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 46 毫秒
1.
The rate of ethanol elimination in fed and fasted rats can be predicted based on the liver content of alcohol dehydrogenase (EC 1.1.1.1), the steady-state rate equation, and the concentrations of substrates and products in liver during ethanol metabolism. The specific activity, kinetic constants, and multiplicity of enzyme forms are similar in fed and fasted rats, although the liver content of alcohol dehydrogenase falls 40% with fasting. The two major forms of the enzyme were separated and found to have very similar kinetic properties. The rat alcohol dehydrogenase is subject to substrate inhibition by ethanol at concentrations above 10 mM and follows a Theorell-Chance mechanism. The steady-state rate equation for this mechanism predicts that the in vivo activity of the enzyme is limited by NADH product inhibition at low ethanol concentrations and by both NADH inhibition and substrate inhibition at high ethanol concentrations. When the steady-state rate equation and the measured concentrations of substrates and products in freeze-clamped liver of fed and fasted rats metabolizing alcohol are employed to calculate alcohol oxidation rates, the values agree very well with the actual rates of ethanol elimination determined in vivo.  相似文献   

2.
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   

3.
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   

4.
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   

5.
Activation parameters for each reaction step in the kinetic mechanism of liver alcohol dehydrogenase have been measured for the oxidation of ethanol and the reduction of acetaldehyde. In the oxidation process, the highest enthalpy of activation, 9.7 kcal/mol, occurs for the turnover of the liver alcohol dehydrogenase-NAD(+)-ethanol ternary complex. To investigate if this enthalpy requirement represents a change in the ionization state of ethanol bound in the ternary complex, inhibition of ethanol oxidation was determined using the following series of small, electronegative alcohols with pKa values ranging from 12.37 to 15.5: 2,2,2-trifluoroethanol, 2,2,2-trichloroethanol, 2,2,2-tribromoethanol, 2,2-dichloroethanol, 2,2-difluoroethanol, propargyl alcohol, 3-hydroxypropionitrile, 2-chloroethanol, 2-iodoethanol, 2-methoxyethanol, ethylene glycol, and methanol. The observed inhibition patterns were analyzed according to several kinetic inhibition models; in each case, the best fit model was used to determine the substrate competitive inhibition constant. A plot of the logarithm of these inhibition constants is shown to be dependent on the pKa values of the inhibiting alcohols with a slope approaching -1, indicating that inhibition is controlled by a proton loss from the alcohol. The observed competitive inhibition behavior, coupled with crystallographic studies depicting a direct ligation of an alcohol oxygen to the catalytic zinc ion, indicates that inhibition is controlled by the formation of a zinc-bound alkoxide. Because the inhibiting alcohols are structurally homologous to ethanol, a relationship between the inhibition constant and the inhibiting alcohol's pKa can be derived to show that the pKa of an alcohol bound in a ternary complex is also dependent on its pKa as a free alcohol. Ternary complex pKa values have been determined for ethanol and the inhibiting alcohols.  相似文献   

6.
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   

7.
A new form of alcohol dehydrogenase, designated mu-alcohol dehydrogenase, was identified in surgical human stomach mucosa by isoelectric focusing and kinetic determinations. This enzyme was anodic to class I (alpha, beta, gamma) and class II (pi) alcohol dehydrogenases on agarose isoelectric focusing gels. The partially purified mu-alcohol dehydrogenase, specifically using NAD+ as cofactor, catalyzed the oxidation of aliphatic and aromatic alcohols with long chain alcohols being better substrates, indicating a barrel-shape hydrophobic binding pocket for substrate. mu-Alcohol dehydrogenase stood out in high Km values for both ethanol (18 mM) and NAD+ (340 microM) as well as in high Ki value (320 microM) for 4-methylpyrazole, a competitive inhibitor for ethanol. mu-Alcohol dehydrogenase may account for up to 50% of total stomach alcohol dehydrogenase activity and appeared to play a significant role in first-pass metabolism of ethanol in human.  相似文献   

8.
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   

9.
Previous studies showed that natural human liver alcohol dehydrogenase gamma exhibits negative cooperativity (substrate activation) with ethanol. Studies with the recombinant gamma(2) isoenzyme now confirm that observation and show that the saturation kinetics with other alcohols are also nonhyperbolic, whereas the kinetics for reactions with NAD(+), NADH, and acetaldehyde are hyperbolic. The substrate activation with ethanol and 1-butanol are explained by an ordered mechanism with an abortive enzyme-NADH-alcohol complex that releases NADH more rapidly than does the enzyme-NADH complex. In contrast, high concentrations of cyclohexanol produce noncompetitive substrate inhibition against varied concentrations of NAD(+) and decrease the maximum velocity to 25% of the value that is observed at optimal concentrations of cyclohexanol. Transient kinetics experiments show that cyclohexanol inhibition is due to a slower rate of dissociation of NADH from the abortive enzyme-NADH-cyclohexanol complex than from the enzyme-NADH complex. Fluorescence quenching experiments confirm that the alcohols bind to the enzyme-NADH complex. The nonhyperbolic saturation kinetics for oxidation of ethanol, cyclohexanol, and 1-butanol are quantitatively explained with the abortive complex mechanism. Physiologically relevant concentrations of ethanol would be oxidized predominantly by the abortive complex pathway.  相似文献   

10.
Acetylation, glycosylation, and methylation, which modify lysine residues of horse liver alcohol dehydrogenase, have been investigated. Acetylation reacted with approximately two-third of the total lysines to induce the greatest structural changes of the enzyme. Glycosylation modified only one lysine residue selectively with indiscernible structural changes. The glycosylation effect was very specific with respect to diastereoisomers for aldopentoses, aldohexoses, and ketohexoses. Methylation produced the largest enhancement in the oxidative activity, which is related to the stability of the modified enzyme to prolonged modification and thermal denaturation. Kinetic studies revealed that a change in the maximal velocity was primarily responsible for the observed activity differences in the modifications.  相似文献   

11.
Potato tubers are shown to contain at least 3 alcohol dehydrogenases, one active with NAD and aliphatic alcohols, one active with NADP and terpene alcohols and one active with NADP and aromatic alcohols. The purification of the aliphatic alcohol dehydrogenase is described and its activity with a wide range of substrates is reported. On the basis of substrate specificity, the enzyme is shown to resemble yeast alcohol dehydrogenase rather than liver alcohol dehydrogenase. The enzyme shows high activity with and high affinity for ethanol, activity and affinity decline as the chain length is increased from ethanol to butanol, but a further increase in chain length leads to increased affinity for the alcohol. The physiological significance of the results is briefly discussed.  相似文献   

12.
Alcohol dehydrogenase isozymes from mouse liver (A2 and B2) and stomach (C2) tissues have been purified to homogeneity using triazine-dye affinity chromatography. The enzymes are dimers with similar but distinct subunit sizes, as determined by SDS/polyacrylamide gel electrophoresis: A, 43000; B, 39000, and C, 47000. Zinc analyses and 1,10-phenanthroline inhibition studies indicated that the A and C subunits each contained two atoms of zinc, with at least one being involved catalytically, whereas the B subunit probably contained a single non-catalytic zinc atom. The isozymes exhibited widely divergent kinetic characteristics. A2 exhibited a Km value for ethanol of 0.15 mM and a broad substrate specificity, with Km values decreasing dramatically with an increase in chain length; C2 also exhibited this broad specificity for alcohols but showed a Km value of 232 mM for ethanol. These isozymes also showed broad substrate specificities as aldehyde reductases. In contrast, B2 showed no detectable activity as an aldehyde reductase for the aldehydes examined, and used ethanol as substrate only at very high concentrations (greater than 0.5 M). The isozyme exhibited low Km and high Vmax values, however, with medium-chain alcohols. Immunological studies showed that A2 was immunologically distinct from the B2 and C2 isozymes. In vitro molecular hybridization studies gave no evidence for association between the alcohol dehydrogenase subunits. The results confirm genetic analyses [Holmes, Albanese, Whitehead and Duley (1981) J. Exp. Zool. 215, 151-157] which are consistent with at least three structural genes encoding alcohol dehydrogenase in the mouse and confirm the role of the major liver isozyme (A2) in ethanol metabolism.  相似文献   

13.
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   

14.
Purification and characterization of human liver sorbitol dehydrogenase   总被引:1,自引:0,他引:1  
W Maret  D S Auld 《Biochemistry》1988,27(5):1622-1628
Sorbitol dehydrogenase from human liver was purified to homogeneity by affinity chromatography on immobilized triazine dyes, conventional cation-exchange chromatography, and high-performance liquid chromatography. The major form is a tetrameric, NAD-specific enzyme containing one zinc atom per subunit. Human liver sorbitol dehydrogenase oxidizes neither ethanol nor other primary alcohols. It catalyzes the oxidation of a secondary alcohol group of polyol substrates such as sorbitol, xylitol, or L-threitol. However, the substrate specificity of human liver sorbitol dehydrogenase is broader than that of the liver enzymes of other sources. The present report describes the stereospecific oxidation of (2R,3R)-2,3-butanediol, indicating a more general function of sorbitol dehydrogenase in the metabolism of secondary alcohols. Thus, the enzyme complements the substrate specificities covered by the three classes of human liver alcohol dehydrogenase.  相似文献   

15.
The NAD+ analogue, 3-aldoxime pyridine adenine dinucleotide, is prepared by transglycosidation. Contrary to the published data, this analogue shows no activity as coenzyme with alcohol dehydrogenase from horse liver or from yeast. This is demonstrated by three methods: no increase of absorption at 331 nm by the enzymic oxidation of ethanol; no increase at 290 nm with cinnamic alcohol; and no exchange reaction. The inhibition by this analogue of the oxidation of ethanol by NAD+ is competitive at pH 7.6 and 9.5 with yeast alcohol dehydrogenase; with liver alcohol dehydrogenase, it is of the mixed type at pH 7.6 and non-competitive at pH 9.5. The lack of activity of the analogue and inhibition of the competitive or mixed type may be explained by the fact that the binary complex does not bind the substrate or that in the ternary complex the hydride shift does not occur. The non-competitive inhibition at pH 9.5 with the horse liver alcohol dehydrogenase may be explained by the existence of binding sites specific for this analogue.  相似文献   

16.
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   

17.
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   

18.
Fatty aldehyde dehydrogenase (FALDH) is an NAD+-dependent oxidoreductase involved in the metabolism of fatty alcohols. Enzyme activity has been implicated in the pathology of diabetes and cancer. Mutations in the human gene inactivate the enzyme and cause accumulation of fatty alcohols in Sj?gren-Larsson syndrome, a neurological disorder resulting in physical and mental handicaps. Microsomal FALDH was expressed in E. coli and purified. Using an in vitro activity assay an optimum pH of approximately 9.5 and temperature of approximately 35 degrees C were determined. Medium- and long-chain fatty aldehydes were converted to the corresponding acids and kinetic parameters determined. The enzyme showed high activity with heptanal, tetradecanal, hexadecanal and octadecanal with lower activities for the other tested substrates. The enzyme was also able to convert some fatty alcohol substrates to their corresponding aldehydes and acids, at 25-30% the rate of aldehyde oxidation. A structural model of FALDH has been constructed, and catalytically important residues have been proposed to be involved in alcohol and aldehyde oxidation: Gln-120, Glu-207, Cys-241, Phe-333, Tyr-410 and His-411. These results place FALDH in a central role in the fatty alcohol/acid interconversion cycle, and provide a direct link between enzyme inactivation and disease pathology caused by accumulation of alcohols.  相似文献   

19.
This paper describes the metabolism of fatty alcohols by microsomal and cytosolic fractions from intestinal mucosa. Microsomes of rabbit intestinal mucosa had a high activity of [1-14C]dodecanol oxidation as did those of liver. The intestinal cytosolic fraction also exhibited oxidation activity to a lesser extent than the microsomes did. The reaction product was determined as lauric acid using thin-layer chromatography. Laurylaldehyde was detected as another product, when semicarbazide was added to the incubation system. Cyclodextrins exhibited a stimulation effect similarly to bovine serum albumin on the microsomal activity. We have compared the stimulatory effects of dimethyl-beta-cyclodextrin, beta-cyclodextrin, gamma-cyclodextrin and alpha-cyclodextrin, which decrease in that order. Effects of NAD+ and dodecanol concentrations, pH and pyrazole on microsomal activity were compared with those on cytosolic activity. Dodecanol oxidation activity was solubilized and reconstituted with a fatty alcohol dehydrogenase and a fatty aldehyde dehydrogenase separated from the intestinal microsomes. These findings indicate that both the dehydrogenases participate in microsomal oxidation of fatty alcohols to fatty acids with fatty aldehydes as intermediates in the reaction.  相似文献   

20.
Alcohol dehydrogenase from horse liver is shown to catalyze ester hydrolysis. Nicotinamide coenzymes do not affect the rate of esterolysis. A kinetic approach to study esterase reaction at low substrate to enzyme ratio is described. Kinetic effects of ester structure, temperature, pH, solvent polarity, and ionic strength were investigated. The liver enzyme enhances the rate of esterolysis by lowering activation energy of reaction according to the Uni-Bi kinetic sequence. Two ionizable groups, cysteine and lysine, are tentatively assigned at the esterolytic site of liver alcohol dehydrogenase from pH-rate profiles and chemical modification studies. A plausible mechanism for the esterase reaction proceeds via the acid-assisted nucleophilic catalysis involving the ammonium ion of lysine and the thiolate of cysteine in the acyl-oxygen cleavage.  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