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.
Malonate exerted a stronger inhibitory effect on brefeldin A production than on mycelia growth in cultures ofCurvularia lunata especially at inhibitory levels of 100 to 200 mm. The extent of 200 mm malonate inhibition of growth and brefeldin A production was greater in cultures treated with malonate prior to inoculation than those treated following 5 days after inoculation. Maleate at levels of 40 to 220 mm activated brefeldin A formation in cultures though exerting variable effects on mycelia growth.  相似文献   

2.
A sterile glucose-mineral salts broth was fortified with equimolar concentrations (10-3 M) of various organic acids and intermediates in the tricarboxylic acid cycle. Appropriate media were neutralized with 2 N NaOH, inoculated with spore suspensions or mycelial pellets ofPenicillium rubrum and incubated quiescently for 14 days or with shaking for 5 days. Rubratoxins were recovered from culture filtrates by ether extraction and resolved by thin-layer chromatography. Toxin formation in quiescent cultures was enhanced by malonate but was not markedly affected by ethyl malonate, shikimate, and acetate or by isocitrate or oxaloacetate added in the presence of malonate. Citrate, cis-aconitate, -ketoglutarate, succinate, fumarate, and malonate when present in the medium alone or in conjunction with malonate caused a 15 to 50% reduction in rubratoxin formation. Acetyl-CoA (10-5 M/flask) caused an 80% increase in toxin yield. Rubratoxin formation in shake cultures was not affected by succinate and malonate. All other combinations of intermediates and malonate caused a 10 to 50% reduction in toxin formation. At 10–3 M, citrate enhanced rubratoxin B formation and stimulated rubratoxin A production by as much as 100%. Above 10–3 M, citrate inhibited toxin production. Incorporation of [2-14C]acetate into rubratoxin was enhanced by malonate, fumarate, and malonate. A combination of pyruvate and malonate produced a 40% increase in [2-14C]acetate incorporation into rubratoxin. The highest reduction of labeled acetate incorporation (36%) was caused by succinate or -ketoglutarate combined with malonate.  相似文献   

3.
The rates of the oxidized (Eox) and reduced (Ered) (by NAD . H through the ubiquinone pool) succinate dehydrogenase inhibition by N-ethyl-maleimide are equal and obey pseudo-first order kinetics. The protection of the enzyme against irreversible alkylation was used to quantitate the dissociation constants for Eox and Ered complexes with fumarate, succinate and malonate under conditions when no intramolecular redox reactions might occur. the membrane-bound succinate dehydrogenase catalyzes the succinate : phenazine-methosulphate reductase reaction in the presence of thenoyltrifluoroacetone by a Slater-Bonner mechanism. A comparison of the constants measured by the protection with those derived from the steady-state kinetics shows that succinate affinity for Eox is about 10 times higher than that for Ered; the reverse relations were found for fumarate, whereas the affinity for malonate only slightly depends on the redox state of the enzyme. The data obtained suggest that the dicarboxylate binding at the active site induces changes in the enzyme redox potential. The surface charge does not contribute significantly to the energy of the dicarboxylate binding to the active site of the membrane-bound enzyme.  相似文献   

4.
1. Metabolism of propionate by sheep-liver mitochondria was stimulated catalytically by alpha-oxoglutarate, pyruvate, citrate and isocitrate. Succinate was stimulatory at higher concentrations, but fumarate and malate were inert. These effects were all independent of the presence of ATP, succinate being less effective when ATP was present. 2. Compared with the metabolism of added succinate, propionate metabolism was resistant to malonate inhibition, but only in the presence of added ATP. In the absence of ATP propionate metabolism was more sensitive to malonate inhibition than was the metabolism of succinate. 3. In the absence of malonate, and at malonate concentrations in the range 5-100mm, alpha-oxoglutarate increased the rate of fixation of [2-(14)C]propionate by about 50% without altering the nature of the fixation products. 4. Metabolism of [1-(14)C]-propionate in the presence of 50mm-malonate was accompanied by accumulation of about half the propionate consumed as succinate. When alpha-oxoglutarate was present in addition part of the alpha-oxoglutarate was metabolized and the rate of propionate consumption was increased. The total succinate that accumulated corresponded to the alpha-oxoglutarate consumed plus about half the propionate metabolized. 5. When [1-(14)C]propionate was metabolized in the absence of malonate about 70% of the generated succinate was oxidized to fumarate or beyond. The addition of malonate decreased the rate of propionate metabolism, and decreased to about half the fraction of generated succinate oxidized. 6. When propionate and 10mm-succinate were metabolized together, the total oxidation of succinate was greater than that with 10mm-succinate alone. The increment in succinate oxidation corresponded to about half the propionate metabolized in the presence or absence of malonate or ATP. 7. It is suggested that the metabolism of propionate is specifically limited by the rate of oxidation of the generated succinate, and that the succinate oxidase concerned is distinct from that responsible for the oxidation of added succinate. 8. The results are discussed in terms of the mode of action of certain stimulants and inhibitors of propionate metabolism. It is suggested that many of these act by stimulation or inhibition of the specific succinate oxidase that limits propionate metabolism.  相似文献   

5.
Nocardia salmonicolor, grown on acetate, commercial D,L-lactate or hydrocarbon substrates, has high isocitrate lyase activities compared with those resulting from growth on other carbon sources. This presumably reflects the anaplerotic role of the glyoxylate cycle during growth on the former substrates. Amongst a variety of compounds tested, including glucose, pyruvate and tricarboxylic acid cycle intermediates, only succinate and fumarate prevented an increase in enzyme activity in the presence of acetate. When acetate (equimolar to the initial sugar concentration) was added to cultures growing on glucose, there followed de novo synthesis of isocitrated lyase and isocitrate dehydrogenase, with increases in growth rate and glucose utilization, and both acetate and glucose were metabolized simultaneously. A minute amount of acetate (40 muM) caused isocitrate lyase synthesis (a three-fold increase in activity within 3 min of addition) when added to glucose-limited continuous cultures, but even large amounts added to nitrogen-limited batch cultures were ineffective. Malonate, at a concentration that was not totally growth-inhibitory (1mM) prevented the inhibition of acetate-stimulated isocitrate lyase synthesis by succinate, but fumarate still inhibited in the presence of malonate. Phosphoenolpyruvate is a non-competitive inhibitor of the enzyme (apparent Ki 1-7 mM). The results are consistent with the induction of isocitrate or a closely related metabolite, and catabolite repression by a C-4 acid of the tricarboxylic acid cycle, possibly fumarate.  相似文献   

6.
In order to improve the production of succinate and malate by the filamentous fungus Aspergillus niger the activity of the glyoxylate bypass pathway was increased by over-expression of the isocitrate lyase (icl) gene. The hypothesis was that when isocitrate lyase was up-regulated the flux towards glyoxylate would increase, leading to excess formation of malate and succinate compared to the wild-type. However, metabolic network analysis showed that an increased icl expression did not result in an increased glyoxylate bypass flux. The analysis did show a global response with respect to gene expression, leading to an increased flux through the oxidative part of the TCA cycle. Instead of an increased production of succinate and malate, a major increase in fumarate production was observed.The effect of malonate, a competitive inhibitor of succinate dehydrogenase (SDH), on the physiological behaviour of the cells was investigated. Inhibition of SDH was expected to lead to succinate production, but this was not observed. There was an increase in citrate and oxalate production in the wild-type strain. Furthermore, in the strain with over-expression of icl the organic acid production shifted from fumarate towards malate production when malonate was added to the cultivation medium.Overall, the icl over-expression and malonate addition had a significant impact on metabolism and on organic acid production profiles. Although the expected succinate and malate formation was not observed, a distinct and interesting production of fumarate and malate was found.  相似文献   

7.
Fumarate reductase activity of Streptococcus faecalis   总被引:1,自引:0,他引:1       下载免费PDF全文
Some characteristics of a fumarate reductase from Streptococcus faecalis are described. The enzyme had a pH optimum of 7.4; optimal activity was observed when the ionic strength of the phosphate buffer was adjusted to 0.088. The K(m) value of the enzyme for reduced flavin mononucleotide was 2 x 10(-4)m as determined with a 26-fold preparation. In addition to fumarate, the enzyme reduced maleate and mesaconate. No succinate dehydrogenase activity was detected, but succinate did act as an inhibitor of the fumarate reductase activity. Other inhibitors were malonate, citraconate, and trans-, trans-muconate. Metal-chelating agents did not inhibit the enzyme. A limited inhibition by sulfhydryl-binding agents was observed, and the preparations were sensitive to air oxidation and storage. Glycine, alanine, histidine, and possibly lysine stimulated fumarate reductase activity in the cell-free extracts. However, growth in media supplemented with glycine did not enhance fumarate reductase activity. The enzymatic activity appears to be constitutive.  相似文献   

8.
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   

9.
Two kinetically distinct states of succinate oxidase have been detected in the mitochondria of Neruospora crassa. One state has a K(m) for succinate of 4.1 x 10(-3)m, and the other has a K(m) for succinate of 3.5 x 10(-4)m. The high K(m) state was found in freshly extracted mitochondria from either 20- or 72-hr mycelium. However, the succinate oxidase activity in mitochondria from 20-hr mycelium rapidly deteriorated in vitro, leaving a stable residual activity with the lower K(m) for succinate. Adenosine triphosphate (ATP) plus Mg(2+) stabilized the high K(m) state in these preparations. The high K(m) state of succinate oxidase was further characterized by a two- to threefold increase in activity over the pH range 6.6 to 8.0 and by classical competitive inhibition by fumarate and malonate. By contrast, the low K(m) state of succinate oxidase showed a relatively flat response to pH over the range 6.6 to 8.0 and a nonclassical pattern of inhibition by fumarate and malonate, as shown by nonlinear plots of reciprocal velocity versus reciprocal substrate concentration in the presence of inhibitor or reciprocal velocity versus inhibitor concentration at fixed substrate concentrations. The relationship of mycelial age to the in vitro stability of succinate oxidase is considered with reference to probable changes in the relative pool sizes of extra- and intramitochondrial ATP in response to changes in the rate of glycolysis.  相似文献   

10.
Stemphyloxin II, a new phytotoxic compound isolated from liquid cultures of Stemphylium botryosum f. sp. lycopersici has been identified as a tricyclic compound derived from stemphyloxin I. Production of stemphyloxins I and II increases by 4-6 fold in the presence of succinate, fumarate or malonate. The secretion of stemphyloxins is iron-regulated and both compounds act as ferric chelates. The phytotoxicity of stemphyloxin I is approximately 100-fold higher than that of stemphyloxin II. The possible role of stemphyloxins in iron acquisition by S. botryosum is discussed.  相似文献   

11.
Oxidation of succinate to fumarate is an energetically difficult step in the biochemical pathway of propionate oxidation by syntrophic methanogenic cultures. Therefore, the effect of fumarate on propionate oxidation by two different propionate-oxidizing cultures was investigated. When the methanogens in a newly enriched propionate-oxidizing methanogenic culture were inhibited by bromoethanesulfonate, fumarate could act as an apparent terminal electron acceptor in propionate oxidation. 13C-nuclear magnetic resonance experiments showed that propionate was carboxylated to succinate while fumarate was partly oxidized to acetate and partly reduced to succinate. Fumarate alone was fermented to succinate and CO2. Bacteria growing on fumarate were enriched and obtained free of methanogens. Propionate was metabolized by these bacteria when either fumarate or Methanospirillum hungatii was added. In cocultures with Syntrophobacter wolinii, such effects were not observed upon addition of fumarate. Possible slow growth of S. wolinii on fumarate could not be demonstrated because of the presence of a Desulfovibrio strain which grew rapidly on fumarate in both the absence and presence of sulfate.  相似文献   

12.
The protective effect of dicarboxylates on the active-site-directed inhibition of the membrane-bound succinate dehydrogenase by N-ethylmaleimide, steady-state kinetics methods for Ki and K, determinations, and equillibrium studies were employed to quantitate the relative the relative affinities of succinate, fumarate, malonate and oxaloacetate to the reduced and oxidixed species of the enzyme. A more than 10-fold difference in the relative affinities of the reduced and oxidized succinate dehydrogenase to succinate, fumarate and oxaloacetate is found, whereas the reactivity of the active-site sulphydryl group does not depend on the redox state of the enzyme. The redox-state-dependent changes in the affinity of the membrane-bound succinate dehydrogenase to oxaloacetate can be quantitatively accounted for by a 10-fold increase in the rate of dissociation of the enzyme-inhibitor complex whch occurs upon reduction of the enzyme. The data obtained give no support for either the existence of a sulphydryl group other than the active-site one important for the catalysis or for the presence of a separate dicarboxylate-specific regulatory site in the succinate dehydrogenase molecule.  相似文献   

13.
Two recombinant plasmid Escherichia coli strains containing amplified fumarate reductase activity converted fumarate to succinate at significantly higher rates and yields than a wild-type E. coli strain. Glucose was required for the conversion of fumarate to succinate, and in the absence of glucose or in cultures with a low cell density, malate accumulated. Two-dimensional gel electrophoretic analysis of proteins from the recombinant DNA and wild-type strains showed that increased quantities of both large and small fumarate reductase subunits were expressed in the recombinant DNA strains.  相似文献   

14.
Two recombinant plasmid Escherichia coli strains containing amplified fumarate reductase activity converted fumarate to succinate at significantly higher rates and yields than a wild-type E. coli strain. Glucose was required for the conversion of fumarate to succinate, and in the absence of glucose or in cultures with a low cell density, malate accumulated. Two-dimensional gel electrophoretic analysis of proteins from the recombinant DNA and wild-type strains showed that increased quantities of both large and small fumarate reductase subunits were expressed in the recombinant DNA strains.  相似文献   

15.
Inducible binding proteins for C4-dicarboxylic acids (DBP) and glucose (GBP) were isolated from Pseudomonas aeruginosa by extraction of exponential-phase cells with 0.2 M MgC12 (pH 8.5) and by an osmotic shock procedure without affecting cell viability. DBP synthesis was induced by growth on aspartate, alpha-ketoglutarate, succinate, fumarate, malate, and malonate but not by growth on acetate, citrate, pyruvate, or glucose. Binding of succinate by DBP was competitively inhibited by 10-fold concentrations of fumarate and malate but not by a variety of related substances. GBP synthesis and transport of methyl alpha-glucoside by whole cells were induced by growth on glucose or pyruvate plus galactose, 2-deoxyglucose, or methyl alpha-glucoside but not by growth on gluconate, succinate, acetate, or pyruvate. The binding of radioactive glucose by GBP was significantly inhibited by 10-fold concentrations of glucose, galactose, and glucose-1-phosphate but not by the other carbohydrates tested. The binding of glucose by GBP or succinate by DBP did not result in any chemical alteration of the substrates.  相似文献   

16.
Malonate was studied for its effect on succinate dehydrogenase activity and conidiation. It Was found to stimulate the succinate dehydrogenase activity and also conidiation of Neurospora crassa. The efficiency of sucrose metabolization for cellular synthesis was improved in malonate supplemented cultures. High Concentration (0.5 M) had a distinct toxic effect on conidiation and economic efficiency. Teratological structures were observed at 0.5 M malonate concentration. A stimulation of the glyoxylate cycle is considered to be one of the factors responsible for the conidiogenic effect.  相似文献   

17.
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   

18.
Molar growth yields were determined from chemostat cultures of Hydrogenomonas eutropha on succinate and on fumarate. The yields from culture on succinate were about 12 g higher than on fumarate. Assuming this difference to be equivalent to 1 molecule of adenosine triphosphate, it is concluded that the oxidation by oxygen of the Hydrogenomonas cytochrome b yields 1 molecule of adenosine triphosphate.  相似文献   

19.
The endogenous metabolites of the coelomic fluid of the earthworm Eisenia veneta were characterised using high-resolution one-dimensional and two-dimensional 1H nuclear magnetic resonance spectroscopy. Signals from common organic acids, such as acetate, fumarate, malonate, malate, formate, and succinate, were identified together with adenosine and nicotinamide mononucleotide. The potential use of this information as a baseline dataset for future toxicological or physiological studies was demonstrated by a metabonomic analysis: a series of earthworms were dosed with the model compound 3-fluoro-4-nitrophenol, and toxic effects followed by multivariate analysis of the spectral data of the coelomic fluid. Relative concentrations of acetate and malonate were decreased in the dosed worms compared to the controls.  相似文献   

20.
Respiratory complex II oxidizes succinate to fumarate as part of the Krebs cycle and reduces ubiquinone in the electron transport chain. Previous experimental evidence suggested that complex II is not a significant contributor to the production of reactive oxygen species (ROS) in isolated mitochondria or intact cells unless mutated. However, we find that when complex I and complex III are inhibited and succinate concentration is low, complex II in rat skeletal muscle mitochondria can generate superoxide or H(2)O(2) at high rates. These rates approach or exceed the maximum rates achieved by complex I or complex III. Complex II generates these ROS in both the forward reaction, with electrons supplied by succinate, and the reverse reaction, with electrons supplied from the reduced ubiquinone pool. ROS production in the reverse reaction is prevented by inhibition of complex II at either the ubiquinone-binding site (by atpenin A5) or the flavin (by malonate), whereas ROS production in the forward reaction is prevented by malonate but not by atpenin A5, showing that the ROS from complex II arises only from the flavin site (site II(F)). We propose a mechanism for ROS production by complex II that relies upon the occupancy of the substrate oxidation site and the reduction state of the enzyme. We suggest that complex II may be an important contributor to physiological and pathological ROS production.  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