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.
The enzyme acetohydroxy acid synthase (AHAS, EC 4.1.3.18) catalyzes two competing reactions of physiological importance: condensation of two molecules of pyruvate to form acetolactate (AL) or condensation of pyruvate and 2-ketobutyrate to form acetohydroxybutyrate (AHB). The activity of AHAS is most frequently analyzed using the Westerfeld method, in which the acetoin formed upon decarboxylation of AL is determined by colorimetric reaction with creatine and alpha-naphthol. However, there has been confusion as to the interpretation of the results of this assay in the presence of both substrates, conditions which lead to formation of both AL and AHB. By applying this assay to enzymatically prepared samples of AL and AHB which have also been analyzed by two other independent methods, we show here that the color yield for AHB in the commonly used assay is 35-40% that for equivalent amounts of acetoin or AL. The relative color yield is not significantly affected by varying the time or temperature of various steps in the color-forming reaction. This information could in principle be used, together with an independent specific assay for AHB, to determine the composition of an AHAS product mixture; it would, however, be less accurate than a simultaneous chromatographic method.  相似文献   

2.
The rates of formation of the two alternative products of acetohydroxy acid synthase (AHAS) have been determined by a new analytical method (N. Gollop, Z. Barak, and D. M. Chipman, Anal. Biochem., 160:323-331, 1987). For each of the three distinct isozymes of AHAS in Escherichia coli and Salmonella typhimurium, a specificity ratio, R, was defined: Formula: see text, which is constant over a wide range of substrate concentrations. This is consistent with competition between pyruvate and 2-ketobutyrate for an active acetaldehyde intermediate formed irreversibly after addition of the first pyruvate moiety to the enzyme. Isozyme I showed no product preference (R = 1), whereas isozymes II and III form acetohydroxybutyrate (AHB) at approximately 180- and 60-fold faster rates, respectively, than acetolactate (AL) at equal pyruvate and 2-ketobutyrate concentrations. R values higher than 60 represent remarkably high specificity in favor of the substrate with one extra methylene group. In exponentially growing E. coli cells (under aerobic growth on glucose), which contain about 300 microM pyruvate and only 3 microM 2-ketobutyrate, AHAS I would produce almost entirely AL and only 1 to 2% AHB. However, isozymes II and III would synthesize AHB (on the pathway to Ile) and AL (on the pathway to valine-leucine) in essentially the ratio required for protein synthesis. The specificity ratio R of any AHAS isozyme was affected neither by the natural feedback inhibitors (Val, Ile) nor by the pH. On the basis of the specificities of the isozymes, the known regulation of AHAS I expression by the catabolite repression system, and the reported behavior of bacterial mutants containing single AHAS isozymes, we suggest that AHAS I enables a bacterium to cope with poor carbon sources, which lead to low endogenous pyruvate concentrations. Although AHAS II and III are well suited to producing the branched-chain amino acid precursors during growth on glucose, they would fail to provide appropriate quantities of AL when the concentration of pyruvate is relatively low.  相似文献   

3.
Acetohydroxy acid synthase (AHAS; EC 4.1.3.18) catalyzes the following two parallel, physiologically important reactions: condensation of two molecules of pyruvate to form acetolactate (AL), in the pathway to valine and leucine, and condensation of pyruvate plus 2-ketobutyrate to form acetohydroxybutyrate (AHB), in the pathway to isoleucine. We have determined the specificity ratio R with regard to these two reactions (where VAHB and VAL are rates of formation of the respective products) as follows: VAHB/VAL = R [2-ketobutyrate]/[pyruvate] for 14 enzymes from 10 procaryotic and eucaryotic organisms. Each organism considered has at least one AHAS of R greater than 20, and some appear to contain but a single biosynthetic AHAS. The implications of this for the design of the pathway are discussed. The selective pressure for high specificity for 2-ketobutyrate versus pyruvate implies that the 2-ketobutyrate concentration is much lower than the pyruvate concentration in all these organisms. It seems important for 2-ketobutyrate levels to be relatively low to avoid a variety of metabolic interferences. These results also reinforce the conclusion that biosynthetic AHAS isozymes of low R (1 to 2) are a special adaptation for heterotrophic growth on certain poor carbon sources. Two catabolic "pH 6 AL-synthesizing enzymes" are shown to be highly specific for AL formation only (R less than 0.1).  相似文献   

4.
An enzyme which catalyzes a decomposition of α-aminoisobutyrate (AIB) was purified and its kinetic properties were investigated. Michaelis constants for AIB decomposing reaction are able to be calculated by Ping Pong initial velocity equation. This enzyme catalyzes also l-alanine: α-ketobutyrate transamination as well as AIB decomposing reaction. Approximately equal values of Michaelis constants were obtained for α-ketobutyrate and pyridoxal 5′-phosphate (PLP), which are common substrates of both reactions.

In higher concentration of the enzyme, transamination between PLP and AIB or l-alanine was detected, whereas the reaction between pyridoxamine 5′-phosphate and pyruvate was not observed. These results are probably ascribed to a difference in affinity of two coenzymes for the enzyme.  相似文献   

5.
In the progress curve of the reaction of the pyruvate dehydrogenase complex, a lag phase was observed when the concentration of thiamin diphosphate was lower than usual (about 0.2-1 mM) in the enzyme assay. The length of the lag phase was dependent on thiamin diphosphate concentration, ranging from 0.2 min to 2 min as the thiamin diphosphate concentration varied from 800 nM to 22 nM. The lag phase was also observed in the elementary steps catalyzed by the pyruvate dehydrogenase component. A Km value of 107 nM was found for thiamin diphosphate with respect to the steady-state reaction rate following the lag phase. The pre-steady-state kinetic data indicate that the resulting lag phase was the consequence of a slow holoenzyme formation from apoenzyme and thiamin diphosphate. The thiamin diphosphate can bind to the pyruvate dehydrogenase complex in the absence of pyruvate, but the presence of 2 mM pyruvate increases the rate constant of binding from 1.4 X 10(4) M-1 S-1 to 1.3 X 10(5) M-1 S-1 and decreases the rate constant of dissociation from 2.3 X 10(-2) S-1 to 4.1 X 10(-3) S-1. On the other hand, the effect of pyruvate on the thiamin diphosphate binding revealed the existence of a thiamin-diphosphate-independent pyruvate-binding site in the pyruvate dehydrogenase complex. Direct evidence was also obtained with fluorescence techniques for the existence of this binding site and the dissociation constant of pyruvate was found to be 0.38 mM. On the basis of these data we have proposed a random mechanism for the binding of pyruvate and thiamin diphosphate to the complex. Binding of substrates to the enzyme complex caused an increase in the fluorescence of the dansylaziridine-labelled pyruvate dehydrogenase complex, showing that binding of substrates to the complex is accompanied by structural changes.  相似文献   

6.
Fo.F1-ATP synthase in inside-out coupled vesicles derived from Paracoccus denitrificans catalyzes Pi-dependent proton-translocating ATPase reaction if exposed to prior energization that relieves ADP.Mg2+ -induced inhibition (Zharova, T.V. and Vinogradov, A.D. (2004) J. Biol. Chem.,279, 12319-12324). Here we present evidence that the presence of medium ADP is required for the steady-state energetically self-sustained coupled ATP hydrolysis. The initial rapid ATPase activity is declined to a certain level if the reaction proceeds in the presence of the ADP-consuming, ATP-regenerating system (pyruvate kinase/phosphoenol pyruvate). The rate and extent of the enzyme de-activation are inversely proportional to the steady-state ADP concentration, which is altered by various amounts of pyruvate kinase at constant ATPase level. The half-maximal rate of stationary ATP hydrolysis is reached at an ADP concentration of 8 x 10(-6) M. The kinetic scheme is proposed explaining the requirement of the reaction products (ADP and Pi), the substrates of ATP synthesis, in the medium for proton-translocating ATP hydrolysis by P. denitrificans Fo.F1-ATP synthase.  相似文献   

7.
Bonete MJ  Ferrer J  Pire C  Penades M  Ruiz JL 《Biochimie》2000,82(12):1143-1150
An NAD-dependent D-2-hydroxyacid dehydrogenase (EC 1.1.1.) was isolated and characterized from the halophilic Archaeon Haloferax mediterranei. The enzyme is a dimer with a molecular mass of 101.4 +/- 3.3 kDa. It is strictly NAD-dependent and exhibits its highest activity in 4 M NaCl. The enzyme is characterized by a broad substrate specificity 2-ketoisocaproate and 2-ketobutyrate being the substrates with the higher Vmax/Km. When pyruvate and 2-ketobutyrate were the substrates the optimal pH was acidic (pH 5) meanwhile for 2-ketoisocaproate maximum activity was achieved at basic pH between 7.5 and 8.5. The optimum temperature was 52 degrees C and at 65 degrees C there was a pronounced activity decrease. This new enzyme can be used for the production of D-2-hydroxycarboxylic acid.  相似文献   

8.
The kinetic properties of a continuous enzyme reactor containing rabbit muscle lactate dehydrogenase, horse liver alcohol dehydrogenase and poly(ethylene glycol)-bound NAD (PEG-NAD) were investigated experimentally and theoretically. The enzymes and PEG-NAD were retained in the reactor with an ultrafiltration membrane, and the substrates (pyruvate and ethanol) were fed continuously. The reactions of the dehydrogenases were coupled by the recycling of the cofactor. The steady-state concentration of L-lactate, one of the products, was measured under different experimental conditions and compared with the corresponding theoretical value. The theoretical value was calculated based on a simplified ordered bi-bi mechanism for the individual enzyme reactions, of which kinetic constants were determined by independent kinetic studies. Differences were found between the kinetic constants of the enzymes for NAD(H) and PEG-NAD(H). The steady-state values obtained by continuous operation were lower than those calculated, possibly due to the simplification made for the kinetic model; but there was general agreement between them in the dependence on the experimental conditions. The steady-state behavior of the enzyme reactor was explained semi-quantitatively by the simple kinetic model.  相似文献   

9.
Protein/histone acetyltransferases (PATs/HATs) have been implicated in a number of cellular functions including gene regulation, DNA synthesis, and repair. This paper reviews methods that can be used to quantitatively determine the activity and ultimately the catalytic/kinetic mechanism of PAT/HATs in vitro. Two methods will be described in detail. The first method is a filter-binding assay that measures the transfer of radiolabeled acetate from acetyl-CoA to protein. The second method is a continuous, spectroscopic, enzyme-coupled assay that links the PAT/HAT reaction to the reduction of NAD+ by pyruvate or alpha-ketoglutarate dehydrogenase. Both methods are highly applicable in determining steady-state reaction rates, and obtaining the kinetic constants Vmax, Km, and V/K from substrate saturation curves. We describe a new application of the filter-binding assay to determine the kinetic parameters for HATs using low concentrations of nucleosomal substrates.  相似文献   

10.
Decarboxylation rates for a series of C-3 to C-6 α-keto acids were determined in the presence of resting cells and cell-free extracts of Streptococcus lactis var. maltigenes. The C-5 and C-6 acids branched at the penultimate carbon atom were converted most rapidly to the respective aldehydes in the manner described for α-carboxylases. Pyruvate and α-ketobutyrate did not behave as α-carboxylase substrates, in that O2 was absorbed when they were reacted with resting cells. The same effect with pyruvate was noted in a nonmalty S. lactis, accounting for CO2 produced by some “homofermentative” streptococci. Mixed substrate reactions indicated that the same enzyme was responsible for decarboxylation of α-ketoisocaproate and α-ketoisovalerate, but it appeared unlikely that this enzyme was responsible for the decarboxylation of pyruvate. Ultrasonic disruption of cells of the malty culture resulted in an extract inactive for decarboxylation of pyruvate in the absence of ferricyanide. Dialyzed cell-free extracts were inactive against all keto acids and could not be reactivated.  相似文献   

11.
The kinetic mechanism of the cytosolic NADP(+)-dependent malic enzyme from cultured human breast cancer cell line was studied by steady-state kinetics. In the direction of oxidative decarboxylation, the initial-velocity and product-inhibition studies indicate that the enzyme reaction follows a sequential ordered Bi-Ter kinetic mechanism with NADP+ as the leading substrate followed by L-malate. The products are released in the order of CO2, pyruvate, and NADPH. The enzyme is unstable at high salt concentration and elevated temperature. However, it is stable for at least 20 min under the assay conditions. Tartronate (2-hydroxymalonate) was found to be a noncompetitive inhibitor for the enzyme with respect to L-malate. The kinetic mechanism of the cytosolic tumor malic enzyme is similar to that for the pigeon liver cytosolic malic enzyme but different from those for the mitochondrial enzyme from various sources.  相似文献   

12.
Acetohydroxy acid synthase (AHAS; EC 2.2.1.6) is a thiamin diphosphate (ThDP)-dependent decarboxylase-ligase that catalyzes the first common step in the biosynthesis of branched-chain amino acids. In the first stage of the reaction, pyruvate is decarboxylated and the reactive intermediate hydroxyethyl-ThDP carbanion/enamine is formed. In the second stage, the intermediate is ligated to another 2-ketoacid to form either acetolactate or acetohydroxybutyrate. AHAS isozyme I from Escherichia coli is unique among the AHAS isozymes in that it is not specific for 2-ketobutyrate (2-KB) over pyruvate as an acceptor substrate. It also appears to have a different mechanism for inhibition by valine than does AHAS III from E. coli. An investigation of this enzyme by directed mutagenesis and knowledge of detailed kinetics using the rapid mixing-quench NMR method or stopped-flow spectroscopy, as well as the use of alternative substrates, suggests that two residues determine most of the unique properties of AHAS I. Gln480 and Met476 in AHAS I replace the Trp and Leu residues conserved in other AHASs and lead to accelerated ligation and product release steps. This difference in kinetics accounts for the unique specificity, reversibility and allosteric response of AHAS I. The rate of decarboxylation of the initially formed 2-lactyl-ThDP intermediate is, in some AHAS I mutants, different for the alternative acceptors pyruvate and 2-KB, putting into question whether AHAS operates via a pure ping-pong mechanism. This finding might be compatible with a concerted mechanism (i.e. the formation of a ternary donor-acceptor:enzyme complex followed by covalent, ThDP-promoted catalysis with concerted decarboxylation-carboligation). It might alternatively be explained by an allosteric interaction between the multiple catalytic sites in AHAS.  相似文献   

13.
Acetohydroxy acid synthase (AHAS) and related enzymes catalyze the production of chiral compounds [(S)-acetolactate, (S)-acetohydroxybutyrate, or (R)-phenylacetylcarbinol] from achiral substrates (pyruvate, 2-ketobutyrate, or benzaldehyde). The common methods for the determination of AHAS activity have shortcomings. The colorimetric method for detection of acyloins formed from the products is tedious and does not allow time-resolved measurements. The continuous assay for consumption of pyruvate based on its absorbance at 333 nm, though convenient, is limited by the extremely small extinction coefficient of pyruvate, which results in a low signal-to-noise ratio and sensitivity to interfering absorbing compounds. Here, we report the use of circular dichroism spectroscopy for monitoring AHAS activity. This method, which exploits the optical activity of reaction products, displays a high signal-to-noise ratio and is easy to perform both in time-resolved and in commercial modes. In addition to AHAS, we examined the determination of activity of glyoxylate carboligase. This enzyme catalyzes the condensation of two molecules of glyoxylate to chiral tartronic acid semialdehyde. The use of circular dichroism also identifies the product of glyoxylate carboligase as being in the (R) configuration.  相似文献   

14.
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   

15.
Direct evidence for an enzyme-bound intermediate in the EPSP synthase reaction pathway has been obtained by rapid chemical quench-flow studies. The transient-state kinetic analysis has led to the following complete scheme: (formula; see text) Values for all 12 rate constants were obtained. Substrate trapping experiments in the forward and reverse reactions established the kinetically preferred order of binding and release of substrates and products and showed that shikimate 3-phosphate (S3P) and 5-enolpyruvoylshikimate 3-phosphate (EPSP) dissociate at rates greater than turnover in each direction. Pre-steady-state bursts of product formation were observed in the reaction in each direction indicating a rate-limiting step following catalysis. Single turnover experiments with enzyme in excess over substrate demonstrated the formation of a transient intermediate in both the forward and reverse reactions. In these experiments, the enzymatic reaction was observed by employing a radiolabel in the enol moiety of either phosphoenol pyruvate (PEP) or EPSP. The separation and quantitation of reaction products were accomplished by HPLC monitoring radioactivity. The intermediate was observed as the transient production of radiolabeled pyruvate, formed due to the breakdown of the intermediate in the acid quench used to stop the reaction. The intermediate was observed within 5-10 ms after the substrates were mixed with enzyme and decayed in a reaction paralleling the formation of product in each direction. Thus, the kinetics demonstrate directly the kinetic competence of the presumed intermediate. No pyruvate was formed, on a time scale which is relevant to catalysis, after incubation of the enzyme with dideoxy-S3P and PEP or with EPSP in the absence of phosphate; and so, the intermediate does not accumulate under these conditions. The intermediate broke down to form PEP and EPSP in addition to pyruvate when the reaction was quenched with base rather than acid; therefore, the intermediate must contain the elements of each product. Other experiments were designed to measure directly the phosphate binding rate and further constrain the PEP binding rate. The overall solution equilibrium constant in the forward direction was determined to be 180 by quantitation of radiolabeled reactants and products in equilibrium after incubation with a low enzyme concentration. The internal, active site equilibrium constant was obtained by incubation of radiolabeled S3P with excess enzyme and high concentrations of phosphate and PEP to provide the ratio of [EPSP]/[S3P] = 2.3, which is largely a measure of K4.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

16.
The steady-state kinetic mechanism of the reaction catalyzed by octopine dehydrogenase [N2-(1-carboxyethyl)-L-arginine:NAD+ oxidoreductase] was investigated at pH 6.9 and 9.2 by studies of substrate inhibition, analogue inhibition, and product inhibition. In the direction of octopine synthesis, the inhibition patterns in the presence of delta- guanidinovalerate and propionate show that NADH binds to the enzyme first followed by L-arginine and pyruvate which bind randomly. In the direction of octopine oxidation, the substrate patterns show that NAD binds to the enzyme before octopine in a rapid equilibrium fashion, and the product inhibition patterns show that the products L-arginine and pyruvate are released in a random fashion. Double, synergistic, substrate inhibition by L-arginine and pyruvate was shown to be due to binding (hypothetically of the imine) to the free enzyme and the enzyme-NAD complex. Furthermore, an alternate minor pathway was demonstrated which includes an enzyme-NADH-octopine complex and an enzyme-octopine complex.  相似文献   

17.
Dimethylglycine oxidase (DMGO) is a covalent flavoenzyme from Arthrobacter globiformis that catalyzes the oxidative demethylation of dimethylglycine to yield sarcosine, formaldehyde, and hydrogen peroxide. Stopped-flow and steady-state kinetic studies have been used to study the reductive and oxidative half-reactions using dimethylglycine and O2 as substrates. The reductive half-reaction is triphasic. The rate of the fast phase is dependent on substrate concentration, involves flavin reduction, and has a limiting rate constant of 244 s(-1). This phase also displays a kinetic isotope effect of 2.9. Completion of the first kinetic phase generates an intermediate with broad spectral signature between 350 and 500 nm, which is attributed to a reduced enzyme-iminium charge-transfer species, similar to the purple intermediate that accumulates in reactions of D-amino acid oxidase (DAAO) with alanine. The second phase (16 s(-1)) is independent of substrate concentration and is attributed to iminium hydrolysis/deprotonation. The third phase (2 s(-1)) is attributed to product release, the rate of which is less than the steady-state turnover rate (10.6 s(-1)). Flavin oxidation of dithionite- and dimethylglycine-reduced enzyme by O2 occurs in a single phase, and the rate shows a linear dependence on oxygen concentration, giving bimolecular rate constants of 342 and 201 mM(-1) x s(-1), respectively. Enzyme-monitored turnover experiments indicate that decay of the reduced enzyme-iminium intermediate is rate-limiting, consistent with rate constants determined from single turnover studies. A minimal kinetic mechanism is presented, which establishes a close relationship to the mechanism of action of DAAO. The covalent flavin in dimethylglycine oxidase is identified as an alphaN1-histidyl48-FAD, and equilibrium titration studies establish a single redox center that displays typical flavoprotein 'oxidase' characteristics.  相似文献   

18.
α-Ketobutyrate decarboxylase encoded in the -methionine catabolism operon of Pseudomonas putida is homologous with the E1 component of pyruvate dehydrogenase complex from gram-negative bacteria. The enzyme was purified to homogeneity from the cell extract of an Escherichia coli transformant. The purified enzyme was homodimeric with a subunit of Mr 93,000 on SDS-PAGE. The enzyme activity was activated by the addition of both thiamine pyrophosphate (TPP) and a divalent cation, such as Mg2+, Mn2+, and Co2+. The enzyme showed high activity for α-ketobutyrate and α-keto-n-valerate rather than pyruvate, but the α-keto acids with increasing length of the side chain as well as branching, such as α-keto-n-caproate and α-keto-3-methylvalerate, were not used by the enzyme. The Km values for α-ketobutyrate and pyruvate were 0.016 and 0.147 mM, respectively, and the kcat/Km value (10.69 s−1 mM−1) for α-ketobutyrate was 29-fold greater than that for pyruvate. Thus, α-ketobutyrate decarboxylase is distinguished from the pyruvate dehydrogenase E1 component with respect to the substrate specificity, although their structural and enzymological properties were similar. These results suggest that the unique substrate specificity of α-ketobutyrate decarboxylase is due to a slight difference in the highly conserved active sites of both enzymes.  相似文献   

19.
The mitochondrial NAD(P)+ malic enzyme [EC 1.1.1.39, L-malate:NAD+ oxidoreductase (decarboxylating)] was purified from rabbit heart to a specific activity of 7 units (mumol/min)/mg at 23 degrees C. A study of the reductive carboxylation reaction indicates that this enzymic reaction is reversible. The rate of the reductive carboxylation reaction appears to be completely inhibited at an NADH concentration of 0.92 mM. A substrate saturation curve of this reaction with NADH as the varied substrate describes this inhibition. The apparent kinetic parameters for this reaction are Ka(NADH) = 239 microM and Vr = 1.1 mumol/min per mg at 23 degrees C. The steady-state product-inhibition patterns for pyruvate and NADH indicate a sequential binding of the substrates: NAD+ followed by L-malate. These data also indicate that NADH is the last product released. A steady-state kinetic model is proposed that incorporates NADH-enzyme dead-end complexes.  相似文献   

20.
Bilophila wadsworthia RZATAU is a Gram-negative bacterium which converts the sulfonate taurine (2-aminoethanesulfonate) to ammonia, acetate and sulfide in an anaerobic respiration. Taurine:pyruvate aminotransferase (Tpa) catalyses the initial metabolic reaction yielding alanine and sulfoacetaldehyde. We purified Tpa 72-fold to apparent homogeneity with an overall yield of 89%. The purified enzyme did not require addition of pyridoxal 5'-phosphate, but highly active enzyme was only obtained by addition of pyridoxal 5'-phosphate to all buffers during purification. SDS/PAGE revealed a single protein band with a molecular mass of 51 kDa. The apparent molecular mass of the native enzyme was 197 kDa as determined by gel filtration, which indicates a homotetrameric structure. The kinetic constants for taurine were: Km = 7.1 mM, Vmax = 1.20 nmol.s-1, and for pyruvate: Km = 0.82 mM, Vmax = 0.17 nmol.s-1. The purified enzyme was able to transaminate hypotaurine (2-aminosulfinate), taurine, beta-alanine and with low activity cysteine and 3-aminopropanesulfonate. In addition to pyruvate, 2-ketobutyrate and oxaloacetate were utilized as amino group acceptors. We have sequenced the encoding gene (tpa). It encoded a 50-kDa peptide, which revealed 33% identity to diaminopelargonate aminotransferase from Bacillus subtilis.  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