1 versus gamma 2), which affects the overall rate of alcohol oxidation since NADH dissociation is the rate-determining step. Residue 47 is His in beta 2 and Arg in the beta 1, gamma 1, and gamma 2 subunits, and in horse liver alcohol dehydrogenase. Both His and Arg can make a hydrogen bond to a phosphate oxygen atom of NAD; hence the lower turnover rate of beta 1 apparently derives from a charge effect. The substitution to Gly in the alpha subunit results in one less hydrogen bond in NAD binding, and consequently in rapid dissociation. This may explain why the overall rate is an order of magnitude faster than that of beta 1. The important difference between gamma 1 and gamma 2 is an exchange at position 271 from Arg to Gln which can give a hydrogen bond from Gln in gamma 2 to the adenine of NAD. The tighter binding to gamma 2 can account for the slower overall catalytic rate in this isozyme. The kinetics and interactions of cyclohexanol and benzyl alcohol with the isozymes were judged by docking experiments using an interactive fitting program.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

18.
Ethanol oxidation by the soluble fraction of a rat hepatoma was compared to that of the liver. Ethanol oxidation by the hepatoma was NAD+-dependent and sensitive to pyrazole, suggesting the presence of alcohol dehydrogenase. At low concentrations of ethanol (10.8 mm) the alcohol dehydrogenase activities of hepatoma and liver supernatant fractions were comparable. When the concentration of ethanol was raised to 108 mm, the activity of the liver enzyme decreased, whereas the activity in hepatoma supernatant fractions was strikingly elevated. m-Nitrobenzaldehyde-reducing activity was also conspicuously higher in hepatoma supernatant fractions. By contrast the ability to metabolize steroids and cyclohexanone was less than that in supernatant fractions of the liver.Electrophoresis of the liver supernatant fractions on ionagar at pH 7.0 revealed only one component that oxidized ethanol. On the other hand, hepatoma supernatant fractions contained two components with alcohol dehydrogenase activity; one with the same electrophoretic mobility as the liver enzyme, the other showing a slower rate of migration. The latter component, which is absent in the liver, is referred to as hepatoma alcohol dehydrogenase. By electrophoresis on starch gels at pH 8.5, it could be demonstrated that the liver and hepatoma enzymes moved in opposite directions.The liver and hepatoma enzymes differ in electrophoretic mobility, susceptibility to heat treatment, pH activity optimum and some catalytic properties. The substrate specificity of the hepatoma enzyme is narrower than that of liver alcohol dehydrogenase; cyclohexanone or 3β-hydroxysteroids of A/B cis configuration and the corresponding 3-ketones are not substrates for the hepatoma enzyme. The overall substrate specificity characteristics are, however, similar to those of the liver enzyme in that the effectiveness of substrates increases with an increase in chain length and introduction of unsaturation or an aromatic group. Both liver and hepatoma alcohol dehydrogenase cross-react with antibody to horse liver alcohol dehydrogenase EE. The Michaelis constant for ethanol with the hepatoma enzyme is 223 mm, compared to 0.3 mm for liver alcohol dehydrogenase; at 1.0 m ethanol the hepatoma enzyme is not fully saturated with substrate. The Michaelis constant for 2-hexene-1-ol is 0.3 mm, indicating that the hepatoma enzyme is better suited for dehydrogenation of longer chain alcohols. Stomach alcohol dehydrogenase has kinetic properties comparable to those of the hepatoma enzyme, as well as similar electrophoretic mobility. The hepatoma enzyme can be detected in the serum of rats bearing hepatomas.  相似文献   

19.
A three-dimensional model of yeast alcohol dehydrogenase, based on the homologous horse liver enzyme, was used to compare the substrate binding pockets of the three isozymes (I, II, and III) from Saccharomyces cerevisiae and the enzyme from Schizosaccharomyces pombe. Isozyme I and the S. pombe enzyme have methionine at position 294 (numbered as in the liver enzyme, corresponding to 270 in yeast), whereas isozymes II and III have leucine. Otherwise the active sites of the S. cerevisiae enzymes are the same. All four wild-type enzymes were produced from the cloned genes. In addition, oligonucleotide-directed mutagenesis was used to change Met-294 in alcohol dehydrogenase I to leucine. The mechanisms for all five enzymes were predominantly ordered with ethanol (but partially random with butanol) at pH 7.3 and 30 degrees C. The wild-type alcohol dehydrogenases and the leucine mutant had similar kinetic constants, except that isozyme II had 10-20-fold smaller Michaelis and inhibition constants for ethanol. Thus, residue 294 is not responsible for this difference. Apparently, substitutions outside of the substrate binding pocket indirectly affect the interactions of the alcohol dehydrogenases with ethanol. Nevertheless, the substitution of methionine with leucine in the substrate binding site of alcohol dehydrogenase I produced a 7-10-fold increase in reactivity (V/Km) with butanol, pentanol, and hexanol. The higher activity is due to tighter binding of the longer chain alcohols and to more rapid hydrogen transfer.  相似文献   

20.
A comparative study of the kinetic be,avior of horse, sheep, chicken, pig, and ox liver carboxylesterases is reported. The enzymes exhibit similar specificites towards a series of phenyl esters in which the acyl group is varied, and towards a series of butyrate esters in which the alcohol group is varied. Non-Michaelis-Menten kinetics are exhibited by the horse enzyme in the hydrolysis of methyl and ethyl butyrates, and by the pig enzyme with ethyl butyrate. Each enzyme exhibits inhibition by one or more substrates. A simple scheme which accounts for both activation and inhibition is discussed. pH-k(cat) profiles for the horse and chicken liver carboxylesterase-catalyzed hydrolyses of phenyl butyrate demonstrate dependencies on pK(a)S of 4.75 and 5.0, respectively.  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