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.
McCormick, N. G. (University of Washington, Seattle), E. J. Ordal, and H. R. Whiteley. Degradation of pyruvate by Micrococcus lactilyticus. I. General properties of the formate-exchange reaction (J. Bacteriol. 83:887-898. 1962.-At an alkaline pH, extracts of Micrococcus lactilyticus(2) catalyze the phosphoroclastic degradation of pyruvate to formate and acetyl phosphate and the rapid exchange of formate into the carboxyl group of pyruvate. At an acid pH, hydrogen, carbon dioxide, and acetyl phosphate are produced, and carbon dioxide is exchanged into the carboxyl group of pyruvate. A concentration of approximately 1 m phosphate is required for the phosphoroclastic reaction and formate exchange; the production of carbon dioxide and hydrogen is greatly inhibited by high concentrations of phosphate. Formate exchange requires a divalent metal ion and is stimulated by reducing agents and an atmosphere of hydrogen. Inhibition by p-chloromercuribenzoate, Zn(++), Cd(++), and arsenite indicates that sulfhydryl groups on the enzyme are involved in the reaction; the inhibition by arsenite and Cd(++) may be relieved by 2,3-dimercaptopropanol, suggesting that vicinal dithiols may be required. Inhibition by hypophosphite may reflect a competition with formate for a site on the enzyme.At an alkaline pH, alpha-ketobutyrate is degraded to propionate and formate, whereas alpha-ketoglutarate is fermented to succinate, propionate, carbon dioxide, hydrogen, and formate. Formate is exchanged into the carboxyl groups of alpha-ketobutyrate and alpha-ketoglutarate under these conditions. Only traces of alpha-ketovalerate and alpha-ketoisovalerate are fermented at an alkaline pH and the exchange of formate into these compounds is very low.The addition of viologen dyes under the conditions used for formate exchange causes a reduction of pyruvate, alpha-ketobutyrate, alpha-ketovalerate, and alpha-ketoisovalerate to the corresponding alpha-hydroxy acids.  相似文献   

2.
Pyruvate metabolism by aminopterin-inhibited Aerobacter aerogenes   总被引:1,自引:1,他引:0  
1. The synthesis and utilization of both alanine (by reductive amination, oxidative deamination and transamination) and valine (by transamination only) in Aerobacter aerogenes are unaffected by aminopterin. These amino acids, which accumulate in aminopterin-treated cultures of this organism, are therefore considered to be formed as secondary products from the excess of pyruvate that also accumulates. 2. Oxidative metabolism of pyruvate and the synthesis of acetylmethylcarbinol by A. aerogenes cells are unaltered by growth in the presence of aminopterin. 3. Cells from static and anaerobic cultures that have been treated with the folic acid antagonist in the early exponential phase have a decreased ability to cleave pyruvate to acetate and formate, and to effect the exchange of formate with the carboxyl group of pyruvate. 4. 3-Methyl-2-oxobutanoate, the keto acid precursor of valine, cannot replace pyruvate as substrate in either the phosphoroclastic or the exchange reaction.  相似文献   

3.
Bisulfite reductase, purified from Desulfovibrio vulgaris, was coupled with the pyruvate phosphoroclastic reaction. Moderate to low reducing conditions resulted in the formation of trithionate; however, when the concentration of reductant was high, a mixture of trithionate and thiosulfate was formed. Sulfide was also a detectable product, but only when the concentration of bisulfite was low. Flavodoxin mediated native coupling between bisulfite reductase and the phosphoroclastic reaction. A model for bisulfite reductase activity is proposed.  相似文献   

4.
Conversion of glucose to pyruvate via reactions homologous to the non-phosphorylated Entner-Doudoroff (non-P ED) pathway could be achieved in the presence of two amino acid catalysts, cysteine and histidine: cystine oxidizes glucose to gluconic acid by the reaction homologous to glucose dehydrogenase and histidine changes gluconic acid to 2-keto-3-deoxy gluconic acid, then to pyruvate by the reaction homologous to gluconic acid dehydratase and 2-keto-3-deoxy gluconate aldolase, respectively. Pyruvate can be converted to acetyl CoA by the reaction with CoA, TPP and FAD in the presence of cysteine and histidine, which resembles pyruvate dehydrogenase reaction. It was found that gluconic acid dehydration alone is non-specific, in contrast to other reactions. The non-P ED pathway is used by some extreme thermophiles in bacteria and archaea, usually thought as the oldest among the contemporary organisms. This study suggests the possible contribution of amino acid to the origin of the glycolytic pathway.  相似文献   

5.
Glucose metabolism and the mechanisms of NADH oxidation by Treponema hyodysenteriae were studied. Under an N2 atmosphere, washed cell suspensions of the spirochete consumed glucose and produced acetate, butyrate, H2, and CO2. Approximately twice as much H2 as CO2 was produced. Determinations of radioactivity in products of [14C]glucose and [14C]pyruvate metabolism and analyses of enzyme activities in cell lysates revealed that glucose was catabolized to pyruvate via the Embden-Meyerhof-Parnas pathway. The results of pyruvate exchange reactions with NaH14CO3 and Na14COOH demonstrated that pyruvate was converted to acetyl coenzyme A (acetyl-CoA), H2, and CO2 by a clostridium-type phosphoroclastic mechanism. NADH:ferredoxin oxidoreductase and hydrogenase activities were present in cell lysates and produced H2 from NADH oxidation. Phosphotransacetylase and acetate kinase catalyzed the formation of acetate from acetyl-CoA. Butyrate was formed from acetyl-CoA via a pathway that involved 3-hydroxybutyryl-coenzyme A (CoA) dehydrogenase, butyryl-CoA dehydrogenase, and butyryl-CoA transferase. T. hyodysenteriae cell suspensions generated less H2 and butyrate under 10% O2-90% N2 than under 100% N2. Cell lysates contained NADH oxidase, NADH peroxidase, and superoxide dismutase activities. These findings indicated there are three major mechanisms that T. hyodysenteriae cells use to recycle NADH generated from the Embden-Meyerhof-Parnas pathway--enzymes in the pathway from acetyl-CoA to butyrate, NADH:ferredoxin oxidoreductase, and NADH oxidase. Versatility in methods of NADH oxidation and an ability to metabolize oxygen could benefit T. hyodysenteriae cells in the colonization of tissues of the swine large bowel.  相似文献   

6.
Glucose metabolism and the mechanisms of NADH oxidation by Treponema hyodysenteriae were studied. Under an N2 atmosphere, washed cell suspensions of the spirochete consumed glucose and produced acetate, butyrate, H2, and CO2. Approximately twice as much H2 as CO2 was produced. Determinations of radioactivity in products of [14C]glucose and [14C]pyruvate metabolism and analyses of enzyme activities in cell lysates revealed that glucose was catabolized to pyruvate via the Embden-Meyerhof-Parnas pathway. The results of pyruvate exchange reactions with NaH14CO3 and Na14COOH demonstrated that pyruvate was converted to acetyl coenzyme A (acetyl-CoA), H2, and CO2 by a clostridium-type phosphoroclastic mechanism. NADH:ferredoxin oxidoreductase and hydrogenase activities were present in cell lysates and produced H2 from NADH oxidation. Phosphotransacetylase and acetate kinase catalyzed the formation of acetate from acetyl-CoA. Butyrate was formed from acetyl-CoA via a pathway that involved 3-hydroxybutyryl-coenzyme A (CoA) dehydrogenase, butyryl-CoA dehydrogenase, and butyryl-CoA transferase. T. hyodysenteriae cell suspensions generated less H2 and butyrate under 10% O2-90% N2 than under 100% N2. Cell lysates contained NADH oxidase, NADH peroxidase, and superoxide dismutase activities. These findings indicated there are three major mechanisms that T. hyodysenteriae cells use to recycle NADH generated from the Embden-Meyerhof-Parnas pathway--enzymes in the pathway from acetyl-CoA to butyrate, NADH:ferredoxin oxidoreductase, and NADH oxidase. Versatility in methods of NADH oxidation and an ability to metabolize oxygen could benefit T. hyodysenteriae cells in the colonization of tissues of the swine large bowel.  相似文献   

7.
Preparations of pyruvate formate-lyase were made from Escherichia coli cells. Net reversal of the "phosphoroclastic split" of pyruvate was readily demonstrated with these preparations. Incubation of acetyl phosphate with formate resulted in the accumulation of pyruvate in concentrations up to 0.5 mm. Catalytic amounts of coenzyme A were essential. Pyruvate was also readily formed from acetyl coenzyme A and formate. The equilibrium constant of the reaction (pyruvate(-) + HPO(4) (2-) --> acetyl phosphate(2-) + formate(-)) has been determined to be about 23 at 37 C.  相似文献   

8.
Pyruvate Metabolism in Sarcina maxima   总被引:6,自引:4,他引:2       下载免费PDF全文
The mechanisms of pyruvate cleavage and hydrogen production by Sarcina maxima were studied. It was found that a phosphoroclastic system for pyruvate oxidation, similar to that occurring in saccharolytic clostridia, is present in S. maxima. Cleavage of pyruvate by extracts of the latter organism resulted in the formation of acetyl phosphate, CO(2), and electrons which were transferred to ferredoxin. Formate was not an intermediate in this system. Pyruvate oxidation was coupled with ferredoxin-dependent nicotinamide adenine dinucleotide phosphate (NADP) reduction. A hydrogenase, active in particulate extracts of S. maxima, did not accept electrons from reduced ferredoxin. Formate was detected as a fermentation product when S. maxima was grown in media buffered with CaCO(3). Whole cells and extracts degraded formate to H(2) and CO(2). The evidence suggests that electrons generated by ferredoxin-linked pyruvate oxidation by S. maxima are not used for H(2) production, but that they serve for the reduction of NADP. Reduced NADP may be utilized by the organisms for synthesis of cell material. Production of H(2) by S. maxima may occur through a pyruvate clastic system similar to that present in coliform bacteria.  相似文献   

9.
Under anaerobic conditions, cells of Entamoeba histolytica grown with bacteria produce H2 and acetate while cells grown axenically produce neither. Aerobically, acetate is produced and O2 is consumed by amebae from either type of cells. Centrifuged extracts, 2.4 x 106 x g x min, from both types of cells contain pyruvate synthase (EC 1.2.7.1) and an acetate thiokinase which, together, form a system capable of converting pyruvate to acetate. Pyruvate synthase catalyzes the reaction: pyruvate + CoA leads to CO2 + acetyl-CoA + 2E. Electron acceptors which function with this enzyme are FAD, FMN, riboflavin, ferredoxin, and methyl viologen, but not NAD or NADP. The amebal acetate thiokinase catalyzes the reaction acetyl-CoA + ADP + Pi leads to acetate + ATP + CoA. For this apparently new enzyme we suggest the trivial name acetyl-CoA-synthetase (ADP-forming). Extracts from axenic amebae do not contain hydrogenase, but extracts from cells grown with bacteria do. It is postulated that in bacteria-grown amebae electrons generated at the pyruvate synthase step are utilized anaerobically to produce H2 via the hydrogenase and that the acetyl-CoA is converted to acetate in an energy-conserving step catalyzed by amebal acetyl-CoA synthetase. Aerobically, cells grown under either regimen may utilize the energy-conserving pyruvate-to-acetate pathway since O2 then serves as the ultimate electron acceptor.  相似文献   

10.
McCormick, N. G. (University of Washington, Seattle), E. J. Ordal, and H. R. Whiteley. Degradation of pyruvate by Micrococcus lactilyticus. II. Studies of cofactors in the formate-exchange reaction. J. Bacteriol. 83:899-906. 1962.-Enzyme preparations from Micrococcus lactilyticus(2) are rendered inactive with respect to formate exchange by treatment with charcoal or Dowex-50, by dialysis, or by fractionation with ammonium sulfate. The activity may be completely restored by a "kochsaft" preparation (BES) obtained from M. lactilyticus and partially restored by similar BES preparations from Escherichia coli and Clostridium butyricum. Diphosphothiamine is required for formate exchange but full activity cannot be restored by known cofactors. Brief exposure to increased temperatures, air, extremes of pH, and absorption with charcoal and Dowex-50 decrease the cofactor activity of BES preparations. The addition of BES preparations from E. coli and Streptococcus faecalis causes a shift in the degradation of pyruvate by extracts of M. lactilyticus from the phosphoroclastic cleavage (to acetyl phosphate and formate) to the dismutation of pyruvate (to lactate, acetate, and carbon dioxide).C. cylindrosporum was found to mediate the formate-exchange reaction; the activity of crude extracts was stimulated by M. lactilyticus and C. butyricum BES preparations. M. lactilyticus BES also increased the formate-exchange activity of extracts of E. coli.  相似文献   

11.
The kinetic properties for the native forward reaction of pyruvate:NADP+ oxidoreductase from Euglena gracilis were determined. The substrate kinetics gave a pattern of a ping-pong mechanism involving a competitive substrate inhibition of CoA against pyruvate. The Km values for pyruvate, CoA, and NADP+ were estimated to be 27, 6.6, and 28 microM, respectively, and the Ki value of CoA against pyruvate was 28 microM. CO2 inhibited noncompetitively against pyruvate and NADP+, and uncompetitively against CoA. Acetyl-CoA showed a competitive inhibition with respect to pyruvate and an uncompetitive inhibition with respect to NADP+. NADPH inhibited competitively versus NADP+, noncompetitively versus CoA, and uncompetitively versus pyruvate. The kinetic behavior is consistent with a two-site ping-pong mechanism involving the substrate inhibition. From the kinetic mechanism, it is proposed that the enzyme has two catalytic sites linked by an intramolecular electron-transport chain. One of these is a thiamine pyrophosphate-containing catalytic site which reacts with pyruvate and CoA to form CO2 and acetyl-CoA, and the other site functions in the reduction of NADP+. In contrast, when methyl viologen was used as an artificial one-electron acceptor substituting for NADP+, the reaction gave a pattern characteristic of an octa uni ping-pong mechanism involving a competitive substrate inhibition of CoA against pyruvate.  相似文献   

12.
Sheep kidney pyruvate carboxylase has been desensitized against its allosteric effector, acetyl CoA, by limited covalent modification with trinitrobenzene sulphonic acid.Trinitrophenylation of the enzyme resulted in a strong inhibition of the rate of the acetyl CoA-stimulated pyruvate carboxylation and enhancement of the rate of the acetyl CoA-independent reaction. A good correlation was found between the requirement for acetyl CoA of the exchange reactions catalysed by the enzyme and the extent of their inhibition by trinitrobenzene sulphonic acid modification.Spectrophotometric data indicated that one to two lysyl residues per monomer were trinitrophenylated. Modification had only a slight effect on the sedimentation properties of the enzyme.  相似文献   

13.
Effect of pyruvate on glucose metabolism in Clostridium acetobutylicum   总被引:1,自引:0,他引:1  
Pyruvate effects on the metabolism of Clostridium acetobutylicum during glucose fermentation were studied. After addition to the culture medium, the pyruvate was rapidly used, provoking several changes in the metabolic pattern of the bacteria. When pyruvate addition occurred early in the fermentation, the glucose utilization decreased and the solventogenic phase was not induced. When pyruvate was added during solventogenesis, glucose consumption was slightly affected and the cells fermented both substrates simultaneously: however, the acidogenic phase started again to the detriment of solvent formation. Usually, during the solvent phase, the cells remetabolized acetic and butyric acids into solvents, but when pyruvate was added, the utilization of acids was stopped and the specific rates of acetate and butyrate formation increased immediately. The acidogenic growth phase was characterized by high levels of acetate and butyrate kinase which dropped during the solvent phase. Addition of pyruvate limited the down shift of these two enzymes and the levels of the activities remained constant during the course of the fermentation. Conversely, the acetoacetate decarboxylase, which is characteristic of the solvent phase, decreased sharply in the presence of pyruvate. The fact that the specific rate of glucose consumption was not decreased by the pyruvate metabolism, a cosubstrate, proves that the phosphoroclastic reaction is not a limiting step. Furthermore, the pyruvate utilization represented a promising approach to obtain useful data on the intracellular compounds implicated in the mechanism for switching from the acidogenic to the solventogenic phase.  相似文献   

14.
Most conditions detected by expanded newborn screening result from deficiency of one of the enzymes that degrade acyl-coenzyme A (CoA) esters in mitochondria. The role of acyl-CoAs in the pathophysiology of these disorders is poorly understood, in part because CoA esters are intracellular and samples are not generally available from human patients. We created a mouse model of one such condition, deficiency of 3-hydroxy-3-methylglutaryl-CoA lyase (HL), in liver (HLLKO mice). HL catalyses a reaction of ketone body synthesis and of leucine degradation. Chronic HL deficiency and acute crises each produced distinct abnormal liver acyl-CoA patterns, which would not be predictable from levels of urine organic acids and plasma acylcarnitines. In HLLKO hepatocytes, ketogenesis was undetectable. Carboxylation of [2-14C] pyruvate diminished following incubation of HLLKO hepatocytes with the leucine metabolite 2-ketoisocaproate (KIC). HLLKO mice also had suppression of the normal hyperglycemic response to a systemic pyruvate load, a measure of gluconeogenesis. Hyperammonemia and hypoglycemia, cardinal features of many inborn errors of acyl-CoA metabolism, occurred spontaneously in some HLLKO mice and were inducible by administering KIC. KIC loading also increased levels of several leucine-related acyl-CoAs and reduced acetyl-CoA levels. Ultrastructurally, hepatocyte mitochondria of KIC-treated HLLKO mice show marked swelling. KIC-induced hyperammonemia improved following administration of carglumate (N-carbamyl-L-glutamic acid), which substitutes for the product of an acetyl-CoA-dependent reaction essential for urea cycle function, demonstrating an acyl-CoA-related mechanism for this complication.  相似文献   

15.
1. A method was devised for preparing pig heart pyruvate dehydrogenase free of thiamin pyrophosphate (TPP), permitting studies of the binding of [35S]TPP to pyruvate dehydrogenase and pyruvate dehydrogenase phosphate. The Kd of TPP for pyruvate dehydrogenase was in the range 6.2-8.2 muM, whereas that for pyruvate dehydrogenase phosphate was approximately 15 muM; both forms of the complex contained about the same total number of binding sites (500 pmol/unit of enzyme). EDTA completely inhibited binding of TPP; sodium pyrophosphate, adenylyl imidodiphosphate and GTP, which are inhibitors (competitive with TPP) of the overall pyruvate dehydrogenase reaction, did not appreciably affect TPP binding. 2. Initial-velocity patterns of the overall pyruvate dehydrogenase reaction obtained with varying TPP, CoA and NAD+ concentrations at a fixed pyruvate concentration were consistent with a sequential three-site Ping Pong mechanism; in the presence of oxaloacetate and citrate synthase to remove acetyl-CoA (an inhibitor of the overall reaction) the values of Km for NAD+ and CoA were 53+/- 5 muM and 1.9+/-0.2 muM respectively. Initial-velocity patterns observed with varying TPP concentrations at various fixed concentrations of pyruvate were indicative of either a compulsory order of addition of substrates to form a ternary complex (pyruvate-Enz-TPP) or a random-sequence mechanism in which interconversion of ternary intermediates is rate-limiting; values of Km for pyruvate and TPP were 25+/-4 muM and 50+/-10 nM respectively. The Kia-TPP (the dissociation constant for Enz-TPP complex calculated from kinetic plots) was close to the value of Kd-TPP (determined by direct binding studies). 3. Inhibition of the overall pyruvate dehydrogenase reaction by pyrophosphate was mixed non-competitive versus pyruvate and competitive versus TPP; however, pyrophosphate did not alter the calculated value for Kia-TPP, consistent with the lack of effect of pyrophosphate on the Kd for TPP. 4. Pyruvate dehydrogenase catalysed a TPP-dependent production of 14CO2 from [1-14C]pyruvate in the absence of NAD+ and CoA at approximately 0.35% of the overall reaction rate; this was substantially inhibited by phosphorylation of the enzyme both in the presence and absence of acetaldehyde (which stimulates the rate of 14CO2 production two- or three-fold). 5. Pyruvate dehydrogenase catalysed a partial back-reaction in the presence of TPP, acetyl-CoA and NADH. The Km for TPP was 4.1+/-0.5 muM. The partial back-reaction was stimulated by acetaldehyde, inhibited by pyrophosphate and abolished by phosphorylation. 6. Formation of enzyme-bound [14C]acetylhydrolipoate from [3-14C]pyruvate but not from [1-14C]acetyl-CoA was inhibited by phosphorylation. Phosphorylation also substantially inhibited the transfer of [14C]acetyl groups from enzyme-bound [14C]acetylhydrolipoate to TPP in the presence of NADH. 7...  相似文献   

16.
Procedures are described for isolating highly purified porcine liver pyruvate and α-ketoglutarate dehydrogenase complexes. Rabbit serum stabilized these enzyme complexes in mitochondrial extracts, apparently by inhibiting lysosomal proteases. The complexes were purified by a three-step procedure involving fractionation with polyethylene glycol, pelleting through 12.5% sucrose, and a second fractionation under altered conditions with polyethylene glycol. Sedimentation equilibrium studies gave a molecular weight of 7.2 × 106 for the liver pyruvate dehydrogenase complex. Kinetic parameters are presented for the reaction catalyzed by the pyruvate dehydrogenase complex and for the regulatory reactions catalyzed by the pyruvate dehydrogenase kinase and pyruvate dehydrogenase phosphatase. For the overall catalytic reaction, the competitive Ki to Km ratio for NADH versus NAD+ and acetyl CoA versus CoA were 4.7 and 5.2, respectively. Near maximal stimulations of pyruvate dehydrogenase kinase by NADH and acetyl CoA were observed at NADH:NAD+ and acetyl CoA:CoA ratios of 0.15 and 0.5, respectively. The much lower ratios required for enhanced inactivation of the complex by pyruvate dehydrogenase kinase than for product inhibition indicate that the level of activity of the regulatory enzyme is not directly determined by the relative affinity of substrates and products of catalytic sites in the pyruvate dehydrogenase complex. In the pyruvate dehydrogenase kinase reaction, K+ and NH+4 decreased the Km for ATP and the competitive inhibition constants for ADP and (β,γ-methylene)adenosine triphosphate. Thiamine pyrophosphate strongly inhibited kinase activity. A high concentration of ADP did not alter the degree of inhibition by thiamine pyrophosphate nor did it increase the concentration of thiamine pyrophosphate required for half-maximal inhibition.  相似文献   

17.
Bifidobacterium breve NCFB 2257 was grown in glucose-limited and nitrogen (N)-limited chemostats at dilution rates (D) from 0.04 to 0.60 h–1, to study the effect of nutrient availability on carbohydrate metabolism. The results showed that D had little effect on fermentation product formation, irrespective of the form of nutrient limitation. However, marked differeces were observed in the distribution of fermentation products, that were attributable to glucose availability. In glucose-limited cultures, formate and acetate were the principal end-products of metabolism. Lactate was never detected under these growth conditions. In contrast, lactate and acetate were mainly formed when glucose was in excess, and formate was not produced. These results are explained by the metabolic fate of pyruvate, which can be dissimilated by either phosphoroclastic cleavage to acetyl phosphate and formate, or alternatively, it may be reduced to lactate. Enzymic studies were made to establish the mechanisms that regulated pyruvate metabolism. The data demonstrated that control was not exercised through regulation of the synthesis and activity of lactate dehydrogenase (LDH), phosphofructokinase or alcohol dehydrogenase. It is possible however, that there was competition for pyruvate by LDH and the phosphoroclastic enzyme, which would determine the levels of lactate and formate produced respectively. These results demonstrate the metabolic flexibility of B. breve, which preferentially uses lactate as an electron sink during N-limited growth, whereas under energy-limitation, carbon flow is directed towards acetyl phosphate to maximise ATP synthesis. Correspondence to: B. A. Degnan  相似文献   

18.
Furdui C  Ragsdale SW 《Biochemistry》2002,41(31):9921-9937
Pyruvate:ferredoxin oxidoreductase (PFOR) catalyzes the coenzyme A (CoA)-dependent oxidative decarboxylation of pyruvate. In many autotrophic anaerobes, PFOR links the Wood-Ljungdahl pathway to glycolysis and to cell carbon synthesis. Herein, we cloned and sequenced the M. thermoacetica PFOR, demonstrating strong structural homology with the structurally characterized D. africanus PFOR, including the presence of three [4Fe-4S] clusters per monomeric unit. The PFOR reaction includes a hydroxyethyl-thiamin pyrophosphate (HE-TPP) radical intermediate, which forms rapidly after PFOR reacts with pyruvate. This step precedes electron transfer from the HE-TPP radical intermediate to an intramolecular [4Fe-4S] cluster. We show that CoA increases the rate of this redox reaction by 10(5)-fold. Analysis by Marcus theory indicates that, in the absence of CoA, this is a true electron-transfer reaction; however, in its presence, electron transfer is gated by an adiabatic event. Analysis by the Eyring equation indicates that entropic effects dominate this rate enhancement. Our results indicate that the energy of binding CoA contributes minimally to the rate increase since the thiol group of CoA lends over 40 kJ/mol to the reaction, whereas components of CoA that afford most of the cofactor's binding energy contribute minimally. Major conformational changes also do not appear to explain the rate enhancement. We propose several ways that CoA can accomplish this rate increase, including formation of a highly reducing adduct with the HE-TPP radical to increase the driving force for electron transfer. We also consider the possibility that CoA itself forms part of the electron-transfer pathway.  相似文献   

19.
The kinetic mechanisms of the 2-oxoglutarate and pyruvate dehydrogenease complexes from pig heart mitochondria were studied at pH 7.5 and 25 degrees. A three-site ping-pong mechanism for the actin of both complexes was proposed on the basis of the parallel lines obtained when 1/v was plotted against 2-oxoglutarate or pyruvate concentration for various levels of CoA and a level of NAD+ near its Michaelis constant value. Rate equations were derived from the proposed mechanism. Michaelis constants for the reactants of the 2-oxoglutarate dehydrogenase complex reaction are: 2-oxoglutarate, 0.220 mM; CoA, 0.025 mM; NAD+, 0.050 mM. Those of the pyruvate dehydrogenase complex are: pyruvate, 0.015 mM; CoA, 0.021 mM; NAD+, 0.079 mM. Product inhibition studies showed that succinyl-CoA or acetyl-CoA was competitive with respect to CoA, and NADH was competitive with respect to NAD+ in both overall reactions, and that succinyl-CoA or acetyl-CoA and NADH were uncompetitive with respect to 2-oxoglutarate or pyruvate, respectively. However, noncompetitive (rather than uncompetitive) inhibition patterns were observed for succinyl-CoA or acetyl-CoA versus NAD+ and for NADH versus CoA. These results are consistent with the proposed mechanisms.  相似文献   

20.
Pyruvate dehydrogenase from Desulfovibrio vulgaris Miyazaki F was partially purified from the soluble fraction of the bacterial sonicate, and characterized. The enzyme catalyzes oxidative decarboxylation of pyruvate to produce acetyl-CoA, in contrast to statements in current review articles in which acetyl phosphate is indicated to be a direct decomposition product of pyruvate in sulfate-reducing bacteria. The established reaction stoichiometry is: pyruvate + CoA + FMN----acetyl-CoA + CO2 + FMNH2. The Km values are 2.9 mM for pyruvate, 32 microM for CoA and 6.7 mumol for FMN. Participation of thiamine diphosphate in the enzymic process was not proven. 2-Oxobutyrate, but not 2-oxoglutarate, can substitute for pyruvate. The three flavin compounds, FMN, FAD, and flavodoxin, as well as clostridial ferredoxin, serve as electron carriers for the enzyme. Thus the enzyme is a kind of pyruvate synthase [EC 1.2.7.1], but acts in the direction of pyruvate degradation in the growing cells. The rate of cytochrome C3 reduction is extremely low, but in the presence of flavodoxin as an electron mediator, the reduction rate of cytochrome C3 becomes faster than the reduction rate of flavodoxin alone. It seems that the physiological electron acceptor for this enzyme is flavodoxin, which might be complexed with cytochrome C3 to produce a very efficient electron transfer system in the cell. The soluble fraction of D. vulgaris cells has been proved to contain, in addition to the pyruvate dehydrogenase, lactate dehydrogenase (Ogata, M., Arihara, K., & Yagi, T. (1981) J. Biochem. 89, 1423-1431), phosphate acetyltransferase and acetate kinase, i.e., all the enzymes necessary to convert lactate to acetate, producing ATP by substrate level phosphorylation.  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