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 15 毫秒
1.
Malate dehydrogenase may interfere with the assay of NAD malic enzyme, as NADH is formed during the conversion of malate to oxaloacetate. During the present study, two additional effects of malate dehydrogenase were investigated; they are evident only if the malate dehydrogenase reaction is allowed to reach equilibrium prior to initiating the malic enzyme reaction. One of these (Outlaw, Manchester 1980 Plant Physiol 65: 1136-1138) might cause an underestimation of NAD reduction by malic enzyme due to the oxidation of NADH during reversal of the malate dehydrogenase reaction. A second effect may result in overestimation of malic enzyme activity, as Mn2+-catalyzed oxaloacetate decarboxylation causes continuing net NADH formation via malate dehydrogenase. These effects were studied by assaying the activity of a partially purified preparation of Amaranthus retroflexus NAD malic enzyme in the presence or absence of purified NAD malate dehydrogenase.  相似文献   

2.
1. When [2-(14)C]pyruvate is injected into rats the C3-position of liver glutamate becomes more heavily labelled than the C2-position, thus establishing that oxaloacetate and fumarate are not in equilibrium in rat liver mitochondria in vivo. The amount of disequilibrium was shown to be simply related to the value that the C3-label/C2-label ratio would have were no label recycled. This ratio, z, was calculated for post-absorptive rats in environmental temperatures of 20 degrees and 30 degrees C from determinations of the distribution of label within glutamate 1, 3 and 10min after intravenous injection of [2-(14)C]pyruvate. The values of z (best estimate and range) were 1.65 (1.60-1.69) in rats at 20 degrees C and 2.43 (2.23-2.63) in rats at 30 degrees C. These values of z imply the following rates of interconversion in mitochondria of fumarate and oxaloacetate (in terms of the oxaloacetate-->citrate flux, R) in rats at 20 degrees C: [Formula: see text] and in rats at 30 degrees C: [Formula: see text] 2. The kinetic parameters of malate dehydrogenase and fumarate hydratase and the intramitochondrial concentrations of NAD(+) and NADH under (as far as could be judged) conditions in vivo were collated. From them and the best estimates of R now available were calculated the rates of interconversion of fumarate, malate and oxaloacetate required to give the found values of z. These rates showed that the fumarate hydratase reaction was nearly in equilibrium, but that the malate dehydrogenase reaction was considerably out of equilibrium. The calculations also led to the following conclusions. 3. In livers of rats at 20 degrees and 30 degrees C mitochondrial malate concentrations were respectively about 5 and 1.5 times mean cellular concentrations. 4. Mitochondrial oxaloacetate concentrations were less than 0.2 of the mean cellular concentrations. They were also only 0.65 and 0.55 of the equilibrium concentrations for the malate dehydrogenase reaction in rats at 20 degrees and 30 degrees C respectively. 5. Malate dehydrogenase activity was low because of the very low oxaloacetate concentrations in the mitochondria and the very small fraction of the enzyme complexed with NAD(+), i.e. in each direction one substrate concentration was very sub-optimal.  相似文献   

3.
Abstract Malate dehydrogenase from the syntrophic propionate-oxidizing bacterium strain MPOB was purified 42-fold. The native enzyme had an apparent molecular mass of 68 kDa and consisted of two subunits of 35 kDa. The enzyme exhibited maximum activity with oxaloacetate at pH 8.5 and 60 °C. The K m for oxaloacetate was 50 μM and for NADH 30 μM. The K m values for l-malate and NAD were 4 and 1.1 mM, respectively. Substrate inhibition was found at oxaloacetate concentrations higher than 250 μM. The N-terminal amino acid sequence of the enzyme was similar to the sequences of a variety of other malate dehydrogenases from plants, animals and micro-organisms.  相似文献   

4.
Malate dehydrogenase from Sulfolobus acidocaldarius has been purified 240-fold to apparent electrophoretic homogeneity. The enzyme shows a specific activity of 277 U/mg and crystallizes readily. The relative molecular mass of the native enzyme is estimated as 128,500 by ultracentrifugation. After cross-linking a relative molecular mass of 134,000 is found by sodium dodecyl sulfate gel electrophoresis. Malate dehydrogenase from S. acidocaldarius is composed of four subunits of identical size with a relative molecular mass of 34,000. Active-enzyme sedimentation in the analytical ultracentrifuge indicates that the tetramer is the catalytically active species. Kinetic studies in the direction of oxaloacetate reduction showed a Km for NADH of 4.1 microM and a Km for oxaloacetate of 52 microM. Oxaloacetate exhibits substrate inhibition at higher concentrations, L-malate, NAD and NADP were found to be product inhibitors. The enzymatic activity is inhibited by 2-oxoglutarate but not by the adenosine nucleotides AMP, ADP and ATP. Only low activity is detected in the direction of malate oxidation. Malate dehydrogenase from S. acidocaldarius utilizes both NADH and NADPH to reduce oxaloacetate. The enzyme shows A-side stereospecificity for both nicotinamide dinucleotides.  相似文献   

5.
Exogenous NAD+ stimulated the rotenone-resistant oxidation of all the NAD+-linked tricarboxylic acid-cycle substrates in mitochondria from Jerusalem artichoke (Helianthus tuberosus L.) tubers. The stimulation was not removed by the addition of EGTA, which is known to inhibit the oxidation of exogenous NADH. It is therefore concluded that added NAD+ gains access to the matrix space and stimulates oxidation by the rotenone-resistant NADH dehydrogenase located on the matrix surface of the inner membrane. Added NAD+ stimulated the activity of malic enzyme and displaced the equilibrium of malate dehydrogenase; both observations are consistent with entry of NAD+ into the matrix space. Analysis of products of malate oxidation showed that rotenone-resistant oxygen uptake only occurred when the concentration of oxaloacetate was low and that of NADH was high. Thus it is proposed that the concentration of NADH regulates the activity of the two internal NADH dehydrogenases. Evidence is presented to suggest that the rotenone-resistant NADH dehydrogenase is engaged under conditions of high phosphorylation potential, which restricts electron flux through the rotenone-sensitive dehydrogenase (coupled to ATP synthesis).  相似文献   

6.
The malate dehydrogenase activity (EC 1.1.1.37), present in the cytoplasm of Pisum sativum root nodules, can be separated by ion-exchange chromatography into four different fractions. Malate dehydrogenase activity present in the cytoplasm of roots elutes mainly as a single peak. During nodule development an increase in malate dehydrogenase activity per gram of material was observed. This increase occurred concomitantly with the increase in nitrogenase activity. The kinetic properties of the separated malate dehydrogenases of root nodule cytoplasm and root cytoplasm were studied. The Km values for malate (2.6 mM), NAD+ (27 microM), oxaloacetate (18 microM) and NADH (13 microM) of the dominant form of the root nodule cytoplasm are much lower than those of the dominant malate dehydrogenase root form (64 mM, 4.4 mM, 89 microM and 70 microM respectively). Binding of malate by the enzyme-NADH complex from root nodules results in an abortive complex, thereby blocking the further reduction of oxaloacetate by NADH. The dominant root malate dehydrogenase does not form the abortive complex. From the kinetic data it is concluded, first, that the root nodule forms of the enzyme are capable of catalysing at a high rate the reduction of oxaloacetate, to meet the demands for malate governed by the bacteroid and the infected plant cell. The second conclusion, drawn from the kinetic data, is that under physiological conditions the conversion of oxaloacetate can be controlled just by the malate concentration. Consequently the major root nodule forms of malate dehydrogenase are able to allow a high flux of malate production from oxaloacetate but also to establish a sufficient oxaloacetate concentration necessary for the assimilation and transport of fixed nitrogen.  相似文献   

7.
A NADH-linked oxygen-tolerant malate dehydrogenase was purified 270-fold from cell extracts of Methanospirillum hungatii. Inhibitors of the enzyme included ADP, alpha-ketoglutarate, and excess NADH. Inhibition patterns for ADP were competitive with respect to NADH and non-competitive with respect to oxalacetate. Inhibition by alpha-ketoglutarate was non-competitive with oxalacetate as variable substrate and uncompetitive with respect to NADH. alpha-Ketoglutarate is surmised to function as an end-product inhibitor of the enzyme in reactions converting oxalacetate to alpha-ketoglutarate. No enzyme activity was detected in the direction of malate conversion to oxalacetate, in keeping with a strictly biosynthetic function of the enzyme. An analysis of variance of intial rate data fit to sequential and ping-pong equations showed that a sequential mechanism was perferred. The malate dehydrogenase of M. hungatii resembles those of many other bacteria and eucaryotic cells respect to molecular weight (61,700) and reaction mechanism, but may be regulated differently.  相似文献   

8.
The particulate activities of Leishmania mexicana mexicana amastigote malate dehydrogenase (L-malate:NAD+ oxidoreductase, EC 1.1.1.37) and phosphoenolpyruvate carboxykinase (ATP:oxaloacetate carboxy-lyase (transphosphorylating) EC 4.1.1.49) have been purified to apparent electrophoretic homogeneity by hydrophobic interaction chromatography using Phenyl-Sepharose CL-4B, affinity chromatography using 5'AMP-Sepharose 4B, and gel filtration using Sephadex G-100. Malate dehydrogenase was purified 150-fold overall with a final specific activity of 1230 units/mg protein and a recovery of 63%. Phosphoenolpyruvate carboxykinase was purified 132-fold with a final specific activity of 30.3 units/mg protein and a recovery of 20%. Molecular weights determined by gel filtration and SDS-gel electrophoresis were 39 800 and 33 300 for malate dehydrogenase and 63 100 and 65 100 for phosphoenolpyruvate carboxykinase, respectively. Kinetic studies with malate dehydrogenase assayed in the direction of oxaloacetic acid reduction showed a Km(NADH) of 41 microM and a Km(oxaloacetic acid) of 39 microM. For malate oxidation there was a Km(malate) of 3.6 mM and a Km(NAD) of 0.79 mM. Oxaloacetic acid exhibited substrate inhibition at concentrations greater than 0.83 mM and malate was found to be a product inhibitor at high concentrations. However, there was no modification of enzyme activity by a number of glycolytic intermediates and cofactors, suggesting that malate dehydrogenase is not a major regulatory enzyme in L. m. mexicana. The results show that these L. m. mexicana amastigote enzymes are in several ways similar to their mammalian counterparts; nevertheless, their apparent importance and unique subcellular organization in the parasite make them potential targets for chemotherapeutic attack.  相似文献   

9.
Hydroxypyrenetrisulfonate binds to pig mitochondrial malate dehydrogenase (L-malate: NAD+ oxidoreductase, EC 1.1.1.37) in the presence and absence of coenzymes with a stoichiometry of one dye molecule/enzyme subunit. Binding is competitive with substrates and known substrate analogs as well as with squaric acid, a newly detected analog forming a ternary complex with enzyme/NAD+ similar to enzyme/NAD+/sulfite. Displacement of hydroxypyrenetrisulfonate by substrates and analogs was used to determine dissociation constants of binary and ternary complexes. Binary complexes form with dissociation constants of about 10 mM. They may be important for kinetic studies at high substrate concentrations where oxaloacetate inhibition and malate activation have been described.  相似文献   

10.
Michel Neuburger  Roland Douce 《BBA》1980,589(2):176-189
Mitochondria isolated from spinach leaves oxidized malate by both a NAD+-linked malic enzyme and malate dehydrogenase. In the presence of sodium arsenite the accumulation of oxaloacetate and pyruvate during malate oxidation was strongly dependent on the malate concentration, the pH in the reaction medium and the metabolic state condition.Bicarbonate, especially at alkaline pH, inhibited the decarboxylation of malate by the NAD+-linked malic enzyme in vitro and in vivo. Analysis of the reaction products showed that with 15 mM bicarbonate, spinach leaf mitochondria excreted almost exclusively oxaloacetate.The inhibition by oxaloacetate of malate oxidation by spinach leaf mitochondria was strongly dependent on malate concentration, the pH in the reaction medium and on the metabolic state condition.The data were interpreted as indicating that: (a) the concentration of oxaloacetate on both sides of the inner mitochondrial membrane governed the efflux and influx of oxaloacetate; (b) the NAD+/NADH ratio played an important role in regulating malate oxidation in plant mitochondria; (c) both enzymes (malate dehydrogenase and NAD+-linked malic enzyme) were competing at the level of the pyridine nucleotide pool, and (d) the NAD+-linked malic enzyme provided NADH for the reversal of the reaction catalyzed by the malate dehydrogenase.  相似文献   

11.
Initial velocity and product inhibition studies were carried out on UDP-glucose dehydrogenase (UDPglucose: NAD+ 6-oxidoreductase, EC 1.1.1.22) from beef liver to determine if the kinetics of the reaction are compatible with the established mechanism. An intersecting initial velocity pattern was observed with NAD+ as the variable substrate and UDPG as the changing fixed substrate. UDPglucuronic acid gave competitive inhibition of UDPG and non-competitive inhibition of NAD+. Inhibition by NADH gave complex patterns.Lineweaver-Burk plots of 1/upsilon versus 1/NAD+ at varied levels of NADH gave highly non-linear curves. At levels of NAD+ below 0.05 mM, non-competitive inhibition patterns were observed giving parabolic curves. Extrapolation to saturation with NAD+ showed NADH gave linear uncompetitive inhibition of UDPG if NAD+ was saturating. However, at levels of NAD+ above 0.10 mM, NADH became a competitive inhibitor of NAD+ (parabolic curves) and when NAD+ was saturating NADH gave no inhibition of UDPG. NADH was non-competitive versus UDPG when NAD+ was not saturating. These results are compatible with a mechanism in which UDPG binds first, followed by NAD+, which is reduced and released. A second mol of NAD+ is then bound, reduced, and released. The irreversible step in the reaction must occur after the release of the second mol of NADH but before the release of UDPglucuronic acid. This is apparently caused by the hydrolysis of a thiol ester between UDPglucoronic acid and the essential thiol group of the enzyme. Examination of rate equations indicated that this hydrolysis is the rate-limiting step in the overall reaction. The discontinuity in the velocities observed at high NAD+ concentrations is apparently caused by the binding of NAD+ in the active site after the release of the second mol of NADH, eliminating the NADH inhibition when NAD+ becomes saturating.  相似文献   

12.
Lactate and malate dehydrogenases (EC 1.1.1.27 and EC 1.1.1.37, respectively) were precipitated with ammonium sulfate, redissolved in 100 mM phosphate buffer, and the kinetic parameters of each enzyme determined. Lactate dehydrogenase: The enzyme preparation had a specific activity of 0.35 μmole NADH oxidized/min/mg protein for pyruvate reduction, and 0.10 μmole NAD reduced/min/mg protein for lactate oxidation. Km values for the substrates and cofactors were as follows: pyruvate = 0.51, mM; lactate = 3.8 mM; NADH = 0.011 mM; and NAD = 0.17 mM. NADPH, NADP, or d(?)-lactate would not replace NADH, NAD, or l(+)-lactate, respectively. The enzyme was relatively stable at 50 C for 45 min, but much less stable at 60 C; repeated freezing and thawing of the enzyme preparation had little effect on LDH activity. Both p-chloromercuribenzoate (p-CMB) and N-ethylmaleimide (NEM) significantly inhibited LDH activity. Polyacrylamide gel electrophoresis demonstrated the presence of at least two LDH isoenzymes in the unpurified enzyme preparation. The molecular weight was estimated at 160,000 by gel chromatography. Malate dehydrogenase: The enzyme preparation had a specific activity of 6.70 μmole NADH oxidized/min/mg protein for oxaloacetate reduction, and 0.52 μmole NAD reduced/ min/mg protein for malate oxidation. Km values for substrates and cofactors were as follows: l-malate = 1.09 mM; oxaloacetate = 0.0059 mM; NADH = 0.017 mM; and NAD = 0.180 mM. NADP and NADPH would not replace NAD and NADH, respectively, d-malate was oxidized slowly when present in high concentrations (>100 mM). Significant substrate inhibition occurred with concentrations of l-malate and oxaloacetate above 40 mM and 0.5 mM, respectively. The enzyme was unstable at temperatures above 40 C, but repeated freezing and thawing of the enzyme preparation had little effect on MDH activity. Only p-CMB inhibited MDH activity. Polyacrylamide gel electrophoresis demonstrated the presence of at least three MDH isoenzymes in the unpurified enzyme preparation, and the molecular weight was estimated at 49,000 by gel chromatography.  相似文献   

13.
Purification and characterization of mitochondrial malate dehydrogenase [EC 1.1.1.37] from unfertilized eggs of the sea urchin, Anthocidaris crassispina, are described. The purification method consisted of dextran sulfate fractionation, Blue Dextran Sepharose chromatography, Phenyl-Sepharose hydrophobic chromatography and DEAE-cellulose chromatography. The enzyme was purified 771-fold with a 7% yield from the crude extract. The purified enzyme appeared homogeneous on polyacrylamide gel electrophoresis under both native and denatured conditions. After incubation at 45 degrees C for 50 min, the enzyme lost about 90% of its activity. In the presence of NADH, however, the enzyme was protected against the heat denaturation. The native enzyme had a molecular weight of about 65,000 and probably consisted of two identical subunits. In the reduction of oxaloacetate with NADH, a broad optimum pH ranging from 8.2 to 9.4 was found with 50 mM Tris-HCl and glycine-NaOH buffers. Sodium phosphate buffer apparently activated the enzyme. The apparent Km values for oxaloacetate and NADH were 19 microM and 30 microM, respectively. The optimum pH for malate oxidation with NAD+ was 10.2 in 50 mM NaHCO3-Na2CO3 buffer. The apparent Km values for malate and NAD+ were 7.0 mM and 0.6 mM, respectively. Zinc ion, sulfite ion, p-chloromercuriphenylsulfonate and adenine nucleotides strongly inhibited the enzyme.  相似文献   

14.
Several key enzymes related to carbohydrate metabolism were assayed in Setaria digitata. In the cytosolic fraction pyruvate kinase, phosphoenolpyruvate carboxykinase, malate dehydrogenase, malic enzyme, aspartate transaminase and alanine transaminase were found. Among the TCA cycle enzymes succinate dehydrogenase, fumarate reductase, fumarase (malate dehydration), malate dehydrogenase (malate oxidation and oxaloacetate reduction) and malic enzyme (malate decarboxylation) were detected in the mitochondrial fraction. Only reduced nicotinamide adenine dinucleotide (NADH) dehydrogenase, NADH oxidase and NADH-cytochrome c reductase were found in the mitochondrial fraction. The significance of these results with respect to the metabolic capabilities of the worm are discussed.  相似文献   

15.
Cytoplasmic malate dehydrogenase from ovine liver Echinococcus granulosus protoscolices was purified 22-fold by QAE- and SP-Sephadex chromatography. The pH optimum of the enzyme was 8.0 in either Tris-HCl or barbital buffer. The κm values of oxaloacetate and NADH were 0.400 ± 0.018 and 0.410 ± 0.038 mM, respectively. The enzyme lost about 90% of its activity when heated for 2 min at 65°C. A 61.4% inhibition of the enzyme was noted at 4 mM concentration of diethyl pyrocarbonate. A 3 mM concentration of fructose 1,6-diphosphate inhibited the enzyme by 76.5%. The inhibition was non-competitive with respect to NADH with a κi value of 0.85 mM. A 75% inhibition of the enzyme was noted at 1 mM concentration of mebendazole that inhibited the enzyme upon competing with NADH with a κi value of 0.176 mM. A 2-mM concentration of citrate almost doubled the enzyme activity. The enzyme was inhibited at high concentrations of either substrate. The enzyme was not inhibited by p-hydroxymercuribenzoate or fumarate. The enzyme was absolutely specific for NADH as a cofactor. The properties of this enzyme are compared with those of the enzyme from the host liver, the cyst fluid and some other animal sources. The results are discussed in terms of the differences among the properties of the host liver, the cyst fluid and the protoscolices enzymes. The biochemical basis for the use of mebendazole in the treatment of echinococcosis is also elucidated.  相似文献   

16.
alpha-Ketoglutarate dehydrogenase has been demonstrated for the first time in cell extracts from the filamentous fungus Aspergillus niger. A minimum protein concentration of 5 mg/ml is necessary for detecting enzyme activity, but a maximum of ca. 0.060 mumol/min per mg of protein is observed only when the protein concentration is above 9 mg/ml. alpha-Ketoglutarate can partly stabilize the enzyme against dilution in the assay system. Neither bovine serum albumin nor a variety of substrates or effectors of the enzyme could stabilize the enzyme against inactivation by dilution. A kinetic analysis of the enzyme revealed Michaelis-Menten kinetics with respect to alpha-ketoglutarate, coenzyme A, and NAD. Thiamine PPi was required for maximal activity. NADH, oxaloacetate, succinate, and cis-aconitate were found to inhibit the enzyme; AMP was without effect. Monovalent cations including NH4+ were inhibitory at high concentrations (greater than 20 mM). The highest enzyme activity was found in rapidly growing mycelia (glucose-NH4+ or glucose-peptone medium). We discuss the possibility that citric acid accumulation is caused by oxaloacetate and NADH inhibition of the alpha-ketoglutarate dehydrogenase of A. niger.  相似文献   

17.
In photorespiration, leaf peroxisomes convert serine to glycerate via serine-glyoxylate aminotransferase and NADH-hydroxypyruvate reductase. We isolated intact spinach leaf peroxisomes in 0.25 M sucrose, and characterized their enzymatic conversion of serine to glycerate using physiological concentrations of substrates and coenzymes. In the presence of glycolate (glyoxylate), and NADH and NAD alone or together in physiological proportions, the rate of serine-to-glycerate conversion was enhanced and sustained by the addition of malate. The rate was similar at 1 and 5 mM serine, but was two to three times higher in 50 mM than 5 mM malate. In the presence of NAD and malate, there was 1:1 stoichiometric formation of glycerate and oxaloacetate. Addition of 1 or 5 mM glutamate resulted in a negligible enhancement of the conversion of hydroxypyruvate to glycerate. Intact peroxisomes produced glycerate from either serine or hydroxypyruvate at a rate two times higher than osmotically lysed peroxisomes. These results suggest that under physiological conditions, the peroxisomal malate dehydrogenase operates independent of aspartate-alpha-ketoglutarate aminotransferase in supplying NADH for hydroxypyruvate reduction. This supply of NADH is the rate-limiting step in the conversion of serine to glycerate. The compartmentation of hydroxypyruvate reductase and malate dehydrogenase in the peroxisomes confers a higher efficiency in the supply of NADH for hydroxypyruvate reduction under a normal, high NAD/NADH ratio in the cytosol.  相似文献   

18.
1. A biochemical comparison was made among cytoplasmic malate dehydrogenase allozymic variants from Drosophila melanogaster. Experiments were carried out on enzyme extracted from six different genotypes: three homozygotes and their respective heterozygotes. 2. The allozyme forms (MDH A, MDH B, MDH C) were indistinguishable in terms of NAD and L-malate optima, while they are distinguishable in terms of NADH and OAA saturation conditions. Activities were inhibited at concentrations greater than 0.36 and 0.40 mM NADH for BB and AA, CC, respectively, while in relation to OAA inhibition was observed at concentrations higher than 3 or 6 mM for the AA, CC and BB, respectively. 3. differences among genotypes were also observed in thermal stability: Km values for OAA, L-malate, NADH and NAD: and pG optima. 4. A simple method is presented for the separation of the cytoplasmic from the mitochondrial malate dehydrogenase.  相似文献   

19.
Initial velocity studies in the absence and presence of product and dead-end inhibitors suggest a steady-state random mechanism for malic enzyme in the direction of reductive carboxylation of pyruvate. For this quadreactant enzymatic reaction (Mn2+ is a pseudoreactant), initial velocity patterns were obtained under conditions in which two substrates were maintained at saturating concentrations while one reactant was varied at several fixed concentrations of the other. Data from the resulting reciprocal plots, analyzed in terms of a bireactant mechanism, are consistent with a sequential mechanism with an obligatory order of addition of metal prior to pyruvate. NAD is competitive against NADH whether pyruvate and CO2 are maintained at low or high concentrations, whereas it is noncompetitive against pyruvate and CO2. Thio-NADH, alpha-ketobutyrate, and nitrite were used as dead-end analogs of NADH, pyruvate, and CO2, respectively. Thio-NADH is competitive against NADH, whereas it is noncompetitive against pyruvate and CO2, in accordance with a random mechanism. alpha-Ketobutyrate and nitrite gave noncompetitive inhibition against all substrates. The noncompetitive patterns observed for alpha-ketobutyrate versus pyruvate and nitrite versus CO2 suggest binding of the inhibitor to both the E.Mn.NADH and E.Mn.NAD complexes. Primary deuterium isotope effects are equal on all kinetic parameters, in agreement with the random mechanism, and suggest equal off-rates for NAD from E.Mn.NAD as well as pyruvate and NADH from E.Mn.NADH.pyruvate. Data are consistent with an overall symmetry in the malic enzyme reaction in the two reaction directions with a requirement for metal bound prior to pyruvate and malate.  相似文献   

20.
Kinetic properties of purified 5-carboxymethyl-2-hydroxymuconate semialdehyde (CHMSA) dehydrogenase (EC 1.2.1.-) in the 4-hydroxyphenylacetate meta-cleavage pathway from Escherichia coli have been studied. The temperature--activity relationship for the enzyme from 27 to 45 degrees C showed an Arrhenius plot with an inflexion at 36 degrees C. When 5-carboxymethyl-2-hydroxymuconic semialdehyde and NAD were used as variable substrates, the double reciprocal plots were all linear and the lines intersected at one point below the horizontal axis, suggesting that a sequential mechanism is operating. From the replots of intercepts and slopes against reciprocal substrate concentrations were calculated Km (CHMSA) = 9.0 +/- 1.02 microM, Km (NAD) = 29.1 +/- 4.65 microM and the value for the dissociation constant of enzyme--NAD complex = 6.3 +/- 1.21 microM. ATP and the product of the reaction (NADH) acted as competitive inhibitors of the enzyme with respect to NAD. Apparent Ki values, estimated from Dixon plots, were 25.0 +/- 3.5 and 88.0 +/- 22.1 microM for NADH and ATP, respectively.  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