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.
l-Alanine dehydrogenase was found in extracts of the antibiotic producer Streptomyces clavuligerus. The enzyme was induced by ammonia, and the level of induction was dependend on the extracellular concentration. l-Alanine was the only amino acid able to induce alanine dehydrogenase. The enzyme was characterized from a 38-fold purified preparation. Pyruvate (K m =1.1 mM), ammonia (K m =20 mM) and NADH (K m =0.14 mM) were required for the reductive amination, and l-alanine (K m =9.1 mM) and NAD (K m =0.5 mM) for the oxidative deaminating reaction. The aminating reaction was inhibited by alanine, serine and NADPH. Alanine inhibited uncompetitively with respect to NADH (K i =1.6 mM) and noncompetitively with respect to ammonia (K i =2.0 mM) and pyruvate (K i =3.0 mM). In the aminating reaction 3-hydroxypyruvate, glyoxylate and 2-oxobutyrate could partially (6–7%) substitute pyruvate. Alanine dehydrogenase from S. clavuligerus differed with respect to its molecular weight (92000) and its kinetic properties from those described for other microorganisms.Abbreviation Alanine-DH l-alanine:NAD oxidoreductase  相似文献   

2.
The binding of NADH to bull semen NAD nucleosidase was observed to be accompanied by a considerable enhancement of the fluorescence of NADH. The fluorescence enhancement observed in the binding of NADH to the enzyme was utilized to study the stoichiometry of binding of this compound to the enzyme. Results obtained from the fluorescence titration of the enzyme with NADH indicated the binding of one mole of NADH per mole of enzyme (36,000 g). The dissociation constant for the enzyme-NADH complex was determined to be 2.52 × 10?6m. NADH was also found to be a very effective competitive inhibitor of the NADase-catalyzed hydrolysis of NAD, and the inhibitor dissociation constant (KI) for the enzyme-NADH complex was determined to be 2.99 × 10?6m which was in good agreement with the value obtained from the fluorescence titration experiments.  相似文献   

3.
The kinetics of NADH oxidation by the outer membrane electron transport system of intact beetroot (Beta vulgaris L.) mitochondria were investigated. Very different values for Vmax and the Km for NADH were obtained when either antimycin A-insensitive NADH-cytochrome c activity (Vmax= 31 ± 2.5 nmol cytochrome c (mg protein)?1 min?1; Km= 3.1 ± 0.8 μM) or antimycin A-insensitive NADH-ferricyanide activity (Vmax= 1.7 ± 0.7 μmol ferricyanide (mg protein)?1 min?1; Km= 83 ± 20 μM) were measured. As ferricyanide is believed to accept electrons closer to the NADH binding site than cytochrome c, it was concluded that 83 ± 20 μM NADH represented a more accurate estimate of the binding affinity of the outer membrane dehydrogenase for NADH. The low Km determined with NADH-cytochrome c activity may be due to a limitation in electron flow through the components of the outer membrane electron transport chain. The Km for NADH of the externally-facing inner membrane NADH dehydrogenase of pea leaf (Pisum sativum L. cv. Massey Gem) mitochondria was 26.7 ± 4.3 μM when oxygen was the electron acceptor. At an NADH concentration at which the inner membrane dehydrogenase should predominate, the Ca2+ chelator, ethyleneglycol-(β-aminoethylether)-N,N,-tetraacetic acid (EGTA), inhibited the oxidation of NADH through to oxygen and to the ubiquinone-10 analogues, duroquinone and ubiquinone-1, but had no effect on the antimycin A-insensitive ferricyanide reduction. It is concluded that the site of action of Ca2+ involves the interaction of the enzyme with ubiquinone and not with NADH.  相似文献   

4.
Formaldehyde dehydrogenase and formate dehydrogenase were purified 45- and 16-fold, respectively, from Hansenula polymorpha grown on methanol. Formaldehyde dehydrogenase was strictly dependent on NAD and glutathione for activity. The K mvalues of the enzyme were found to be 0.18 mM for glutathione, 0.21 mM for formaldehyde and 0.15 mM for NAD. The enzyme catalyzed the glutathine-dependent oxidation of formaldehyde to S-formylglutathione. The reaction was shown to be reversible: at pH 8.0 a K mof 1 mM for S-formylglutathione was estimated for the reduction of the thiol ester with NADH. The enzyme did not catalyze the reduction of formate with NADH. The NAD-dependent formate dehydrogenase of H. polymorpha showed a low affinity for formate (K mof 40 mM) but a relatively high affinity for S-formylglutathione (K mof 1.1 mM). The K mvalues of formate dehydrogenase in cell-free extracts of methanol-grown Candida boidinii and Pichia pinus for S-formylglutathione were also an order of magnitude lower than those for formate. It is concluded that S-formylglutathione rather than free formate is an intermediate in the oxidation of methanol by yeasts.  相似文献   

5.
The reaction mechanism of the F1 isozyme of horse liver aldehyde dehydrogenase (EC 1.2.1.3) was investigated using both steady-state and rapid kinetic techniques. Using the steady-state substrate velocity patterns, the NADH inhibition patterns at several aldehyde concentrations, and the substrate analog (adenosine diphosphoribose and chloral hydrate) inhibition patterns, the enzymic catalysis was shown to involve ordered addition of NAD followed by aldehyde. This mechanism was confirmed using the kinetics of the hydrolysis of p-nitrophenyl acetate as an indicator of the dehydrogenase substrate binding. Steady-state experiments with deuteroacetaldehyde showed the V to be unchanged, but the Km increased (Km CH3CDOKm CH3CHO = 3.4). Stopped flow experiments where E-NAD was rapidly mixed with aldehyde showed a burst of NADH formation followed by slower steady-state turnover. This result clearly indicates that the rate limiting step lies after NAD reduction. The NADH off rate (0.7 s?1) as estimated by displacement of NADH from the E-NADH complex upon rapid addition of NAD was found to be very close to the steady-state site turnover number (0.3 s?1). This fact and the relatively small effect of aldehyde R-group on maximal velocity suggest that the slow rate of NADH release contributes significantly to limitation of the enzyme catalytic velocity.  相似文献   

6.
Active pharmaceutical ingredients (APIs) such as l-sugars and keto acids are favorably accessed through selective oxidation of sugar alcohols and amino acids, respectively, catalyzed by NAD(P)-dependent dehydrogenases. Cofactor regeneration from NAD(P)H conveniently is achieved via water-forming NAD(P)H oxidases (nox2), which only need molecular oxygen as co-substrate. Turnover-dependent overoxidation of the conserved cysteine residue in the active site of water-forming NADH oxidases is the presumed cause of the limited nox2 stability.We present a novel NAD(P)H oxidase, NoxV from Lactobacillus plantarum, with specific activity of 167 U/mg and apparent kinetic constants at air saturation and 25 °C of kcat,app = 212 s−1 and KM,app = 50.2 μM in the broad pH optimum from 5.5 to 8.0. The enzyme features a higher stability than other NAD(P)H oxidases against overoxidation, as is evidenced by a higher total turnover number, in the presence (168,000) and, most importantly, also in the absence (128,000) of exogenously added reducing agents. While the native enzyme shows exclusively activity on NADH, we engineered the substrate binding pocket to generate variants, G178K,R and L179K,R,H that accommodate and oxidize both NADH and NADPH as substrates.  相似文献   

7.
Acetoacetyl-CoA was found to strongly inhibit the dehydrogenation of L-3-hydroxybutyryl-CoA catalyzed by L-3-hydroxyacyl-CoA dehydrogenase from pig heart. The inhibition constant (Ki) was determined to be 7.7 × 10?6 M, a value which is similar to the Km value of 12 × 10?6 M obtained for acetoacetyl-CoA in its NADH-dependent reduction catalyzed by the same enzyme. A suggested ordered BiBi mechanism for this enzyme, with NAD binding to the enzyme first, explains the observed noncompetitive nature of this inhibition. The possible effect of this inhibition on fatty acid oxidation is discussed.  相似文献   

8.
Phosphoribulokinase from the nonsulfur purple bacterium Rhodopseudomonas acidophila has been purified to apparent homogeneity, using affinity chromatography on Cibacron Blue-agarose and AMP-agarose. The relative molar mass of the enzyme was determined by sucrose density gradient centrifugation to be M r=248,000 with a sedimentation coefficient of s 20,w=10.9 S. Dodecyl sulfate polyacrylamide gel electrophoresis revealed that the enzyme consists of identical size subunits of M r=32,000, suggesting an octameric structure of the holoenzyme. The enzyme cross-reacted with heterologous antibodies raised against phosphoribulokinase from the hydrogen bacterium Alcaligenes eutrophus. The pH optimum of the enzyme was shifted from 8.4 in the absence of the activator NADH to 7.6 in the presence of the effector. Mg2+ ions were the most effective divalent cations required for activity. Specificity of the enzyme for the sugar phosphate substrate ribulose 5-phosphate was high whereas a variety of nucleoside triphosphates besides ATP could serve as phosphate donors. NADH was a strong activator of the enzyme (K a=0.05 mM) that primarily affected the maximal reaction velocity in a pH-dependent manner. The only other effector identified was phosphoenolpyruvate. It moderately inhibited the enzyme (I 0.5=0.32 mM).Abbreviation PRK phosphoribulokinase Dedicated to Prof. Dr. H. G. Schlegel on the occasion of his 60th birthday  相似文献   

9.
Bacterial 2,3-butanediol dehydrogenases   总被引:3,自引:0,他引:3  
Enterobacter aerogenes, Aeromonas hydrophila, Serratia marcescens and Staphylococcus aureus possessing L(+)-butanediol dehydrogenase produced mainly meso-butanediol and small amounts of optically active butanediol; Acetobacter suboxydans, Bacillus polymyxa and Erwinia carotovora containing D(-)-butanediol dehydrogenase produced more optically active butanediol than meso-butanediol. Resting and growing cells of these organisms oxidized only one enantiomer of racemic butanediol. The D(-)-butanediol dehydrogenase from Bacillus polymyxa was partially purified (30-fold) with a specific activity of 24.5. Except NAD and NADH no other cofactors were required. Optimum pH-values for oxidation and reduction were pH 9 and pH 7, respectively. The optimum temperature was about 60°C. The molecular weight was 100000 to 107000. The K m-values were 3.3 mM for D(-)-butanediol, 6.25 mM for meso-butanediol, 0.53 mM for acetoin, 0.2 mM for NAD, 0.1 mM for NADH, 87 mM for diacetyl, 38 mM for 1,2-propanediol; 2,3-pentanedion was not a substrate for this enzyme. The L(+)-butanediol dehydrogenase from Serratia marcescens was purified 57-fold (specific activity 22.3). Besides NAD or NADH no cofactors were required. The optimum value for oxidation was about pH 9 and for reduction pH 4.5. The optimum temperature was 32–36°C. The molecular weight was 100000 to 107000. The K m-values were 5 mM for meso-butanediol, 10 mM for racemic butanediol, 6.45 for acetoin, 1 mM for NAD, 0.25 mM for NADH, 2.08 mM for diacetyl, 16.7 mM for 2,3-pentanedion and 11.8 mM for 1,2-propanediol.Abbreviations Bud 2,3-butanediol - DH dehydrogenase  相似文献   

10.
Yeast alcohol dehydrogenase (EC 1.1.1.1) catalyzes the novel reduction of p-nitro-so-N,N-dimethylaniline with NADH as a cofactor. Apparent kinetic constants for this enzymatic reaction are: V 2=2.1 s–1, K Q=456 M, K iQ=119 M, and K P=1.47 mM, at pH 8.9, 25 °C. This reaction is especially useful for the quantitative determination of NAD+ and NADH by enzymatic cycling.  相似文献   

11.
35Cl nmr relaxation rate measurements have been used to study anion-binding sites in pig heart lactate dehydrogenase. These studies reveal two types of sites, one is intimately associated with the active site, the other is not. The nonactive site has been ascribed to a subunit site in analogy with crystallographic results from the dogfish M4 enzyme. The binding of either the reduced or the oxidized form of NAD results in an increase in the 35Cl nmr relaxation rate by a factor of 1.8–2. The enhanced nmr relaxation rate of the binary lactate dehydrogenase-NAD complex is reduced on binding of the substrate inhibitor molecules oxamate or oxalate to a value less than that exhibited by lactate dehydrogenase alone. The enhancement of the nmr relaxation rate is attributed to a decrease in the dissociation constant of Cl for the enzyme. The Kp values for Cl binding to the active center site of lactate dehydrogenase is 0.85 m and for lactate dehydrogenase-NADH is 0.25 m. The ratio of these constants, 3.4, agrees well with the measured enhancement value 3.7. The effect of coenzyme analogs on the 35Cl nmr relaxation rate has been examined. 3-Acetylpyridine NAD produces an enhancement of 4.3, thionicotinamide NAD of 2.3, whereas 3-pyridinealdehyde, adenosinediphosphoribose, and adenosine diphosphate do not affect the nmr relaxation state of Cl bound to lactate dehydrogenase.  相似文献   

12.
Oxidative deamination of putrescine, the precursor of polyamines, gives rise to γ-aminobutyraldehyde (ABAL). In this study an aldehyde dehydrogenase, active on ABAL, has been purified to electrophoretic homogeneity from rat liver cytoplasm and its kinetic behaviour investigated. The enzyme is a dimer with a subunit molecular weight of 51,000. It is NAD+-dependent, active only in the presence of sulphhydryl compounds and has a pH optimum in the range 7.3–8.4. Temperatures higher than 28°C promote slow activation and the process is favoured by the presence of at least one substrate. Km for aliphatic aldehydes decreases from 110 μM for ABAL and acetaldehyde to 2–3 μM for capronaldehyde. The highest relative V-values have been observed with ABAL (100) and isobutyraldehyde (64), and the lowest with acetaldehyde (14). Affinity for NAD+ is affected by the aldehyde present at the active site: Km for NAD+ is 70 μM with ABAL, 200 μM with isobutyraldehyde and capronaldehyde, and>800 μM with acetaldehyde. The kinetic behaviour at 37°C is quite complex; according to enzymatic models, NAD+ activates the enzyme (Kact 500 μM) while NADH competes for the regulatory site (Kin 70 μM). In the presence of high NAD+ concentrations (4 mM), ABAL promotes further activation by binding to a low-affinity regulatory site (Kact 10 mM). The data show that the enzyme is probably an E3 aldehyde dehydrogenase, and suggest that it can effectively metabolize aldehydes arising from biogenic amines.  相似文献   

13.
《Experimental mycology》1982,6(3):274-282
Isocitrate dehydrogenase (threo-ds-isocitrate: NAD oxidoreductase (decar☐ylating) EC 1.1.1.41) from Dictyostelium dicoideum was purified 161-fold. The purified enzyme was NAD specific and required Mn2+ for activity. Isocitrate consumption and 2-oxoglutarate and NADH production were stoichiometric; no NADH oxidase or glutamate dehydrogenase activities were detected. The pH optimum range for activity was pH 7.5–8.5. Reductive car☐ylation of 2-oxoglutarate with NADH could not be demonstrated. Lineweaver - Burk plots of data from initial velocity studies were linear. There was no evidence of allosteric control by reported effectors (AMP, ADP, citrate) of isocitrate dehydrogenase activity. The reaction was inhibited by NADH. The inhibition by NADH was competitive when either isocitrate or NAD was the variable substrate. 2-Oxoglutarate was not inhibitory at concentrations below 4 mm. The Michaelis constant (Km) and dissociation constant (Kib) for isocitrate were 0.16 mm; and Km and dissociation constant (Kia) for NAD were 0.34 mm. The inhibition constant for NADH was 0.02 mm. The data are consistent with a rapid equilibrium random bi-bi reaction mechanism (Cleland nomenclature). The NAD-linked isocitrate dehydrogenase activity was also demonstrated in crude extracts of isolated mitochondria.  相似文献   

14.
Using isolated chloroplasts or purified thylakoids from photoautotrophically grown cells of the chromophytic alga Pleurochloris meiringensis (Xanthophyceae) we were able to demonstrate a membrane bound NAD(P)H dehydrogenase activity. NAD(P)H oxidation was detectable with menadione, coenzyme Q0, decylplastoquinone and decylubiquinone as acceptors in an in vitro assay. K m-values for both pyridine nucleotides were in the molar range (K m[NADH]=9.8 M, K m[NADPH]=3.2 M calculated according to Lineweaver-Burk). NADH oxidation was optimal at pH 9 while pH dependence of NADPH oxidation showed a main peak at 9.8 and a smaller optimum at pH 7.5–8. NADH oxidation could be completely inhibited with rotenone, an inhibitor of mitochondrial complex I dehydrogenase, while NADPH oxidation revealed the typical inhibition pattern upon addition of oxidized pyridine nucleotides reported for ferredoxin: NADP+ reductase. Partly-denaturing gel electrophoresis followed by NAD(P)H dehydrogenase activity staining showed that NADPH and NADH oxidizing proteins had different electrophoretic mobilities. As revealed by denaturing electrophoresis, the NADH oxidizing enzyme had one main subunit of 22 kDa and two further polypeptides of 29 and 44 kDa, whereas separation of the NADPH depending protein yielded five bands of different molecular weight. Measurement of oxygen consumption due to PS I mediated methylviologen reduction upon complete inhibition of PS II showed that the NAD(P)H dehydrogenase is able to catalyze an input of electrons from NADH to the photosynthetic electron transport chain in case of an oxidized plastoquinone-pool. We suggest ferredoxin: NADP+ reductase to be the main NADPH oxidizing activity while a thylakoidal NAD(P)H: plastoquinone oxidoreductase involved in the chlororespiratory pathway in the dark acts mainly as an NADH oxidizing enzyme.Abbreviations Coenzyme Q0-2,3-dimethoxy-5-methyl-1,4-benzoquinone - FNR ferredoxin: NADP+ reductase - MD menadione - MV methylviologen - NDH NAD(P)H dehydrogenase - PQ plastoquinone - PQ10 decylplastoquinone - SDH succinate dehydrogenase - UQ10 decylubiquinone (2,3-dimethoxy-5-methyl-6-decyl-1,4-benzoquinone)  相似文献   

15.
The kinetics and other characteristics of nitrate reductase (NR, EC 1.6.6.1) in cowpea [Vigna unguiculata (L.) Walp.] seedlings irradiated with biologically effective UV-B radiation (280-320 nm, 3.2 W m-2 s-1) were recorded. The in vivo and in vitro NR activities were inhibited by 34 and 41 % under UV-B treatment, respectively. Both Vmax and Km for the substrate were enhanced by UV-B radiation. The Km for nitrate increased from 1.2 to 1.7 mM after the UV-B irradiation. The change in Km for NADH was from 0.12 to 0.17 mM. The increases in Km indicate that UV-B radiation seriously changes the topology of NR, particularly with respect to the nitrate and NADH binding sites. The rate of NR turnover indicates the extent of damage inflicted by UV-B radiation on the nitrate metabolism. The half-life (t1/2) of NR was reduced from 7 to 4 h in the UV-B treated seedlings. UV-B also inhibited the kinetics of nitrate uptake by plants: its Km increased from 0.08 to 0.12 mM. This revised version was published online in June 2006 with corrections to the Cover Date.  相似文献   

16.
Using a novel coupled enzyme activity assay, with a partially purified preparation of spinach leaf nitrate reductase, the apparent Km for NADH was determined as 1.4 μM. These measurements were carried out in the presence of 0.5 mM NAD, which is within the physiological range found in the cytosol of a leaf cell. The results show that an NADH/NAD ratio of 3 × 10?3 is sufficient for a half maximal rate of nitrate reductase.  相似文献   

17.
The binding of TNP-ATP (2 or 3-O-(2,4,6-trinitrophenyl)-ATP) to cytochrome c oxidase (COX) from bovine heart and liver and to the two-subunit COX of Paracoccus denitrificans was measured by its change of fluorescence. Three binding sites, two with high (dissociation constant Kd = 0.2 µM) and one with lower affinity (Kd = 0.9 µM), were found at COX from bovine heart and liver, while the Paracoccus enzyme showed only one binding site (Kd = 3.6 µM). The binding of [35S]ATPaS was measured by equilibrium dialysis and revealed seven binding sites at the heart enzyme (Kd = 7.5 µM) and six at the liver enzyme (Kd = 12 µM). The Paracoccus enzyme had only one binding site (Kd = 16 µM). The effect of variable intraliposomal ATP/ADP ratios, but at constant total concentration of [ATP + ADP] = 5 mM, on the H+/e- stoichiometry of reconstituted COX from bovine heart and liver were studied. Above 98% ATP the H+/e- stoichiometry of the heart enzyme decreased to about half of the value measured at 100% ATP. In contrast, the H+/e- stoichiometry of the liver enzyme was not influenced by the ATP/ADP ratio. It is suggested that high intramitochondrial ATP/ADP ratios, corresponding to low cellular work load, will decrease the efficiency of energy transduction and result in elevated thermogenesis for the maintenance of body temperature. (Mol Cell Biochem 174: 131–135, 1997)  相似文献   

18.
Tania Bizouarn  Tina Bhakta 《BBA》2005,1708(3):404-410
Transhydrogenase (E.C. 1.6.1.1) couples the redox reaction between NAD(H) and NADP(H) to the transport of protons across a membrane. The enzyme is composed of three components. The dI and dIII components, which house the binding site for NAD(H) and NADP(H), respectively, are peripheral to the membrane, and dII spans the membrane. We have estimated dissociation constants (Kd values) for NADPH (0.87 μM), NADP+ (16 μM), NADH (50 μM), and NAD+ (100-500 μM) for intact, detergent-dispersed transhydrogenase from Escherichia coli using micro-calorimetry. This is the first complete set of dissociation constants of the physiological nucleotides for any intact transhydrogenase. The Kd values for NAD+ and NADH are similar to those previously reported with isolated dI, but the Kd values for NADP+ and NADPH are much larger than those previously reported with isolated dIII. There is negative co-operativity between the binding sites of the intact, detergent-dispersed transhydrogenase when both nucleotides are reduced or both are oxidised.  相似文献   

19.
A particulate (Na + K)-ATPase preparation from dog kidney bound [48V]-ortho-vanadate rapidly at 37°C through a divalent cation-dependent process. In the presence of 3 mM MgCl2 theK d was 96 nM; substituting MnCl2 decreased theK d to 12 nM but the maximal binding remained the same, 2.8 nmol per mg protein, consistent with 1 mol vanadate per functional enzyme complex. Adding KCl in the presence of MgCl2 increased binding, with aK 0.5 for KCl near 0.5 mM; the increased binding was associated with a drop inK d for vanadate to 11 nM but with no change in maximal binding. Adding NaCl in the presence of MgCl2 decreased binding markedly, with anI 50 for NaCl of 7 mM. However, in the presence of MnCl2 neither KCl nor NaCl affected vanadate binding appreciably. Both the nonhydrolyzable, ,-imido analog of ATP and nitrophenyl phosphate, a substrate for the K-phosphatase reaction that this enzyme also catalyzes, decreased vanadate binding at concentrations consistent with their acting at the low-affinity substrate site of the enzyme; the presence of KCl increased the concentration of each required to decrease vanadate binding. Oligomycin decreased vanadate binding in the presence of MgCl2, whereas dimethyl sulfoxide and ouabain increased it. With inside-out membrane vesicles from red blood cells vanadate inhibited both the K-phosphatase and (Na + K)-ATPase reactions; however, with the K-phosphatase reaction extravesicular K+ (corresponding to intracellular K+) both stimulated catalysis and augmented vanadate inhibition, whereas with the (Na + K)-ATPase reaction intravesicular K+ (corresponding to extracellular K+) both stimulated catalysis and augmented vanadate binding.  相似文献   

20.
Nicotinamide adenine dinucleotide-specific glutamate dehydrogenase (NAD-GDH; EC 1.4.1.3) from Amphibacillus xylanus DSM 6626 was enriched 100-fold to homogeneity. The molecular mass was determined by native polyacrylamide electrophoresis and by gel filtration to be 260 kDa (±25 kDa); the enzyme was composed of identical subunits of 45 (±5) kDa, indicating that the native enzyme has a hexameric structure. NAD-GDH was highly specific for the coenzyme NAD(H) and catalyzed both the formation and the oxidation of glutamate. Apparent K m -values of 56 mM glutamate, 0.35 mM NAD (oxidative deamination) and 6.7 mM 2-oxoglutaric acid, 42 mM NH4Cl and 0.036 mM NADH (reductive amination) were measured. The enzyme was unusually resistant towards variation of pH, chaotropic agents, organic solvents, and was stable at elevated temperature, retaining 50% activity after 120 min incubation at 85°C.  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