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 203 毫秒
1.
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   

2.
3.
Malate oxidation supported C2H2 reduction by bacteroids isolated from Sesbania rostrata stem nodules. Optimal activity reached 7.5 nanomoles per minute per milligram of dry weight and was in the same order of magnitude as that observed with succinate but always required a lower O2 tension. Malate dehydrogenase (EC 1.1.1.37), purified 66-fold from bacteroids, actively oxidized malate (Km = 0.19 millimolar). Malic enzyme (EC 1.1.1.39) from Sesbania bacteroids had a lower affinity for malate (Km = 2.32 millimolar). Both enzymes exclusively required NAD+ as cofactor and required an alkaline pH for optimal activity. 2-Oxoglutarate and oxalate, inhibiting malate dehydrogenase and malic enzyme, respectively, were used to specifically block each malate oxidation pathway in bacteroids. The predominance of malate dehydrogenase activity to support bacteroid N2 fixation was demonstrated. The inhibition of O2 consumption by 2-oxoglutarate confirmed the importance of the malate dehydrogenase pathway in malate oxidation. It is proposed that the utilization of malate, with regard to O2, is important in a general strategy of this legume to maintain N2 fixation under O2 limited conditions.  相似文献   

4.
Pea leaf mitochondria showed complex kinetics for malate metabolism. O2 uptake increased as malate concentration increased from 0 to 10 mm, reached a plateau between 10 and 20 mm malate, and then increased again up to 40 mm malate. Analysis of the products of malate oxidation by high-performance liquid chromatography revealed that the first phase of O2 uptake coincided with the synthesis of both pyruvate and oxalacetate (OAA) while the second phase of O2 uptake at higher malate levels usually occurred with a large increase in OAA formation. The biphasic response in O2 uptake and the changing ratios of pyruvate and OAA synthesis did not appear to be the direct result of the differing Km values of malate dehydrogenase and malic enzyme. Rather, they resulted from thermodynamic properties of these two malate oxidases and the kinetics of the two NADH dehydrogenases found in plant mitochondria. At low malate concentrations the rotenone-sensitive NADH dehydrogenase was active and could accept electrons from both malate oxidases. This NADH dehydrogenase became saturated at about 10 mm malate. At higher malate concentrations the rotenone-insensitive NADH dehydrogenase was increasingly important and its increased electron transport capacity was best exploited by malate dehydrogenase. At the higher malate concentrations an increasing portion of the electrons from malate reduce O2 through the alternative oxidase. Although this coincided with the second phase of malate-dependent O2 uptake it was not required for this phase to be seen.  相似文献   

5.
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   

6.
The glyceollin inhibition of electron transport by isolated soybean and corn mitochondria was similar to that of rotenone, acting at site I between the internal NADH dehydrogenase and coenzyme Q. Coupled state 3 malate oxidation was inhibited by glyceollin and rotenone with apparent Ki values of about 15 and 5 micromolar, respectively. Carbonylcyanide m-chlorophenyl hydrazone uncoupled state 4 malate oxidation was also inhibited by glyceollin and rotenone, but uncoupled succinate and exogenous NADH state 4 oxidation was only slightly inhibited by both compounds. Glyceollin also inhibited ferricyanide reduction with malate as the electron donor, with an apparent Ki of 5.4 micromolar, but failed to inhibit such reduction with succinate or externally added NADH as electron donors. Glyceollin did not inhibit state 4 oxidation of malate, succinate, or exogenous NADH. Glyceollin did not act as a classical uncoupler or as an inhibitor of oxidative phosphorylation.  相似文献   

7.
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   

8.
Mitochondria isolated from pea (Pisum sativum L.) leaves are able to transport the keto acid, oxaloacetate, from the reaction medium into he mitochondrial matrix at high rates. The rate of uptake by the mitochondria was measured as the rate of disappearance of oxaloacetate from the reaction medium as it was reduced by matrix malate dehydrogenase using NADH provided by glycine oxidation. The oxaloacetate transporter was identifed as being distinct from the dicarboxylate and the α-ketoglutarate transporters because of its inhibitor sensitivities and its inability to interact with other potential substrates. Phthalonate and phthalate were competitive inhibitors of oxaloacetate transport with Ki values of 60 micromolar and 2 millimolar, respectively. Butylmalonate, an inhibitor of the dicarboxylate and α-ketoglutarate transporters, did not alter the rate of oxaloacetate transport. In addition, a 1000-fold excess of malate, malonate, succinate, α-ketoglutarate, or phosphate had little effect on the rate of oxaloacetate transport. The Km for the oxaloacetate transporter was about 15 micromolar with a maximum velocity of over 500 nanomoles per milligram mitochondrial protein/min at 25°C. No requirement for a counter ion to move against oxaloacetate was detected and the highest rates of uptake occurred at alkaline pH values. An equivalent transporter has not been reported in animal mitochondria.  相似文献   

9.
Dicarboxylate transport in maize mesophyll chloroplasts   总被引:7,自引:0,他引:7  
Evidence is presented for high rates of carrier-mediated dicarboxylate anion transport in maize mesophyll chloroplasts. Radioactively labeled malate is transported across the chloroplast envelope leading to accumulation in the stroma. Malate in the stroma will exchange for external malate, oxaloacetate, glutamate, aspartate, and oxoglutarate. At 4 °C the V of malate uptake is 50 μmol·h?1·mg Chl?1 and the Km for malate is 0.5 mm. Oxaloacetate competitively inhibits malate uptake with a Ki estimated to be 0.3 mm. The temperature dependence of malate uptake indicates an activation energy of 12 kcal/mol, and extrapolation using this value gives a rate of transport at 30 °C of approximately 300 μmol·h?1·mg Chl?1. This rate approximates the rates of photosynthetic malate production by these chloroplasts.  相似文献   

10.
The characteristics of malate transport into aerobically grown cells of the purple photosynthetic bacterium Rhodobacter capsulatus were determined. A single transport system was distinguished kinetically which displayed a Kt value of 2.9 ± 1.2 μM and Vmax of 43 ± 6 nmol · min-1 · mg-1 protein. Competition experiments indicated that the metabolically related C4-dicarboxylates succinate and fumarate are also transported by this system. Malate uptake was sensitive to osmotic shock and evidence from the binding of radiolabelled malate and succinate to periplasmic protein fractions indicated that transport is mediated by a dicarboxylate binding protein. The activity of the transport system was studied as a function of external and internal pH and it was found that a marked activation of uptake occurred at intracellular pH values greater than 7. The use of a high affinity binding protein dependent system to transport a major carbon and energy source suggests that Rhodobacter capsulatus would be capable of obtaining growth sustaining quantities of C4-dicarboxylates even if these were present at very low concentrations in the environment.  相似文献   

11.
Cations were generally ineffective in stimulating succinate transport in a succinate dehydrogenase mutant of Bacillus subtilis unless accompanied by polyvalent anions; phosphate and sulfate being particularly active. The Km values for the phosphate or sulfate requirement were approx. 3 mM.Biphasic kinetics were characteristic of both the succinate (Km values 0.1 and 1 mM), and inorganic phosphate (Km values 0.1 and 3 mM) transport system(s). The phosphate transport system(s) was repressed by high inorganic phosphate and a coordinate increase in the transport of phosphate, arsenate, and phosphate-stimulated succinate transport accompanied growth in low phosphate media.A class of arsenate resistant mutants were simultaneously defective in the transport of arsenate, phosphate and succinate when cells were repressed for phosphate transport, however, the transport of these ions was regained in these mutants when grown in low phosphate media. Organic phosphate esters did not stimulate succinate transport in arsenate resistant mutants but were effective after growth in low phosphate media. Growth under phosphate limitation permitted the simultaneous regain of both phosphate and sulfate dependent succinate transport activities whereas sulfate limitation alone was ineffective.Succinate was not transported by an anion exchange diffusion mechanism since phosphate efflux was low or absent during succinate transport.The transport of C4-dicarboxylates in B. subtilis is strongly stimulated by intracellular polyvalent anions. The absence of an anion permeability mechanism precludes succinate transport but partial escape from this restriction is mediated by the derepression of a phosphate transport system.  相似文献   

12.
In vivo and in vitro studies on asparagine biosynthesis in soybean seedlings   总被引:15,自引:0,他引:15  
The biosynthesis of asparagine in plants was investigated by feeding radioactive metabolites to soybean cotyledons and by extracting an asparagine synthetase from the same tissue. Soybean cotyledon slices were supplied with radioactive succinate, malate, or aspartate in the presence or absence of various unlabeled metabolites for periods of up to 80 min. Neither aspartate nor malate was rapidly converted to asparagine; labeled aspartate was converted largely to malate. Labeled succinate was rapidly converted to asparagine, and several lines of evidence suggested that fumarate, malate, and aspartate are intermediates. The results suggest that asparagine biosynthesis in plant cells is compartmentalized beginning with succinate. Although results were also consistent with asparagine formation via aspartate, metabolism of a mixture of [14C] plus [3H]succinate resulted in a lower 14C3H ratio in asparagine than aspartate, suggesting that some asparagine may be formed via another pathway. Demonstration of a glutamine-linked asparagine synthetase in soybean cotyledons supports the idea that asparagine is formed via aspartate. The enzyme requires aspartate (Km = 2.2 mm), glutamine (Km = 0.12 mm), ATP (Km = 0.066 mm), magnesium ion, and sulfhydryl protection. It has a pH optimum of 7.7 and is not located in mitochrondria. A small amount of asparagine was formed when ammonium ion was substituted for glutamine, but the Km of the enzyme for ammonium ion was about 25-fold greater than the Km for glutamine suggesting that glutamine is the physiologically important substrate. Soybean cotyledons actively convert [14C]-cyanide to asparagine, apparently via β-cyanoalanine. However, malate was also rapidly labeled from [14C]cyanide and this result cannot be explained by known metabolic pathways.  相似文献   

13.
Experiments are reported in which the uptake of 86Rb+, used as an analog of K+, into cultured cells of Arabidopsis thaliana is investigated. A single transport system is found with Km = 0.34 millimolar and Vmax = 14 nmoles per milligram of protein per hour. This system is blocked by the metabolic inhibitor carbonyl cyanide m-chlorophenyl hydrazone (CCCP) and by cold. At high concentrations of external K+ (above 1 millimolar), a significant fraction of total uptake is energy-independent. No evidence is found for more than one energy-dependent uptake system or for concentration-dependent modifications of a carrier as postulated in multiphasic transport models.  相似文献   

14.
Fructose transport was not apparently affected in a number of Pseudomonas putida strains with deranged activity of a common glucose-gluconate uptake system, indicating the existence of an independent fructose uptake system. Fructose uptake by glucose-gluconate uptake mutants was induced by fructose and obeyed saturation kinetics (apparent K m =0.3 mM). The fructose uptake system serves to transport glucose in addition to fructose. The entry of fructose into P. putida cells appears to be mediated also by the glucose-gluconate uptake system, as shown by the ability to accumulate fructose of wild type cells grown on glucose, a substrate that induces the glucose-gluconate uptake system but not the fructose uptake system. In addition, fructose was found to be an inducer of the glucose-gluconate uptake system. The physiological significance of these observations is not clear because the fructose uptake system can provide the cell with a high enough internal concentration of fructose to support maximum growth rate on this hexose, as shown by following the growth course of glucose-gluconate uptake mutants on fructose.  相似文献   

15.
《Anaerobe》2000,6(3):187-196
To investigate electron transport in the dissimilatory iron-reducing isolate Geobacter metallireducens strain GS-15, assays for redox enzymes and characterizations of cytochromes were performed. G. metallireducens produced 1.56 g dry cell weight per mol etransferred when grown on benzoate and contained the following citric acid cycle enzymes (activities in nkat per mg cell protein); isocitrate dehydrogenase (0.84), coenzyme A-dependent 2-oxoglutarate: methyl viologen oxidoreductase (2.80), succinate dehydrogenase (0.80), and malate dehydrogenase (8.35). An oxygen-sensitive, soluble coenzyme A-dependent 2-oxoglutarate: ferredoxin oxidoreductase (0.14) with no NAD(P)-activity was observed. In cell suspensions NADPH, but not NADH, could reduce methyl viologen (2.45). Isocitrate and malate dehydrogenase activities were soluble enzymes that coupled with NADP and NAD, respectively. NADPH (0.94) and NADH (1.85) oxidation activities were observed in detergent solubilized, whole-cell suspensions using the artificial electron acceptor menadione. Menaquinone was observed at 1.2 μmol per g cell protein. The triheme c7cytochrome was purified and 37 amino acids were determined. The mass observed by mass spectroscopy was 9684±10 Da. The average mid-point potential for the three hemes was measured at −91 mV. The growth yield, redox reactions, and electron transfer components are discussed with regards to possible sites of energy conservation during growth on iron(III).  相似文献   

16.
《BBA》1986,850(1):64-71
NAD+ supplied to purified Solanum tuberosum mitochondria caused progressive inhibition of succinate oxidation in State 3. This inhibition was especially pronounced at alkaline pH and at low succinate concentrations. Glutamate counteracted the inhibition. NAD+ promoted oxaloacetate accumulation in State 3; supplied oxaloacetate inhibited O2 uptake in the presence of succinate much more severely in State 3 than in State 4. NAD reduction linked to succinate oxidation by ATP-dependent reverse electron transport was likewise inhibited by oxaloacetate. We conclude that NAD+-induced inhibition of succinate oxidation is due to an inhibition of succinate dehydrogenase resulting from increased accumulation of oxaloacetate generated from malate oxidation via malate dehydrogenase. The results are discussed in the context of the known regulatory characteristics of plant succinate dehydrogenase.  相似文献   

17.
The localization of a series of enzymes involved in the anaerobic photodissimilation of acetate in Chlamydomonas reinhardtii F-60 adapted to a hydrogen metabolism was determined through the enzymic analyses of the chloroplastic, cytoplasmic, and mitochondrial fractions obtained with a cellular fractionation procedure that incorporated cell wall removal by treatment with autolysine, digestion of the plasmalemma with the detergent digitonin, and fractionation by differential centrifugation on a Percoll step gradient. The sequence of events leading to the photoevolution of H2 from acetate includes the conversion of acetate into succinate via the extraplastidic glyoxylate cycle, the oxidation of succinate to fumarate by chloroplastic succinate dehydrogenase, and the oxidation of malate to oxaloacetate in the chloroplast by NAD dependent malate dehydrogenase. The level of potential activity for the enzymes assayed were sufficient to accommodate the observed rate of the photoanaerobic dissimilation of acetate and the photoevolution of H2.  相似文献   

18.
Aspartate or glutamate stimulated the rate of light-dependent malate decarboxylation by isolated Zea mays bundle sheath chloroplasts. Stimulation involved a decrease in the apparent Km (malate) and an increased maximum velocity of decarboxylation. In the presence of glutamate other dicarboxylates (succinate, fumarate) competitively inhibited malate decarboxylation by intact chloroplasts with respect to malate with an apparent Ki of about 6 millimolar. For comparison the Ki for inhibition of nicotinamide adenine dinucleotide phosphate-malic enzyme from freshly lysed chloroplasts by these dicarboxylates was 15 millimolar. A range of compounds structurally related to aspartate stimulated malate decarboxylation by intact chloroplasts. Ka values for stimulation at 5 millimolar malate were 1.7, 5, and 10 millimolar for l-glutamate, l-aspartate, and β-methyl-dl-aspartate, respectively. Certain compounds, notably cysteic acid, which stimulated malate decarboxylation by intact chloroplasts inhibited malate decarboxylation by nicotinamide adenine dinucleotide phosphate-malic enzyme obtained from lysed chloroplasts and assayed under comparable conditions. It was concluded that aspartate, glutamate, and related compounds affect the transport of malate into the intact chloroplasts and that malate translocation does not take place on the general dicarboxylate translocator previously reported for higher plant chloroplasts.  相似文献   

19.
In comparative study of respiratory metabolism, it was established that the relative proportions of respiratory end-products (succinic, acetic and lactic acids) differed consistently in two strains of Hymenolepis diminuta (Toronto and ANU). The ANU strain produced more lactic acid and less succinic acid under aerobic and anaerobic conditions. In the shift from aerobic to anaerobic conditions both strains compensated by increasing their outputs of succinic acid. The ANU strain possessed significantly higher activities of hexokinase, pyruvate kinase, lactate dehydrogenase, cytosolic and mitochondrial malic enzyme and cytosolic α-glycerophosphate dehy drogenase. The Toronto strain had significantly higher activities of fumarase, succinate dehydrogenase, and fumarate reductase. There were no significant differences in the activities of phosphoenolpyruvate carboxykinase and malic dehydrogenase between strains. The fumarase activity in the Toronto strain was 16 times that of the ANU strain, its Km (malate) was 0.8mM, as opposed to 2.5 mM, and it was less sensitive to inhibition by NAD or ATP. These observations are consistent with the patterns of end-product formation in the two strains. Ratios of end-products and calculations of approximate redox balance suggest that the Toronto strain may have a greater capacity for aerobic metabolism.  相似文献   

20.
Day DA  Hanson JB 《Plant physiology》1977,59(2):139-144
A study was made to determine conditions under which malate oxidation rates in corn (Zea mays L.) mitochondria are limited by transport processes. In the absence of added ADP, inorganic phosphate increased malate oxidation rates by processes inhibited by mersalyl and oligomycin, but phosphate did not stimulate uncoupled respiration. However, the uncoupled oxidation rates were inhibited by butylmalonate and mersalyl. When uncoupler was added prior to substrate, subsequent O2 uptake rates were reduced when malate and succinate, but not exogenous NADH, were used. Uncoupler and butylmalonate also inhibited swelling in malate solutions and malate accumulation by these mitochondria, which were found to have a high endogenous phosphate content. Addition of uncoupler after malate or succinate produced an initial rapid oxidation which declined as the mitochondria lost solute and contracted. This decline was not affected by addition of ADP or AMP, and was not observed when exogenous NADH was substrate. Increasing K+ permeability with valinomycin increased the P-trifluoromethoxy (carboxylcyanide)phenyl hydrazone inhibition. Kinetic studies showed the slow rate of malate oxidation in the presence of uncoupler to be characterized by a high Km and a low Vmax, probably reflecting a diffusion-limited process.  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