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.
The transport of the tricarboxylic acid cycle C(4)-dicarboxylic acids was studied in both the wild-type strain and tricarboxylic acid cycle mutants of Bacillus subtilis. Active transport of malate, fumarate, and succinate was found to be inducible by these dicarboxylic acids or by precursors to them, whereas glucose or closely related metabolites catabolite-repressed their uptake. l-Malate was found to be the best dicarboxylic acid transport inducer in succinic dehydrogenase, fumarase, and malic dehydrogenase mutants. Succinate and fumarate are accumulated over 100-fold in succinic dehydrogenase and fumarase mutants, respectively, whereas mutants lacking malate dehydrogenase were unable to accumulate significant quantities of the C(4)-dicarboxylic acids. The stereospecificity of this transport system was studied from a comparison of the rates of competitive inhibition of both succinate uptake and efflux in a succinate dehydrogenase mutant by utilizing thirty dicarboxylic acid analogues. The system was specific for the C(4)-dicarboxylic acids of the tricarboxylic acid cycle, neither citrate nor alpha-ketoglutarate were effective competitive inhibitors. Of a wide variety of metabolic inhibitors tested, inhibiors of oxidative phosphorylation and of the formation of proton gradients were the most potent inhibitors of transport. From the kinetics of dicarboxylic acid transport (K(m) approximately 10(-4) M for succinate or fumarate in succinic acid dehydrogenase and fumarase mutants) and from the competitive inhibition studies, it was concluded that an inducible dicarboxylic acid transport system mediates the entry of malate, fumarate, or succinate into B. subtilis. Mutants devoid of alpha-ketoglutarate dehydrogenase were shown to accumulate both alpha-ketoglutarate and glutamate, and these metabolites subsequently inhibited the transport of all the C(4)-dicarboxylic acids, suggesting a regulatory role.  相似文献   

2.
Malate and succinate were taken up rapidly by isolated, intact peribacteroid units (PBUs) from soybean (Glycine max (L.) Merr.) root nodules and inhibited each other in a competitive manner. Malonate uptake was slower and was severely inhibited by equimolar malate in the reaction medium. The apparent Km for malonate uptake was higher than that for malate and succinate uptake. Malate uptake by PBUs was inhibited by (in diminishing order of severity) oxaloacetate, fumarate, succinate, phthalonate and oxoglutarate. Malonate and butylmalonate inhibited only slightly and pyruvate,isocitrate and glutamate not at all. Of these compounds, only oxaloacetate, fumarate and succinate inhibited malate uptake by free bacteroids. Malate uptake by PBUs was inhibited severely by the uncoupler carbonylcyanidem-chlorophenyl hydrazone and the respiratory poison KCN, and was stimulated by ATP. We conclude that the peribacteroid membrane contains a dicarboxylate transport system which is distinct from that on the bacteroid membrane and other plant membranes. This system can catalyse the rapid uptake of a range of dicarboxylates into PBUs, with malate and succinate preferred substrates, and is likely to play an important role in symbiotic nitrogen fixation. Energization of both the bacteroid and peribacteroid membranes controls the rate of dicarboxylate transport into peribacteroid units.  相似文献   

3.
Succinate transport in Rhizobium leguminosarum.   总被引:19,自引:13,他引:6       下载免费PDF全文
The transport of succinate was studied in an effective streptomycin-resistant strain of Rhizobium leguminosarum. High levels of succinate transport occurred when cells were grown on succinate, fumarate, or malate, whereas low activity was found when cells were grown on glucose, sucrose, arabinose, or pyruvate as the sole carbon source. Because of the rapid metabolism of succinate after transport into the cells, a succinate dehydrogenase-deficient mutant was isolated in which intracellular succinate accumulated to over 400 times the external concentration. Succinate transport was completely abolished in the presence of metabolic uncouplers but was relatively insensitive to sodium arsenate. Succinate transport was a saturable function of the succinate concentration, and the apparent Km and Vmax values for transport were determined in both the parent and the succinate dehydrogenase mutant. Malate and fumarate competitively inhibited succinate transport, whereas citrate and malonate had no effect. Succinate transport mutants were isolated by transposon (Tn5) mutagenesis. These mutants were unable to transport succinate or malate and were unable to grow on succinate, malate, or fumarate as the sole carbon source. The mutants grew normally on pyruvate, oxaloacetate, citrate, or arabinose, and revertants isolated on succinate minimal medium had regained the ability to grow on malate and fumarate. From these data, we conclude that R. leguminosarum possesses a C4-dicarboxylic acid transport system which is inducible and mediates the active transport of succinate, fumarate, and malate into the cell.  相似文献   

4.
Transport of succinate by Pseudomonas putida   总被引:9,自引:0,他引:9  
Induced succinate uptake and transport (defined as transport of a compound followed by its metabolism and transport in the absence of subsequent metabolism) by Pseudomonas putida are active processes resulting in intracellular succinate concentrations 10-fold that of the initial extracellular concentration. Uptake was studied with the wild-type strain P. putida P2 and transport with a mutant deficient in succinate dehydrogenase activity. Addition of succinate, fumarate, or malate to the growth medium induces both processes above a basal level. Induction is dependent on protein synthesis and subject to catabolite repression. When extracts of induced and noninduced wild-type cells were assayed for succinate dehydrogenase, fumarase, and malate dehydrogenase only malate dehydrogenase increased in specific activity. Transport is inhibited by iodoacetamide, KCN, NaN3, and 2,4-dinitrophenol and shows pH and temperature optima of 6.2 and 30 °C. Kinetic parameters are: basal uptake (cells grown on glutamate) Km 11.6 μm, v 0.32 nmoles per min per mg dry cell mass; induced uptake (cells grown on succinate plus NH4Cl) Km 12.5 μm, v 5.78 nmoles per min per mg dry cell mass; induced transport (cells grown on nutrient broth plus succinate) Km 10 μm, V 0.98 nmoles per min per mg dry cell mass. It was not possible to determine the kinetic parameters of basal transport. Malate and fumarate were the only compounds exhibiting competitive inhibition of uptake and transport suggesting common transport system for all three compounds. The Ki values for competitive inhibition and the Km for succinate indicate the order of affinity for both uptake and transport are succinate > malate > fumarate. Data from kinetic parameters of uptake and transport and studies on succinate metabolism provide evidence consistent with concurrent increases in transport and metabolism to account for induced succinate uptake by P. putida.  相似文献   

5.
Malate transport in Schizosaccharomyces pombe.   总被引:2,自引:0,他引:2       下载免费PDF全文
The transport of malate was studied in a Schizosaccharomyces pombe wild-type strain and in mutant strains unable to utilize malic acid. Two groups of such mutants, i.e., malic enzyme-deficient and malate transport-defective mutants, were differentiated by a 14C-labeled L-malate transport assay and by starch gel electrophoresis followed by activity staining for malic enzyme (malate dehydrogenase [oxaloacetate decarboxylating] [NAD+]; 1.1.1.38) and malate dehydrogenase (1.1.1.37). Transport of malate in S. pombe was constitutive and strongly inhibited by inhibitors of oxidative phosphorylation and of the formulation of proton gradients. Transport was a saturable function of the malate concentration. The apparent Km and Vmax values for transport by the parent were 3.7 mM and 40 nmol/min per mg of protein, respectively, while those of the malic enzyme-deficient mutant were 5.7 mM and 33 nmol/min per mg of protein, respectively. Malate transport was pH and temperature dependent. The specificity of transport was studied with various substrates, including mono- and dicarboxylic acids, and the possibility of a common transport system for dicarboxylic acids is discussed.  相似文献   

6.
The uptake of hexoses by Chlorella vulgaris is accompanied by the uptake of protons. For 6-deoxyglucose a stoichiometry of one proton taken up per sugar molecule has been measured, whereas for 1-deoxyglucose approximately two protons are taken up per sugar molecule. It was found that in the presence of 1-deoxyglucose a considerable proportion of "carrier" catalyzes the transport of protons without the concomitant transport of sugar. Presumably, the binding of sugar initiates the translocation of the carrier-proton-sugar complex, but whereas 1-deoxyglucose can still dissociate from the complex at the external side of the cytoplasmic membrane, the translocation of the carrier-proton complex continues. This conclusion was reached since (a) the composition of the translocated carrier-proton-sugar complex is the same for both sugar. Its formation is a first order reaction with respect to protons. (b) When 6-deoxyglucose, present inside cells, is exchanged for external sugar, the exchange ratio is two to one when the external sugar is 1-deoxyglucose, two molecules of 6-deoxyglucose are lost for each molecule of 1-deoxyglucose entering. This result indicates that during uptake of 1-deoxyglucose statistically only each second carrier molecule appearing at the internal side of the cytoplasmic membrane is carrying sugar.  相似文献   

7.
Transport of dicarboxylic acids in castor bean mitochondria   总被引:1,自引:1,他引:0       下载免费PDF全文
Mitochondria from castor bean (Ricinus communis cv Hale) endosperm, purified on sucrose gradients, were used to investigate transport of dicarboxylic acids. The isolated mitochondria oxidized malate and succinate with respiratory control ratios greater than 2 and ADP/O ratios of 2.6 and 1.7, respectively. Net accumulation of 14C from [14C]malate or [14C]succinate into the mitochondrial matrix during substrate oxidation was examined by the silicone oil centrifugation technique. In the presence of ATP, there was an appreciable increase in the accumulation of 14C from [14C]malate or [14C]succinate accompanied by an increased oxidation rate of the respective dicarboxylate. The net accumulation of dicarboxylate in the presence of ATP was saturable with apparent Km values of 2 to 2.5 millimolar. The ATP-stimulated accumulation of dicarboxylate was unaffected by oligomycin but inhibited by uncouplers (2,4-dinitrophenol and carbonyl cyanide m-chlorophenylhydrazone) and inhibitors of the electron transport chain (antimycin A, KCN). Dicarboxylate accumulation was also inhibited by butylmalonate, benzylmalonate, phenylsuccinate, mersalyl and N-ethylmaleimide. The optimal ATP concentration for stimulation of dicarboxylate accumulation was 1 millimolar. CTP was as effective as ATP in stimulating dicarboxylate accumulation, and other nucleotide triphosphates showed intermediate or no effect on dicarboxylate accumulation. Dicarboxylate accumulation was phosphate dependent but, inasmuch as ATP did not increase phosphate uptake, the ATP stimulation of dicarboxylate accumulation was apparently not due to increased availability of exchangeable phosphate.

The maximum rate of succinate accumulation (14.5 nanomoles per minute per milligram protein) was only a fraction of the measured rate of oxidation (100-200 nanomoles per minute per milligram protein). Efflux of malate from the mitochondria was shown to occur at high rates (150 nanomoles per minute per milligram protein) when succinate was provided, suggesting dicarboxylate exchange. The uptake of [14C]succinate into malate or malonate preloaded mitochondria was therefore determined. In the absence of phosphate, uptake of [14C]succinate into mitochondria preloaded with malate was rapid (27 nanomoles per 15 seconds per milligram protein at 4°C) and inhibited by butylmalonate, benzylmalonate, and phenylsuccinate. Uptake of [14C]succinate into mitochondria preloaded with malonate showed saturation kinetics with an apparent Km of 2.5 millimolar and Vmax of 250 nanomoles per minute per milligram protein at 4°C. The measured rates of dicarboxylate-dicarboxylate exchange in castor bean mitochondria are sufficient to account for the observed rates of substrate oxidation.

  相似文献   

8.
The uptake of hexoses by Chlorella vulgaris is accompanied by the uptake of protons. For 6-deoxyglucose a stoichiometry of one proton taken up per sugar molecule has been measured, whereas for 1-deoxyglucose approximately two protons are taken up per sugar molecule.It was found that in the presence of 1-deoxyglucose a considerable proportion of “carrier” catalyzes the transport of protons without the concomitant transport of sugar. Presumably the binding of sugar initiates the translocation of the carrier-proton-sugar complex, but whereas 1-deoxyglucose can still dissociate from the complex at the external side of the cytoplasmic membrane, the translocation of the carrier-proton complex continues.This conclusion was reached since (a) the composition of the translocated carrier-proton-sugar complex is the same for both sugar. Its formation is a first order reaction with respect to protons.(b) When 6-deoxyglucose, present inside cells, is exchanged for external sugar, the exchange ratio is two to one when the external sugar is 1-deoxyglucose, two molecules of 6-deoxyglucose are lost for each molecule of 1-deoxyglucose entering. This result indicates that during uptake of 1-deoxyglucose statistically only each second carrier molecule appearing at the internal side of the cytoplasmic membrane is carrying sugar.  相似文献   

9.
Succinate, fumarate, and malate are valuable four-carbon (C4) dicarboxylic acids used for producing plastics and food additives. C4 dicarboxylic acid is biologically produced by heterotrophic organisms. However, current biological production requires organic carbon sources that compete with food uses. Herein, we report C4 dicarboxylic acid production from CO2 using metabolically engineered Synechocystis sp. PCC 6803. Overexpression of citH, encoding malate dehydrogenase (MDH), resulted in the enhanced production of succinate, fumarate, and malate. citH overexpression increased the reductive branch of the open cyanobacterial tricarboxylic acid (TCA) cycle flux. Furthermore, product stripping by medium exchanges increased the C4 dicarboxylic acid levels; product inhibition and acidification of the media were the limiting factors for succinate production. Our results demonstrate that MDH is a key regulator that activates the reductive branch of the open cyanobacterial TCA cycle. The study findings suggest that cyanobacteria can act as a biocatalyst for converting CO2 to carboxylic acids.  相似文献   

10.
The uptake of C4 dicarboxylates by cells from exponential cultures of Rhodopseudomonas spheroides followed saturation kinetics at concentrations below 100 muM with Km values for succinate, malate, and fumarate of 2.7, 2.3, and 0.8, respectively. Corresponding Vmax values of 50, 52, and 67.5 nmol/min per mg of protein at 20 C were obtained. Each of these compounds interfered competitively with uptake of the others, and a common transport system appears to be involved. Fructose-grown cells took up C4 dicarboxylates only at very low rates, and pyruvate-grown cells took up C4 dicarboxylates at one-third the rates found with succinate-grown cultures. Malonate and maleate inhibited uptake less severely, and aspartate and alpha-ketoglutarate had no effect at 100-fold excess. Divalent metals stimulated uptake. Light or respiration was required for uptake, and entered materials were rapidly converted to other metabolities, notably amino acids. Pyruvate entry appeared to be mediated by several systems, of which only one could be resolved kinetically. This system had a Km of 13 muM and Vmax of 5.6 nmol/min per mg of protein at 20 C. A number of related mono- and dicarboxylates interfered with pyruvate uptake. The pyruvate uptake system was distinguishable from the C4 dicarboxylate system by the absence of divalent cation stimulation and by substrate and inhibitor specificity.  相似文献   

11.
Coenzyme A (CoA) transport was studied in isolated rat heart mitochondria. Uptake of CoA was assayed by determining [3H]CoA associated with mitochondria under various conditions. Various oxidizable substrates including alpha-ketoglutarate, succinate, or malate stimulated CoA uptake. The membrane proton (delta pH) and electrical (delta psi) gradients, which dissipated with time in the absence of substrate, were maintained at their initial levels throughout the incubation in the presence of substrate. Addition of phosphate caused a concentration-dependent decrease of both delta pH and CoA uptake. Nigericin also dissipated the proton gradient and prevented CoA uptake. Valinomycin also prevented CoA uptake into mitochondria. Although the proton gradient was unaffected, the electrical gradient was completely abolished in the presence of valinomycin. Addition of 5 mM phosphate 10 min after the start of incubation prevented further uptake of CoA into mitochondria. A rapid dissipation of the proton gradient upon addition of phosphate was observed. Addition of nigericin or valinomycin 10 min after the start of incubation also resulted in no further uptake of CoA into with mitochondria; valinomycin caused an apparent efflux of CoA from mitochondria. Uptake was found to be sensitive to external pH displaying a pH optimum at pHext 8.0. Although nigericin significantly inhibited CoA uptake over the pHext range of 6.75-8, maximal transport was observed around pHext 8.0-8.25. Valinomycin, on the other hand, abolished transport over the entire pH range. The results suggest that mitochondrial CoA transport is determined by the membrane electrical gradient. The apparent dependence of CoA uptake on an intact membrane pH gradient is probably the result of modulation of CoA transport by matrix pH.  相似文献   

12.
The effect of bivalent cations on phosphate uptake by Saccharomyces cerevisiae was investigated. Phosphate uptake via the Na+-dependent transport system at pH 7.2 is stimulated by bivalent cations. The apparent affinity of phosphate for the transport mechanism is increased, but the apparent affinity for Na+ is decreased. Uptake of phosphate via the Na+-independent transport system is accompanied by a net proton influx of 2H+ and an efflux of 1 K+ for each phosphate ion taken up. At pH 4.5 phosphate uptake via the Na+-independent system is stimulated by bivalent cations, whereas at pH 7.2 uptake is inhibited. The effect of bivalent cations on phosphate uptake can be ascribed to a decrease in the surface potential.  相似文献   

13.
In Saccharomyces cerevisiae, L-malic acid transport is not carrier mediated and is limited to slow, simple diffusion of the undissociated acid. Expression in S. cerevisiae of the MAE1 gene, encoding Schizosaccharomyces pombe malate permease, markedly increased L-malic acid uptake in this yeast. In this strain, at pH 3.5 (encountered in industrial processes), L-malic acid uptake involves Mae1p-mediated transport of the monoanionic form of the acid (apparent kinetic parameters: Vmax = 8.7 nmol/mg/min; Km = 1.6 mM) and some simple diffusion of the undissociated L-malic acid (Kd = 0.057 min(-1)). As total L-malic acid transport involved only low levels of diffusion, the Mae1p permease was further characterized in the recombinant strain. L-Malic acid transport was reversible and accumulative and depended on both the transmembrane gradient of the monoanionic acid form and the DeltapH component of the proton motive force. Dicarboxylic acids with stearic occupation closely related to L-malic acid, such as maleic, oxaloacetic, malonic, succinic and fumaric acids, inhibited L-malic acid uptake, suggesting that these compounds use the same carrier. We found that increasing external pH directly inhibited malate uptake, resulting in a lower initial rate of uptake and a lower level of substrate accumulation. In S. pombe, proton movements, as shown by internal acidification, accompanied malate uptake, consistent with the proton/dicarboxylate mechanism previously proposed. Surprisingly, no proton fluxes were observed during Mae1p-mediated L-malic acid import in S. cerevisiae, and intracellular pH remained constant. This suggests that, in S. cerevisiae, either there is a proton counterflow or the Mae1p permease functions differently from a proton/dicarboxylate symport.  相似文献   

14.
The fermentation of fumarate and L-malate by Clostridium formicoaceticum was investigated. Growing and nongrowing cells degraded fumarate by dismutation to succinate, acetate, and CO2; on the other hand, only small amounts of succinate were detected when the organism was grown on L-malate. This dicarboxylic acid was mainly converted to acetate and CO2. The fermentation balances were modified if bicarbonate or formate were present in the medium. When C. formicoaceticum was grown in the presence of both dicarboxylic acids, fumarate was consumed before L-malate. The latter was mainly converted to acetate, whereas fumarate was fermented to acetate and succinate. Molar growth yields were determined to be 6 g of dry weight per mol of fumarate and 8 g of dry weight per mol of L-malate fermented.  相似文献   

15.
Characterization of Glutathione Uptake in Broad Bean Leaf Protoplasts   总被引:11,自引:2,他引:9       下载免费PDF全文
Transport of reduced glutathione (GSH) and oxidized glutathione (GSSG) was studied with broad bean (Vicia faba L.) leaf tissues and protoplasts. Protoplasts and leaf discs took up GSSG at a rate about twice the uptake rate of GSH. Detailed studies with protoplasts indicated that GSH and GSSG uptake exhibited the same sensitivity to the external pH and to various chemical reagents. GSH uptake was inhibited by GSSG and glutathione conjugates. GSSG uptake was inhibited by GSH and GS conjugates, and the uptake of metolachlor-GS was inhibited by GSSG. Various amino acids (L-glutamic acid, L-glutamine, L-cysteine, L-glycine, L-methionine) and peptides (glycine-glycine, glycine-glycine-glycine) affected neither the transport of GSH nor GSSG. Uptake kinetics indicate that GSH is taken up by a single saturable transporter, with an apparent Km of 0.4 mM, whereas GSSG uptake exhibits two saturable phases, with an apparent Km of 7 [mu]M and 3.7 mM. It is concluded that the plasma membrane of leaf cells contains a specific transport system for glutathione, which takes up GSSG and GS conjugates preferentially over GSH. Proton flux measurements and electrophysiological measurements indicate that GSH and GSSG are taken up with proton symport. However, a detailed analysis of these measurements suggests that the ion movements induced by GSSG differ from those induced by GSH.  相似文献   

16.
Derivatives of Escherichia coli C were engineered to produce primarily succinate or malate in mineral salts media using simple fermentations (anaerobic stirred batch with pH control) without the addition of plasmids or foreign genes. This was done by a combination of gene deletions (genetic engineering) and metabolic evolution with over 2,000 generations of growth-based selection. After deletion of the central anaerobic fermentation genes (ldhA, adhE, ackA), the pathway for malate and succinate production remained as the primary route for the regeneration of NAD+. Under anaerobic conditions, ATP production for growth was obligately coupled to malate dehydrogenase and fumarate reductase by the requirement for NADH oxidation. Selecting strains for improved growth co-selected increased production of these dicarboxylic acids. Additional deletions were introduced as further improvements (focA, pflB, poxB, mgsA). The best succinate biocatalysts, strains KJ060(ldhA, adhE, ackA, focA, pflB) and KJ073(ldhA, adhE, ackA, focA, pflB, mgsA, poxB), produce 622-733 mM of succinate with molar yields of 1.2-1.6 per mole of metabolized glucose. The best malate biocatalyst, strain KJ071(ldhA, adhE, ackA, focA, pflB, mgsA), produced 516 mM malate with molar yields of 1.4 per mole of glucose metabolized.  相似文献   

17.
Membrane vesicles isolated from Bacillus subtilis W23 catalyze active transport of the C4 dicarboxylic acids L-malate, fumarate, and succinate under aerobic conditions in the presence of the electron donor reduced beta-nicotinamide adenine dinucleotide or the non-physiological electron donor system ascorbate-phenazine methosulfate. The dicarboxylic acids are accumulated in unmodified form. Inhibitors of the respiratory chain, sulfhydryl reagents, and uncoupling agents inhibit the accumulation of the dicarboxylic acids. The affinity constants for transport of L-malate, fumarate, and succinate are 13.5, 7.5, and 4.3 muM, respectively; these values are severalfold lower than those reported previously for whole cells. Active transport of these dicarboxylic acids occurs via one highly specific transport system as is indicated by the following observations. (i) Each dicarboxylic acid inhibits the transport of the other two dicarboxylic acids competitively. (ii) The affinity constants determined for the inhibitory action are very similar to those determined for the transport process. (iii) Each dicarboxylic acid exchanges rapidly with a previously accumulated dicarboxylic acid. (iv) Other metabolically and structurally related compounds do not inhibit transport of these dicarboxylic acids significantly, except for L-aspartate and L-glutamate. However, transport of these dicarboxylic amino acids is mediated by independent system because membrane vesicles from B. subtilis 60346, lacking functional dicarboxylic amino acid transport activity, accumulate the C4 dicarboxylic acids at even higher rates than vesicles from B. subtilis W 23. (v) A constant ratio exists between the initial rates of transport of L-malate, fumarate, and succinate in all membrane vesicle preparations isolated from cells grown on various media. This high-affinity dicarboxylic acid transport system seems to be present constitutively in B. subtilis W23.  相似文献   

18.
Enzyme activities forming extracellular products from succinate, fumarate, and malate were examined using washed cell suspensions of Pseudomonas fluorescens from chemostat cultures. Membrane-associated enzyme activities (glucose, gluconate, and malate dehydrogenases), producing large accumulations of extracellular oxidation products in carbon-excess environments, have previously been found in P. fluorescens. Investigations carried out here have demonstrated the presence in this microorganism of a malic enzyme activity which produces extracellular pyruvate from malate in carbon-excess environments. Although the three membrane dehydrogenase enzymes decrease significantly in carbon-limited chemostat cultures, malic enzyme activity was found to increase fourfold under these conditions. The regulation of malate dehydrogenase and malic enzyme by malate or succinate was similar. Malate dehydrogenase increased and malic enzyme decreased in carbon-excess cultures. The opposite effect was observed in carbon-limited cultures. When pyruvate or glucose was used as the carbon source, malate dehydrogenase was regulated similarly by the available carbon concentration, but malic enzyme activity producing extracellular pyruvate was not detected. While large accumulations of extracellular oxalacetate and pyruvate were produced in malate-excess cultures, no extracellular oxidation products were detected in succinate-excess cultures. This may be explained by the lack of detectable activity for the conversion of added external succinate to extracellular fumarate and malate in cells from carbon-excess cultures. In cells from carbon-limited (malate or succinate) cultures, very active enzymes for the conversion of succinate to extracellular fumarate and malate were detected. Washed cell suspensions from these carbon-limited cultures rapidly oxidized added succinate to extracellular pyruvate through the sequential action of succinate dehydrogenase, fumarase, and malic enzyme. Succinate dehydrogenase and fumarase activities producing extracellular products were not detected in cells from chemostat cultures using pyruvate or glucose as the carbon source. Uptake activities for succinate, malate, and pyruvate also were found to increase in carbon-limited (malate or succinate) and decrease in carbon-excess cultures. The role of the membrane-associated enzymes forming different pathways for carbon dissimilation in both carbon-limited and carbon-excess environments is discussed.  相似文献   

19.
The location of the catalytic site of the membrane-bound respiratory fumarate reductase of Escherichia coli was investigated using mutants and inhibitors of dicarboxylic acid transport. Comparison of apparent Km and Vmax values for fumarate in intact cells and in inverted membrane vesicles showed that externally added fumarate was required to be transported across the cytoplasmic membrane prior to reduction. The catalytic site of fumarate reductase must therefore be located on the cytoplasmic face of the membrane.  相似文献   

20.
Succinate:ubiquinone reductase was shown to catalyze the oxidation of L- and D-stereoisomers of malate by artificial electron acceptors and ubiquinone. The rate of malate oxidation by succinate:ubiquinone reductase is by two orders of magnitude lower than that for the natural substrate--succinate. The values of kinetic constants for the oxidation of D- and L-stereoisomers of malate are equal to: V infinity = 0.1 mumol/min/mg protein, Km = 2 mM and V infinity = 0.05 mumol/min/mg protein, Km = 2 mM, respectively. The malate dehydrogenase activity is fully inhibited by the inhibitors of the dicarboxylate-binding site of the enzyme, i.e., N-ethylmaleimide and malonate and is practically insensitive to carboxin, a specific inhibitor of the ubiquinone-binding center. The enol form of oxaloacetate was shown to be the product of malate oxidation by succinate:ubiquinone reductase. The kinetics of inhibition of the enzyme activity by the ketone and enol forms of oxaloacetate was studied. Both forms of oxaloacetate effectively inhibit the succinate:ubiquinone reductase reaction.  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