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

2.
Hexose-6-phosphate dehydrogenase (refers to hexose-6-phosphate dehydrogenase from any species in general) has been purified to apparent homogeneity from the teleost fish Fundulus heteroclitus. The enzyme was characterized for native (210 kDa) and subunit molecular mass (54 kDa), isoelectric point (6.65), amino acid composition, substrate specificity, and metal dependence. Glucose 6-phosphate, galactose 6-phosphate, 2-deoxyglucose 6-phosphate, glucose 6-sulfate, glucosamine 6-phosphate, and glucose were found to be substrates in the reaction with NADP+, but only glucose was a substrate when NAD+ was used as coenzyme. A unique reaction mechanism for the forward direction was found for this enzyme when glucose 6-phosphate and NADP+ were used as substrates; ordered with glucose 6-phosphate binding first. NAD+ was found to be a competitive inhibitor toward NADP+ and an uncompetitive inhibitor with regard to glucose 6-phosphate in this reaction; Vmax = 7.56 mumol/min/mg, Km(NADP+) = 1.62 microM, Km(glucose 6-phosphate) = 7.29 microM, Kia(glucose 6-phosphate) = 8.66 microM, and Ki(NAD+) = 0.49 microM. The use of alternative substrates confirmed this result. This type of reaction mechanism has not been previously reported for a dehydrogenase.  相似文献   

3.
The kinetic properties for the native forward reaction of pyruvate:NADP+ oxidoreductase from Euglena gracilis were determined. The substrate kinetics gave a pattern of a ping-pong mechanism involving a competitive substrate inhibition of CoA against pyruvate. The Km values for pyruvate, CoA, and NADP+ were estimated to be 27, 6.6, and 28 microM, respectively, and the Ki value of CoA against pyruvate was 28 microM. CO2 inhibited noncompetitively against pyruvate and NADP+, and uncompetitively against CoA. Acetyl-CoA showed a competitive inhibition with respect to pyruvate and an uncompetitive inhibition with respect to NADP+. NADPH inhibited competitively versus NADP+, noncompetitively versus CoA, and uncompetitively versus pyruvate. The kinetic behavior is consistent with a two-site ping-pong mechanism involving the substrate inhibition. From the kinetic mechanism, it is proposed that the enzyme has two catalytic sites linked by an intramolecular electron-transport chain. One of these is a thiamine pyrophosphate-containing catalytic site which reacts with pyruvate and CoA to form CO2 and acetyl-CoA, and the other site functions in the reduction of NADP+. In contrast, when methyl viologen was used as an artificial one-electron acceptor substituting for NADP+, the reaction gave a pattern characteristic of an octa uni ping-pong mechanism involving a competitive substrate inhibition of CoA against pyruvate.  相似文献   

4.
Intra- and intermolecular electron transfer processes in redox proteins   总被引:2,自引:0,他引:2  
Initial velocity and product inhibition experiments were performed to characterize the kinetic mechanism of branched chain ketoacid dehydrogenase (the branched chain complex) activity. The results were directly compared to predicted patterns for a three-site ping-pong mechanism. Product inhibition experiments confirmed that NADH is competitive versus NAD+ and isovaleryl CoA is competitive versus CoA. Furthermore, both NADH and isovaleryl CoA were uncompetitive versus ketoisovaleric acid. These results are consistent with a ping-pong mechanism and are similar to pyruvate dehydrogenase and alpha-ketoglutarate dehydrogenase. However, inhibition patterns for isovaleryl CoA versus NAD+ and NADH versus CoA are not consistent with a ping-pong mechanism. These patterns may result from a steric interaction between the flavoprotein and transacetylase subunits of the complex. To determine the kinetic mechanism of the substrates and feedback inhibitors (NADH and isovaleryl CoA) of the branched chain complex, it was necessary to define the interaction of the inhibitors at nonsaturating fixed substrate (CoA and NAD+) concentrations. While the competitive inhibition patterns were maintained, slope replots for NADH versus NAD+ at nonsaturating CoA concentrations were parabolic. This unexpected finding resembles a linear mixed type of inhibition where the inhibition is a combination of pure competitive and noncompetitive inhibition.  相似文献   

5.
Indanol dehydrogenase was purified to apparent homogeneity from monkey liver cytosol. The enzyme was a monomer with a molecular weight of 36,000 and pI of 8.7. The amino acid composition was determined. The enzyme oxidized alicyclic alcohols including transdihydrodiols of benzene and naphthalene in the presence of both NADP+ and NAD+, and reduced several xenobiotic carbonyl compounds in the presence of NADPH, the 4-pro-R hydrogen atom of which was transferred to the substrate. The results of fluorometric binding and kinetic studies are consistent with an ordered sequential mechanism with NADP+ binding first. The enzyme was inhibited competitively versus NADP+ and uncompetitively versus 1-indanol not only by chelating agents such as 1,10-phenanthroline and 2,2'-bipyridine but also by a nonchelating isomer, 4,4'-bipyridine, which suggests hydrophobic interaction of the aromatic compounds with the enzyme, which did not contain zinc. The enzyme was also inhibited by Cibacron blue dye, synthetic estrogens, and delta 4-3-ketosteroids. The inhibition by Cibacron blue was competitive versus NADP+ and noncompetitive versus 1-indanol, whereas those by hexestrol, medroxyprogesterone acetate, and progesterone were uncompetitive versus NADP+ and competitive versus 1-indanol, corraborating the ordered addition of the coenzyme prior to 1-indanol.  相似文献   

6.
Initial velocity, product inhibition, and substrate inhibition studies suggest that the endogenous lactate dehydrogenase activity of duck epsilon-crystallin follows an order Bi-Bi sequential mechanism. In the forward reaction (pyruvate reduction), substrate inhibition by pyruvate was uncompetitive with inhibition constant of 6.7 +/- 1.7 mM. In the reverse reaction (lactate oxidation), substrate inhibition by L-lactate was uncompetitive with inhibition constant of 158 +/- 25 mM. The cause of these inhibitions may be due to epsilon-crystallin-NAD(+)-pyruvate and epsilon-crystallin-NADH-L-lactate abortive ternary complex formation as suggested by the multiple inhibition studies. Pyruvate binds to free enzyme very poorly, with a very large dissociation constant. Bromopyruvate, fluoropyruvate, pyruvate methyl ester, and pyruvate ethyl ester are alternative substrates for pyruvate. 3-Acetylpyridine adenine dinucleotide, nicotinamide 1,N6-ethenoadenine dinucleotide, and nicotinamide hypoxanthine dinucleotide serve as alternative coenzymes for epsilon-crystallin. All the above alternative substrates or coenzymes showed an intersecting initial-velocity pattern conforming to the order Bi--Bi kinetic mechanism. Nicotinic acid adenine dinucleotide, thionicotinamide adenine dinucleotide, and 3-aminopyridine adenine dinucleotide acted as inhibitors for this enzymatic crystallin. The inhibitors were competitive versus NAD+ and noncompetitive versus L-lactate. alpha-NAD+ was a noncompetitive inhibitor with respect to the usual beta-NAD+. D-Lactate, tartronate, and oxamate were strong dead-end inhibitors for the lactate dehydrogenase activity of epsilon-crystallin. Both D-lactate and tartronate were competitive inhibitors versus L-lactate while oxamate was a competitive inhibitor versus pyruvate. We conclude that the structural requirements for the substrate and coenzyme of epsilon-crystallin are similar to those of other dehydrogenases and that the carboxamide carbonyl group of the nicotinamide moiety is important for the coenzyme activity.  相似文献   

7.
During growth of the halophilic archaeon Haloarcula marismortui on D-xylose, a specific D-xylose dehydrogenase was induced. The enzyme was purified to homogeneity. It constitutes a homotetramer of about 175 kDa and catalyzed the oxidation of xylose with both NADP+ and NAD+ as cosubstrates with 10-fold higher affinity for NADP+. In addition to D-xylose, D-ribose was oxidized at similar kinetic constants, whereas D-glucose was used with about 70-fold lower catalytic efficiency (kcat/Km). With the N-terminal amino acid sequence of the subunit, an open reading frame (ORF)-coding for a 39.9-kDA protein-was identified in the partially sequenced genome of H. marismortui. The function of the ORF as the gene designated xdh and coding for xylose dehydrogenase was proven by its functional overexpression in Escherichia coli. The recombinant enzyme was reactivated from inclusion bodies following solubilization in urea and refolding in the presence of salts, reduced and oxidized glutathione, and substrates. Xylose dehydrogenase showed the highest sequence similarity to glucose-fructose oxidoreductase from Zymomonas mobilis and other putative bacterial and archaeal oxidoreductases. Activities of xylose isomerase and xylulose kinase, the initial reactions of xylose catabolism of most bacteria, could not be detected in xylose-grown cells of H. marismortui, and the genes that encode them, xylA and xylB, were not found in the genome of H. marismortui. Thus, we propose that this first characterized archaeal xylose dehydrogenase catalyzes the initial step in xylose degradation by H. marismortui.  相似文献   

8.
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   

9.
The kinetic mechanism of pigeon liver fatty acid synthetase action has been studied using steady state kinetic analysis. Initial velocity studies are consistent with an earlier suggestion that the enzyme catalyzes this reaction by a seven-site ping-pong mechanism. Although the range of substrate concentrations that could be used was limited by several factors, the initial velocity patterns showing the relationship between the substrates acetyl coenzyme CoA, malonyl-CoA, and NADPH appear to be a series of parallel lines, regardless of which substrate is varied at fixed levels of a second substrate. However, two of the substrates, acetyl-CoA and malonly-CoA, apparently exhibit a competitive substrate inhibition with respect to each other, but NADPH shows no inhibition of any kind. Product inhibition patterns suggest that free CoA is competitive versus acetyl-CoA and malonyl-CoA and is uncompetitive versus NADPH, and that NADP+ is competitive versus NADPH and uncompetitive versus acetyl-CoA or malonyl-CoA. These results are consistent with a seven-site ping-pong mechanism with intermediates covalently bound to 4'-phosphopantetheine (part of acyl carrier protein). Double competitive substrate inhibition by acetyl-CoA and malonyl-CoA is consistent with the rate equation derived for the over-all mechanism. The kinetic mechanism developed from these results is capable of explaining the formation of fatty acids from malonyl-CoA and NADPH alone (Katiyar, S. S., Briedis, A. V., and Porter, J. W. (1974) Arch. Biochem. Biophys. 162, 412-420) and also the formation of triacetic acid lactone from either malonyl-CoA alone or acetyl-CoA plus malonyl-CoA.  相似文献   

10.
Data on initial velocity and isotope exchange at equilibrium suggest a nonclassical ping-pong mechanism for the dihydropyrimidine dehydrogenase from pig liver. Initial velocity patterns in the absence of inhibitors appeared parallel at low reactant concentration, with substrate inhibition by NADPH that is competitive with uracil and with substrate inhibition by uracil that is uncompetitive with NADPH. The Km values for both uracil (1 microM) and NADPH (7 microM) are low. As a result, it was difficult to determine whether the initial velocity pattern in the absence of added inhibitors was parallel. Thus, the pattern was redetermined in the presence of the dead-end inhibitor 2,6-dihydroxypyridine, which binds to both sites. This treatment effectively eliminates the inhibition by both substrates and increases their Km values, giving a strictly parallel pattern. Product and dead-end inhibition patterns are consistent with a mechanism in which NADPH reduces the enzyme at site 1 and electrons are transferred to site 2 to reduce uracil to dihydrouracil. The predicted mechanism is corroborated by exchange between [14C] NADP and NADPH as well as [14C]thymine and dihydrothymine in the absence of the other substrate-product pair.  相似文献   

11.
Glucose-6-phosphate dehydrogenase from sporangiophores of Phycomyces blakesleeanus NRRL 1555 (-) was partially purified. The enzyme showed a molecular weight of 85 700 as determined by gel-filtration. NADP+ protected the enzyme from inactivation. Magnesium ions did not affect the enzyme activity. Glucose-6-phosphate dehydrogenase was specific for NADP+ as coenzyme. The reaction rates were hyperbolic functions of substrate and coenzyme concentrations. The Km values for NADP+ and glucose 6-phosphate were 39.8 and 154.4 microM, respectively. The kinetic patterns, with respect to coenzyme and substrate, indicated a sequential mechanism. NADPH was a competitive inhibitor with respect to NADP+ (Ki = 45.5 microM) and a non-competitive inhibitor with respect to glucose 6-phosphate. ATP inhibited the activity of glucose-6-phosphate dehydrogenase. The inhibition was of the linear-mixed type with respect to NADP+, the dissociation constant of the enzyme-ATP complex being 2.6 mM, and the enzyme-NADP+-ATP dissociation constant 12.8 mM.  相似文献   

12.
1. N10-Formyltetrahydrofolate dehydrogenase was purified to homogeneity from rat liver with a specific activity of 0.7--0.8 unit/mg at 25 degrees C. The enzyme is a tetramer (Mw = 413,000) composed of four similar, if not identical, substrate addition and give the Km values as 4.5 micron [(-)-N10-formyltetrahydrofolate] and 0.92 micron (NADP+) at pH 7.0. Tetrahydrofolate acts as a potent product inhibitor [Ki = 7 micron for the (-)-isomer] which is competitive with respect to N10-formyltetrahydrofolate and non-competitive with respect to NADP+. 3. Product inhibition by NADPH could not be demonstrated. This coenzyme activates N10-formyltetrahydrofolate dehydrogenase when added at concentrations, and in a ratio with NADP+, consistent with those present in rat liver in vivo. No effect of methionine, ethionine or their S-adenosyl derivatives could be demonstrated on the activity of the enzyme. 4. Hydrolysis of N10-formyltetrahydrofolate is catalysed by rat liver N10-formyltetrahydrofolate dehydrogenase at 21% of the rate of CO2 formation based on comparison of apparent Vmax. values. The Km for (-)-N10-folate is a non-competitive inhibitor of this reaction with respect to N10-formyltetrahydrofolate, with a mean Ki of 21.5 micron for the (-)-isomer. NAD+ increases the maximal rate of N10-formyltetrahydrofolate hydrolysis without affecting the Km for this substrate and decreases inhibition by tetrahydrofolate. The activator constant for NAD+ is obtained as 0.35 mM. 5. Formiminoglutamate, a product of liver histidine metabolism which accumulates in conditions of excess histidine load, is a potent inhibitor of rat liver pyruvate carboxylase, with 50% inhibition being observed at a concentration of 2.8 mM, but has no detectable effect on the activity of rat liver cytosol phosphoenolpyruvate carboxykinase measured in the direction of oxaloacetate synthesis. We propose that the observed inhibition of pyruvate carboxylase by formiminoglutamate may account in part for the toxic effect of excess histidine.  相似文献   

13.
In cell extracts of Methanosarcina barkeri, the methylcoenzyme M methylreductase system with H2 as the electron donor was inhibited by NAD+ and NADP+, but NADH and NADPH had no effect on enzyme activity. NAD+ (4 and 8 mM) shifted the saturation curve for methylcoenzyme M from hyperbolic (Hill coefficient [nH] = 1.0; concentration of substrate giving half maximal velocity [Km] = 0.21 mM) to sigmoidal (nH = 1.5 and 2.0), increased Km (Km = 0.25 and 0.34 mM), and slightly decreased Vmax. Similarly NADP+ at 4m and 8 mM increased nH to 1.6 and 1.85 respectively, but the Km values (0.3 and 0.56 mM) indicated that NADP+ was a more efficient inhibitor than NAD+.  相似文献   

14.
Saccharomyces cerevisiae mitochondria contain an NADH:Q6 oxidoreductase (internal NADH dehydrogenase) encoded by NDI1 gene in chromosome XIII. This enzyme catalyzes the transfer of electrons from NADH to ubiquinone without the translocation of protons across the membrane. From a structural point of view, the mature enzyme has a single subunit of 53 kDa with FAD as the only prosthetic group. Due to the fact that S. cerevisiae cells lack complex I, the expression of this protein is essential for cell growth under respiratory conditions. The results reported in this work show that the internal NADH dehydrogenase follows a ping-pong mechanism, with a Km for NADH of 9.4 microM and a Km for oxidized 2,6-dichorophenolindophenol (DCPIP) of 6.2 microM. NAD+, one of the products of the reaction, did not inhibit the enzyme while the other product, reduced DCPIP, inhibited the enzyme with a Ki of 11.5 microM. Two dead-end inhibitors, AMP and flavone, were used to further characterize the kinetic mechanism of the enzyme. AMP was a linear competitive inhibitor of NADH (Ki = 5.5 mM) and a linear uncompetitive inhibitor of oxidized DCPIP (Ki = 11.5 mM), in agreement with the ping-pong mechanism. On the other hand, flavone was a partial inhibitor displaying a hyperbolic uncompetitive inhibition regarding NADH, and a hyperbolic noncompetitive inhibition with respect to oxidized DCPIP. The apparent intercept inhibition constant (Kii = 5.4 microM) and the slope inhibition constant (Kis = 7.1 microM) were obtained by non linear regression analysis. The results indicate that the ternary complex F-DCPIPox-flavone catalyzes the reduction of DCPIP, although with lower efficiency. The effect of pH on Vmax was studied. The Vmax profile shows two groups with pKa values of 5.3 and 7.2 involved in the catalytic process.  相似文献   

15.
Benzaldehyde dehydrogenase I was purified from Acinetobacter calcoaceticus by DEAE-Sephacel, phenyl-Sepharose and f.p.l.c. gel-filtration chromatography. The enzyme was homogeneous and completely free from the isofunctional enzyme benzaldehyde dehydrogenase II, as judged by denaturing and non-denaturing polyacrylamide-gel electrophoresis. The subunit Mr value was 56,000 (determined by SDS/polyacrylamide-gel electrophoresis). Estimations of the native Mr value by gel-filtration chromatography gave values of 141,000 with a f.p.l.c. Superose 6 column, but 219,000 with Sephacryl S300. Chemical cross-linking of the enzyme subunits indicated that the enzyme is tetrameric. Benzaldehyde dehydrogenase I was activated more than 100-fold by K+, Rb+ and NH4+, and the apparent Km for K+ was 11.2 mM. The pH optimum in the presence of K+ was 9.5 and the pI of the enzyme was 5.55. The apparent Km values for benzaldehyde and NAD+ were 0.69 microM and 96 microM respectively, and the maximum velocity was approx. 110 mumol/min per mg of protein. Various substituted benzaldehydes were oxidized at significant rates, and NADP+ was also used as cofactor, although much less effectively than NAD+. Benzaldehyde dehydrogenase I had an NAD+-activated esterase activity with 4-nitrophenol acetate as substrate, and the dehydrogenase activity was inhibited by a range of thiol-blocking reagents. The absorption spectrum indicated that there was no bound cofactor or prosthetic group. Some of the properties of the enzyme are compared with those of other aldehyde dehydrogenases, specifically the very similar isofunctional enzyme benzaldehyde dehydrogenase II from the same organism.  相似文献   

16.
Abstract— The overall steady state kinetic mechanism of pyruvate dehydrogenase multienzyme complex purified from rat brain has been investigated. Initial rate patterns were a series of parallel lines regardless of which substrate was varied at several fixed concentrations of other substrates. Product inhibition patterns showed that acetyl CoA is competitive vs CoA, that NADH is competitive vs NAD, and that both acetyl CoA and NADH are uncompetitive vs pyruvate. Both acetyl CoA and NADH are noncompetitive vs NAD and CoASH, respectively. These results are inconsistent with classical 'hexa uni' ping-pong mechanisms, but are consistent with a non-classical 3-site ping-pong mechanism.  相似文献   

17.
1. Pig heart pyruvate dehydrogenase complex is inactivated by phosphorylation (MgATP2-) of an alpha-chain of the decarboxylase component. Three serine residues may be phosphorylated, one of which (site 1) is the major inactivating site. 2. The relative rates of phosphorylation are site 1 greater than 2 greater than site 3. 3. The kinetics of the inactivating phosphorylation were investigated by measuring inactivation of the complex with MgATP2-. The apparent Km for the Mg complex of ATP was 25.5 microM; ADP was a competitive inhibitor (Ki 69.8 microM) and sodium pyruvate an uncompetitive inhibitor (Ki 2.8 microM). Inactivation was accelerated by increasing concentration ratios of NADH/NAD+ and of acetyl-CoA/CoA. 4. The kinetics of additional phosphorylations (predominantly site 2 under these conditions) were investigated by measurement of 32P incorporation into non-radioactive pyruvate dehydrogenase phosphate containing 3-6% of active complex, and assumed from parrallel experiments with 32P labelling to contain 91% of protein-bound phosphate in site 1 and 9% in site 2. 5. The apparent Km for the Mg complex of ATP was 10.1 microM; ADP was a competitive inhibitor (Ki 31.5 microM) and sodium pyruvate an uncompetitive inhibitor (Ki 1.1 mM). 6. Incorporation was accelerated by increasing concentration ratios of NADH/NAD+ and of acetyl-CoA/CoA, although it was less marked at the highest ratios.  相似文献   

18.
At low concentrations of Mg2+ or Mn2+ the reaction catalyzed by isocitrate dehydrogenase from bovine adrenal cortex proceeds with a lag period which disappears as a result of the enzyme saturation with Mn2+ or Mg2+. The nu o versus D,L-isocitrate concentration curve is non-hyperbolic, which may be interpreted either by the presence of two active sites with different affinity for the substrate (K'mapp = 2.3 and 63 microM) within the enzyme molecule or by the "negative" cooperativity of these sites. The apparent Km value for NADP lies within the range of 3.6-9 microM. High concentrations of NADP inhibit isocitrate dehydrogenase (Ki = 1.3 mM). NADP.H inhibits the enzyme in a mixed manner with respect to NADP (Ki = 0.32 mM). In the presence of NADP.H the curve nu o dependence on NADP concentration shows a "negative" cooperativity between NADP binding sites. The reverse enzyme-catalyzed reaction of reductive carboxylation of 2-oxoglutarate does not exhibit any significant deviations from the Michaelis-Menten kinetics. The Km value for 2-oxoglutarate is 120 microM, while that for NADP.H is 10 microM.  相似文献   

19.
Valine dehydrogenase (VDH) from Streptomyces coelicolor A3(2) was purified from cell-free extracts to apparent homogeneity. The enzyme had an Mr 41,000 in denaturing conditions and an Mr 70,000 by gel filtration chromatography, indicating that it is composed of two identical subunits. It oxidized L-valine and L-alpha-aminobutyric acid efficiently, L-isoleucine and L-leucine less efficiently, and did not act on D-valine. It required NAD+ as cofactor and could not use NADP+. Maximum dehydrogenase activity with valine was at pH 10.5 and the maximum reductive amination activity with 2-oxoisovaleric acid and NH4Cl was at pH 9. The enzyme exhibited substrate inhibition in the forward direction and a kinetic pattern with NAD+ that was consistent with a sequential ordered mechanism with non-competitive inhibition by valine. The following Michaelis constants were calculated from these data: L-valine, 10.0 mM; NAD+, 0.17 mM; 2-oxoisovalerate, 0.6 mM; and NADH, 0.093 mM. In minimal medium, VDH activity was repressed in the presence of glucose and NH4+, or glycerol and NH4+ or asparagine, and was induced by D- and L-valine. The time required for full induction was about 24 h and the level of induction was 2- to 23-fold.  相似文献   

20.
We tested various methods of assaying the ADP-ribosyltransferase activity of cholera toxin using artificial acceptors of the ADP-ribosyl group. Any of several proteins or poly(L-arginine) could be used with [adenine-14C]NAD+ as ADP-ribosyl donor, but this method was not ideal because of the heterogeneity of potential acceptor groups and the necessity of using costly labeled NAD+. We, therefore, developed an alternative assay using a synthetic low molecular weight acceptor, 125I-N-guanyltyramine (125I-GT). 125I-GT was specifically ADP-ribosylated by thiol-treated cholera toxin or its A1 peptide in the presence of beta-NAD. ADP-ribosyl-125I-GT was quantified after separation from unreacted 125I-GT by batch absorption of the latter to cation exchange resins. Analysis of the kinetics of ADP-ribosylation of 125I-GT indicated that the reaction proceeds by a sequential rather than a ping-pong mechanism. The Km values for NAD+ and 125I-GT were 3.6 mM and 44 microM, respectively. L-Arginine was a competitive inhibitor of 125I-GT (KI = 75 mM), but was at least 1000-fold less active than 125I-GT as an ADP-ribose acceptor.  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