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.
The kinetic mechanism of action of octopine dehydrogenase was investigated. This enzyme catalyses the reversible dehydrogenation of D-octopine to L-arginine and pyruvate, in the presence of nicotinamide-adenine dinucleotide. Initial velocity and product inhibition studies were carried out in both directions. Most of the results are consistent with a bi-ter sequential mechanism where NAD+ binds first to the enzyme followed by D-octopine, and the products are released in the order L-arginine, pyruvate and NADH. Various kinetic parameters were determined for each reactant at 33 degrees C, at pH 9.6 for NAD reduction, at pH 6.6 for NADH oxidation.  相似文献   

2.
The kinetic scheme of octopine dehydrogenase of Pecten maximus L., a monomeric enzyme obeying a bi-ter sequential mechanism, was completed, essentially in the forward reaction, by steady-state studies over a wide range of substrate concentration at pH 7.0. Deviation from the Michaelis-Menten behavior with respect to NAD+ and other significant kinetic data led us to ascribe for octopine dehydrogenase mechanism the mnemonical enzyme concept. In addition, another regulatory behavior can be envisaged involving the formation of two dead-end complexes enzyme.NADH.D-octopine and enzyme.NAD+.pyruvate.L-arginine.  相似文献   

3.
The kinetics of pyruvate reduction by lactate dehydrogenase from Phycomyces blakesleeanus NRRL 1555 (-) have been determined at pH 6.0. Initial rate studies performed in the pyruvate reduction direction suggest that a sequential mechanism is operating. Product inhibition studies with NAD+ and L(+)-lactate are consistent with an ordered sequential mechanism if we considered that NAD+ mimics the NADH that binds cooperatively on the enzyme and also the existence of dead-end complex responsible for substrate inhibition by pyruvate at this pH value.  相似文献   

4.
Alanine dehydrogenase (L-alanine: NAD+ oxidoreductase, deaminating) was simply purified to homogeneity from a thermophile, Bacillus sphaericus DSM 462, by ammonium sulfate fractionation, red-Sepharose 4B chromatography and preparative slab gel electrophoresis. The enzyme had a molecular mass of about 230 kDa and consisted of six subunits with an identical molecular mass of 38 kDa. The enzyme was much more thermostable than that from a mesophile, B. sphaericus, and retained its full activity upon heating at 75 degrees C for at least 60 min and with incubation in pH 5.5-9.5 at 75 degrees C for 10 min. The enzyme can be stored without loss of its activity in a frozen state (-20 degrees C, at pH 7.2) for over 5 months. The optimum pH for the L-alanine deamination and pyruvate amination were around 10.5 and 8.2, respectively. The enzyme exclusively catalyzed the oxidative deamination of L-alanine in the presence of NAD+, but showed low amino acceptor specificity; hydroxypyruvate, oxaloacetate, 2-oxobutyrate and 3-fluoropyruvate are also aminated as well as pyruvate in the presence of NADH and ammonia. Initial velocity and product inhibition studies showed that the reductive amination proceeded through a sequential mechanism containing partially random binding. NADH binds first to the enzyme, and then pyruvate and ammonia bind in a random fashion. The products are sequentially released from the enzyme in the order L-alanine then NAD+. A dead-end inhibition by the formation of an abortive ternary complex which consists of the enzyme, NAD+ and pyruvate was included in the reaction. A possible role of the dead-end inhibition is to prevent the enzyme from functioning in the L-alanine synthesis. The Michaelis constants for the substrates were as follows: NADH, 0.10 mM; pyruvate, 0.50 mM; ammonia, 38.0 mM; L-alanine, 10.5 mM and NAD+, 0.26 mM.  相似文献   

5.
In the direction of reductive condensation of alpha-ketoglutarate and lysine, saccharopine dehydrogenase (N6-(glutar-2-yl)-L-lysine:NAD oxidoreductase (lysine-forming) is inhibited by high concentrations of alpha-ketoglutarate and lysine, but not by NADH. NAD+ and saccharopine show no substrate inhibition in the reverse direction. Substrate inhibition by alpha-ketoglutarate and lysine is linear uncompetitive versus NADH. However, when the inhibition is examined with alpha-ketoglutarate or lysine as the variable substrate, the double reciprocal plots show a family of curved lines concave up. The curvature is more pronounced with increasing concentrations of the inhibitory substrate, suggesting an interaction of variable substrate with the enzyme form carrying the inhibitory substrate. These inhibition patterns, the lack of interaction of structural analogs of lysine such as ornithine and norleucine with the E-NAD+ complex (Fujioka M., and Nakatani, Y. (1972) Eur. J. Biochem. 25, 301-307), the identity of values of inhibition constants of alpha-ketoglutarate and lysine obtained with either one as the substrate inhibitor, and the substrate inhibition data in the presence of a reaction product, NAD+, are consistent with the mechanism that substrate inhibition results from the formation of a dead-end E-NAD+-alpha-ketoglutarate complex followed by the addition of lysine to this abortive complex.  相似文献   

6.
Octopine dehydrogenase (OcDH) from the adductor muscle of the great scallop, Pecten maximus, catalyzes the NADH dependent, reductive condensation of L-arginine and pyruvate to octopine, NAD+, and water during escape swimming and/or subsequent recovery. The structure of OcDH was recently solved and a reaction mechanism was proposed which implied an ordered binding of NADH, L-arginine and finally pyruvate. Here, the order of substrate binding as well as the underlying conformational changes were investigated by NMR confirming the model derived from the crystal structures. Furthermore, the crystal structure of the OcDH/NADH/agmatine complex was determined which suggests a key role of the side chain of L-arginine in protein cataylsis. Thus, the order of substrate binding to OcDH as well as the molecular signals involved in octopine formation can now be described in molecular detail.  相似文献   

7.
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   

8.
The occ and noc regions of pTiAch5 (octopine) and pTiC58 (nopaline) Ti plasmids are responsible for the catabolic utilization of octopine and nopaline in Agrobacterium spp. The first enzymatic step is the oxidative cleavage into L-arginine and pyruvate or 2-ketoglutarate, respectively, by membrane-bound opine oxidases requiring two polypeptides (subunits B and A) for function. The DNA sequences showed that the subunits of pTiAch5 and pTiC58 are related, but none of the proteins revealed significant similarities to the biosynthetic enzymes expressed in transformed plant cells. The four proteins had no extensive overall similarity to other proteins, but the 35 N-terminal amino acids contained motifs found in many enzymes utilizing flavin adenine dinucleotide, flavin mononucleotide, or NAD(P)+ as cofactors. However, the activities were completely independent of added cofactors, and the nature of the electron acceptor remained unclear. Membrane solubilization led to complete loss of enzyme activity. The nopaline oxidase accepted nopaline and octopine (Vmax ratio, 5:1) with similar Km values (1.1 mM). The octopine oxidase had high activity with octopine (Km = 1 mM) and barely detectable activity with nopaline. The subunits from the occ and the noc regions were exchangeable. The combinations ooxB-noxA and noxB-ooxA both produced active enzymes which oxidized octopine and nopaline at similar rates, suggesting that both subunits contributed to the substrate specificity. These experiments also showed that the formation of functional enzyme required close proximity of the subunit genes on the same plasmid and that even a reversal of the gene order (A-B instead of B-A) led to reduced activity.  相似文献   

9.
The preceding paper in this journal has reported that pyruvate could be substituted for 2-oxo-glutarate as a substrate of saccharopine dehydrogenase [epsilon-N-(L-glutaryl-2)-L-lysine:NAD oxidoreductase (L-lysine-forming) in the direction of reductive condensation. In the present communication, the kinetic mechanism of saccharopine dehydrogenase reaction with NADH, L-lysine and pyruvate as reactants is reported. The results of initial velocity study, inhibition studies with lysine analogs and a reaction product, NAD+, are consistent with an ordered mechanism with the coenzyme binding first and pyruvate last. The reaction mechanism is at variance with that of the normal reaction in which 2-oxoglutarate is the substrate, in that the order of addition of the amino and oxo acid substrates is reversed. This fact suggests that there exists a small degree of randomness in the binding of amino and oxo acid substrates. From a product inhibition study, NAD+ was shown to be the last reactant released. Saccharopine [epsilon-N-(L-glutaryl-2)-L-lysine] was found to act as a potent dead-end inhibitor of the condensation reactions (of lysine and 2-oxoglutarate, and of lysine and pyruvate) by forming an abortive E. NADH. saccharopine complex.  相似文献   

10.
The reduction of both NO2- and hydroxylamine by the NADH-dependent nitrite reductase of Escherichia coli K 12 (EC 1.6.6.4) appears to follow Michaelis-Menten kinetics over a wide range of NADH concentrations. Substrate inhibition can, however, be detected at low concentrations of the product NAD+. In addition, NAD+ displays mixed product inhibition with respect to NADH and mixed or uncompetitive inhibition with respect to hydroxylamine. These inhibition characteristics are consistent with a mechanism in which hydroxylamine binds during catalysis to a different enzyme form from that generated when NAD+ is released. The apparent maximum velocity with NADH as varied substrate increases as the NAD+ concentration increases from 0.05 to 0.7 mM with 1 mM-NO2- or 100 mM-hydroxylamine as oxidized substrate. This increase is more marked for hydroxylamine reduction than for NO2- reduction. Models incorporating only one binding site for NAD can account for the variation in the Michaelis-Menten parameters for both NADH and hydroxylamine with [NAD+] for hydroxylamine reduction. According to these models, activation of the reaction occurs by reversal of an over-reduction of the enzyme by NADH. If the observed activation of the enzyme by NAD+ derives both from activation of the generation of the enzyme-hydroxylamine complex from the enzyme-NO2- complex during NO2- reduction and from activation of the reduction of the enzyme-hydroxylamine complex to form NH4+, then the variation of Vapp. for NO2- or hydroxylamine with [NAD+] is consistent with the occurrence of the same enzyme-hydroxylamine complex as an intermediate in both reactions.  相似文献   

11.
The formation of the ternary complex of lactate dehydrogenase (L-lactate:NAD+ oxidoreductase, EC 1.1.1.27) from pig heart and skeletal muscle with the adduct of pyruvate to NAD", spin-labeled at N6 was studied by ultraviolet spectroscopy and ESR techniques. According to ultraviolet measurements we found identical binding characteristics for the natural coenzyme and its spin-labeled analog. The rate by which the ESR signal of free spin-labeled NAD+ decreased upon addition of pyruvate to the binary complexes was substantially different in the two isozymes. With the heart type an initial drop followed by a further linear decrease, zero order in the enzyme and coenzyme concentration was observed. In case of the skeletal muscle isozyme no immediate reaction and a first order process occurred. The initial reaction can be attributed to a non-covalent enzyme/spin-labeled NAD+/pyruvate complex with a dissociation constant for pyruvate of 11 +/- 1 mM, thus explaining the well-known substrate inhibition in the heart isozyme above 2 mM pyruvate. The further reaction is then determined by the buffer dependent enolization of pyruvate. In the muscle isozyme formation of the covalent adduct is not assisted by prior binding of pyruvate in a non-covalent ternary complex and therefore the rate depends on the binary complex concentration.  相似文献   

12.
Two octopine dehydrogenases in crown-gall tumor tissue   总被引:1,自引:0,他引:1  
Extracts from four crown-gall tumor tissue culture lines, originally induced by two octopine-type strains of Agrobacterium on three plant species, converted l-arginine-[5-3H] to a compound which co-migrated with octopine on electrophoresis. Synthesis showed dependence on added pyruvate and reduced pyridine nucleotide. Both NADH and NADPH were active and mixtures of the two coenzymes, when tested with Vinca strain W1 tumor extracts, were more effective than either coenzyme at comparable concentrations. Addition of an NADH-consuming enzyme system to reaction mixtures containing NADPH had little effect on this activity. Products formed by Vinca rosea strain W1 tumor extracts and Phaseolus vulgaris strain B6 tumor extracts in reaction mixtures containing pyruvate plus NADH or NADPH co-eluted with unlabeled octopine on ion exchange chromatography. The product from the Vinca reaction mixtures co-migrated with an octopine standard in three TLC systems. Permanganate treatment of the enzymatically formed tritiated product and of unlabeled octopine gave compounds with Rf, similar to arginine and γ-guanidinobutyric acid, the products expected from permanganate degradation of octopine. The Vinca W1 extracts catalyzed the oxidative cleavage of octopine, with the formation of arginine, in the presence of NAD or NADP. Two octopine dehydrogenases were concluded to be present in these tissues, one dependent on NAD, the second on NADP.  相似文献   

13.
This work reports on the interaction of the fluorescent nicotinamide 1,N6-ethenoadenine dinucleotide (epsilonNAD+) with horse liver alcohol dehydrogenase, octopine dehydrogenase, and glyceraldehyde-3-phosphate dehydrogenase from different sources (yeast, lobster muscle, and rabbit muscle). The coenzyme fluorescence is enhanced by a factor of 10-13 in all systems investigated. It is shown that this enhancement cannot be due to changes in the polarity of the environment upon binding, and that it must be rather ascribed to structural properties of the bound coenzyme. Although dynamic factors could also be important for inducing changes in the quantum yield of epsilonNAD+ fluorescence, the close similarity of the fluorescence enhancement factor in all cases investigated indicates that the conformation of bound coenzyme is rather invariant in the different enzyme systems and overwhelmingly shifted toward an open form. Dissociation constants for epsilonNAD+-dehydrogenases complexes can be determined by monitoring the coenzyme fluorescence enhancement or the protein fluorescence quenching. In the case of yeast glyceraldehyde-3-phosphate dehydrogenase at pH 7.0 and t = 20 degrees the binding plots obtained by the two methods are coincident, and show no cooperativity. The affinity of epsilonNAD+ is generally lower than that of NAD+, although epsilonNAD+ maintains most of the binding characteristics of NAD+. For example, it forms a tight complex with horse liver alcohol dehydrogenase and pyrazole, and with octopine dehydrogenase saturated by L-arginine and pyruvate. One major difference in the binding behavior of NAD+ and epsilonNAD+ seems to be present in the muscle glyceraldehyde-3-phosphate dehydrogenase. In fact, no difference was found for epsilon NAD+ between the affinities of the third and fourth binding sites. The results and implications of this work are compared with those obtained recently by other authors.  相似文献   

14.
Octopine dehydrogenase (ODH) is electrophoretically polymorphic in the gastropod mollusk Strombus luhuanus. The frequencies of the six electrophoretic phenotypes in the Heron Island population, together with the molecular weight values of 38,000 obtained for each of the three forms of the enzyme, demonstrate that the monomeric enzyme is encoded by three codominant alleles at a single locus. The purified allozymes are indistinguishable in terms of K m values for substrates, product inhibition by octopine and NAD, pH optima, and substrate inhibition by pyruvate. No statistically significant correlations were found between the ODH phenotype and the maximum activities of ODH or alanopine dehydrogenase, the capacity for anaerobic muscle work, or the accumulation of octopine or strombine/alanopine during exercise. It would appear that the ODH allozymes may be functionally equivalent both in vitro and in vivo.  相似文献   

15.
The steady state kinetic properties of a simple model for an enzyme catalyzed group transfer reaction between two substrates have been calculated. One substrate is assumed to bind slowly and the other rapidly to the enzyme. Apparent substrate inhibition or substrate activation by the rapidly binding substrate may result if the slowly binding substrate binds at unequal rates to the free enzyme and to the complex between the enzyme and the rapidly binding substrate. Competitive inhibition by each product with respect to its structurally analogous substrate is to be expected if both substrates are in rapid equilibrium with their enzyme-substrate complexes. This product inhibition pattern, however, may also be observed when one substrate binds slowly. Noncompetitive inhibition with respect to the rapidly binding substrate by its structurally analogous product may result if the slowly binding substrate binds more slowly to the enzyme-product complex than to the free enzyme. Inhibition by substrate analogs which are not products should follow the same rules as inhibition by products. Thus substrate analog inhibition experiments are not particularly informative. The form of inhibition by "transition state analog" inhibitors should reveal which substrate binds slowly. There is no sharp conceptual distinction between ordered and random "kinetic mechanisms". I therefore suggest that the use of these concepts should be abandoned.  相似文献   

16.
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   

17.
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   

18.
The kinetic mechanism of homoisocitrate dehydrogenase from Saccharomyces cerevisiae was determined using initial velocity studies in the absence and presence of product and dead end inhibitors in both reaction directions. Data suggest a steady state random kinetic mechanism. The dissociation constant of the Mg-homoisocitrate complex (MgHIc) was estimated to be 11 +/- 2 mM as measured using Mg2+ as a shift reagent. Initial velocity data indicate the MgHIc complex is the reactant in the direction of oxidative decarboxylation, while in the reverse reaction direction, the enzyme likely binds uncomplexed Mg2+ and alpha-ketoadipate. Curvature is observed in the double-reciprocal plots for product inhibition by NADH and the dead-end inhibition by 3-acetylpyridine adenine dinucleotide phosphate when MgHIc is the varied substrate. At low concentrations of MgHIc, the inhibition by both nucleotides is competitive, but as the MgHIc concentration increases, the inhibition changes to uncompetitive, consistent with a steady state random mechanism with preferred binding of MgHIc before NAD. Release of product is preferred and ordered with respect to CO2, alpha-ketoadipate, and NADH. Isocitrate is a slow substrate with a rate (V/E(t)) 216-fold slower than that measured with HIc. In contrast to HIc, the uncomplexed form of isocitrate and Mg2+ bind to the enzyme. The kinetic mechanism in the direction of oxidative decarboxylation of isocitrate, on the basis of initial velocity studies in the absence and presence of dead-end inhibitors, suggests random addition of NAD and isocitrate with Mg2+ binding before isocitrate in rapid equilibrium, and the mechanism approximates rapid equilibrium random. The Keq for the overall reaction measured directly using the change in NADH as a probe is 0.45 M.  相似文献   

19.
Steady state initial velocity studies were carried out to determine the kinetic mechanism of human liver aldehyde dehydrogenase. Intersecting double reciprocal plots obtained in the absence of inhibitors demonstrated that the dehydrogenase reaction proceeded by sequential addition of both substrates prior to release of products. Dead end inhibition patterns obtained with coenzyme and substrate analogues (e.g. thionicotinamide-AD+ and chloral hydrate) indicated that NAD+ and aldehyde can bind in random fashion. The patterns of inhibition by the product NADH and of substrate inhibition by glyceraldehyde were also consistent with this mechanism. However, comparisons between kinetic constants associated with the dehydrogenase and esterase activities of this enzyme suggested that most of the dehydrogenase reaction flux proceeds via formation of an initial binary NAD+-enzyme complex over a wide range of substrate and coenzyme concentrations.  相似文献   

20.
Tauropine dehydrogenase (tauropine:NAD oxidoreductase) was purified from the shell adductor muscle of the ormer, Haliotis lamellosa. The enzyme was found to utilize stoichiometrically NADH as co-enzyme and pyruvate and taurine as substrates producing tauropine [rhodoic acid; N-(D-1-carboxyethyl)-taurine]. The enzyme was purified to a specific activity of 463 units/mg protein using a combination of ammonium sulphate fractionation, ion-exchange and affinity chromatography. The relative molecular mass was 38,000 +/- 1000 when assessed by gel filtration on Ultrogel AcA 54 and 42,000 +/- 150 by electrophoresis on 5-10% polyacrylamide gels in the presence of 1% sodium dodecyl sulphate; the data suggest a monomeric structure. Tauropine and pyruvate were found to be the preferred substrates. Among the amino acids tested for activity with the enzyme, only alanine is used as an alternative substrate, but with a rate less than 6% of the enzyme activity with taurine. Of the oxo acids tested, 2-oxobutyrate and 2-oxovalerate were also found to be substrates. Apparent Km values for the substrates NADH, pyruvate and taurine are 0.022 +/- 0.003 mM, 0.64 +/- 0.07 mM and 64.7 +/- 5.4 mM, respectively, at pH 7.0 and for the products, NAD+ and tauropine, are 0.29 +/- 0.01 mM and 9.04 +/- 1.27 mM, respectively, at pH 8.3. Apparent Km values for both pyruvate and taurine decrease with increasing co-substrate (taurine or pyruvate) concentration. NAD+ and tauropine were found to be product inhibitors of the forward reaction. NAD+ was a competitive inhibitor of NADH, whereas tauropine gave a mixed type of inhibition with respect to pyruvate and taurine. Succinate was found to inhibit non-competitively with respect to taurine and pyruvate with an apparent Ki value in the physiological range of this anaerobic end product. The inhibition by L-lactate, not an end product in the ormer, was competitive with respect to pyruvate. The physiological role or tauropine dehydrogenase during anaerobiosis is discussed.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号