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.
Decarboxylation of oxalacetate by pyruvate carboxylase   总被引:1,自引:0,他引:1  
P V Attwood  W W Cleland 《Biochemistry》1986,25(25):8191-8196
The decarboxylation of oxalacetate by pyruvate carboxylase in the absence of ADP and Pi is stimulated 400-fold by the presence of oxamate, which is an inhibitory analogue of pyruvate. The observation of substrate inhibition when either oxamate or oxalacetate is varied at a fixed concentration of the other indicates that both molecules bind at the same site on the enzyme. The pH profiles for this reaction show no evidence of the involvement of an enzymic acid-base catalyst, suggesting that the proton and CO2 units may be exchanged directly between the reactants (although CO2 sequestered in the active site may be an intermediate in the process). The pH profiles of the full reverse reaction of pyruvate carboxylase in which oxalacetate decarboxylation is coupled to ATP formation and where Pi is the variable substrate do, however, indicate that such an acid-base catalyst is involved in the other partial reaction of the enzyme in proton transfer to and from biotin. The enzyme also displays two oxamate-independent oxalacetate decarboxylating activities, one of which is biotin-dependent and the other is independent of biotin.  相似文献   

2.
S H Park  B G Harris  P F Cook 《Biochemistry》1986,25(13):3752-3759
Both chicken liver NADP-malic enzyme and Ascaris suum NAD-malic enzyme catalyze the metal-dependent decarboxylation of oxalacetate. Both enzymes catalyze the reaction either in the presence or in the absence of dinucleotide. The presence of dinucleotide increases the affinity of oxalacetate for the chicken liver NADP-malic enzyme, but this information could not be obtained in the case of A. suum NAD-malic enzyme because of the low affinity of free enzyme for NAD. The kinetic mechanism for oxalacetate decarboxylation by the chicken liver NADP-malic enzyme is equilibrium ordered at pH values below 5.0 with NADP adding to enzyme first. The Ki for NADP increases by a factor of 10 per pH unit below pH 5.0. An enzyme residue is required protonated for oxalacetate decarboxylation (by both enzymes) and pyruvate reduction (by the NAD-malic enzyme), but the beta-carboxyl of oxalacetate must be unprotonated for reaction (by both enzymes). The pK of the enzyme residue of the chicken liver NADP-malic enzyme decreases from a value of 6.4 in the absence of NADP to about 5.5 with Mg2+ and 4.8 with Mn2+ in the presence of NADP. The pK value of the enzyme residue required protonated for either oxalacetate decarboxylation or pyruvate reduction for the A. suum NAD-malic enzyme is about 5.5-6.0. Although oxalacetate binds equally well to protonated and unprotonated forms of the NADP-enzyme, the NAD-enzyme requires that oxalacetate or pyruvate selectively bind to the protonated form of the enzyme. Both enzymes prefer Mn2+ over Mg2+ for oxalacetate decarboxylation.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

3.
Studies were performed to elucidate factors involved in the regulation of pyruvate dehydrogenase activity in rat brain synaptosomes during membrane depolarization. Addition of 24 mM-KCl to synaptosomes resulted in increases in rates of O2 consumption (90%) and [1-(14)C]pyruvate decarboxylation (85%) and in the active/total ratio of extractable pyruvate dehydrogenase (90--100%) within 10 s. Neither pyruvate (10 mM) nor dichloroacetate (10 mM) affected the activation state of the enzyme complex. Also, the activation state of pyruvate dehydrogenase was unaffected by addition of 1 mM-octanoate, L-(--)-carnitine, 3-hydroxybutyrate, glutamate, citrate, lactate, L-malate, acetate, acetaldehyde or ethanol. Removal of Ca2+ by using EGTA lowered the active/total ratio to about 70%, although the rate of O2 consumption and pyruvate decarboxylation was unaffected. Rates of pyruvate decarboxylation in the presence of carbonyl cyanide p-trifluoromethoxyphenylhydrazone in the presence and absence of NaF and EGTA demonstrated a linear correlation with changes in the activity of the enzyme complex. This observation indicated that a change in the activation state of pyruvate dehydrogenase from 90 to 100% active could result in a 27% increase in the rate of pyruvate decarboxylation. It is suggested that the pyruvate dehydrogenase complex is an important site for the regulation of substrate utilization in rat brain synaptosomes. Further, the phosphorylation/dephosphorylation system and direct feedback-inhibitory effects on the enzyme complex both play a significant role in rapidly adapting pyruvate decarboxylation to changes in the requirements for mitochondrial energy production.  相似文献   

4.
Highly purified pyruvic decarboxylase (EC 4.1.1.1) from wheat germ catalyses the decarboxylation of hydroxypyruvate. A kinetic analysis of the activity of the enzyme with pyruvate and hydroxypyruvate as substrates suggests that a single enzyme is involved. The kinetics of decarboxylation are autocatalytic. The time lag before maximum activity is reached is affected by the concentration of hydroxypyruvate and the pH. The question whether or not hydroxypyruvate is a natural substrate for the enzyme remains unresolved, but it may be significant that at physiological pH (ca 7.5) the enzyme shows optimum activity with hydroxypyruvate, but negligible activity with pyruvate.  相似文献   

5.
In this paper, we report for the first time on the identification, purification, and characterization of the alpha-ketoisovalerate decarboxylase from Lactococcus lactis, a novel enzyme responsible for the decarboxylation into aldehydes of alpha-keto acids derived from amino acid transamination. The kivd gene consisted of a 1647 bp open reading frame encoding a putative peptide of 61 kDa. Analysis of the deduced amino acid sequence indicated that the enzyme is a non-oxidative thiamin diphosphate (ThDP)-dependent alpha-keto acid decarboxylase included in the pyruvate decarboxylase group of enzymes. The active enzyme is a homo-tetramer that showed optimum activity at 45 degrees C and at pH 6.5 and exhibited an inhibition pattern typical for metal-dependant enzymes. In addition to Mg(2+), activity was observed in presence of other divalent cations such as Ca(2+), Co(2+) and Mn(2+). The enzyme showed the highest specific activity (80.7 Umg(-1)) for alpha-ketoisovalerate, an intermediate metabolite in valine and leucine biosynthesis. On the other side, decarboxylation of indole-3-pyruvate and pyruvate only could be detected by a 100-fold increase in the enzyme concentration present in the reaction.  相似文献   

6.
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   

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.
The effect of the mitochondrial pyruvate transport inhibitors, α-cyanocinnamate and α-cyano-4-hydroxycinnamate, on the regulation of the pyruvate dehydrogenase multienzyme complex was investigated in the isolated perfused rat heart. Metabolic flux through pyruvate dehydrogenase was monitored by measuring 14CO2 production from [1-14C]pyruvate infused into the heart. A stepwise increase in the concentration of the inhibitor in the influent perfusate effected a stepwise reduction of the flux through the enzyme complex at all pyruvate concentrations tested. However, the magnitude of the α-cyanocinnamate-insensitive flux through pyruvate dehydrogenase increased markedly as the infused pyruvate concentration was elevated. The inhibition of pyruvate decarboxylation in the heart was nearly completely reversed following cessation of the inhibitor infusion. α-Cyanocinnamate was nearly 10 times more potent than α-cyano-4-hydroxycinnamate as an inhibitor of the flux through pyruvate dehydrogenase. Maximally inhibiting levels of α-cyano-4-hydroxycinnamate caused an increase in the ratio of the active form of pyruvate dehydrogenase to the total extractable enzyme complex from a value of 0.5 at 1 mm infused pyruvate (in the absence of the inhibitor) to a value of near unity. This result indicated that the intramitochondrial pyruvate concentration was severely depleted by the infusion of the inhibitor and that the enzyme complex was interconverted to its active form under these conditions. Removal of the inhibitor from the perfusion medium again lowered the ratio of the active/total pyruvate dehydrogenase to near its original level of 0.5 and restored the original flux through the enzyme complex indicating that mitochondrial pyruvate transport has been restored. The results of this study indicate that α-cyanocinnamate and its derivatives are effective inhibitors of pyruvate transport in the perfused heart and that carrier-mediated pyruvate transport can be an important parameter in the regulation of the activation state and the metabolic flux through the pyruvate dehydrogenase multienzyme complex in the heart.  相似文献   

9.
A kinetic study of "malic' enzyme (EC 1.1.1.40) from potato suggests that the mechanism is Ordered Bi Ter with NADP+ binding before malate, and NADPH binding before pyruvate and HCO3-. The analysis is complicated by the non-linearity that occurs in some of the plots. meso-Tartrate is shown to inhibit the oxidative decarboxylation of malate but to activate the reductive carboxylation of pyruvate. To explain these unidirectional effects it is suggested that the control site of "malic' enzyme binds organic acids (including meso-tartrate) which activate the enzyme. meso-Tartrate, however, competes with malate for the active site and thus inhibits the oxidative decarboxylation of malate. Because meso-tartrate does not compete effectively with pyruvate for enzyme-NADPH, its binding at the control site leads to a stimulation of the carboxylation of pyruvate. A similar explanation is advanced for the observation that malic acid stimulates its own synthesis.  相似文献   

10.
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   

11.
J L Wyatt  R F Colman 《Biochemistry》1977,16(7):1333-1342
Rabbit muscle pyruvate kinase is irreversibly inactivated upon incubation with the adenine nucleotide analogue, 5'-p-fluorosulfonylbenzoyladenosine. A plot of the time dependence of the logarithm of the enzymatic activity at a given time divided by the initial enzymatic activity(logE/Eo) reveals a biphasic rate of inactivation, which is consistent with a rapid reaction to form partially active enzyme having 54% of the original activity, followed by a slower reaction to yield totally inert enzyme. In addition to the pyruvate kinase activity of the enzyme, modification with 5'-p-fluorosulfonylbenzoyladenosine also disrupts its ability to catalyze the decarboxylation of oxaloacetate and the ATP-dependent enolization of pyruvate. In correspondence with the time dependence of inactivation, the rate of incorporation of 5'-p-[14C]fluorosulfonylbenzoyladenosine is also biphasic. Two moles of reagent per mole of enzyme subunit are bound when the enzyme is completely inactive. The pseudo-first-order rate constant for the rapid rate is linearly dependent on reagent concentration, whereas the constant for the slow rate exhibits saturation kinetics, suggesting that the reagent binds reversibly to the second site prior to modification. The adenosine moiety is essential for the effectiveness of 5'-p-fluorosulfonylbenzoyladenosine, since p-fluorosulfonylbenzoic acid does not inactivate pyruvate kinase at a significant rate. Thus, the reaction of 5'-p-fluorosulfonylbenzoyladenosine with pyruvate kinase exhibits several of the characteristics of affinity labeling of the enzyme. Protection against inactivation by 5'-p-fluorosulfonylbenzoyladenosine is provided by the addition to the incubation mixture of phosphoenolpyruvate. Mg-ADP or Mg2+. In contrast, the addition of pyruvate, Mg-ATP, or ADP and ATP alone has no effect on the rate of inactivation. These observations are consistent with the postulate that the 5'-p-fluorosulfonylbenzoyladenosine specifically labels amino acid residues in the binding region of Mg2+ and the phosphoryl group of phosphoenolpyruvate which is transferred during the catalytic reaction. The rate of inactivation increases with increasing pH, and k1 depends on the unprotonated form of an amino acid residue with pK = 8.5. On the basis of the pH dependence of the reaction of pyruvate kinase with 5'-p-fluorosulfonylbenzoyladenosine and the elimination of cysteine residues as possible sites of reaction, it is postulated that lysyl or tyrosyl residues are the most probably candidates for the critical amino acids.  相似文献   

12.
The isotope effect at C-1 on the H2O2-catalysed decarboxylation of pyruvate (used as a model reaction for the enzymic reaction) increases between pH 3 and 10 from 1.0007 +/- 0.0004 to 1.0283 +/- 0.0014 (25 degrees C). This result indicates a change in the rate-determining step from formation of the tetrahedral intermediate to decarboxylation of this intermediate. Practically no isotope fractionation at C-1 (1.0011 +/- 0.0002, pH 6.0, 25 degrees C) is found in the lactate oxidase-catalysed decarboxylation of lactate, which is indicative for the existence of an irreversible O2-dependent step prior to the enzyme-catalysed decarboxylation. In addition, the result provides further evidence that dissociation of pyruvate and H2O2 from the enzyme can be excluded. The isotope effect at C-2 of lactate in the enzymic reaction (1.0048 +/- 0.0004) is attributed to the hydrogen transfer step from lactate to the coenzyme.  相似文献   

13.
Growth of Acinetobacter calcoaceticus strain BD413 in malate-mineral medium resulted in the excretion of large quantities of oxalacetate. Malate was virtually depleted by the time the cell density reached 60% of its final value; most of the remaining growth took place at the expense of oxalacetate. Experiments in which oxalacetate was used as the initial substrate showed that pyruvate was not utilized until most of the oxalacetate disappeared. The generation time for growth on malate or oxalacetate was approximately 40 min; the generation time for growth on pyruvate was 62 min, which implies that pyruvate transport may be rate limiting. Oxalacetate and pyruvate, however, supported approximately the same growth yield. These observations suggested that the first step in the utilization of oxalacetate as an energy source consisted of an enzymatic decarboxylation of the keto acid to pyruvate and CO(2). Three enzyme reactions that carry out this decarboxylation have been detected in extracts of A. calcoaceticus. The first, which functioned maximally at pH 4.8, was attributable to the oxalacetate decarboxylase activity of oxidized diphosphopyridine nucleotide-malic enzyme. The second and third, which functioned in the neutral pH range, resulted from coupling of oxidized diphosphopyridine nucleotide-malic enzyme to reduced diphosphopyridine nucleotide-dependent malic dehydrogenase, and oxidized triphosphopyridine nucleotide-malic enzyme to a reduced triphosphopyridine nucleotide-dependent malic dehydrogenase. The efficiency of these coupled reactions was high enough so that the overall reaction could be physiologically significant.  相似文献   

14.
L M Abell  M H O'Leary 《Biochemistry》1988,27(16):5933-5939
The decarboxylation of histidine by the pyruvate-dependent histidine decarboxylase of Lactobacillus 30a shows a carbon isotope effect of k12/k13 = 1.0334 +/- 0.0005 and a nitrogen isotope effect k14/k15 = 0.9799 +/- 0.0006 at pH 4.8, 37 degrees C. The carbon isotope effect is slightly increased by deuteriation of the substrate and slightly decreased in D2O. The observed nitrogen isotope effect indicates that the imine nitrogen in the substrate-Schiff base intermediate complex is ordinarily protonated, and the pH dependence of the carbon isotope effect indicates that both protonated and unprotonated forms of this intermediate are capable of undergoing decarboxylation. As with the pyridoxal 5'-phosphate dependent enzyme, Schiff base formation and decarboxylation are jointly rate-limiting, with the intermediate histidine-pyruvate Schiff base showing a decarboxylation/Schiff base hydrolysis ratio of 0.5-1.0 at pH 4.8. The decarboxylation transition state is more reactant-like for the pyruvate-dependent enzyme than for the pyridoxal 5'-phosphate dependent enzyme. These studies find no particular energetic or catalytic advantage to the use of pyridoxal 5'-phosphate over covalently bound pyruvate in catalysis of the decarboxylation of histidine.  相似文献   

15.
The biotin-containing oxaloacetate decarboxylase from Klebsiella aerogenes catalyzed the Na+-dependent decarboxylation of oxaloacetate to pyruvate and bicarbonate (or CO2) but not the reversal of this reaction, not even in the presence of an oxaloacetate trapping system. The enzyme catalyzed an avidin-sensitive isotopic exchange between [1-14C]pyruvate and oxaloacetate, which indicated the intermediate formation of a carboxybiotin enzyme. Sodium ions were not required for this partial reaction, but promoted the second partial reaction, the decarboxylation of the carboxybiotin enzyme, thus accounting for the Na+ requirement of the overall reaction. Therefore, the 14CO2-enzyme which was formed upon incubation of the decarboxylase with [4-15C]oxaloacetate, could only be isolated if Na+ ions were excluded. Preincubation of the decarboxylase with avidin also prevented its labelling with 14CO2. The isolated 14CO2-labelled oxaloacetate decarboxylase revealed the following properties. It was slowly decarboxylated at neutral pH and rapidly upon acidification. The 14CO2 residues of the 14CO2-enzyme could be transferred to pyruvate yielding [4-14C]oxaloacetate. In the presence of Na+ this 14CO2 transfer was repressed by the simultaneous decarboxylation of the 14CO2-enzyme. However, Na+ alone was insufficient as a cofactor for the decarboxylation of the isolated 14CO2-enzyme, since this required pyruvate in addition to Na+. It is therefore concluded that the decarboxylation of oxaloacetate proceeds over a CO2-enzyme--pyruvate complex and that free CO2-enzyme is an abortive reaction intermediate. The activation energy of the enzymic decarboxylation of oxaloacetate changed with temperature and was about 113 kJ below 11 degrees C, 60 kJ between 11 degrees C and 31 degrees C and 36 kJ between 31--45 degrees C.  相似文献   

16.
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   

17.
Considerable evidence indicates that acetaldehyde is released from the leaves of a variety of plants. The conventional explanation for this is that ethanol formed in the roots is transported to the leaves where it is converted to acetaldehyde by the alcohol dehydrogenase (ADH) found in the leaves. It is possible that acetaldehyde could also be formed in leaves by action of pyruvate decarboxylase (PDC), an enzyme with an uncertain metabolic role, which has been detected, but not characterized, in cottonwood leaves. We have found that leaf PDC is present in leaf veins and petioles, as well as in non-vein tissues. Veins and petioles contained measurable pyruvate concentrations in the range of 2 mM. The leaf vein form of the enzyme was purified approximately 143-fold, and, at the optimum pH of 5.6, the Km value for pyruvate was 42 μM. This Km is lower than the typical millimolar range seen for PDCs from other sources. The purified leaf PDC also decarboxylates 2-ketobutyric acid (Km = 2.2 mM). We conclude that there are several possible sources of acetaldehyde production in cottonwood leaves: the well-characterized root-derived ethanol oxidation by ADH in leaves, and the decarboxylation of pyruvate by PDC in leaf veins, petioles, and other leaf tissues. Significantly, the leaf vein form of PDC with its high affinity for pyruvate, could function to shunt pyruvate carbon to the pyruvate dehydrogenase by-pass and thus protect the metabolically active vascular bundle cells from the effects of oxygen deprivation.  相似文献   

18.
Pyruvate decarboxylase [2-oxo acid carboxy-lyase, EC 4.1.1.1] was isolated from sweet potato roots and was partially purified from healthy and diseased tissues. There was no appreciable difference in properties between the enzymes from healthy and diseased tissues. The molecular weight of the enzyme was found to be 240,000 by polyacrylamide gel electrophoresis. Since sodium dodecyl sulfate polyacrylamide gel electrophoresis gave a molecular weight of 60,000 for the monomeric form of the enzyme, it is likely that sweet potato pyruvate decarboxylase contains 4 single polypeptide chains. The optimal pH of the decarboxylation reaction was 6.1--6.6. The Lineweaver-Burk double reciprocal plot curved upward, and the Hill coefficient was more than 1, with low concentrations of pyruvate. The enzyme was localized in the cytosol fraction. The activity of the enzyme increased in response to black-rot fungus infection, but decreased in response to cutting.  相似文献   

19.
The catalytic efficiency (kcat/Km) of Escherichia coli flavin pyruvate oxidase can be stimulated 450-fold either by the addition of lipid activators or by limited proteolytic hydrolysis. Previous studies have shown that a functional lipid binding site is a mandatory prerequisite for the in vivo functioning of this enzyme (Grabau, C., and Cronan, J. E., Jr. (1986) Biochemistry 25, 3748-3751). The effect of activation on the transient state kinetics of partial reactions in the overall oxidative conversion of pyruvate to acetate and CO2 has now been examined. The rate of decarboxylation of pyruvate to form CO2 and hydroxyethylthiamin pyrophosphate for both activated and unactivated forms of the enzyme is identical within experimental error. The decarboxylation step was measured using substrate concentrations of the enzyme in the absence of an electron acceptor. The pseudo-first order rate constant for the decarboxylation step is 60-80 s-1. The rate of oxidation of hydroxyethylthiamin pyrophosphate and concomitant enzyme-bound flavin reduction was analyzed by stopped-flow methods utilizing synthetic hydroxyethylthiamin pyrophosphate. The pseudo-first order rate for this step with unactivated enzyme was 2.85 s-1 and increased 145-fold for lipid-activated enzyme to 413 s-1 and 61-fold for the proteolytically activated enzyme to 173 s-1. The analysis of a third reaction step, the reoxidation of enzyme-bound FADH, was also investigated by stopped-flow techniques utilizing ferricyanide as the electron acceptor. The rate of oxidation of enzyme.FADH is very fast for both unactivated (1041 s-1) and activated enzyme (645 s-1). The data indicate that the FAD reduction step is the rate-limiting step in the overall reaction for unactivated enzyme. Alternatively, the rate-limiting step in the overall reaction with the activated enzyme shifts to one of the partial steps in the decarboxylation reaction.  相似文献   

20.
The regulation of the pyruvate dehydrogenase multienzyme complex was investigated during alpha-adrenergic stimulation with phenylephrine in the isolated perfused rat liver. The metabolic flux through the pyruvate dehydrogenase reaction was monitored by measuring the production of 14CO2 from infused [1-14C] pyruvate. In livers from fed animals perfused with a low concentration of pyruvate (0.05 mM), phenylephrine infusion significantly inhibited the rate of pyruvate decarboxylation without affecting the amount of pyruvate dehydrogenase in its active form. Also, phenylephrine caused no significant effect on tissue NADH/NAD+ and acetyl-CoA/CoASH ratios or on the kinetics of pyruvate decarboxylation in 14CO2 washout experiments. Phenylephrine inhibition of [1-14C]pyruvate decarboxylation was, however, closely associated with a decrease in the specific radioactivity of perfusate lactate, suggesting that the pyruvate decarboxylation response simply reflected dilution of the labeled pyruvate pool due to phenylephrine-stimulated glycogenolysis. This suggestion was confirmed in additional experiments which showed that the alpha-adrenergic-mediated inhibitory effect on pyruvate decarboxylation was reduced in livers perfused with a high concentration of pyruvate (1 mM) and was absent in livers from starved rats. Thus, alpha-adrenergic agonists do not exert short term regulatory effects on pyruvate dehydrogenase in the liver. Furthermore, the results suggest either that the rat liver pyruvate dehydrogenase complex is insensitive to changes in mitochondrial calcium or that changes in intramitochondrial calcium levels as a result of alpha-adrenergic stimulation are considerably less than suggested by others.  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