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.
Kinetic studies were carried out on mitochondrial aldehyde dehydrogenase (EC 1.2.1.3) isolated from sheep liver. Steady-state studies over a wide range of acetaldehyde concentrations gave a non-linear double-reciprocal plot. The dissociation of NADH from the enzyme was a biphasic process with decay constants 0.6s-1 and 0.09s-1. Pre-steady-state kinetic data with propionaldehyde as substrate could be fitted by using the same burst rate constant (12 +/- 3s-1) over a wide range of propionaldehyde concentrations. The quenching of protein fluorescence on the binding of NAD+ to the enzyme was used to estimate apparent rate constants for binding (2 X 10(4) litre.mol-1.s-1) and dissociation (4s-1). The kinetic properties of the mitochondrial enzyme, compared with those reported for the cytoplasmic aldehyde dehydrogenase from sheep liver, show significant differences, which may be important in the oxidation of aldehydes in vivo.  相似文献   

2.
The displacement of NADH from cytoplasmic aldehyde dehydrogenase (EC 1.2.1.3) from sheep liver was studied by using NAD+, 1,10-phenanthroline, ADP-ribose, deamino-NAD+ and pyridine-3-aldehyde-adenine dinucleotide as displacing agents, by following the decrease in fluorescence as a function of time. The data obtained could be fitted by assuming two first-order processes were occurring, a faster process with an apparent rate constant of 0.85 +/- 0.20 s-1 and a relative amplitude of 60 +/- 10% and a slower process with an apparent rate constant of 0.20 +/- 0.05 s-1 and a relative amplitude of 40 +/- 10% (except for pyridine-3-aldehyde-adenine dinucleotide, where the apparent rate constant for the slow process was 0.05 s-1). The displacement rates did not change significantly when the pH was varied from 6.0 to 9.0. Kinetic data are also reported for the dependence of the rate of binding of NADH to the enzyme on the total concentration of NADH. Detailed arguments are presented based on the isolation and purification procedures, the equilibrium coenzyme-binding studies and the kinetic data, which lead to the following model for the release of NADH from the enzyme: (formula: see article). The parameters that best fit the data are: k + 1 = 0.2 s-1; k - 1 = 0.05 s-1; k + 2 = 0.8 s-1 and k - 2 = 5 X 10(5)litre-mol-1-s-1. The slow phase of the NADH release is similar to the steady-state turnover number for substrates such as acetaldehyde and propionaldehyde and appears to contribute significantly to the limitation of the steady-state rate.  相似文献   

3.
Vinyl acetate is subject to microbial degradation in the environment and by pure cultures. It was hydrolyzed by samples of soil, sludge, and sewage at rates of up to 6.38 and 1 mmol/h per g (dry weight) under aerobic and anaerobic conditions, respectively. Four yeasts and thirteen bacteria that feed aerobically on vinyl acetate were isolated. The pathway of vinyl acetate degradation was studied in bacterium V2. Vinyl acetate was degraded to acetate as follows: vinyl acetate + NAD(P)+----2 acetate + NAD(P)H + H+. The acetate was then converted to acetyl coenzyme A and oxidized through the tricarboxylic acid cycle and the glyoxylate bypass. The key enzyme of the pathway is vinyl acetate esterase, which hydrolyzed the ester to acetate and vinyl alcohol. The latter isomerized spontaneously to acetaldehyde and was then converted to acetate. The acetaldehyde was disproportionated into ethanol and acetate. The enzymes involved in the metabolism of vinyl acetate were studied in extracts. Vinyl acetate esterase (Km = 6.13 mM) was also active with indoxyl acetate (Km = 0.98 mM), providing the basis for a convenient spectrophotometric test. Substrates of aldehyde dehydrogenase were formaldehyde, acetaldehyde, propionaldehyde, and butyraldehyde. The enzyme was equally active with NAD+ or NADP+. Alcohol dehydrogenase was active with ethanol (Km = 0.24 mM), 1-propanol (Km = 0.34 mM), and 1-butanol (Km = 0.16 mM) and was linked to NAD+. The molecular sizes of aldehyde dehydrogenase and alcohol dehydrogenase were 145 and 215 kilodaltons, respectively.  相似文献   

4.
M Nieder  B Sunarko    O Meyer 《Applied microbiology》1990,56(10):3023-3028
Vinyl acetate is subject to microbial degradation in the environment and by pure cultures. It was hydrolyzed by samples of soil, sludge, and sewage at rates of up to 6.38 and 1 mmol/h per g (dry weight) under aerobic and anaerobic conditions, respectively. Four yeasts and thirteen bacteria that feed aerobically on vinyl acetate were isolated. The pathway of vinyl acetate degradation was studied in bacterium V2. Vinyl acetate was degraded to acetate as follows: vinyl acetate + NAD(P)+----2 acetate + NAD(P)H + H+. The acetate was then converted to acetyl coenzyme A and oxidized through the tricarboxylic acid cycle and the glyoxylate bypass. The key enzyme of the pathway is vinyl acetate esterase, which hydrolyzed the ester to acetate and vinyl alcohol. The latter isomerized spontaneously to acetaldehyde and was then converted to acetate. The acetaldehyde was disproportionated into ethanol and acetate. The enzymes involved in the metabolism of vinyl acetate were studied in extracts. Vinyl acetate esterase (Km = 6.13 mM) was also active with indoxyl acetate (Km = 0.98 mM), providing the basis for a convenient spectrophotometric test. Substrates of aldehyde dehydrogenase were formaldehyde, acetaldehyde, propionaldehyde, and butyraldehyde. The enzyme was equally active with NAD+ or NADP+. Alcohol dehydrogenase was active with ethanol (Km = 0.24 mM), 1-propanol (Km = 0.34 mM), and 1-butanol (Km = 0.16 mM) and was linked to NAD+. The molecular sizes of aldehyde dehydrogenase and alcohol dehydrogenase were 145 and 215 kilodaltons, respectively.  相似文献   

5.
1. Aldehyde dehydrogenase subcellular distribution studies were performed in a heterogeneous stock (HS) of male and female mice (Mus musculus) with propionaldehyde (5 mM and 50 microM) and formaldehyde (1 mM) and NAD+ or NADP+. 2. The relative percents of distribution were: cytosolic 55-68%, mitochondrial 12-20%, microsomal 9-18% and lysosomal 3-15% for both propionaldehyde concentrations and NAD+. 3. Kinetic experiments using propionaldehyde and acetaldehyde with NAD+ revealed two separate enzymes, Enzyme I (low Km) and Enzyme II (high Km) in the cytosolic and mitochondrial fractions. 4. The kinetic data also indicated a spectrum of cytosolic low Km values that exhibited a bimodal distribution with one congruent to 40 microM and one congruent to 5 microM. 5. It was concluded that there was no significant difference in aldehyde-metabolizing capability between male and female HS mice, compared on a per gram of liver basis. The cytosolic low Km enzyme plays a major role in aldehyde oxidation at moderate to low aldehyde concentrations.  相似文献   

6.
Two isozymes of horse liver aldehyde dehydrogenase (aldehyde, NAD oxidoreductase (EC 1.2.1.3)), F1 and F2, have been purified to homogeneity using salt fractionation followed by ion exchange and gel filtration chromatography. The specific activities of the two isozymes in a pH 9.0 system with propionaldehyde as substrate were approximately 0.35 and 1.0 mumol of NADH/min/mg of protein for the F1 and F2 isozymes, respectively. The multiporosity polyacrylamide gel electrophoresis molecular weights of the F1 and F2 isozymes were approximately 230,000 and 240,000 respectively. Sodium dodecyl sulfate-polyacrylamide gel electrophoresis gave subunit molecular weight estimates of 52,000 and 53,000 for the F1 and F2 isozymes, respectively. The amino acid compositions of the two isozymes were found to be similar; the ionizable amino acid contents being consistent with the electrophoretic and chromatographic behavior of the two isozymes. Both isozymes exhibited a broad aldehyde specificity, oxidizing a wide variety of aliphatic and aromatic aldehydes and utilized NAD as coenzyme, but at approximately 300-fold higher coenzyme concentration could use NADP. The F1 isozyme exhibited a very low Km for NAD (3 muM) and a higher Km for acetaldehyde (70 muM), while the F2 isozyme was found to have a higher Km for NAD (30 muM) and a low Km for acetaldehyde (0.2 muM). The two isozymes showed similar chloral hydrate and p-chloromercuribenzoate inhibition characteristics, but the F1 isozyme was found to be several orders of magnittude more sensitive to disulfiram, a physiological inhibitor of acetaldehyde oxidation. Based on its disulfiram inhibition characteristics, it has been suggested that the F1 isozyme may be the primary enzyme for oxidizing the acetyldehyde produced during ethanol oxidation in vivo.  相似文献   

7.
The free-living protist Euglena gracilis showed an enhanced growth when cultured in the dark with high concentrations of ethanol as carbon source. In a medium containing glutamate/malate plus 1% ethanol, E. gracilis reached a density of 3 x 10(7) cells/ml after 100 h of culture, which was 5 times higher than that attained with glutamate/malate or ethanol separately. This observation suggested the involvement of a highly active aldehyde dehydrogenase in the metabolism of ethanol. Purification of the E. gracilis aldehyde dehydrogenase from the mitochondrial fraction by affinity chromatography yielded an enrichment of 34 times and recovery of 33% of the total mitochondrial activity. SDS-PAGE and molecular exclusion chromatography revealed a native tetrameric protein of 160 kDa. Kinetic analysis showed Km values of 5 and 50 microM for propionaldehyde and NAD(+), respectively, and a Vm value of 1,300 nmol (min x mg protein)(-1). NAD(+) and NADH stimulated the esterase activity of the purified aldehyde dehydrogenase. The present data indicated that the E. gracilis aldehyde dehydrogenase has kinetic and structural properties similar to those of human aldehyde dehydrogenases class 1 and 2.  相似文献   

8.
Yeast alcohol dehydrogenase (EC 1.1.1.1) is able to catalyze the oxidation of acetaldehyde by NAD+ with a concomitant formation of ethanol, at pH 8.8 and pH 7.1; the stoichiometry of aldehyde oxidation vs. ethanol formation is 2:1. This enzymatic reaction obeys the Michaelis-Menten kinetics and was characterized by a high KM for acetaldehyde (68 mM) and a low kcat (2.3 s–1), at pH 8.8, 22°C. There is no visible burst of NADH during the reaction, from pH 7.1–10.1. Therefore, we have concluded that the enzyme catalyzes an apparent dismutation of two molecules of acetaldehyde into a molecule of acetic acid and a molecule of ethanol.  相似文献   

9.
ABSTRACT. An enzyme that oxidizes ethanol to acetaldehyde in the presence of NADP (but not NAD) and reduces acetaldehyde to ethanol in the presence of NADPH (but not NADH) is present in Leishmania donovani promastigotes. The activity is present only in the supernatant fraction obtained from sonication of the cells and high speed centrifugation. The Km and Vm values were evaluated for propanol and propionaldehyde as well as for ethanol and acetaldehyde in cells obtained from late log and 3-day stationary phase cultures. There was no significant change in Km or Vm values for any of these four substrates with culture age. Since the Km values for ethanol and propanol are much higher than for the corresponding aldehydes and higher than any physiological range of alcohol concentration likely to be encountered, this enzyme is considered to function as an aldehyde reductase.  相似文献   

10.
Individuals who carry the most active alcohol dehydrogenase (ADH) isoforms are protected against alcoholism. This work addresses the mechanism by which a high ADH activity leads to low ethanol intake in animals. Male and female ethanol drinker rats (UChB) were allowed access to 10% ethanol for 1 h. Females showed 70% higher hepatic ADH activity and displayed 60% lower voluntary ethanol intake than males. Following ethanol administration (1 g/kg ip), females generated a transient blood acetaldehyde increase ("burst") with levels that were 2.5-fold greater than in males (P < 0.02). Castration of males led to 1) an increased ADH activity (+50%, P < 0.001), 2) the appearance of an acetaldehyde burst (3- to 4-fold vs. sham), and 3) a reduction of voluntary ethanol intake comparable with that of na?ve females. The ADH inhibitor 4-methylpyrazole blocked the appearance of arterial acetaldehyde and increased ethanol intake. Since the release of NADH from the ADH.NADH complex constitutes the rate-limiting step of ADH (but not of ALDH2) activity, endogenous NADH oxidizing substrates present at the time of ethanol intake may contribute to the acetaldehyde burst. Sodium pyruvate given at the time of ethanol administration led to an abrupt acetaldehyde burst and a greatly reduced voluntary ethanol intake. Overall, a transient surge of arterial acetaldehyde occurs upon ethanol administration due to 1) high ADH levels and 2) available metabolites that can oxidize hepatic NADH. The acetaldehyde burst is strongly associated with a marked reduction in ethanol intake.  相似文献   

11.
A study was made of the effect of chronic administration of the hypolipidemic drug clofibrate on the activity and intracellular localization of rat liver aldehyde dehydrogenase. The enzyme was assayed using several aliphatic and aromatic aldehydes. Clofibrate treatment caused a 1.5 to 2.3-fold increase in the liver specific aldehyde dehydrogenase activity. The induced enzyme has a high Km for acetaldehyde and was found to be located in peroxisomes and microsomes. Clofibrate did not alter the enzyme activity in the cytoplasmic fraction. The total peroxisomal aldehyde dehydrogenase activity increased 3 to 4-fold under the action of clofibrate. Disruption of the purified peroxisomes by the hypotonic treatment or in the alkaline conditions resulted in the release of catalase from the broken organelles, while aldehyde dehydrogenase as well as nucleoid-bound urate oxidase and the peroxisomal membrane marker NADH:cytochrome c reductase remained in the peroxisomal 'ghosts'. At the same time, treatment by Triton X-100 led to solubilization of the membrane-bound NADH:cytochrome c reductase and aldehyde dehydrogenase from intact peroxisomes and their 'ghosts'. These results indicate that aldehyde dehydrogenase is located in the peroxisomal membrane. The peroxisomal aldehyde dehydrogenase is active with different aliphatic and aromatic aldehydes, except for formaldehyde and glyceraldehyde. The enzyme Km values lie in the millimolar range for acetaldehyde, propionaldehyde, benzaldehyde and phenylacetaldehyde and in the micromolar range for nonanal. Both NAD and NADP serve as coenzymes for the enzyme. Aldehyde dehydrogenase was inhibited by disulfiram, N-ethylmaleimide and 5,5'-dithiobis(2-nitrobenzoic)acid. According to its basic kinetic properties peroxisomal aldehyde dehydrogenase seems to be similar to a clofibrate-induced microsomal enzyme. The functional role of both enzymes in the liver cells is discussed.  相似文献   

12.
1. Mitochondrial aldehyde dehydrogenase is purified to near homogeneity by hydroxylapatite-, affinity- and hydrophobic interaction-chromatography. 2. The enzyme is an oligomeric protein and its molecular weight, as determined by gel-filtration, is 117,000 +/- 5000. 3. Active only in the presence of exogenous sulfhydryl compounds and NAD(+)-dependent, aldehyde dehydrogenase works optimally with linear-chain aliphatic aldehydes and is practically inactive with benzaldehyde. The pH-optimum is at about pH 8.5. 4. Km-Values for aliphatic aldehydes (C2-C6) range between 0.17 and 0.32 microM. The Km for NAD+ increases from 16 microM with acetaldehyde to 71 microM with capronaldehyde. 5. Millimolar concentrations of Mg2+ promote high increases of both V and Km for NAD+. At the same time, saturation curves with C4-C6 aldehydes can be simulated with a substrate inhibition model. 6. Inhibition by NADH is competitive: with capronaldehyde, the inhibition constant for NADH is 52 microM in the absence of Mg2+ and 14 microM in the presence of 4 mM Mg2+; with acetaldehyde, the inhibition constant is about three times higher (36 and 159 microM, respectively).  相似文献   

13.
Nidetzky B  Klimacek M  Mayr P 《Biochemistry》2001,40(34):10371-10381
Microbial xylose reductase, a representative aldo-keto reductase of primary sugar metabolism, catalyzes the NAD(P)H-dependent reduction of D-xylose with a turnover number approximately 100 times that of human aldose reductase for the same reaction. To determine the mechanistic basis for that physiologically relevant difference and pinpoint features that are unique to the microbial enzyme among other aldo/keto reductases, we carried out stopped-flow studies with wild-type xylose reductase from the yeast Candida tenuis. Analysis of transient kinetic data for binding of NAD(+) and NADH, and reduction of D-xylose and oxidation of xylitol at pH 7.0 and 25 degrees C provided estimates of rate constants for the following mechanism: E + NADH right arrow over left arrow E.NADH right arrow over left arrow E.NADH + D-xylose right arrow over left arrow E.NADH.D-xylose right arrow over left arrow E.NAD(+).xylitol right arrow over left arrow E.NAD(+) right arrow over left arrow E.NAD(+) right arrow over left arrow E + NAD(+). The net rate constant of dissociation of NAD(+) is approximately 90% rate limiting for k(cat) of D-xylose reduction. It is controlled by the conformational change which precedes nucleotide release and whose rate constant of 40 s(-)(1) is 200 times that of completely rate-limiting E.NADP(+) --> E.NADP(+) step in aldehyde reduction catalyzed by human aldose reductase [Grimshaw, C. E., et al. (1995) Biochemistry 34, 14356-14365]. Hydride transfer from NADH occurs with a rate constant of approximately 170 s(-1). In reverse reaction, the E.NADH --> E.NADH step takes place with a rate constant of 15 s(-1), and the rate constant of ternary-complex interconversion (3.8 s(-1)) largely determines xylitol turnover (0.9 s(-1)). The bound-state equilibrium constant for C. tenuis xylose reductase is estimated to be approximately 45 (=170/3.8), thus greatly favoring aldehyde reduction. Formation of productive complexes, E.NAD(+) and E.NADH, leads to a 7- and 9-fold decrease of dissociation constants of initial binary complexes, respectively, demonstrating that 12-fold differential binding of NADH (K(i) = 16 microM) vs NAD(+) (K(i) = 195 microM) chiefly reflects difference in stabilities of E.NADH and E.NAD(+). Primary deuterium isotope effects on k(cat) and k(cat)/K(xylose) were, respectively, 1.55 +/- 0.09 and 2.09 +/- 0.31 in H(2)O, and 1.26 +/- 0.06 and 1.58 +/- 0.17 in D(2)O. No deuterium solvent isotope effect on k(cat)/K(xylose) was observed. When deuteration of coenzyme selectively slowed the hydride transfer step, (D)()2(O)(k(cat)/K(xylose)) was inverse (0.89 +/- 0.14). The isotope effect data suggest a chemical mechanism of carbonyl reduction by xylose reductase in which transfer of hydride ion is a partially rate-limiting step and precedes the proton-transfer step.  相似文献   

14.
The final two steps in the dmp operon-encoded meta-cleavage pathway for phenol degradation in Pseudomonas sp. strain CF600 involve conversion of 4-hydroxy-2-ketovalerate to pyruvate and acetyl coenzyme A (acetyl-CoA) by the enzymes 4-hydroxy-2-ketovalerate aldolase and aldehyde dehydrogenase (acylating) [acetaldehyde:NAD+ oxidoreductase (CoA acetylating), EC 1.2.1.10]. A procedure for purifying these two enzyme activities to homogeneity is reported here. The two activities were found to copurify through five different chromatography steps and ammonium sulfate fractionation, resulting in a preparation that contained approximately equal proportions of two polypeptides with molecular masses of 35 and 40 kDa. Amino-terminal sequencing revealed that the first six amino acids of each polypeptide were those deduced from the previously determined nucleotide sequences of the corresponding dmp operon-encoded genes. The isolated complex had a native molecular mass of 148 kDa, which is consistent with the presence of two of each polypeptide per complex. In addition to generating acetyl-CoA from acetaldehyde, CoA, and NAD+, the dehydrogenase was shown to acylate propionaldehyde, which would be generated by action of the meta-cleavage pathway enzymes on the substrates 3,4-dimethylcatechol and 4-methylcatechol. 4-Hydroxy-2-ketovalerate aldolase activity was stimulated by the addition of Mn2+ and, surprisingly, NADH to assay mixtures. The possible significance of the close physical association between these two polypeptides in ensuring efficient metabolism of the short-chain aldehyde generated by this pathway is discussed.  相似文献   

15.
Pre-steady state, stopped flow analysis of Escherichia coli D-3-phosphoglycerate dehydrogenase was performed by following the fluorescence of protein tryptophan and the fluorescence resonance energy transfer from protein tryptophan to bound NADH. The results indicate that binding of substrates is ordered, with coenzyme, NADH, binding first. Furthermore, the analysis indicated that there are two sets of sites on the tetrameric enzyme that can be differentiated by their kinetic behavior. NADH binding was consistent with an initial binding event followed by a slow conformational change for each site. The slow conformational change is responsible for the apparent tight binding of NADH to the apoenzyme but is too slow to participate in the catalytic cycle when the enzyme is rapidly turning over. Subsequent binding of the substrate, alpha-ketoglutarate, was characterized by a rapid equilibrium binding event followed by a conformational change for each site. Catalysis in the direction of NAD(+) reduction showed a distinct burst of activity followed by a slow rate of turnover, indicating that the rate-limiting step is after hydride transfer. Catalysis in the direction of NADH oxidation did not display burst kinetics, indicating that the rate-limiting step is at or before the hydride transfer step. The burst data indicated that the rate of NAD(+) reduction (3.8 s(-1)) is similar to the k(cat) of the enzyme (2-3 s(-1)) in that direction. However, analysis of the reaction with deuterated NADH failed to show an effect on the velocity of the reaction with a V(H)/V(D)=1.07+/-0.06. None of the other rates determined by stopped flow analysis could account for the k(cat) of the enzyme in either direction (forward k(cat)=0.01 s(-1), reverse k(cat)=2-3 s(-1)), suggesting that the rate-limiting step in both directions is a conformational change in the enzyme that is not detected optically.  相似文献   

16.
The effects of modifiers (NAD+, NADH, propionaldehyde, chloral hydrate, diethylstilboestrol and p-nitrobenzaldehyde) on the hydrolysis of p-nitrophenyl (PNP) pivalate (PNP trimethylacetate) catalysed by cytoplasmic aldehyde dehydrogenase are reported. In each case a different inhibition pattern is obtained to that observed when the substrate is PNP acetate; for example, propionaldehyde and chloral hydrate competitively inhibit the hydrolysis of PNP acetate, but are mixed inhibitors with PNP pivalate. The kinetic results can be rationalized in terms of different rate-determining steps: acylation of the enzyme in the case of the pivalate but acyl-enzyme hydrolysis for the acetate. This is confirmed by stopped-flow studies, in which a burst of p-nitrophenoxide is observed when the substrate is PNP acetate, but not when it is the pivalate. PNP pivalate inhibits the dehydrogenase activity of the enzyme competitively with the aldehyde substrate; this is most simply explained if the esterase and dehydrogenase reactions occur at a common enzymic site.  相似文献   

17.
The kinetics of the NAD+-dependent oxidation of aldehydes, catalysed by aldehyde dehydrogenase purified from sheep liver mitochondria, were studied in detail. Lag phases were observed in the assays, the length of which were dependent on the enzyme concentration. The measured rates after the lag phase was over were directly proportional to the enzyme concentration. If enzyme was preincubated with NAD+, the lag phase was eliminated. Double-reciprocal plots with aldehyde as the variable substrate were non-linear, showing marked substrate activation. With NAD+ as the variable substrate, double-reciprocal plots were linear, and apparently parallel. Double-reciprocal plots with enzyme modified with disulfiram (tetraethylthiuram disulphide) or iodoacetamide, such that at pH 8.0 the activity was decreased to 50% of the control value, showed no substrate activation, and the plots were linear. At pH 7.0, the kinetic parameters Vmax. and Km NAD+- for the oxidation of acetaldehyde and butyraldehyde by the native enzyme are almost identical. Formaldehyde and propionaldehyde show the same apparent maximum rate. Aldehyde dehydrogenase is able to catalyse the hydrolysis of p-nitrophenyl esters. This esterase activity was stimulated by both NAD+ and NADH, the maximum rate for the NAD+ stimulated esterase reaction being roughly equal to the maximum rate for the oxidation of aldehydes. The mechanistic implications of the above behaviour are discussed.  相似文献   

18.
The enzyme previously considered as an isozyme (E4, ALDH IV) of human liver aldehyde dehydrogenase (NAD+) (EC 1.2.1.3) has been purified to homogeneity by the use of ion exchange chromatography on CM-Sephadex and affinity chromatography on Blue Sepharose CL-6B and 5'-AMP Sepharose 4B and identified as glutamic gamma-semialdehyde dehydrogenase, or more precisely 1-pyrroline-5-carboxylate dehydrogenase (EC 1.5.1.12). Glutamic gamma-semialdehyde dehydrogenase was never previously purified to homogeneity from any mammalian species. The homogeneous enzyme is seen on isoelectric focusing gels as two fine bands separated by 0.12 pH units: pI = 6.89 and 6.77. In addition, the enzyme also appears as two bands in gradient gels; however, in polyacrylamide gels containing sodium dodecyl sulfate the enzyme migrates as one band, indicating that its subunits are of identical size. Because the enzyme molecule is considerably smaller (Mr approximately 142,000-170,000) than that of aldehyde dehydrogenases (EC 1.2.1.3) (Greenfield, N. J., and Pietruszko, R. (1977) Biochim. Biophys. Acta 483, 35-45; Mr approximately 220,000) and its subunit weight is different (70,600 versus approximately 54,000 for E1 and E2 isozymes), the enzyme is not an isozyme of aldehyde dehydrogenase previously described. The Michaelis constants for glutamic gamma-semialdehyde dehydrogenase with acetaldehyde and propionaldehyde are in the millimolar range. Its substrate specificity within the straight chain aliphatic aldehyde series is essentially confined to that of acetaldehyde and propionaldehyde with butyraldehyde and longer chain length aldehydes being considerably less active. Other substrates include succinic, glutaric, and adipic semialdehydes in addition to glutamic gamma-semialdehyde. The reaction velocity with glutamic gamma-semialdehyde is at least an order of magnitude larger than with carboxylic acid semialdehydes. Aspartic beta-semialdehyde is not a substrate. The reaction catalyzed appears to be irreversible. Although NADP can be used, NAD is the preferred coenzyme. The enzyme also exhibits an unusual property of being subject to substrate inhibition by NAD.  相似文献   

19.
The kcat value for the oxidation of propionaldehyde by sheep liver cytosolic aldehyde dehydrogenase increased 3-fold, from 0.16 s-1 at pH 7.6 to 0.49 s-1 at pH 5.2, in parallel with the increase in the rate of displacement of NADH from binary enzyme.NADH complexes. A burst in nucleotide fluorescence was observed at all pH values consistent with the rate of isomerization of binary enzyme.NADH complexes constituting the rate-limiting step in the steady state. No substrate activation by propionaldehyde was observed at pH 5.2, but the enzyme exhibited dissociation/association behavior. The inactive dissociated form of the enzyme was favored by low enzyme concentration, low pH, and low ionic strength. Propionaldehyde protected the enzyme against dissociation.  相似文献   

20.
Cytoplasmic aldehyde dehydrogenase catalyses the hydrolysis of methyl p-nitrophenyl (PNP) carbonate at an appreciable rate that is markedly stimualted by NAD+ or NADH. The nuleotides accelerate the rate-limiting hydrolysis of the acyl-enzyme intermediate while slowing the observed burst of p-nitrophenoxide production. With PNP dimethylcarbamate the enzyme catalyses the slow release of approx. 1 molecule of p-nitrophenoxide per tetrameric enzyme molecule; during the reaction the enzyme becomes effectively inactivated, as the rate of hydrolysis of the acyl-enzyme is virtually zero. The presence of NAD+, NADH, propionaldehyde, chloral hydrate, diethylstilboestrol or disulfiram slows the reaction of enzyme with PNP dimethylcarbamate. The reaction appears to be dependent on a group of pKa 7.6, possibly a cysteine residue. The effect of PNP dimethylcarbamate on the dehydrogenase activity of the enzyme is consistent with there being a single type of active site for the enzyme's dehydrogenase and esterase activities. Steric and electronic factors that govern reaction of the enzyme with PNP substrates are discussed.  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