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.
The kinetics of enzymatic oxidation of ethanol in the presence of alcohol dehydrogenase within a wide range of ethanol and NAD concentrations (pH 6.0--11.5) were studied. It was shown that high concentrations of ethanol (greater than 0.7--5 mM, depending on pH) and NAD (greater than 0.4--0.8 mM) activate alcohol dehydrogenase from horse liver within the pH range of 6.0--7.9. A mechanism of activation based on negative cooperativity of ADH subunits for binding of ethanol and NAD was proposed. The catalytic and Michaelis constants for alcohol dehydrogenase were calculated from ethanol and NAD at all pH values studied. The changes resulting from the subunit cooperativity were revealed. The nature of ionogenic groups of alcohol dehydrogenase, which affect the formation of complexes between the enzyme and NAD and ethanol, and the rate constants for catalytic oxidation of ethanol was assumed. The biological significance of the enzyme capacity for activation by high concentrations of ethanol within the physiological range of pH in the blood under excessive use of alcohol is discussed.  相似文献   

2.
1. Produced inhibition by ethanol of the acetaldehyde-NADH reaction, catalysed by the alcohol dehydrogenases from yeast and horse liver, was studied at 25 degrees C and pH 6-9. 2. The results with yeast alcohol dehydrogenase are generally consistent with the preferred-pathway mechanism proposed previously [Dickenson & Dickinson (1975) Biochem. J. 147, 303-311]. The observed hyperbolic inhibition by ethanol of the maximum rate of acetaldehyde reduction confirms the existence of the alternative pathway involving an enzyme-ethanol complex. 3. The maximum rate of acetaldehyde reduction with horse liver alcohol dehydrogenase is also subject to hyperbolic inhibition by ethanol. 4. The measured inhibition constants for ethanol provide some of the information required in the determination of the dissociation constant for ethanol from the active ternary complex. 5. Product inhibition by acetaldehyde of the ethanol-NAD+ reaction with yeast alcohol dehydrogenase was examined briefly. The results are consistent with the proposed mechanism. However, the nature of the inhibition of the maximum rate cannot be determined within the accessible range of experimental conditions. 6. Inhibition of yeast alcohol dehydrogenase by trifluoroethanol was studied at 25 degrees C and pH 6-10. The inhibition was competitive with respect to ethanol in the ethanol-NAD+ reaction. Estimates were made of the dissociation constant for trifluoroethanol from the enzyme-NAD+-trifluoroethanol complex in the range pH6-10.  相似文献   

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

4.
The in vivo and in vitro activity of alcohol dehydrogenase from baker's yeast maintained under aerobic and anaerobic conditions was measured. In vivo measurements were made in cells "permeabilised" with toluene. Michaelis constants (NAD+ as substrate) were found to be almost identical as those reported for purified preparations. In addition the Km of the enzyme from cells incubated under anaerobic conditions was virtually identical to that from cells from aerobic conditions. The activity of the enzyme was found to be greater (in both "permeabilised" cells and extracts) in cells maintained under nitrogen than air. Cells metabolizing glucose in N2 produced greater levels of ethanol than in air and the rate of NAD+ reduction was also found to be greater in N2 than in air. The results indicate that it was feasible to determine rates of this enzyme in vivo and that the difference in activity of alcohol dehydrogenase under N2 and air may conceivably account for differences in rates of glucose utilisation, ethanol production and NAD+ reduction in air and nitrogen.  相似文献   

5.
K G Bitar  M Cochran  J C Warren 《Steroids》1979,34(2):189-198
The activity of 17 beta-estradiol dehydrogenase (E.C. 1.1.1.62) was measured, and its distribution in the subcellular fractions of bovine placenta was compared. Assay of activity was based on the formation of radioactive estrone from 17 beta[4(-14)C]-estradiol. Either NAD+ or NADP+ can serve as cofactor for the enzyme. The nuclear and microsomal fractions of the placental homogenate exhibited the highest specific enzymatic activities before and after treatment with Triton X-100. Electron micrographs of these two fractions prior to treatment with Triton X-100 showed satisfactory purity. 17 beta-estradiol dehydrogenase from bovine placenta exhibits a pH optimum of about 9.5-10.5, and is activated by 5 x 10(-6)M ZnCl2; comparable concentrations of CaCl2 and MgCl2 inactivate the enzyme. The apparent Michaelis constants, Km, for 17 beta-estradiol and NAD+ are 1.4 x 10(-6)M and 5.5 x 10(-5)M respectively. No 17 alpha-estradiol dehydrogenase activity was demonstrable when using 17 alpha-estradiol as substrate.  相似文献   

6.
Further support for hypotheses proposed earlier for the genetic control and subunit composition of the alcohol dehydrogenase of Triticum has been obtained through the purification and partial characterization of the enzyme. The alcohol dehydrogenase of the wheat T. monococcum was purified 103-fold to a specific activity of 55,900 units/mg. Purification was achieved using streptomycin sulfate precipitation, gel filtration chromatography, DEAE-cellulose anion-exchange chromatography, and preparative isoelectric focusing. The native enzyme has a molecular weight of 116,000 and a dimeric subunit structure. The apparent Michaelis constants are 1.2 × 10?2m for ethanol and 1 × 10?4m for NAD. The substrate specificity of wheat alcohol dehydrogenase differs significantly from the substrate specificities of the enzymes of horse and yeast.  相似文献   

7.
1. The disappearance of pyridine nucleotides during incubation with mosquito homogenates proceeds through the hydrolysis of the pyrophosphate linkage of these compounds as demonstrated by the formation of NMN and AMP from NAD(+). This reaction was also demonstrated by the loss in the coenzyme functioning property of NAD(+) (yeast alcohol dehydrogenase reaction) without a concomitant loss in reactivity towards cyanide. Transglycosidase activity was not observed in the mosquito homogenates, and low concentrations of nicotinamide did not inhibit the NAD(+) splitting activity of these homogenates. These observations are all in accord with the presence in these homogenates of a NAD(+) pyrophosphatase rather than a NADase. 2. The NAD(+) pyrophosphatase is destroyed by boiling, is not heat-activated, and has a pH optimum at pH8.75. In addition to NAD(+), other dinucleotides such as NADP(+), the 3-acetylpyridine and thionicotinamide analogues of NAD(+) and the thionicotinamide analogue of NADP(+), function as substrates in the hydrolysis catalysed by the pyrophosphatase. 3. A decrease in the specific activity of NAD(+) pyrophosphatase was observed during larval development, and a barely detectable activity was found in the pupa and adult. 4. Enzyme activity per organism increased in the larva but decreased to a very low value in the pupa and adult. These results indicate that the decrease in specific activity was due to a decrease in enzyme concentration rather than an increase in amounts of protein.  相似文献   

8.
Rat hearts perfused with oxygenated buffer containing [1-14C]ethanol metabolized small amounts of the ethanol to carbon dioxide. Very sensitive techniques are required to separate the resulting 14CO2 from the ethanol. This metabolism is not inhibited by levels of pyrazole which markedly inhibit NAD dependent liver alcohol dehydrogenase (EC 1.1.1.1). In vitro studies suggest that NADP functions as a cofactor for the rat heart alcohol dehydrogenase activity of crude heart homogenates. The kinetics parameters, the specific activity, and the pH dependence of the enzyme activity measured in these experiments suggest that it may have a minor role in ethanol metabolism by the rat.  相似文献   

9.
An NAD-dependent alcohol dehydrogenase has been purified to apparent homogeneity from cell suspension cultures of Lithospermum erythrorhizon Sieb. et Zucc. (Boraginaceae), using protamine sulphate and ammonium sulphate precipitation and chromatography on DEAE-Sephacel, Superdex 200, hydroxyapatite and HiTrap blue. The enzyme is a homodimer with a Mr of ca. 77,000. Each subunit with a Mr of 40,000 contains two zinc atoms. Its isoelectric point was found at pH 5.0. The best alcohol substrate of the enzyme is ethanol. The pH optimum for ethanol oxidation is at pH 8.7 and for acetaldehyde reduction at pH 4.6. The Michaelis constants for ethanol and NAD are 2.49 and 0.05 (pH 8.7), and for acetaldehyde and NADH 2.2 and 0.078 mM (pH 4.6), respectively. Partial amino acid sequences of the purified enzyme showed high homology to alcohol dehydrogenases from other plants.Abbreviations ADH alcohol dehydrogenase - DTT dithiothreitol - PMSF dephenylmethylsulfonyl fluoride - PVPP polyvinylpolypyrrolidone - IAA indole-3-acetic acid - TFA trifluoroacetic acid  相似文献   

10.
F Fan  J A Lorenzen  B V Plapp 《Biochemistry》1991,30(26):6397-6401
In the three-dimensional structures of enzymes that bind NAD or FAD, there is an acidic residue that interacts with the 2'- and 3'-hydroxyl groups of the adenosine ribose of the coenzyme. The size and charge of the carboxylate might repel the binding of the 2'-phosphate group of NADP and explain the specificity for NAD. In the NAD-dependent alcohol dehydrogenases, Asp-223 (horse liver alcohol dehydrogenase sequence) appears to have this role. The homologous residue in yeast alcohol dehydrogenase I (residue 201 in the protein sequence) was substituted with Gly, and the D223G enzyme was expressed in yeast, purified, and characterized. The wild-type enzyme is specific for NAD. In contrast, the D223G enzyme bound and reduced NAD+ and NADP+ equally well, but, relative to wild-type enzyme, the dissociation constant for NAD+ was increased 17-fold, and the reactivity (V/K) on ethanol was decreased to 1%. Even though catalytic efficiency was reduced, yeast expressing the altered or wild-type enzyme grew at comparable rates, suggesting that equilibration of NAD and NADP pools is not lethal. Asp-223 participates in binding NAD and in excluding NADP, but it is not the only residue important for determining specificity for coenzyme.  相似文献   

11.
The binding to yeast alcohol dehydrogenase of NAD+ and its five derivatives (N6-[2-[N-[2-[N-(2-methacrylamidoethyl)carbamoyl]ethyl] carbamoyl]ethyl]-NAD (I), N6-[N-[2-[N-(2-methacrylamidoethyl) carbamoyl]ethyl]carbamoylmethyl]-NAD (II), copolymer of I with acrylamide (PA-I), copolymer of II with acrylamide (PA-II), and copolymer of I with N,N-dimethylacrylamide (PDMA-I] were studied statically and kinetically by the stopped-flow method by using the quenching of the enzyme fluorescence in the presence of pyrazole. Apparent dissociation constants and apparent rate constants were determined therefrom. It was concluded that (1) the N6-CH2CH2CO group (of I) is effective in making the derivative bind more strongly as well as faster than NAD+, while the N6-CH2CO group (of II) is not; and (2) the binding of the polymer derivatives of NAD+ to the enzyme is not essentially weaker and slower than that of native NAD+, but is even faster in some cases. The coenzymic activities of the above compounds were also determined with yeast alcohol dehydrogenase, pig heart malate dehydrogenase, and rabbit muscle lactate dehydrogenase.  相似文献   

12.
The visible absorption bands in the region 525-575 nm of the catalytic cobalt ion in cobalt(II) horse liver alcohol dehydrogenase show characteristic pH-dependent changes both in the free enzyme and its complexes with nicotinamide adenine dinucleotide (NAD+) and NAD+ plus ethanol or 2,2,2-trifluoroethanol. In the free enzyme, the change of the coordination environment has an apparent pK of about 9.4. In the binary complex with NAD+ the spectral changes are complex, indicating changes in the coordination sphere in a lower pH range with an estimated pK value of about 7.9. The ternary complexes enzyme X NAD+ X ethanol and enzyme X NAD+ X 2,2,2-trifluoroethanol exhibit very similar, characteristic spectral features; their apparent pK values are 6.3 and less than 4, respectively. We ascribe these pK values to the ionization of the alcohol bound in the ternary complexes. The results demonstrate that the catalytic cobalt ion is sensing changes of the ionization state of the protein when going from low pH forms to high pH forms both in the absence and presence of coenzyme and substrate/inhibitor.  相似文献   

13.
The kinetic mechanisms of the 2-oxoglutarate and pyruvate dehydrogenease complexes from pig heart mitochondria were studied at pH 7.5 and 25 degrees. A three-site ping-pong mechanism for the actin of both complexes was proposed on the basis of the parallel lines obtained when 1/v was plotted against 2-oxoglutarate or pyruvate concentration for various levels of CoA and a level of NAD+ near its Michaelis constant value. Rate equations were derived from the proposed mechanism. Michaelis constants for the reactants of the 2-oxoglutarate dehydrogenase complex reaction are: 2-oxoglutarate, 0.220 mM; CoA, 0.025 mM; NAD+, 0.050 mM. Those of the pyruvate dehydrogenase complex are: pyruvate, 0.015 mM; CoA, 0.021 mM; NAD+, 0.079 mM. Product inhibition studies showed that succinyl-CoA or acetyl-CoA was competitive with respect to CoA, and NADH was competitive with respect to NAD+ in both overall reactions, and that succinyl-CoA or acetyl-CoA and NADH were uncompetitive with respect to 2-oxoglutarate or pyruvate, respectively. However, noncompetitive (rather than uncompetitive) inhibition patterns were observed for succinyl-CoA or acetyl-CoA versus NAD+ and for NADH versus CoA. These results are consistent with the proposed mechanisms.  相似文献   

14.
1. The reduction of NAD(+), by an enzyme preparation from rat liver, in the presence of 2-mercaptoethanol is reported. 2. It is suggested that the NAD(+)-linked alcohol dehydrogenase in the extract transfers hydrogen from 2-mercaptoethanol to NAD(+). 3. Both yeast and horse-liver alcohol dehydrogenases were observed to reduce NAD(+) in the presence of 2-mercaptoethanol. 4. Some interactions of 2-mercaptoethanol, cysteine or hydroxylamine with the alcohol dehydrogenases from rat liver, horse liver and yeast are discussed.  相似文献   

15.
An enzymatic method for the measurement of ethanol has been developed to permit analyses with unneutralized acid extracts of blood, liver, cell suspensions, or other biological materials. Components of the assay mixture include NAD, yeast alcohol dehydrogenase, tris(hydroxymethyl)aminomethane (Tris), and lysine. Tris is a trapping agent for the reaction product, acetaldehyde. Lysine is used to maintain the pH at 9.7 where oxidation of ethanol is quantitative and most rapid, even when as much as 0.2 ml of 0.5 n HClO4 is added. Lysine also causes the reaction to be 2 to 4 times faster than it is when either glycine or 2-amino-2-methyl-1-propanol is used as the buffer. The assay is linear up to an ethanol concentration of 0.125 mm in the reaction mixture and is complete by 4 min. By substituting ethanol for NAD in the reagents, the assay performs equally well in measuring NAD.  相似文献   

16.
The activity and the kinetic properties of horse liver alcohol dehydrogenase have been studied in water-in-oil microemulsions containing sodium dodecyl sulfate (SDS) or hexadecyl trimethylammonium bromide (CTAB), 1-butanol or 1-pentanol or 1-hexanol or t-butanol, water and cyclohexane alone or with octane. In the anionic microemulsions (i.e. containing sodium dodecyl sulfate), the enzyme quickly lost its activity, but was efficiently protected by the coenzyme and some adenine nucleotides. In the cationic microemulsions (i.e. containing hexadecyl trimethylammonium bromide), the enzyme activity was more stable and with higher alcohols was stable for at least 20 min. The Michaelis constant of NAD+ calculated with respect to the water content was nearly constant and higher than in water. The maximum velocity in anionic microemulsions depends on the water content whereas in cationic microemulsions, the maximum velocity did not show a clear dependence on the water content and was close to the maximum velocity found in water. The pH dependence of Km and Vmax in these microemulsions was similar to that observed in water. The kinetic data for a hydrophobic substrate, cinnamyl alcohol, showed that this alcohol partitions between the pseudo-phases and thus the apparent Michaelis constant and the concentration at which substrate-excess inhibition appeared were increased. The catalytic properties of the enzyme in microemulsions were illustrated by the preparative reduction of cinnamaldehyde with cofactor recycling. The rate determination of NAD+ reduction and of 1-butanol/cinnamaldehyde redox reaction showed that at low water content (2.8%), the NAD+ reduction rate was close to zero whereas the redox reaction rate was about half of the rate at higher water content. Probably at low water content the coenzyme binding-dissociation rates are reduced much more than the binding-dissociation rates of the substrates and the rates of the ternary complex interconversion. The cationic microemulsions seemed to be very favorable medium for enzyme activity, the tetraalkyl ammonium surfactant causing less denaturation than the anionic detergent dodecyl sulfate.  相似文献   

17.
Alkylation at the N-1 position of the adenine moiety of NAD+, ADP or ATP with 2,3-epoxypropyl acrylate, followed by polymerization with or without acrylamide at pH 8, gave water-soluble polymers of NAD+ and ADP where the alkyl chain was located at the exocyclic adenine C-6 amino group. Cofactor incorporations were good to high: 145-447 mumol NAD+/g polymer and 667 mumol ADP/g polymer. About 30% of the bound NAD+ could be reduced with rabbit muscle lactae dehydrogenase, yeast alcohol dehydrogenase and Bacillus subtilis alanine dehydrogenase; 84% of the bound ADP was phosphorylated with rabbit muscle creatine kinase. High cofactor activities were obtained with polymerized NAD+ with alcohol dehydrogenase as enzyme: the initial rate of NAD+ polymer reduction was 35-81% that of free NAD+. These values remained substantially high with agarose-immobilized alcohol dehydrogenase (15-36%) and should eventually allow their use in continuous enzymatic reactors. Enzymatic phosphorylation of ADP polymer by creatine kinase gave an ATP polymer with high biological activity: 480 mumol ATP/g polymer were transformed with yeast hexokinase.  相似文献   

18.
In a microemulsion made with Triton X-100, the stability of the enzymatic activity was higher than in ionic microemulsions. The stability increased with water content. The kinetic constants (Michaelis constant of NAD+ and maximum velocity) were close to those found in the previously studied microemulsions. The Michaelis constant of NAD+ expressed with respect to the buffer volume was higher than in water. The pH dependence of the kinetic constants in this microemulsion was determined. The activity determined by NAD+ reduction decreased with water content, whereas the redox activity determined via butanol oxidation coupled to retinal reduction was only slightly reduced.  相似文献   

19.
1. The inactivation of horse liver alcohol dehydrogenase by pyridoxal 5'-phosphate in phosphate buffer, pH8, at 10 degrees C was investigated. Activity declines to a minimum value determined by the pyridoxal 5'-phosphate concentration. The maximum inactivation in a single treatment is 75%. This limit appears to be set by the ratio of the first-order rate constants for interconversion of inactive covalently modified enzyme and a readily dissociable non-covalent enzyme-modifier complex. 2. Reactivation was virtually complete on 150-fold dilution: first-order analysis yielded an estimate of the rate constant (0.164min-1), which was then used in the kinetic analysis of the forward inactivation reaction. This provided estimates for the rate constant for conversion of non-covalent complex into inactive enzyme (0.465 min-1) and the dissociation constant of the non-covalent complex (2.8 mM). From the two first-order constants, the minimum attainable activity in a single cycle of treatment may be calculated as 24.5%, very close to the observed value. 3. Successive cycles of modification followed by reduction with NaBH4 each decreased activity by the same fraction, so that three cycles with 3.6 mM-pyridoxal 5'-phosphate decreased specific activity to about 1% of the original value. The absorption spectrum of the enzyme thus treated indicated incorporation of 2-3 mol of pyridoxal 5'-phosphate per mol of subunit, covalently bonded to lysine residues. 4. NAD+ and NADH protected the enzyme completely against inactivation by pyridoxal 5'-phosphate, but ethanol and acetaldehyde were without effect. 5. Pyridoxal 5'-phosphate used as an inhibitor in steady-state experiments, rather than as an inactivator, was non-competitive with respect to both NADH and acetaldehyde. 6. The partially modified enzyme (74% inactive) showed unaltered apparent Km values for NAD+ and ethanol, indicating that modified enzyme is completely inactive, and that the residual activity is due to enzyme that has not been covalently modified. 7. Activation by methylation with formaldehyde was confirmed, but this treatment does not prevent subsequent inactivation with pyridoxal 5'-phosphate. Presumably different lysine residues are involved. 8. It is likely that the essential lysine residue modified by pyridoxal 5'-phosphate is involved either in binding the coenzymes or in the catalytic step. 9. Less detailed studies of yeast alcohol dehydrogenase suggest that this enzyme also possesses an essential lysine residue.  相似文献   

20.
Initial reactions involved in the bacterial degradation of polycyclic aromatic hydrocarbons (PAHs) include a ring-dihydroxylation catalyzed by a dioxygenase and a subsequent oxidation of the dihydrodiol products by a dehydrogenase. In this study, the dihydrodiol dehydrogenase from the PAH-degrading Sphingomonas strain CHY-1 has been characterized. The bphB gene encoding PAH dihydrodiol dehydrogenase (PDDH) was cloned and overexpressed as a His-tagged protein. The recombinant protein was purified as a homotetramer with an apparent Mr of 110,000. PDDH oxidized the cis-dihydrodiols derived from biphenyl and eight polycyclic hydrocarbons, including chrysene, benz[a]anthracene, and benzo[a]pyrene, to corresponding catechols. Remarkably, the enzyme oxidized pyrene 4,5-dihydrodiol, whereas pyrene is not metabolized by strain CHY-1. The PAH catechols produced by PDDH rapidly auto-oxidized in air but were regenerated upon reaction of the o-quinones formed with NADH. Kinetic analyses performed under anoxic conditions revealed that the enzyme efficiently utilized two- to four-ring dihydrodiols, with Km values in the range of 1.4 to 7.1 microM, and exhibited a much higher Michaelis constant for NAD+ (Km of 160 microM). At pH 7.0, the specificity constant ranged from (1.3 +/- 0.1) x 10(6) M(-1) s(-1) with benz[a]anthracene 1,2-dihydrodiol to (20.0 +/- 0.8) x 10(6) M(-1) s(-1) with naphthalene 1,2-dihydrodiol. The catalytic activity of the enzyme was 13-fold higher at pH 9.5. PDDH was subjected to inhibition by NADH and by 3,4-dihydroxyphenanthrene, and the inhibition patterns suggested that the mechanism of the reaction was ordered Bi Bi. The regulation of PDDH activity appears as a means to prevent the accumulation of PAH catechols in bacterial cells.  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