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 156 毫秒
1.
Regulation of `malic'' enzyme of Solanum tuberosum by metabolites   总被引:5,自引:1,他引:4       下载免费PDF全文
A purification of ;malic' enzyme from potato is described. The purified enzyme is specific for NADP and requires a bivalent cation for activity. At pH values below 7 the plot of rate versus malate concentration approximates to normal Michaelis-Menten kinetics. At pH values above 7 the plot of rate versus malate concentration is sigmoid. A number of dicarboxylic acids activate the enzyme and remove the sigmoidicity. The enzyme is inhibited by phosphate, triose phosphates and AMP. In general, effectors of the oxidative decarboxylation of malate behave in the same manner in the reductive carboxylation of pyruvate. The response of the enzyme to energy charge is reported and the physiological significance of the response to metabolites is discussed in relation to the proposed role of the enzyme in the control of pH.  相似文献   

2.
The maximum velocity of the malic enzyme (L-malate: NADP+ oxidoreductase (oxaloacetate-decarboxylating), EC 1.1.1.40) reductive carboxylation of pyruvate and V/KCO2 are pH-independent from pH 5.5 to pH 8.5. V/K for pyruvate exhibits pK values values of 6.50 +/- 0.25 and 7.25 +/- 0.25. These data suggest that the binding of pyruvate locks the protonation state of enzyme. In addition, the pK values are within experimental error identical for the pH dependence of V/Kmalate and V/Kpyruvate. Thus, the catalytic groups appear to have reverse protonation states in the two reaction directions. The ratio of (V/Kmalate)/(V/Kpyruvate) is 100, suggesting that the protonation state of enzyme is optimum in the malate oxidative decarboxylation direction. Thus, the group with a pK of about 6 is unprotonated and the group with a pK of 7.5 is protonated for malate decarboxylation, and the opposite is true for pyruvate reductive carboxylation.  相似文献   

3.
The mitochondrial NADP-dependent malic enzyme (EC 1.1.1.40) was purified about 300-fold from cod Gadus morhua heart to a specific activity of 48 units (mumol/min)/mg at 30 degrees C. The possibility of the reductive carboxylation of pyruvate to malate was studied by determination of the respective enzyme properties. The reverse reaction was found to proceed at about five times the velocity of the forward rate at a pH 6.5. The Km values determined at pH 7.0 for pyruvate, NADPH and bicarbonate in the carboxylation reaction were 4.1 mM, 15 microM and 13.5 mM, respectively. The Km values for malate, NADP and Mn2+ in the decarboxylation reaction were 0.1 mM, 25 microM and 5 microM, respectively. The enzyme showed substrate inhibition at high malate concentrations for the oxidative decarboxylation reaction at pH 7.0. Malate inhibition suggests a possible modulation of cod heart mitochondrial NADP-malic enzyme by its own substrate. High NADP-dependent malic enzyme activity found in mitochondria from cod heart supports the possibility of malate formation under conditions facilitating carboxylation of pyruvate.  相似文献   

4.
Karsten WE  Cook PF 《Biochemistry》2006,45(29):9000-9006
An isothermal titration calorimetric study of the binding of substrates and inhibitors to different complexes of tartrate dehydrogenase (TDH) from Pseudomonas putida was carried out. TDH catalyzes the nicotinamide adenine dinucleotide (NAD)-dependent oxidative decarboxylation of d-malate and has an absolute requirement for both a divalent and monovalent metal ion for activity. The ligands Mn(2+), meso-tartrate, oxalate, and reduced nicotinamide adenine dinucleotide (NADH) bound to all TDH complexes with a stoichiometry of 1 per enzyme dimer. The exception is NAD, which binds to E/K(+), E/K(+)/Mn(2+), and E/K(+)/Mg(2+) complexes with a stoichiometry of two per enzyme dimer. The binding studies suggest a half-of-the-sites mechanism for TDH. No significant heat changes were observed for d-malate in the presence of the E/K(+)/Mn(2+) complex, suggesting that it did not bind. In contrast, meso-tartrate does bind to E/K(+)/Mn(2+) but gives no significant heat change in the presence of E/Mn(2+), suggesting that K(+) is required for meso-tartrate binding. meso-Tartrate also binds with a large DeltaC(p) value and likely binds via a different binding mode than d-malate, which binds only in the presence of NAD. In contrast to all of the other ligands tested, the binding of Mn(2+) is entropically driven, likely the result of the entropically favored disruption of ordered water molecules coordinated to Mn(2+) in solution that are lost upon binding to the enzyme. Oxalate, a competitive inhibitor of malate, binds with the greatest affinity to E/K(+)/Mn(2+)/NADH, and its binding is associated with the uptake of a proton. Overall, with d-malate as the substrate, data are consistent with a random addition of K(+), Mn(2+), and NAD followed by the ordered addition of d-malate; there is significant synergism in the binding of NAD and K(+). Although the binding of meso-tartrate also requires enzyme-bound K(+) and Mn(2+), the binding of meso-tartrate and NAD is random.  相似文献   

5.
Aktas DF  Cook PF 《Biochemistry》2008,47(8):2539-2546
The mitochondrial NAD-malic enzyme catalyzes the oxidative decarboxylation of malate to pyruvate and CO2. The role of the dinucleotide substrate in oxidative decarboxylation is probed in this study using site-directed mutagenesis to change key residues that line the dinucleotide binding site. Mutant enzymes were characterized using initial rate kinetics, and isotope effects were used to obtain information on the contribution of these residues to binding energy and catalysis. Results obtained for the N479 mutant enzymes indicate that the hydrogen bond donated by N479 to the carboxamide side chain of the nicotinamide ring is important for proper orientation in the hydride transfer step. The stepwise oxidative decarboxylation mechanism observed for the wt enzyme changed to a concerted one, which is totally rate limiting, for the N479Q mutant enzyme. In this case, it is likely that the longer glutamine side chain causes reorientation of malate such that it binds in a conformation that is optimal for concerted oxidative decarboxylation. Converting N479 to the shorter serine side chain gives very similar values of KNAD, Kmalate, and isotope effects relative to wt, but V/Et is decreased 2 000-fold. Data suggest an increased freedom of rotation, resulting in nonproductively bound cofactor. Changes were also made to two residues, S433 and N434, which interact with the nicotinamide ribose of NAD. In addition, N434 donates a hydrogen bond to the beta-carboxylate of malate. The KNAD for the S433A mutant enzyme increased by 80-fold, indicating that this residue provides significant binding affinity for the dinucleotide. With N434A, the interaction of the residue with malate is lost, causing the malate to reorient itself, leading to a slower decarboxylation step. The longer glutamine and methionine side chains stick into the active site and cause a change in the position of malate and/or NAD resulting in more than a 104-fold decrease in V/Et for these mutant enzymes. Overall, data indicate that subtle changes in the orientation of the cofactor and substrate dramatically influence the reaction rate.  相似文献   

6.
Liu D  Hwang CC  Cook PF 《Biochemistry》2002,41(40):12200-12203
The NAD-malic enzyme from Ascaris suum will utilize L-aspartate, (2S,3R)-tartrate, and meso-tartrate as substrates with V/K values 10(-4)-10(-5) with respect to malate. There is a strict requirement for the 2S stereochemistry for all of these reactants. Since aspartate is unique as an amino acid reactant for malic enzyme, it was informative to determine the details of its mechanism of oxidative decarboxylation. The initial rate of NADH appearance is directly proportional to the concentration of aspartate, and saturation is difficult to achieve. The pH dependence of V/K(aspartate)E(t) shows a decrease at low pH, giving a pK of 5.7. The pH-independent value of V/K(aspartate)E(t) is 3 M(-1) s(-1), 12500-fold lower than that obtained with L-malate. The dissociation constant for aspartate as a competitive inhibitor of malate is 60 mM at neutral pH, allowing an estimate of about 0.18 s(-1) for V/E(t) with L-aspartate compared to a value of 39 s(-1) obtained with L-malate. The deuterium isotope effect on V/K(aspartate) is pH independent over the range 5.1-6.9 with an average value of 3.3. Data suggest that the monoanion of L-aspartate binds to enzyme and that the same general base, general acid mechanism that is responsible for the oxidative decarboxylation of malate to pyruvate applies to the oxidative decarboxylation of aspartate to iminopyruvate. In addition, the oxidation step appears to be largely rate determining with aspartate as the substrate.  相似文献   

7.
In human liver, almost 90% of malic enzyme activity is located within the extramitochondrial compartment, and only approximately 10% in the mitochondrial fraction. Extramitochondrial malic enzyme has been isolated from the post-mitochondrial supernatant of human liver by (NH4)2SO4 fractionation, chromatography on DEAE-cellulose, ADP-Sepharose-4B and Sephacryl S-300 to apparent homogeneity, as judged from polyacrylamide gel electrophoresis. The specific activity of the purified enzyme was 56 mumol.min-1.mg protein-1, which corresponds to about 10,000-fold purification. The molecular mass of the native enzyme determined by gel filtration is 251 kDa. SDS/polyacrylamide gel electrophoresis showed one polypeptide band of molecular mass 63 kDa. Thus, it appears that the native protein is a tetramer composed of identical-molecular-mass subunits. The isoelectric point of the isolated enzyme was 5.65. The enzyme was shown to carboxylate pyruvate with at least the same rate as the forward reaction. The optimum pH for the carboxylation reaction was at pH 7.25 and that for the NADP-linked decarboxylation reaction varied with malate concentration. The Km values determined at pH 7.2 for malate and NADP were 120 microM and 9.2 microM, respectively. The Km values for pyruvate, NADPH and bicarbonate were 5.9 mM, 5.3 microM and 27.9 mM, respectively. The enzyme converted malate to pyruvate (at optimum pH 6.4) in the presence of 10 mM NAD at approximately 40% of the maximum rate with NADP. The Km values for malate and NAD were 0.96 mM and 4.6 mM, respectively. NAD-dependent decarboxylation reaction was not reversible. The purified human liver malic enzyme catalyzed decarboxylation of oxaloacetate and NADPH-linked reduction of pyruvate at about 1.3% and 5.4% of the maximum rate of NADP-linked oxidative decarboxylation of malate, respectively. The results indicate that malic enzyme from human liver exhibits similar properties to the enzyme from animal liver.  相似文献   

8.
Malic enzymes catalyze the oxidative decarboxylation of L-malate to pyruvate and CO(2) with the reduction of the NAD(P)(+) cofactor in the presence of divalent cations. We report the crystal structures at up to 2.1 A resolution of human mitochondrial NAD(P)(+)-dependent malic enzyme in different pentary complexes with the natural substrate malate or pyruvate, the dinucleotide cofactor NAD(+) or NADH, the divalent cation Mn(2+), and the allosteric activator fumarate. Malate is bound deep in the active site, providing two ligands for the cation, and its C4 carboxylate group is out of plane with the C1-C2-C3 atoms, facilitating decarboxylation. The divalent cation is positioned optimally to catalyze the entire reaction. Lys183 is the general base for the oxidation step, extracting the proton from the C2 hydroxyl of malate. Tyr112-Lys183 functions as the general acid-base pair to catalyze the tautomerization of the enolpyruvate product from decarboxylation to pyruvate.  相似文献   

9.
Karsten WE  Pais JE  Rao GS  Harris BG  Cook PF 《Biochemistry》2003,42(32):9712-9721
The kinetic mechanism of activation of the mitochondrial NAD-malic enzyme from the parasitic roundworm Ascaris suum has been studied using a steady-state kinetic approach. The following conclusions are suggested. First, malate and fumarate increase the activity of the enzyme in both reaction directions as a result of binding to separate allosteric sites, i.e., sites that exist in addition to the active site. The binding of malate and fumarate is synergistic with the K(act) decreasing by >or=10-fold at saturating concentrations of the other activator. Second, the presence of the activators decreases the K(m) for pyruvate 3-4-fold, and the K(i) (Mn) >or=20-fold in the direction of reductive carboxylation; similar effects are obtained with fumarate in the direction of oxidative decarboxylation. The greatest effect of the activators is thus expressed at low reactant concentrations, i.e., physiologic concentrations of reactant, where activation of >or=15-fold is observed. A recent crystallographic structure of the human mitochondrial NAD malic enzyme [13] shows fumarate bound to an allosteric site. Site-directed mutagenesis was used to change R105, homologous to R91 in the fumarate activator site of the human enzyme, to alanine. The R105A mutant enzyme exhibits the same maximum rate and V/K(NAD) as does the wild-type enzyme, but 7-8-fold decrease in both V/K(malate) and V/K(Mg), indicating the importance of this residue in the activator site. In addition, neither fumarate nor malate activates the enzyme in either reaction direction. Finally, a change in K143 (a residue in a positive pocket adjacent to that which contains R105), to alanine results in an increase in the K(act) for malate by about an order of magnitude such that it is now of the same magnitude as the K(m) for malate. The K143A mutant enzyme also exhibits an increase in the K(act) for fumarate (in the absence of malate) from 200 microM to about 25 mM.  相似文献   

10.
P A Tipton  J Peisach 《Biochemistry》1990,29(7):1749-1756
Tartrate dehydrogenase (TDH) has been purified to apparent homogeneity from Pseudomonas putida and has been demonstrated to catalyze three different NAD(+)-dependent reactions. TDH catalyzes the oxidation of (+)-tartrate to form oxaloglycolate and the oxidative decarboxylation of D-malate to form pyruvate and CO2. D-Glycerate and CO2 are formed from meso-tartrate in a reaction that is formally a decarboxylation with no net oxidation or reduction. The steady-state kinetics of the first two reactions have been investigated and found to follow primarily ordered mechanisms. The pH dependence of V and V/K was determined and indicates that catalysis requires that a base on the enzyme with a pK of 6.7 be unprotonated. TDH activity requires a divalent and a monovalent cation. Kinetic data suggest that the cations function in substrate binding and facilitation of the decarboxylation of beta-ketoacid intermediates.  相似文献   

11.
The effect of anaplerotic pathways activation on CO2-dependent anaerobic glucose utilization by Escherichia coli strains deficient in the main fermentation pathways and possessing a modified system of glucose transport and phosphorylation was studied. Intracellular CO2 generation in the strains was ensured resulting from oxidative decarboxylation of pyruvic acid by pyruvate dehydrogenase. Sodium bicarbonate dissolved in the medium was used as an external source of CO2. The genes of heterologous pyruvate carboxylase and native NADH-dependent malic enzyme were overexpressed in the strains to allow anaplerotic carboxylation of pyruvic acid to oxaloacetic or malic acid. The ability of the strains to reoxidize NADH utilizing carboxylation products was additionally increased due to enhanced expression of malate dehydrogenase gene. In the case of endogenous CO2 formation, the activation of anaplerotic pathways did not cause a notable increase in the anaerobic glucose consumption by the constructed strains. At the same time, the expression of pyruvate carboxylase led to a pronounced decrease in the secretion of pyruvic acid with the concomitant increase in the yield of four-carbon metabolites. Further enhancement of NADH-dependent malic enzyme expression provoked activation of a pyruvate–oxaloacetate–malate–pyruvate futile cycle in the strains. The availability in the medium of the external CO2 source sharply increased the anaerobic utilization of glucose by strains expressing pyruvate carboxylase. The activity of the futile cycle has raised with the increased malic enzyme expression and dropped upon enhancement of malate dehydrogenase expression. As a result, the efficiency of CO2-dependent anaerobic glucose utilization coupled to the formation of four-carbon carboxylation products increased in the studied strains resulting from the primary anaplerotic conversion of pyruvic acid into oxaloacetic acid followed by the involvement of the precursor formed in NADH-consuming biosynthetic reactions dominating over the reactions of the revealed futile cycle.  相似文献   

12.
BACKGROUND: Malic enzymes catalyze the oxidative decarboxylation of malate to pyruvate and CO2 with the concomitant reduction of NAD(P)+ to NAD(P)H. They are widely distributed in nature and have important biological functions. Human mitochondrial NAD(P)+-dependent malic enzyme (mNAD-ME) may have a crucial role in the metabolism of glutamine for energy production in rapidly dividing cells and tumors. Moreover, this isoform is unique among malic enzymes in that it is a cooperative enzyme, and its activity is controlled allosterically. RESULTS: The crystal structure of human mNAD-ME has been determined at 2.5 A resolution by the selenomethionyl multiwavelength anomalous diffraction method and refined to 2.1 A resolution. The structure of the monomer can be divided into four domains; the active site of the enzyme is located in a deep cleft at the interface between three of the domains. Three acidic residues (Glu255, Asp256 and Asp279) were identified as ligands for the divalent cation that is required for catalysis by malic enzymes. CONCLUSIONS: The structure reveals that malic enzymes belong to a new class of oxidative decarboxylases. The tetramer of the enzyme appears to be a dimer of dimers. The active site of each monomer is located far from the tetramer interface. The structure also shows the binding of a second NAD+ molecule in a pocket 35 A away from the active site. The natural ligand for this second binding site may be ATP, an allosteric inhibitor of the enzyme.  相似文献   

13.
14.
meso-Tartrate inhibited the growth of non-meso-tartrate-utilizing strains of Salmonella typhimurium in peptone water media and mineral salts media with some, but not all, carbon sources. C-R intermediates of the tricarboxylic acid cycle or compounds readily converted to them and substrates metabolized independently of the C-6 part of the cycle spared bacteria from the inhibitory effects of meso-tartrate when added to cultures along with meso-tartrate. Experiments with cell-free extracts of non-meso-tartrate-utilizing strains from batch and continuous cultures showed that meso-tartrate was a competitive inhibitor of isocitrate dehydrogenase and isocitrate lyase activities and also inhibited citrate synthase and malate synthase activities. The synthesis of these enzymes was not inhibited by meso-tartrate. The isocitrate enzymes of meso-tartrate-utilizing strains of S. typhimurium were similarly inhibited by meso-tartrate, but inhibition of the growth of meso-tartrate-utilizing strains was demonstrable only in uninduced cultures in which the intracellular concentrations of meso-tartrate were high.  相似文献   

15.
Arabidopsis thaliana is a plant species that accumulates high levels of organic acids and uses them as carbon, energy and reducing power sources. Among the enzymes that metabolize these compounds, one of the most important ones is malic enzyme (ME). A. thaliana contains four malic enzymes (NADP-ME 1–4) to catalyze the reversible oxidative decarboxylation of malate in the presence of NADP. NADP-ME2 is the only one located in the cell cytosol of all Arabidopsis organs providing most of the total NADP-ME activity. In the present work, the regulation of this key enzyme by fumarate was investigated by kinetic assays, structural analysis and a site-directed mutagenesis approach. The final effect of this metabolite on NADP-ME2 forward activity not only depends on fumarate and substrate concentrations but also on the pH of the reaction medium. Fumarate produced an increase in NADP-ME2 activity by binding to an allosteric site. However at higher concentrations, fumarate caused a competitive inhibition, excluding the substrate malate from binding to the active site. The characterization of ME2-R115A mutant, which is not activated by fumarate, confirms this hypothesis. In addition, the reverse reaction (reductive carboxylation of pyruvate) is also modulated by fumarate, but in a different way. The results indicate pH-dependence of the fumarate modulation with opposite behavior on the two activities analyzed. Thereby, the coordinated action of fumarate over the direct and reverse reactions would allow a precise and specific modulation of the metabolic flux through this enzyme, leading to the synthesis or degradation of C4 compounds under certain conditions. Thus, the physiological context might be exerting an accurate control of ME activity in planta, through changes in metabolite and substrate concentrations and cytosolic pH.  相似文献   

16.
Mitochondria isolated from the Crassulacean acid metabolism plant Sedum praealtum were demonstrated to decarboxylate added malate at basal rates of 30–50 μmol mg?1 original chlorophyll h?1. The basal rate could be stimulated markedly by the addition of ADP, oxaloacetic acid, an uncoupler of oxidative phosphorylation, or NAD, with maximum rates of 70–100 μmol mg?1 original chlorophyll h?1 observed. These observed rates were high enough to account for a large proportion of the estimated rate of malate decarboxylation in vivo. The major products of malate oxidation by the mitochondria in most cases were found to be pyruvate and CO2, indicating that malate oxidation in these mitochondria proceeds mainly through NAD malic enzyme rather than NAD malate dehydrogenase. Under conditions employed little of the pyruvate formed was further oxidized, suggesting a fate other than oxidation (conversion to starch) for this pyruvate. Malate decarboxylation by mitochondria and by partially purified NAD malic enzyme was markedly inhibited by NaHCO3. A possible physiological role is suggested for this inhibition as a feedback control on the enzyme.  相似文献   

17.
Isocitrate dehydrogenase was purified from Hydrogenobacter thermophilus, and the corresponding gene was cloned and sequenced. The enzyme had similar structural properties to the isocitrate dehydrogenase of Escherichia coli, but differed in its catalytic properties, such as coenzyme specificity, pH dependency and kinetic parameters. Notably, the enzyme catalysed the oxidative decarboxylation of isocitrate, but not the reductive carboxylation of 2-oxoglutarate. The carboxylation reaction required the addition of cell extract and ATP-Mg, suggesting the existence of additional carboxylation factor(s). Further analysis of the carboxylation factor(s) resulted in the purification of two polypeptides. N-terminal amino acid sequencing revealed that the two polypeptides are homologues of pyruvate carboxylase with a biotinylated subunit, but do not catalyse pyruvate carboxylation. Pyruvate carboxylase was also purified, but was not active in stimulating isocitrate dehydrogenase. Isocitrate dehydrogenase, the novel biotin protein, ATP-Mg and NADH were essential for the reductive carboxylation of 2-oxoglutarate. These observations indicate that the novel biotin protein is an ATP-dependent factor, which is involved in the reverse (carboxylating) reaction of isocitrate dehydrogenase.  相似文献   

18.
Mitochondrial malate dehydrogenase (mMDH) and malic enzyme (mME) of a filarial worm Setaria digitata were studied. mMDH exhibited the highest activities in the oxidation and reduction reactions at pH 9.5 and pH 6.2, respectively, while mME did so in the malate decarboxylation reaction at pH 6.8. mME showed no detectable activity on the pyruvate carboxylation direction. The Km values for malate (1.7 mM) and oxaloacetate (0.17 mM) and the ratio of Vmax oxidation: Vmax reduction (2.73) tend to favor the oxaloacetate reduction by mMDH. mME showed a relatively high Km value of 8.3 mM, for malate decarboxylation. A drug, diethylcarbamazine citrate (DEC-C), did not change appreciably the activity of either mMDH or mME, while filarin (a drug of herbal origin) effectively inhibited mMDH. The leaf extracts of Ocimum sanctum, Lawsonia inermis and Calotropis gigantea and leaf and flower extracts of Azadirachta indica were, however, found to inhibit both mMDH and mME.  相似文献   

19.
The activity of the biotin-dependent enzyme pyruvate carboxylase from many organisms is highly regulated by the allosteric activator acetyl-CoA. A number of X-ray crystallographic structures of the native pyruvate carboxylase tetramer are now available for the enzyme from Rhizobium etli and Staphylococcus aureus. Although all of these structures show that intersubunit catalysis occurs, in the case of the R. etli enzyme, only two of the four subunits have the allosteric activator bound to them and are optimally configured for catalysis of the overall reaction. However, it is apparent that acetyl-CoA binding does not induce the observed asymmetrical tetramer conformation and it is likely that, under normal reaction conditions, all of the subunits have acetyl-CoA bound to them. Thus the activation of the enzyme by acetyl-CoA involves more subtle structural effects, one of which may be to facilitate the correct positioning of Arg353 and biotin in the biotin carboxylase domain active site, thereby promoting biotin carboxylation and, at the same time, preventing abortive decarboxylation of carboxybiotin. It is also apparent from the crystal structures that there are allosteric interactions induced by acetyl-CoA binding in the pair of subunits not optimally configured for catalysis of the overall reaction.  相似文献   

20.
Karsten WE  Tipton PA  Cook PF 《Biochemistry》2002,41(40):12193-12199
Tartrate dehydrogenase catalyzes the divalent metal ion- and NAD-dependent oxidative decarboxylation of D-malate to yield CO(2), pyruvate, and NADH. The enzyme also catalyzes the metal ion-dependent oxidation of (+)-tartrate to yield oxaloglycolate and NADH. pH-rate profiles and isotope effects were measured to probe the mechanism of this unique enzyme. Data suggest a general base mechanism with likely general acid catalysis in the oxidative decarboxylation of D-malate. Of interest, the mechanism of oxidative decarboxylation of D-malate is stepwise with NAD(+) or the more oxidizing thio-NAD(+). The mechanism does not become concerted with the latter as observed for the malic enzyme, which catalyzes the oxidative decarboxylation of L-malate [Karsten, W. E., and Cook, P. F. (1994) Biochemistry 33, 2096-2103]. It appears the change in mechanism observed with malic enzyme is specific to its transition state structure and not a generalized trait of metal ion- and NAD(P)-dependent beta-hydroxy acid oxidative decarboxylases. The V/K(malate) pH-rate profile decreases at low and high pH and exhibits pK(a) values of about 6.3 and 8.3, while that for V/K(tartrate) (measured from pH 7.5 to pH 9) exhibits a pK(a) of 8.6 on the basic side. A single pK(a) of 6.3 is observed on the acid side of the V(max) pH profile, but the pK(a) seen on the basic side of the V/K pH profiles is not observed in the V(max) pH profiles. Data suggest the requirement for a general base that accepts a proton from the 2-hydroxyl group of either substrate to facilitate hydride transfer. A second enzymatic group is also required protonated for optimum binding of substrates and may also function as a general acid to donate a proton to the enolpyruvate intermediate to form pyruvate. The (13)C isotope effect, measured on the decarboxylation of D-malate using NAD(+) as the dinucleotide substrate, decreases from a value of 1.0096 +/- 0.0006 with D-malate to 1.00787 +/- 0.00006 with D-malate-2-d, suggesting a stepwise mechanism for the oxidative decarboxylation of D-malate. Using thio-NAD(+) as the dinucleotide substrate the (13)C isotope effects are 1.0034 +/- 0.0007 and 1.0027 +/- 0.0002 with D-malate and D-malate-2-d, 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