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.
Cytosolic malate dehydrogenase from human liver was isolated and its physical and kinetic properties were determined. The enzyme had a molecular weight of 72,000 ± 2000 and an amino acid composition similar to those of malate dehydrogenases from other species. The kinetic behaviour of the enzyme was consistent with an Ordered Bi Bi mechanism. The following values (μm) of the kinetic parameters were obtained at pH 7.4 and 37 °C: Ka, 17; Kia, 3.6; Kb, 51; Kib, 68; Kp, 770; Kip, 10,700; Kq, 42; Kiq, 500, where a, b, p, and q refer to NADH, oxalacetate, malate, and NAD+, respectively. The maximum velocity of the enzyme in human liver homogenates was 102 μmol/min/g wet wt of liver for oxalacetate reduction and 11.2 μmol/min/g liver for malate oxidation at pH 7.4 and 37 °C. Calculations using these parameters showed that, under conditions in vivo, the rate of NADH oxidation by the enzyme would be much less than the maximum velocity and could be comparable to the rate of NADH production during ethanol oxidation in human liver. The rate of NADH oxidation would be sensitive to the concentrations of NADH and oxalacetate; this sensitivity can explain the change in cytosolic NAD+NADH redox state during ethanol metabolism in human liver.  相似文献   

2.
If liver alcohol dehydrogenase were rate-limiting in ethanol metabolism, inhibitors of the enzyme should inhibit the metabolism with the same type of kinetics and the same kinetic constants in vitro and in vivo. Against varied concentrations of ethanol, 4-methylpyrazole is a competitive inhibitor of purified rat liver alcohol dehydrogenase (Kis = 0.11 microM, in 83 mM potassium phosphate and 40 mM KCl buffer, pH 7.3, 37 degrees C) and is competitive in rats (with Kis = 1.4 mumol/kg). Isobutyramide is essentially an uncompetitive inhibitor of purified enzyme (Kii = 0.33 mM) and of metabolism in vivo (Kii = 1.0 mmol/kg). Low concentrations of both inhibitors decreased the rate of metabolism as a direct function of their concentrations. Qualitatively, therefore, alcohol dehydrogenase activity appears to be a major rate-limiting factor in ethanol metabolism. Quantitatively, however, the constants may not agree because of distribution in the animal or metabolism of the inhibitors. At saturating concentrations of inhibitors, ethanol is eliminated by inhibitor-insensitive pathways, at about 10% of the total rate at a dose of ethanol of 10 mmol/kg. Uncompetitive inhibitors of alcohol dehydrogenase should be especially useful for inhibiting the metabolism of alcohols since they are effective even at saturating levels of alcohol, in contrast to competitive inhibitors, whose action is overcome by saturation with alcohol.  相似文献   

3.
The level of aspartate aminotransferase in liver mitochondria was found to be approximately 140 microM, or 2-3 orders of magnitude higher than its dissociation constant in complexes with the inner mitochondrial membrane and the high molecular weight enzymes (M(r) = 1.6 x 10(5) to 2.7 x 10(6)) carbamyl-phosphate synthase I, glutamate dehydrogenase, and the alpha-ketoglutarate dehydrogenase complex. The total concentration of aminotransferase-binding sites on these structures in liver mitochondria was more than sufficient to accommodate all of the aminotransferase. Therefore, in liver mitochondria, the aminotransferase could be associated with the inner mitochondrial membrane and/or these high molecular weight enzymes. The aminotransferase in these hetero-enzyme complexes could be supplied with oxalacetate because binding of aminotransferase to the high molecular weight enzymes can enhance binding of malate dehydrogenase, and binding of both malate dehydrogenase and the aminotransferase facilitated binding of fumarase. The level of malate dehydrogenase was found to be so high (140 microM) in liver mitochondria, compared with that of citrate synthase (25 microM) and the pyruvate dehydrogenase complex (0.3 microM), that there would also be a sufficient supply of oxalacetate to citrate synthase-pyruvate dehydrogenase.  相似文献   

4.
In extension of a previous study with yeast glucose-6-P dehydrogenase (Kawaguchi, A., and Bloch, K. (1974) J. Biol. Chem. 249, 5793-5800), the structural changes accompanying the inhibition of glutamate dehydrogenase and several malate dehydrogenases by palmitoyl-CoA and by sodium dodecyl sulfate have been investigated. Palmitoyl-CoA converts liver glutamate dehydrogenase to enzymatically inactive dimeric subunits (Mr = 1.2 X 10(5)) and tightly binds to the dissociated enzyme. Removal of the inhibitor from the palmitoyl-CoA-dimer complex fails to regenerate enzyme activity. The Ki values for palmitoyl-CoA inhibition of malate dehydrogenases (oxalacetate reduction) are, for the enzyme from pig heart mitochondria, 1.8 muM, 500 muM from pig heart supernatant, and 10 muM from chicken heart supernatant. These inhibitions are readily reversible. Palmitoyl-CoA does not alter the quaternary structure of any of the malate dehydrogenases and binds only weakly to these enzymes. Mitochondrial malate dehydrogenase assayed in the direction malate to oxalacetate is much less sensitive to palmitoyl-CoA, with Ki values of 50 muM at pH 10 and greater than 50 muM at pH 7.4. While the differences in palmitoyl-CoA sensitivity in the forward and backward reactions catalyzed by mitochondrial dehydrogenase are unexplained, a physiological rationale for these differential effects is offered. Sodium dodecyl sulfate dissociates the various dehydrogenases to monomeric subunits in contrast to the more selective effects of palmitoyl-CoA.  相似文献   

5.
The activities of the mitochondrial enzymes citrate synthase (citrate oxaloacetatelyase, EC 4.1.3.7), NADP-linked isocitrate dehydrogenase (threo-Ds-isocitrate:NADP+ oxidoreductase (decarboxylating), EC 1.1.1.42), and succinate dehydrogenase (succinate: FAD oxidoreductase, EC 1.3.99.1) as well as their kinetic behavior in the two developmental forms of Trypanosoma cruzi at insect vector stage, epimastigotes and infective metacyclic trypomastigotes, were studied. The results presented in this work clearly demonstrate a higher mitochondrial metabolism in the metacyclic forms as is shown by the extraordinary enhanced activities of metacyclic citrate synthase, isocitrate dehydrogenase, and succinate dehydrogenase. In epimastigotes, the specific activities of citrate synthase at variable concentrations of oxalacetate and acetyl-CoA were 24.6 and 26.6 mU/mg of protein, respectively, and the Michaelis constants were 7.88 and 6.84 microM for both substrates. The metacyclic enzyme exhibited the following kinetic parameters: a specific activity of 228.4 mU/mg and Km of 3.18 microM for oxalacetate and 248.5 mU/mg and 2.75 microM, respectively, for acetyl-CoA. NADP-linked isocitrate dehydrogenase specific activities for epimastigotes and metacyclics were 110.2 and 210.3 mU/mg, whereas the apparent Km's were 47.9 and 12.5 microM, respectively. No activity for the NAD-dependent isozyme was found in any form of T. cruzi differentiation. The particulated succinate dehydrogenase showed specific activities of 8.2 and 39.1 mU/mg for epimastigotes and metacyclic trypomastigotes, respectively, although no significant changes in the Km (0.46 and 0.48 mM) were found. The cellular role and the molecular mechanism that probably take place during this significant shift in the mitochondrial metabolism during the T. cruzi differentiation have been discussed.  相似文献   

6.
Current evidence suggests that mitochondrial matrix enzymes exist in solid-state, multienzyme complexes in vivo. Addition of polyethylene glycol to a solution containing malate dehydrogenase and citrate synthase generates such a solid-state, enzyme complex in vitro at enzyme concentrations permitting kinetic measurements. Suspensions of the isolated, solid-state, hetero-complex of these enzymes were used to study the coupled reactions of citrate synthesis from malate, NAD, and CoASAc. The particles appear to be about 1 microgram in diameter. Considering the ratio of enzyme to oxalacetate molecules in or at the surface of the solid-state particles, one would expect oxalacetate to be converted to citrate within a few molecular distances of the site of oxalacetate generation. This model of "substrate channeling" (or alternatively a direct transfer of oxalacetate between enzymes) is supported by experiments with excess aspartate aminotransferase and glutamate added to the solution phase to give a reaction competing with the synthase for bulk phase oxalacetate. Quantities of aminotransferase that reduce the citrate reaction rate with soluble dehydrogenase and synthase by 90% do not significantly affect rates with comparable amounts of the dehydrogenase-synthase complex. We suggest that similar substrate channeling can occur in vivo and discuss the possible advantages provided thereby.  相似文献   

7.
We purified the psychrophilic and thermolabile malate dehydrogenase to homogeneity from a novel psychrotolerant, Flavobacterium frigidimaris KUC-1, isolated from Antarctic seawater. The enzyme was a homotetramer with a molecular weight of about 123 k and that of the subunit was about 32 k. The enzyme required NAD(P)(+) as a coenzyme and catalyzed the oxidation of L-malate and the reduction of oxalacetate specifically. The reaction proceeded through an ordered bi-bi mechanism. The enzyme was highly susceptible to heat treatment, and the half-life time at 40 degrees C was estimated to be 3.0 min. The k(cat)/K(m) (microM(-1).s(-1)) values for L-malate and NAD(+) at 30 degrees C were 289 and 2,790, respectively. The enzyme showed pro-R stereospecificity for hydrogen transfer at the C4 position of the nicotinamide moiety of the coenzyme. The enzyme contained 311 amino acid residues and much lower numbers of proline and arginine residues than other malate dehydrogenases.  相似文献   

8.
1. A method is described for the estimation of in vivo activity of mitochondrial Acyl-CoA dehydrogenase, without the addition of extra ETF (electron transferring flavoprotein). 2. Using Fumarase as a marker enzyme, the activity of Acyl-CoA dehydrogenase was calculated per gram wet weight. 3. The enzyme displays negative cooperativity in fish muscle, but positive cooperativity in pig liver. S0.5 values for the substrate octanoyl-CoA are between 3 and 5 microM.  相似文献   

9.
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   

10.
Alcohol dehydrogenase was purified in 14 h from male Fischer-344 rat livers by differential centrifugation, (NH4)2SO4 precipitation, and chromatography over DEAE-Affi-Gel Blue, Affi-Gel Blue, and AMP-agarose. Following HPLC more than 240-fold purification was obtained. Under denaturing conditions, the enzyme migrated as a single protein band (Mr congruent to 40,000) on 10% sodium dodecyl sulfate-polyacrylamide gels. Under nondenaturing conditions, the protein eluted from an HPLC I-125 column as a symmetrical peak with a constant enzyme specific activity. When examined by analytical isoelectric focusing, two protein and two enzyme activity bands comigrated closely together (broad band) between pH 8.8 and 8.9. The pure enzyme showed pH optima for activity between 8.3 and 8.8 in buffers of 0.5 M Tris-HCl, 50 mM 2-(N-cyclohexylamino)ethanesulfonic acid (CHES), and 50 mM 3-(cyclohexylamino)-1-propanesulfonic acid (CAPS), and above pH 9.0 in 50 mM glycyl-glycine. Kinetic studies with the pure enzyme, in 0.5 M Tris-HCl under varying pH conditions, revealed three characteristic ionization constants for activity: 7.4 (pK1); 8.0-8.1 (pK2), and 9.1 (pK3). The latter two probably represent functional groups in the free enzyme; pK1 may represent a functional group in the enzyme-NAD+ complex. Pure enzyme also was used to determine kinetic constants at 37 degrees C in 0.5 M Tris-HCl buffer, pH 7.4 (I = 0.2). The values obtained were Vmax = 2.21 microM/min/mg enzyme, Km for ethanol = 0.156 mM, Km for NAD+ = 0.176 mM, and a dissociation constant for NAD+ = 0.306 mM. These values were used to extrapolate the forward rate of ethanol oxidation by alcohol dehydrogenase in vivo. At pH 7.4 and 10 mM ethanol, the rate was calculated to be 2.4 microM/min/g liver.  相似文献   

11.
The malate dehydrogenase activity (EC 1.1.1.37), present in the cytoplasm of Pisum sativum root nodules, can be separated by ion-exchange chromatography into four different fractions. Malate dehydrogenase activity present in the cytoplasm of roots elutes mainly as a single peak. During nodule development an increase in malate dehydrogenase activity per gram of material was observed. This increase occurred concomitantly with the increase in nitrogenase activity. The kinetic properties of the separated malate dehydrogenases of root nodule cytoplasm and root cytoplasm were studied. The Km values for malate (2.6 mM), NAD+ (27 microM), oxaloacetate (18 microM) and NADH (13 microM) of the dominant form of the root nodule cytoplasm are much lower than those of the dominant malate dehydrogenase root form (64 mM, 4.4 mM, 89 microM and 70 microM respectively). Binding of malate by the enzyme-NADH complex from root nodules results in an abortive complex, thereby blocking the further reduction of oxaloacetate by NADH. The dominant root malate dehydrogenase does not form the abortive complex. From the kinetic data it is concluded, first, that the root nodule forms of the enzyme are capable of catalysing at a high rate the reduction of oxaloacetate, to meet the demands for malate governed by the bacteroid and the infected plant cell. The second conclusion, drawn from the kinetic data, is that under physiological conditions the conversion of oxaloacetate can be controlled just by the malate concentration. Consequently the major root nodule forms of malate dehydrogenase are able to allow a high flux of malate production from oxaloacetate but also to establish a sufficient oxaloacetate concentration necessary for the assimilation and transport of fixed nitrogen.  相似文献   

12.
Formation of a bienzyme complex of pig heart mitochondrial malate dehydrogenase and citrate synthase in a buffered system is demonstrated by means of a covalently attached fluorescent probe to citrate synthase. Assuming 1:1 stoichiometry of the enzymes in the complex, an apparent dissociation constant of 10(-6) M was calculated from fluorescence anisotropy measurements. The effect of various metabolites on the interaction was tested. NAD+, oxalacetate, citrate, ATP, and L(-)- or D(+)-malate had no effect on the association of the two enzymes, whereas alpha-ketoglutarate increased and NADH decreased it. The interaction of mitochondrial citrate synthase with cytosolic malate dehydrogenase was found to be much weaker, whereas interaction of citrate synthase with another cytosolic enzyme, aldolase, could not be detected. In kinetic experiments, the activation of malate dehydrogenase by citrate synthase was observed. The effect of pyridine nucleotides and alpha-ketoglutarate is discussed in relation to the direction of the metabolic flow of oxalacetate.  相似文献   

13.
Metabolism of lactate as a carbon source by Pseudomonas citronellolis occurred via a nicotinamide adenine dinucleotide (NAD)-independent L-lactate dehydrogenase, which was present in cells grown on DL-lactate but was not present in cells grown on acetate, aspartate, citrate, glucose, glutamate, or malate. The cells also possessed a constitutive, NAD-independent malate dehydrogenase instead of the conventional NAD-dependent malate dehydrogenase instead of the conventional NAD-dependent enzyme in the tricarboxylic acid cycle. Both enzymes were particulate and used dichlorophenolindo-phenol or oxygen as an electron acceptor. In acetate-grown cells, the activity of pyruvate dehydrogenase and NAD phosphate-linked malate enzyme decreased, cells grown on glucose or lactate. This was consistent with the need to maintain a supply of oxalacetate for metabolism of acetate via the tricarboxylic acid cycle. Changes in enzyme activities suggest that gluconeogenesis from noncarbohydrate carbon sources occurs via the malate enzyme (when oxalacetate decarboxylase is inhibited) or a combination of the NAD-independent malate dehydrogenase and oxalacetate decarboxylase.  相似文献   

14.
The roles of the pyruvate decarboxylation pathway and TCA metabolic cycle in activation of anaerobic metabolism in ripening Hamlin oranges were investigated. Oranges were harvested weekly from October to February during the 1980–81 and 1981–82 growing season. Juice vesicles from each weekly sample were assayed for pyruvate decarboxylase, alcohol dehydrogenase, malic enzyme, phosphoenolpyruvate carboxylase, malate dehydrogenase, citrate synthase, isocitrate dehydrogenase and cytochrome oxidase. Also, juice was assayed for ethanol, acetaldehyde, pyruvate, oxalacetate, malate and citrate. In December when ethanol accumulated rapidly in the fruit, pyruvate decarboxylase and alcohol dehydrogenase increased markedly. During the same month, the pyruvate level declined, suggesting that the increases in enzyme levels activated the conversion of pyruvate to ethanol.  相似文献   

15.
Isolated pea leaf mitochondria oxidatively decarboxylate added glycine. This decarboxylation could be linked to the respiratory chain (in which case it was coupled to three phosphorylations) or to mitochondrial malate dehydrogenase when oxalacetate was supplied. Decarboxylation rates measured as O2 uptake, or CO2 and NH3 release were adequate to account for whole leaf photorespiration. Oxalacetate-supported glycine decarboxylation, measured by linking malate efflux to added malic enzyme, yielded rates considerably less than the electron transport rates. Butylmalonate inhibited malate efflux but not oxalacetate entry; phthalonate inhibited oxalacetate entry but had little effect on malate or α-ketoglutarate oxidation. It is suggested that oxalacetate and malate transport are catalyzed by separate carrier systems of the mitochondrial membrane.  相似文献   

16.
Binding experiments indicate that mitochondrial aspartate aminotransferase can associate with the alpha-ketoglutarate dehydrogenase complex and that mitochondrial malate dehydrogenase can associate with this binary complex to form a ternary complex. Formation of this ternary complex enables low levels of the alpha-ketoglutarate dehydrogenase complex, in the presence of the aminotransferase, to reverse inhibition of malate oxidation by glutamate. Thus, glutamate can react with the aminotransferase in this complex without glutamate inhibiting production of oxalacetate by the malate dehydrogenase in the complex. The conversion of glutamate to alpha-ketoglutarate could also be facilitated because in the trienzyme complex, oxalacetate might be directly transferred from malate dehydrogenase to the aminotransferase. In addition, association of malate dehydrogenase with these other two enzymes enhances malate dehydrogenase activity due to a marked decrease in the Km of malate. The potential ability of the aminotransferase to transfer directly alpha-ketoglutarate to the alpha-ketoglutarate dehydrogenase complex in this multienzyme system plus the ability of succinyl-CoA, a product of this transfer, to inhibit citrate synthase could play a role in preventing alpha-ketoglutarate and citrate from accumulating in high levels. This would maintain the catalytic activity of the multienzyme system because alpha-ketoglutarate and citrate allosterically inhibit malate dehydrogenase and dissociate this enzyme from the multienzyme system. In addition, citrate also competitively inhibits fumarase. Consequently, when the levels of alpha-ketoglutarate and citrate are high and the multienzyme system is not required to convert glutamate to alpha-ketoglutarate, it is inactive. However, control by citrate would be expected to be absent in rapidly dividing tumors which characteristically have low mitochondrial levels of citrate.  相似文献   

17.
Two highly purified proteins with quite different properties capable of oxaloacetate keto-enol-tautomerase activity (oxaloacetate keto-enol-isomerase, EC 5.3.2.2) were isolated from the bovine heart mitochondrial matrix. The first protein has an apparent molecular mass of 37 kDa as determined by SDS-gel electrophoresis and Sephacryl SF-200 gel filtration. It is quite stable upon storage at 40 degrees C and reaches the maximal catalytic activity at pH 8.5 with a half-maximal activity at pH 7.0. The enzyme is specifically inhibited by oxalate and diethyloxaloacetate. When assayed in the enol----ketone direction at 25 degrees C (pH 9.0), the enzyme obeys a simple substrate saturation kinetics with Km and Vmax values of 45 microM and 74 units per mg of protein, respectively; the latter value corresponds to the turnover number of 2700 min-1. The second protein has an apparent molecular mass of 80 kDa as determined by SDS-gel electrophoresis and Sephacryl SF-300 gel filtration. The enzyme is rapidly inactivated at 40 degrees C and shows a sharp pH optimum of activity at pH 9.0. The enzyme can be completely protected from thermal inactivation by oxaloacetate and dithiothreitol. The kinetic parameters of the enzyme as assayed in the enol----ketone direction at 25 degrees C (pH 9.0) are: Km = 220 microM and Vmax = 20 units per mg of protein; the latter corresponds to the turnover number of 1600 min-1. The enzyme activity is specifically inhibited by maleate and pyrophosphate. About 30% of the total oxaloacetate tautomerase activity in crude mitochondrial matrix is represented by the 37 kDa enzyme and about 70% by the 80 kDa protein.  相似文献   

18.
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   

19.
Since ethanol consumption decreases hepatic aminotransferase activities in vivo, mechanisms of ethanol-mediated transaminase inhibition were explored in vitro using mitochondria-depleted rat liver homogenates. When homogenates were incubated at 37 degrees with 50 mM ethanol for 1 hr, alanine aminotransferase decreased by 20%, while aspartate aminotransferase was unchanged. After 2 hr, aspartate aminotransferase decreased by 20% and by 3 hr, alanine and aspartate aminotransferases were decreased by 31 and 23%, respectively. Levels of acetaldehyde generated during ethanol oxidation were 525 +/- 47 microM at 1 hr, 855 +/- 14 microM at 2 hr, and 1293 +/- 140 microM at 3 hr. Although inhibition of alcohol oxidation with methylpyrazole or cyanide markedly decreased ethanol-mediated transaminase inhibition, neither incubation with acetate nor generation of reducing equivalents by oxidation of lactate, malate, xylitol, or sorbitol altered the activity of either enzyme. However, semicarbazide, an aldehyde scavenger, prevented inhibition of both aminotransferases by ethanol. Moreover, incubation with 5 mM acetaldehyde for 1 hr inhibited alanine and aspartate aminotransferases by 36 and 26%, respectively. Cyanamide, an aldehyde dehydrogenase inhibitor, had little effect on ethanol-mediated transaminase inhibition. Thus, metabolism of ethanol by rat liver homogenates produces transaminase inhibition similar to that described in vivo and this effect requires acetaldehyde generation but not acetaldehyde oxidation. Since addition of pyridoxal 5'-phosphate to assay mixes did not reverse ethanol effects, aminotransferase inhibition does not result from displacement of vitamin B6 coenzymes.  相似文献   

20.
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   

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