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 250 毫秒
1.
In recent investigations we were able to demonstrate that the NADP-dependent aldehyde dehydrogenase of Acinetobacter calcoaceticus is an inducible enzyme localized in intracytoplasmic membranes limiting alkane inclusions. Long-chain aliphatic hydrocarbons and alkanols are inducers of the enzyme. It was purified by us and now kinetically characterized using the enzyme-micelle form, which contains bacterial phospholipids and a detergent (sodium cholate), too. The pH optimum of aldehyde dehydrogenase was determined to be at pH 10. The enzyme showed substrate inhibition (by aldehyde excess). The Ks and Km values of the leading substrate NADP+ were found to be 8.6 X 10(-5) and 10.3 X 10(-5)M independent of the chain-length of the aldehydes. The Km values of the aldehydes decreased depending on increasing chain-length (butanal: 1.6 X 10(-3), decanal: 1.5 X 10(-6)M). The Ki values (for inhibition by aldehyde excess) showed a similar behaviour (butanal: 7.5 X 10(-3), decanal: 3.5 X 10(-5)M) as well as the optimal aldehyde concentrations inducing the "maximal" reaction velocity (butanal: 5mM, decanal: 6 microM). The number of inhibiting aldehyde molecules per enzyme-substrate complex was determined to be n = 1. NADPH showed product inhibition kinetics (Ki(NADPH) = 2.2 X 10(-4)M), fatty acids did not. We were unable to measure a reverse reaction. The following ions and organic compounds were non-competitive inhibitors of the enzyme: Sn2+, Fe2+, Cu2+, BO3(3-), CN-, EDTA, o-phenanthroline, p-chloromercuri-benzoate, mercaptoethanol, phenylmethylsulfonyl fluoride, and diisopropylfluorophosphate; iodoacetate did not influence enzyme activity. Chloral hydrate was a competitive inhibitor of the aldehydes. Ethyl butyrate activates the enzyme, dependent on the chain-length of the aldehyde substrates.  相似文献   

2.
A bifunctional enzyme, L-(+)-tartrate dehydrogenase-D-(+)-malate dehydrogenase (decarboxylating) (EC 1.1.1.93 and EC 1.1.1. . . , respectively), was discovered in cells of Rhodopseudomonas sphaeroides Y, which accounts for the ability of this organism to grow on L-(+)-malate. The enzyme was purified 110-fold to homogeneity with a yield of 51%. During the course of purification, including ion-exchange chromatography and preparative gel electrophoresis, both enzyme activities appeared to be in association. The ratio of their activities remained almost constant [1:10, L-(+)-tartrate dehydrogenase/D-(+)-malate dehydrogenase (decarboxylating)] throughout all steps of purification. Analysis by polyacrylamide gel electrophoresis revealed the presence of a single protein band, the position of which was coincident with both L-(+)-tartrate dehydrogenase and D-(+)-malate dehydrogenase (decarboxylating) activities. The apparent molecular weight of the enzyme was determined to be 158,000 by gel filtration and 162,000 by ultracentrifugation. Sodium dodecyl sulfate-polyacrylamide gel electrophoresis yielded a single polypeptide chain with an estimated molecular weight of 38,500, indicating that the enzyme consisted of four subunits of identical size. The isoelectric point of the enzyme was between pH 5.0 and 5.2. The enzyme catalyzed the NAD-linked oxidation of L-(+)-tartrate as well as the oxidative decarboxylation of D-(+)-malate. For both reactions, the optimal pH was in a range from 8.4 to 9.0. The activation energy of the reaction (delta Ho) was 71.8 kJ/mol for L-(+)-tartrate and 54.6 kJ/mol for D-(+)-malate. NAD was required as a cosubstrate, and optimal activity depended on the presence of both Mn2+ and NH4+ ions. The reactions followed Michaelis-Menten kinetics, and the apparent Km values of the individual reactants were determined to be: L-(+)-tartrate, 2.3 X 10(-3) M; NAD, 2.8 X 10(-4) M; and Mn2+, 1.6 X 10(-5) M with respect to L-(+)-tartrate; and D-(+)-malate, 1.7 X 10(-4) M; NAD, 1.3 X 10(-4); and Mn2+, 1.6 X 10(-5) M with respect to D-(+)-malate. Of a variety of compounds tested, only meso-tartrate, oxaloacetate, and dihydroxyfumarate were effective inhibitors. meso-Tartrate and oxaloacetate caused competitive inhibition, whereas dihydroxyfumarate caused noncompetitive inhibition. The Ki values determined for the inhibitors were, in the above sequence, 1.0, 0.014, and 0.06 mM with respect to L-(+)-tartrate and 0.28, 0.012, and 0.027 mM with respect to D-(+)-malate.  相似文献   

3.
NADP-linked aldehyde dehydrogenase (aldehyde : NADP+ oxidoreductase, EC 1.2.1.4) was purified from Proteus vulgaris to the stage of homogeneity as judged by ultracentrifugation and polyacrylamide gel electrophoresis. The molecular weight of the purified enzyme was estimated to be 130000 by gel filtration. The enzyme which was crystallized from ammonium sulfate solution, lost its activity. The enzyme did not require coenzyme A, and the reaction was completely dependent on ammonium ions which could be partially replaced by Rb+ or K+. The optimum pH was about 9. Broad substrate specificity was observed and Km values for propionaldehyde, acetaldehyde and isovaleraldehyde were 1.7 - 10(-5), 4 - 10(-5) and 3 - 10(-5) M, respectively. The physiological role of the enzyme in living cells is obscure, but might account for another degradative pathway of L-leucine in P. vulgaris differing from the established pathway.  相似文献   

4.
Cell-free extracts of Hyphomicrobium X showed NAD-dependent aldehyde dehydrogenase activity, provided that NAD addition preceded that of aldehyde. Activity was lost rather rapidly, especially during purification attempts, but this could be partially masked by including a time-dependent restoration step with thiol compounds in the protocol. The nature of the assay buffer appeared to be critical and stimulation occurred on incorporation of K+ ions in the mixture. An even higher specific activity could be achieved by 1,4-dithiothreitol (DTT) treatment of the preparation, followed by removal of DTT, and assaying in the absence of thiol compounds under anaerobic conditions. Exposure of such a preparation to O2 led to a significant decrease in activity within a couple of hours. Immediate inactivation occurred on addition of H2O2, but this could be prevented completely by prior addition of NAD. Since GSH does not participate in the reaction and no stimulating factor was detected, the role of thiol compounds is most probably confined to restoration or prevention of damage to an O2-sensitive, necessary thiol group. Since the same features were found for cell-free extract as for the partially purified enzyme, only one enzyme type seems to be present. Although the enzyme is a general aldehyde dehydrogenase, the kinetic parameters and the specific activity of the cell-free extract for formaldehyde indicate that it may play a role in formaldehyde dissimilation by Hyphomicrobium X. The NAD-linked, GSH- and factor-independent aldehyde dehydrogenase described here appears to be different in several respects from the formaldehyde dehydrogenase of Pseudomonas putida (EC 1.2.1.46) (despite showing similar behavior toward coenzymes and factors) but resembles the aldehyde dehydrogenase from baker's yeast (EC 1.2.1.5).  相似文献   

5.
A 250- to 300-fold purification of a nicotinamide adenine denucleotide phosphate (NADP)-dependent glutamate dehydrogenase (GDH, E.C. 1.4.1.4) with a yield of 60% from a thermophilic bacillus is described. More than one NADP-specific GDH was detected by polyacrylamide gel electrophoresis. The enzyme is of high molecular weight (approximately 2 X 10-6), similar to that of the beef and frog liver GDH. The pI of the thermophilic GDH is at pH 5.24. The enzyme is highly thermostable at the pH range of 5.8 to 9.0. The purified GDH, unlike the crude enzyme, was very labile at subzero temperatures. An unidentified factor(s) from the crude cell-free extract prevented the inactivation of the purified GDH at -70 C. Various reactants of the GDH system and D-glutamate also protected, to some extent, the enzyme from inactivation at -70 C. From the Michaelis constants for glutamate (1.1 X 10-2M), NADP (3 X 10-4M), ammonia (2.1 X 10-2M), alpha-ketoglutarate (1.3 X 10-3M), and reduced NADP (5.3 X 10-5M), it is suggested that the enzyme catalyzes in vivo the formation of glutamate from ammonia and alpha-ketoglutarate. The amination of alpha-ketoglutarate and deamination of glutamate by the thermophilic GDH are optimal at the pH values of 7.2 and 8.4, respectively.  相似文献   

6.
N-Hydroxy-2-acetylaminofluorene has been found to be an effective inhibitor of aldehyde oxidase. At concentrations of 1 X 10(-6) M and 1 X 10(-5) M, 38% and 88% inhibition was observed on the oxidase activity towards N1-methylnicotinamide. The inhibition was of noncompetitive type and had a Ki value of 4.4 X 10(-6) M. In contrast, little inhibition of the enzyme was observed with 2-aminofluorene, 2-acetylaminofluorene and acetohydroxamic acid even at a concentration of 1 X 10(-4) M.  相似文献   

7.
Two acidic and three basic forms of monomeric dihydrodiol dehydrogenase with molecular weights in the range of 36,000-39,000 were purified from human liver. One acidic enzyme (pI 5.2), which was specific for NADP- and dihydrodiols of benzene and naphthalene, was immunologically identified as aldehyde reductase. The other four enzymes oxidized alicyclic alcohols as well as the dihydrodiols using both NADP+ and NAD+ as cofactors, but showed differences in specificity for hydroxysteroids and inhibitor sensitivity. Two of the basic enzymes (pI 9.7 and 9.1) exhibited a 20 alpha-hydroxysteroid dehydrogenase activity and sensitivity to 1,10-phenanthroline, whereas the third basic enzyme (pI 7.6) oxidized some 3 alpha-hydroxysteroids at low rates and was inhibited by cyclopentane-1,1-diacetic acid. Another acidic enzyme, which accounted for the largest amount of enzyme activity in the tissue and appeared in two heterogenous forms with pI values of 5.9 and 5.4, showed a high 3 alpha-hydroxysteroid dehydrogenase activity and was the most sensitive to inhibition by medroxyprogesterone acetate. The Km values of the enzymes, except the pI 5.2 enzyme, for hydroxysteroids (10(-6) to 10(-7) M) were lower than those for xenobiotic alcohols.  相似文献   

8.
Phosphoglycerate mutase has been purified from methanol-grown Hyphomicrobium X and Pseudomonas AMI by acid precipitation, heat treatment, ammonium sulphate fractionation, Sephadex G-50 gel filtration and DEAE-cellulose column chromatography. The purification attained using the Hyphomicrobium X extract was 72-fold, and using the Pseudomonas AMI extract, 140-fold. The enzyme purity, as shown by analytical polyacrylamide gel electrophoresis, was 50% from Hyphomicrobium X and 40% from Pseudomonas AMI. The enzyme activity was associated with one band. The purified preparations did not contain detectable amounts of phosphoglycerate kinase, phosphopyruvate hydratase, phosphoglycerate dehydrogenase or glycerate kinase activity. The molecular weight of the enzymic preparation was 32000 +/- 3000. The enzyme from both organisms was stable at low temperatures and, in the presence of 2,3-diphosphoglyceric acid, could withstand exposure to high temperatures. The enzyme from Pseudomonas AMI has a broad pH optimum at 7-0 to 7-6 whilst the enzyme from Hyphomicrobium X has an optimal activity at pH 7-3. The cofactor 2,3-diphosphoglyceric acid was required for maximum enzyme activity and high concentrations of 2-phosphoglyceric acid were inhibitory. The Km values for the Hyphomicrobium X enzyme were: 3-phosphoglyceric acid, 6-0 X 10(-3) M: 2-phosphoglyceric acid, 6-9 X 10(-4) M; 2,3-diphosphoglyceric acid, 8-0 X 10(-6) M; and for the Pseudomonas AMI ENzyme: 3-4 X 10(-3) M, 3-7 X 10(-4) M and 10 X 10(-6) M respectively. The equilibrium constant for the reaction was 11-3 +/- 2-5 in the direction of 2-phosphoglyceric acid to 3-phosphoglyceric acid and 0-09 +/- 0-02 in the reverse direction. The standard free energy for the reaction proceeding from 2-phosphoglyceric acid to 3-phosphoglyceric acid was -5-84 kJ mol(-1) and in the reverse direction +5-81 kJ mol(-1).  相似文献   

9.
The possibility of a functional complex formation between glyceraldehyde-3-phosphate dehydrogenase (EC 1.2.1.12) and 3-phosphoglycerate kinase (EC. 2.7.2.3), enzymes catalysing two consecutive reactions in glycolysis has been investigated. Kinetic analysis of the coupled enzymatic reaction did not reveal any kinetic sign of the assumed interaction up to 4 X 10(-6) M kinase and 10(-4) M dehydrogenase. Fluorescence anisotrophy of 10(-7) M or 2 X 10(-5) M glyceraldehyde-3-phosphate dehydrogenase labeled with fluorescein isothiocynate did not change in the presence of non-labeled 3-phosphoglycerate kinase (up to 4 X 10(-5) M). The frontal gel chromatographic analysis of a mixture of the two enzymes (10(-4) M dehydrogenase) could not reveal any molecular species with the kinase activity having a molecular weight higher than that of 3-phosphoglycerate kinase. Both types of physicochemical measurements were also performed in the presence of substrates of the kinase and gave the same results. The data seem to invalidate the hypothesis that there is a complex between purified pig muscle glyceraldehyde-3-phosphate dehydrogenase and 3-phosphoglycerate kinase.  相似文献   

10.
Phosphorylation of ankyrin decreases its affinity for spectrin tetramer   总被引:5,自引:0,他引:5  
The effects of phosphorylation on the interaction between spectrin and ankyrin were investigated. Spectrin and ankyrin were phosphorylated using purified human erythrocyte membrane and cytosolic (casein kinase A) kinases. These two kinases have similar properties as well as activities toward spectrin and ankyrin. Both kinases catalyzed the incorporation of about 2 mol of phosphate/mol of spectrin and about 7 mol of phosphate/mol of ankyrin. These phosphates were incorporated primarily into seryl and threonyl residues of the proteins. The phosphopeptide maps of ankyrin phosphorylated by the membrane kinase and casein kinase A were identical. Binding studies indicate that ankyrin exhibits different affinities for spectrin dimers (KD = 2.5 +/- 0.9 X 10(-6) M) and tetramers (KD = 2.7 +/- 0.8 X 10(-7) M). These dissociation constants were not appreciably affected by the phosphorylation of spectrin. On the other hand, phosphorylation of ankyrin was found to significantly reduce its affinity for either phosphorylated or unphosphorylated spectrin tetramers (KD = 1.2 +/- 0.1 X 10(-6) M) but not spectrin dimers (KD = 2.5 +/- 0.4 X 10(-6) M). The same results were obtained using either the membrane kinase or casein kinase A as the phosphorylating enzyme. The above observation suggests that ankyrin phosphorylation may provide an important mechanism for the regulation of the erythrocyte membrane cytoskeletal network.  相似文献   

11.
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   

12.
Human erythrocyte aldehyde dehydrogenase (aldehyde:NAD+ oxidoreductase, EC 1.2.1.3) was purified to apparent homogeneity. The native enzyme has a molecular weight of about 210,000 as determined by gel filtration, and SDS-polyacrylamide gel electrophoresis of this enzyme yields a single protein and with a molecular weight of 51,500, suggesting that the native enzyme may be a tetramer. The enzyme has a relatively low Km for NAD (15 microM) and a high sensitivity to disulfiram. Disulfiram inhibits the enzyme activity rapidly and this inhibition is apparently of a non-competitive nature. In kinetic characteristic and sensitivity to disulfiram, erythrocyte aldehyde dehydrogenase closely resembles the cytosolic aldehyde dehydrogenase found in the liver of various species of mammalians.  相似文献   

13.
Castellaniella defragrans is a Betaproteobacterium capable of coupling the oxidation of monoterpenes with denitrification. Geraniol dehydrogenase (GeDH) activity was induced during growth with limonene in comparison to growth with acetate. The N-terminal sequence of the purified enzyme directed the cloning of the corresponding open reading frame (ORF), the first bacterial gene for a GeDH (geoA, for geraniol oxidation pathway). The C. defragrans geraniol dehydrogenase is a homodimeric enzyme that affiliates with the zinc-containing benzyl alcohol dehydrogenases in the superfamily of medium-chain-length dehydrogenases/reductases (MDR). The purified enzyme most efficiently catalyzes the oxidation of perillyl alcohol (k(cat)/K(m) = 2.02 × 10(6) M(-1) s(-1)), followed by geraniol (k(cat)/K(m) = 1.57 × 10(6) M(-1) s(-1)). Apparent K(m) values of <10 μM are consistent with an in vivo toxicity of geraniol above 5 μM. In the genetic vicinity of geoA is a putative aldehyde dehydrogenase that was named geoB and identified as a highly abundant protein during growth with phellandrene. Extracts of Escherichia coli expressing geoB demonstrated in vitro a geranial dehydrogenase (GaDH) activity. GaDH activity was independent of coenzyme A. The irreversible formation of geranic acid allows for a metabolic flux from β-myrcene via linalool, geraniol, and geranial to geranic acid.  相似文献   

14.
1. The properties and distribution of the NAD-linked unspecific aldehyde dehydrogenase activity (aldehyde: NAD+ oxidoreductase EC 1.2.1.3) has been studied in isolated cytoplasmic, mitochondrial and microsomal fractions of rat liver. The various types of aldehyde dehydrogenase were separated by ion exchange chromatography and isoelectric focusing. 2. The cytoplasmic fraction contained 10-15, the mitochondrial fraction 45-50 and the microsomal fraction 35-40% of the total aldehyde dehydrogenase activity, when assayed with 6.0 mM propionaldehyde as substrate. 3. The cytoplasmic fraction contained two separable unspecific aldehyde dehydrogenases, one with high Km for aldehydes (in the millimolar range) and the other with low Km for aldehydes (in the micromolar range). The latter can, however, be due to leakage from mitochondria. The high-Km enzyme fraction contained also all D-glucuronolactone dehydrogenase activity of the cytoplasmic fraction. The specific formaldehyde and betaine aldehyde dehydrogenases present in the cytoplasmic fraction could be separated from the unspecific activities. 4. In the mitochondrial fraction there was one enzyme with a low Km for aldehydes and another with high Km for aldehydes, which was different from the cytoplasmic enzyme. 5. The microsomal aldehyde dehydrogenase had a high Km for aldehydes and had similar properties as the mitochondrial high-Km enzyme. Both enzymes have very little activity with formaldehyde and glycolaldehyde in contrast to the other aldehyde dehydrogenases. They are apparently membranebound.  相似文献   

15.
Pyridoxine dehydrogenase (1.1.1.65) (pyridoxal reductase), purified to homogeneity from baker's yeast, is a monomer of Mr approximately 33,000. It catalyzes the reversible oxidation of pyridoxine by NADP to yield pyridoxal and NADPH; equilibrium lies far in the direction of pyridoxine formation (Keq approximately 1.4 X 10(11) l/mol at 25 degrees C). Reduction of pyridoxal occurs most rapidly at pH 6.0-7.0; oxidation of pyridoxine is optimal at pH 8.6. NAD and NADH do not replace NADP and NADPH as substrates; pyridoxine, pyridoxal and pyridoxal 5'-phosphate are the only naturally occurring cosubstrates found. Several other aromatic aldehydes also are reduced, but substrate specificity and other properties of the enzyme distinguish it clearly from other alcohol dehydrogenases or aldehyde reductases. Between pH 6.3 and 7.1 (the intracellular pH of yeast), V/Km with pyridoxal and NADPH as substrates is greater than 600 times that observed with pyridoxine and NADPH as substrates is greater than 600 times that observed with pyridoxine and NADP as substrates. These and other considerations strongly indicate that the dehydrogenase functions in vivo to reduce pyridoxal to pyridoxine, which is the preferred substrate for pyridoxal (pyridoxine) kinase in yeast.  相似文献   

16.
Dihydrodiol dehydrogenase activity was detected in the cytosol of various mouse tissues, among which kidney exhibited high specific activity comparable to the value for liver. The enzyme activity in the kidney cytosol was resolved into one major and three minor peaks by Q-Sepharose chromatography: one minor form cross-reacted immunologically with hepatic 3 alpha-hydroxysteroid dehydrogenase and another with aldehyde reductase. The other minor form was partially purified and the major form was purified to homogeneity. These two forms, although different in their charges, were monomeric proteins with the same molecular weight of 39,000 and had similar catalytic properties. They oxidized cis-benzene dihydrodiol and alicyclic alcohols as well as trans-dihydrodiols of benzene and naphthalene in the presence of NADP+ or NAD+, and reduced several xenobiotic aldehydes and ketones with NAD(P)H as a cofactor. The enzymes also catalyzed the oxidation of 3 alpha-hydroxysteroids and epitestosterone, and the reduction of 3- and 17-ketosteroids, showing much lower Km values (10(-7)-10(-6) M) for the steroids than for the xenobiotic alcohols. The results of mixed substrate experiments, heat stability, and activity staining on polyacrylamide gel electrophoresis suggested that, in the two enzymes, both dihydrodiol dehydrogenase and 3(17)alpha-hydroxysteroid dehydrogenase activities reside on a single enzyme protein. Thus, dihydrodiol dehydrogenase existed in four forms in mouse kidney cytosol, and the two forms distinct from the hepatic enzymes may be identical to 3(17)alpha-hydroxysteroid dehydrogenases.  相似文献   

17.
A unique phosphatase that selectively hydrolyzed phosphotyrosine and 2'-AMP at alkaline pH and p-nitrophenylphosphate at neutral pH was isolated from a cytosolic fraction of rat brain. The purified enzyme appeared homogenous on SDS-polyacrylamide gel electrophoresis and its molecular weight was estimated to be 42,000. The molecular weight of the native enzyme was 45,000 as determined by molecular sieve chromatography. These findings indicate that the native enzyme is a monomer protein. At pH 8.6, the enzyme hydrolyzed L-phosphotyrosine, D-phosphotyrosine, 2'-AMP, p-nitrophenylphosphate, 3'-AMP, 2'-GMP, and 3'-GMP; the ratio of its activities with these substrates was 100:96:115:68:39:25:16. Its Km values for L-phosphotyrosine, 2'-AMP, and p-nitrophenylphosphate were 0.8 X 10(-4) M, 1.4 X 10(-4) M, and 1.7 X 10(-4) M, respectively. At pH 7.4, the enzyme hydrolyzed p-nitrophenylphosphate, L-phosphotyrosine, and D-phosphotyrosine; the ratio of its activities with these compounds was 100:17:17, and its Km values for L-phosphotyrosine and p-nitrophenylphosphate were 1.8 X 10(-4) M and 2.0 X 10(-4) M, respectively. The enzyme activity was dependent on Mn2+ or Mg2+, and was strongly inhibited by 5'-nucleotides, pyrophosphate, and Zn2+. The enzyme was not sensitive to inhibitors of some well-characterized phosphatases such as NaF, molybdate, L(+)tartrate, tetramisole, vanadate, and lithium salt. The physiological role of the enzyme is discussed with respect to its activities toward phosphotyrosine, 2'-AMP, and p-nitrophenylphosphate.  相似文献   

18.
Saccharopine dehydrogenase (N6-(glutar-2-yl)-L-ly-sine:NAD oxidoreductase (L-lysine-forming)) from baker's yeast was purified to homogenicity. The overall purification was about 1,200-fold over the crude extract with a yield of about 24%. The purified enzyme had a sedimentation coefficient (S20,w) of 3.0 S. The molecular weight determinations by sedimentation equilibrium, Sephadex G-100 gel filtration, and sodium dodecyl sulfate-polyacrylamide gel electrophoresis gave a value of about 39,000 and, therefore, saccharopine dehydrogenase is a single polypeptide chain enzyme. A Stokes radius of 27 A and a diffusion constant of 7.9 X 10(-7) cm2 s-1 were obtained from Sephadex gel filtration chromatography. The enzyme had a high isoelectric pH of 10.1. The NH2-terminal sequence was Ala-Ala----. The enzyme possessed 3 cysteine residues/molecule; no disulfide bond was present. Incubation of saccharopine dehydrogenase with p-chloromercuribenzoate or iodoacetate resulted in complete loss of enzyme activity. Whereas the coenzyme and substrates were ineffective in protecting from inactivation by p-chloromercuribenzoate, iodoacetate inhibition was protected by excess coenzyme.  相似文献   

19.
The dissociation of the aldehyde dehydrogenase X NADH complex was studied by displacement with NAD+. The association reaction of enzyme and NADH was also studied. These processes are biphasic, as shown by McGibbon, Buckley & Blackwell [(1977) Biochem. J. 165, 455-462], but the details of the dissociation reaction are significantly different from those given by those authors. Spectral and kinetic experiments provide evidence for the formation of abortive complexes of the type enzyme X NADH X aldehyde. Kinetic studies at different wavelengths with transcinnamaldehyde as substrate provide evidence for the formation of an enzyme X NADH X cinnamoyl complex. Hydrolysis of the thioester relieves a severe quenching effect on the fluorescence of enzyme-bound NADH.  相似文献   

20.
The process of isolation and purification of malate dehydrogenase (decarboxylating) (EC 1.1.1.40) from the mycelium of the actinomycete Streptomyces aureofaciens has been worked out. The enzyme was purified 35 fold. The kinetic characters of the purified enzyme are very similar to the figures for malate dehydrogenase (decarboxylating) from other sources. Km for L-malate = 2.1 X 10(-3)M, Km for NADP = 4.6 X 10(-5)M (at pH 7.4). The reaction requires metal divalent ions, Mn2+ being more effective than Mg2+. The enzyme reaches its maximal activity at pH 8.75.  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