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 46 毫秒
1.
Anaerobically, Escherichia coli cannot grow using either glycerol or citrate as sole carbon and energy source. However, it has been reported that a mixture of glycerol and citrate will support growth. We have found that wild-type strains of E. coli K-12 do not grow on glycerol plus citrate anaerobically. However, growth eventually occurs due to the frequent appearance of mutants. We found that such Cit+ mutants were defective in anaerobic respiration with nitrate or trimethylamine-N-oxide and were chlorate resistant (i.e. molybdenum cofactor deficient). Conversely, well characterized mutants in any of chlA, B, D, E, G and N were also able to use citrate anaerobically. No anaerobic growth differences between wild type and chl mutants were observed either with fermentable sugars or with glycerol plus fumarate or glycerol plus tartrate. Citrate lyase was induced anaerobically by citrate and repressed by glucose in both wild type strains and chl mutants. Furthermore, levels of citrate lyase, fumarate reductase, malate dehydrogenase, fumarase and alcohol dehydrogenase were similar in both types of strains under anaerobic conditions. It is conceivable that a functioning molybdenum cofactor prevents use of citrate by keeping citrate lyase in the inactive form.  相似文献   

2.
Citrate metabolism by Enterococcus faecium ET C9 and Enterococcus durans Ov 421 was studied as sole energy source and in presence of glucose or lactose. Both strains utilized citrate as the sole energy source. Enterococcus faecium ET C9 showed diauxic growth in the presence of a limiting concentration of glucose. Neither strain used citrate until glucose was fully metabolized. The strains showed co-metabolism of citrate and lactose. Lactate, acetate, formate, and flavour compounds (diacetyl, acetoin, and 2,3-butanediol) were detected in both strains. The highest production of flavour compounds was detected during growth of E. durans Ov 421 in media supplemented with citrate-glucose and citrate-lactose. Citrate lyase was inducible in both strains. Acetate kinase activities presented the highest values in LAPTc medium, with E. faecium ET C9 displaying a specific activity 2.4-fold higher than E. durans. The highest levels of alpha-acetolactate synthase specific activity were detected in E. durans grown in LAPTc+g, in accordance with the maximum production of flavour compounds detected in this medium. Diacetyl and acetoinreductases displayed lower specific activity values in the presence of citrate. Enterococcus faecium and E. durans displayed citrate lyase, acetate kinase, alpha-acetolactate synthase, and diacetyl and acetoin reductase activities. These enzymes are necessary for conversion of citrate to flavour compounds that are important in fermented dairy products.  相似文献   

3.
Information on the factors influencing citrate metabolism in lactobacilli is limited and could be useful in understanding the growth of lactobacilli in ripening cheese. Citrate was not used as an energy source by either Lactobacillus casei ATCC 393 or Lact. plantarum 1919 and did not affect the growth rate when co-metabolized with glucose or galactose. In growing cells, metabolism of citrate was minimal at pH 6 but significant at pH 4·5 and was greater in cells co-metabolizing galactose than in those co-metabolizing glucose or lactose. In non-growing cells, optimum utilization of citrate also occurred at pH 4·5 and was not increased substantially by the presence of fermentable sugars. In both growing and non-growing cells, acetate and acetoin were the major products of citrate metabolism; pyruvate was also produced by non-growing cells and was transformed to acetoin once the citrate was exhausted. Citrate was metabolized more rapidly than sugar by non-growing cells; the reverse was true of growing cells. Citrate metabolism by Lact. plantarum 1919 and Lact. casei ATCC 393 increased six- and 22-fold, respectively, when the cells were pre-grown on galactose plus citrate than when pre-grown on galactose only. This was probably due to induction of citrate lyase by growth on citrate plus sugar. These results imply that lactobacilli, if present in large enough numbers, can metabolize citrate in ripening cheese in the absence of an energy source.  相似文献   

4.
Summary The formation of acetate, formate and succinate was studied in Lactobacillus pentosus. These compounds were produced in addition to lactic acid when cells were exposed to anaerobic growth conditions with limited carbohydrates and in the presence of citrate. Citrate was metabolised via oxalacetate serving as an H-acceptor in a joint process together with lactate. The metabolism of citrate resulted in stoichiometric amounts of succinate and acetate. Lactate was degraded to formate and acetate in a reaction catalysed by pyruvate formate lyase. These fermentation products can potentially affect the flavour of fermented food but ecological factors in fermenting meat, e.g. the presence of glucose, nitrate or nitrite prevent this reaction. Offprint requests to: G. Wolf  相似文献   

5.
Citrate metabolism in anaerobic bacteria   总被引:4,自引:0,他引:4  
Abstract The regulation of anaerobic citrate metabolism is very diverse among different groups of bacteria. In organisms like Streptococcus lactis and Clostridium sporosphaeroides which lack citrate synthase, the activity of its antagonistic enzyme, citrate lyase, need not be regulated. Many anaerobes like Rhodocyclus gelatinosus and Clostridium sphenoides are able to synthesize their own l -glutamate and contain citrate synthase. In these bacteria the activity of citrate metabolizing enzymes which are involved in a cascade system are under strict control. In Rc. gelatinosus activation/inactivation of citrate lyase is controlled by acetylation/deacetylation which is catalyzed by its corresponding regulatory enzymes, citrate lyase ligase and citrate lyase deacetylase. In C. sphenoides inactivation of citrate lyase is accomplished by deacetylation as well as by changing in the enzyme conformation. Activation of citrate lyase is catalyzed by citrate lyase ligase whose activity in addition is modulated by phosphorylation/dephosphorylation. Further, electron transport process also seems to play a role in the inactivation of citrate metabolizing enzymes in enteric bacteria.  相似文献   

6.
Anaerobic growth of Klebsiella aerogenes NCDO 711 (NCTC 418) on citrate was dependent on the presence of Na+ in the medium, and fermentation of citrate was mediated via the fermentation pathway enzymes, citrate lyase and a Na+-dependent oxalacetate decarboxylase. This confirms the previous findings on strain NCTC 418. Growth under aerobic conditions was independent of Na+. The mean generation time for cells grown aerobically on either Na+ or K+ citrate medium was about 60 min, with a molar growth yield of about 40 g (dry weight) of cells per mol of citrate utilized. Citrate was apparently metabolized aerobically in both the Na+ and K+ citrate cells via the citric acid cycle, since cell extracts contained alpha-ketoglutarate dehydrogenase but not the citrate fermentation enzymes. The presence of theother enzymes of the citric acid cycle in K. aerogenes was shown in earlier studies. Under aerated conditions (no detectable oxygen tension in the culture), growth was faster on the Na+ citrate medium (mean generation time, 85 min) than on the K+ citrate medium (mean generation time, 120 min). Both cultures grew slower than under aerobic conditions, presumably because of oxygen limitation. Despite the faster growth rate, the molar growth yield of the aerated Na+ citrate culture was one-half that observed for the aerated K+ citrate culture. Citrate was metabolized via the citric acid cycle in cells grown in the K+ citrate medium under aerated conditions since alpha-ketoglutarate dehydrogenase, but not the fermentation enzymes, was detected in extracts prepared from these cells. Metabolism of citrate in the Na+ citrate medium under aerated conditions occurred via both the fermentation pathway (approximately 75 percent) and the citric acid cycle (about 25 percent), as evidenced by (i) the presence of the fermentation enzymes and alpha-ketoglutarate dehydrogenase in extracts of cells grown under these conditions, (ii) a molar growth yield which was intermediate between that obtained for anaerobic and aerated K+ citrate cultures, and (iii) the excretion of acetate, which also occurred in anaerobic cultures but not in aerated K+ citrate or aerobic cultures.  相似文献   

7.
Citrate lyase ligase was shown to be present in Clostridium sphenoides actively degrading citrate. In contrast to citrate lyase ligase from C. sporosphaeroides and Streptococcus lactis, the enzyme from C. sphenoides was under stringent regulatory control. The alteration of the kinetic properties of the enzyme after depletion of citrate suggested the presence of two different enzyme species in different phases of growth: active and partially active citrate lyase ligase. These enzymes were purified from in vivo 32P-labeled C. sphenoides cells, which were grown on low-phosphate medium containing 40 mM citrate and 1 mCi [32]orthophosphate. During enzyme purification only the active form of citrate lyase ligase was shown to be radioactively labeled. Growth experiments with 14C-labeled precursors of purines and pyrimidines and subsequent purification of active citrate lyase ligase indicated that the 32P labeling of the enzyme was not due to the incorporation of a nucleotide. Inactivation of the ligase after its treatment with acid phosphatase also suggested that the active form of the enzyme is phosphorylated. Citrate lyase ligase, therefore, is the first known enzyme in an anaerobic bacterium whose activity is modulated by phosphorylation/dephosphorylation.  相似文献   

8.
Citrate transport in Klebsiella pneumoniae   总被引:5,自引:0,他引:5  
Sodium ions were specifically required for citrate degradation by suspensions of K. pneumoniae cells which had been grown anaerobically on citrate. The rate of citrate degradation was considerably lower than the activities of the citrate fermentation enzymes citrate lyase and oxaloacetate decarboxylase, indicating that citrate transport is rate limiting. Uptake of citrate into cells was also Na+ -dependent and was accompanied by its rapid metabolism so that the tricarboxylic acid was not accumulated in the cells to significant levels. The transport could be stimulated less efficiently by LiCl. Li+ ions were cotransported with citrate into the cells. Transport and degradation of citrate were abolished with the uncoupler [4-(trifluoromethoxy)phenylhydrazono]propanedinitrile (CCFP). After releasing outer membrane components and periplasmic binding proteins by cold osmotic shock treatment, citrate degradation became also sensitive towards monensin and valinomycin. The shock procedure had no effect on the rate of citrate degradation indicating that the transport is not dependent on a binding protein. Citrate degradation and transport were independent of Na+ ions in K. pneumoniae grown aerobically on citrate and in E. coli grown anaerobically on citrate plus glucose. An E. coli cit+ clone obtained by transformation of K. pneumoniae genes coding for citrate transport required Na specifically for aerobic growth on citrate indicating that the Na-dependent citrate transport system is operating. Na+ and Li+ were equally effective in stimulating citrate degradation by cell suspensions of E. coli cit+. Citrate transport in membrane vesicles of E. coli cit+ was also Na+ dependent and was energized by the proton motive force (delta micro H+). Dissipation of delta micro H+ or its components delta pH or delta psi by ionophores either totally abolished or greatly inhibited citrate uptake. It is suggested that the systems energizing citrate transport under anaerobic conditions are provided by the outwardly directed cotransport of metabolic endproducts with protons yielding delta pH and by the decarboxylation of oxaloacetate yielding delta pNa+ and delta psi. In citrate-fermenting K. pneumoniae an ATPase which is activated by Na+ was not found. The cells contain however a proton translocating ATPase and a Na+/H+ antiporter in their membrane.  相似文献   

9.
Hu Y  Holden JF 《Journal of bacteriology》2006,188(12):4350-4355
The hyperthermophilic archaeon Pyrobaculum islandicum uses the citric acid cycle in the oxidative and reductive directions for heterotrophic and autotrophic growth, respectively, but the control of carbon flow is poorly understood. P. islandicum was grown at 95 degrees C autotrophically, heterotrophically, and mixotrophically with acetate, H2, and small amounts of yeast extract and with thiosulfate as the terminal electron acceptor. The autotrophic growth rates and maximum concentrations of cells were significantly lower than those in other media. The growth rates on H2 and 0.001% yeast extract with and without 0.05% acetate were the same, but the maximum concentration of cells was fourfold higher with acetate. There was no growth with acetate if 0.001% yeast extract was not present, and addition of H2 to acetate-containing medium greatly increased the growth rates and maximum concentrations of cells. P. islandicum cultures assimilated 14C-labeled acetate in the presence of H2 and yeast extract with an efficiency of 55%. The activities of 11 of 19 enzymes involved in the central metabolism of P. islandicum were regulated under the three different growth conditions. Pyruvate synthase and acetate:coenzyme A (CoA) ligase (ADP-forming) activities were detected only in heterotrophically grown cultures. Citrate synthase activity decreased in autotrophic and acetate-containing cultures compared to the activity in heterotrophic cultures. Acetylated citrate lyase, acetate:CoA ligase (AMP forming), and phosphoenolpyruvate carboxylase activities increased in autotrophic and acetate-containing cultures. Citrate lyase activity was higher than ATP citrate synthase activity in autotrophic cultures. These data suggest that citrate lyase and AMP-forming acetate:CoA ligase, but not ATP citrate synthase, work opposite citrate synthase to control the direction of carbon flow in the citric acid cycle.  相似文献   

10.
Growth of Enterobacter cloacae on K+ citrate under aerated conditions (no detectable oxygen tension in the medium even though it was aerated) was slower (mean generation time, 130 min) than under aerobic conditions (mean generation time, 72 min), but with a faster utilization of citrate, resulting in a molar growth yield of 10.6 g (dry weight) of cells per mol of citrate utilized versus 40 g (dry weight) of cells per mol of citrate utilized for aerobic growth. The rapid utilization of citrate under aerated conditions was apparently due to the induction of citrate lyase and was supported by the finding that cells excreted acetate and a small amount of oxalacetate under aerated conditions, but not under aerobic conditions when the cells were devoid of citrate lyase activity. The activity of oxalacetate decarboxylase in aerated cells was slightly lower than in aerobic cells, indicating that little of the oxalacetate produced by the citrate lyase was metabolized by the decarboxylase. Oxalacetate was probably metabolized by malate dehydrogenase, previously shown to be present in anaerobic and aerobic cells. Thus, about 70% of the citrate was cleaved by the citrate lyase, resulting in little or no production of energy for growth. The remaining citrate was metabolized via the citric acid cycle under aerated conditions, since the cells contained alpha-ketoglutarate dehydrogenase at the same level as in aerobically grown cells. The presence of the other enzymes of the cycle was shown in earlier studies.  相似文献   

11.
Regulatory citrate lyase mutants of Salmonella typhimurium   总被引:2,自引:1,他引:1  
Citrate lyase, the key enzyme of anaerobic citrate catabolism, could not be deleted from Salmonella typhimurium. The only class of mutants found had a mode of covalent regulation that strongly resembled the Escherichia coli system: citrate lyase was only active, i.e., acetylated, when a cosubstrate was present.  相似文献   

12.
1. The enzymes citrate lyase and isocitrate lyase catalyse similar reactions in the cleavage of citrate to acetate plus oxaloacetate and of isocitrate to succinate plus glyoxylate, respectively. 2. Nevertheless, the mechanism of action of each enzyme appears to be different from each other. Citrate lyase is an acyl carrier protein-containing enzyme complex whereas isocitrate lyase is not. The active form of citrate lyase is an acetyl-S-enzyme but that of isocitrate lyase is not a corresponding succinyl-S-enzyme. 3. In contrast to citrate lyase, the isocitrate enzyme is not inhibited by hydroxylamine nor does it acquire label if treated with appropriately labelled radioactive substrate. 4. Isotopic exchange experiments performed in H18-2O with isocitrate as a substrate produced no labelling in the product succinate. This was shown by mass-spectrometric analysis. 5. The conclusion drawn from these results is that no activation of succinate takes place on the enzyme through transient formation of succinic anhydride or a covalently-linked succinyl-enzyme, derived from this anhydride.  相似文献   

13.
Rat liver ATP citrate lyase was inactivated by 2, 3-butanedione and phenylglyoxal. Phenylglyoxal caused the most rapid and complete inactivation of enzyme activity in 4-(2-hydroxyethyl)-1-piperazine-ethanesulphonic acid buffer, pH 8. Inactivation by both butanedione and phenylglyoxal was concentration-dependent and followed pseudo- first-order kinetics. Phenylglyoxal also decreased autophosphorylation (catalytic phosphate) of ATP citrate lyase. Inactivation by phenylglyoxal and butanedione was due to the modification of enzyme arginine residues: the modified enzyme failed to bind to CoA-agarose. The V declined as a function of inactivation, but the Km values were unaltered. The substrates, CoASH and CoASH plus citrate, protected the enzyme significantly against inactivation, but ATP provided little protection. Inactivation with excess reagent modified about eight arginine residues per monomer of enzyme. Citrate, CoASH and ATP protected two to three arginine residues from modification by phenylglyoxal. Analysis of the data by statistical methods suggested that the inactivation was due to modification of one essential arginine residue per monomer of lyase, which was modified 1.5 times more rapidly than were the other arginine residues. Our results suggest that this essential arginine residue is at the CoASH binding site.  相似文献   

14.
Citrate Metabolism in Aerobacter cloacae   总被引:3,自引:3,他引:0       下载免费PDF全文
Growth of Aerobacter cloacae on citrate either anaerobically or aerobically did not require and was not stimulated by the presence of Na(+) in the medium. Citrate was metabolized anaerobically via the fermentation pathway as evidenced by the (i) presence of oxalacetate decarboxylase, (ii) induction of citrate lyase, and (iii) repression of alpha-ketoglutarate dehydrogenase under anaerobic conditions. Thus, although all the other enzymes of the citric acid cycle were present in anaerobic cells, this pathway was not available for the metabolism of citrate. Citrate was metabolized aerobically via the citric acid cycle, since (i) citrate lyase but not oxalacetate decarboxylase was repressed and (ii) alpha-ketoglutarate dehydrogenase was induced under these conditions. The presence of Na(+) in the medium did not lead to a repression of alpha-ketoglutarate dehydrogenase as in the case of Aerobacter aerogenes. The oxalacetate decarboxylase was a soluble, constitutive enzyme, not activated by Na(+) nor inhibited by avidin. It was slightly inhibited by ethylenediaminetetraacetate but was not stimulated by Mg(2+) or Mn(2+). Thus, this enzyme differed markedly in its properties from the same enzyme found in citrate-grown A. aerogenes.  相似文献   

15.
1. Citrate inhibits the activities of phosphofructokinase from muscles and nervous tissues from different animals across the Animal Kingdom except for the insects. The enzymes from the flight muscle of nine different insects and the cerebral ganglion of the locust were investigated: no inhibition by citrate was observed. Inhibition was observed with the enzymes from both aerobic (e.g. pectoral muscle of pigeon) and anaerobic (e.g. fish muscle, pectoral muscle of the game birds) muscles. It is suggested that this inhibition is of physiological importance in decreasing the rate of glucose utilization in skeletal muscle of animals during starvation and/or prolonged exercise. 2. The rates of glucose utilization by the sartorius and gastrocnemius muscles of the frog were markedly decreased by ketone bodies. The latter elevated the glucose 6-phosphate and citrate contents of the gastrocnemius muscle, indicating that citrate inhibition of phosphofructokinase could be, in part, responsible for the decreased rate of glycolysis. 3. These findings provide evidence that the concept of the glucose-fatty acid-ketone-body cycle involves both aerobic and anaerobic skeletal muscle and nervous tissue from a wide range of animals except the insects. In the latter the concept of the cycle may not be applicable.  相似文献   

16.
The metabolically versatile purple bacterium Rhodobacter sphaeroides 2.4.3 is a denitrifier whose genome contains two periplasmic nitrate reductase-encoding gene clusters. This work demonstrates nonredundant physiological roles for these two enzymes. One cluster is expressed aerobically and repressed under low oxygen while the second is maximally expressed under low oxygen. Insertional inactivation of the aerobically expressed nitrate reductase eliminated aerobic nitrate reduction, but cells of this strain could still respire nitrate anaerobically. In contrast, when the anaerobic nitrate reductase was absent, aerobic nitrate reduction was detectable, but anaerobic nitrate reduction was impaired. The aerobic nitrate reductase was expressed but not utilized in liquid culture but was utilized during growth on solid medium. Growth on a variety of carbon sources, with the exception of malate, the most oxidized substrate used, resulted in nitrite production on solid medium. This is consistent with a role for the aerobic nitrate reductase in redox homeostasis. These results show that one of the nitrate reductases is specific for respiration and denitrification while the other likely plays a role in redox homeostasis during aerobic growth.  相似文献   

17.
Anaerobic l-lactate degradation by Lactobacillus plantarum   总被引:5,自引:0,他引:5  
Abstract Lactobacillus plantarum strains used as silage inoculants were investigated for their ability to metabolize lactic acid anaerobically after prolonged incubation (7–30 days) when glucose was absent from the medium. When citrate was present in the medium together with glucose during the initial fermentation, the lactic acid produced was degraded. Citrate was concomitantly degraded, resulting in accumulation of formic, acetic and succinic acids along with CO2. The anaerobic degradation was confirmed by the use of l 14C(U) labelled lactate. The existence of pyruvate formate lyase in L. plantarum was indicated by using 14C-labelled pyruvate and HPLC identification of end-products. The 1-14C-carboxylic acid group of pyruvate was converted to formic acid, and the 3-14C was found in acetic acid. The key enzyme(s) in this metabolic pathway appears to require anaerobic conditions and induction by citrate.  相似文献   

18.
The genome annotations of all sequenced Dehalococcoides strains lack a citrate synthase, although physiological experiments have indicated that such an activity should be encoded. We here report that a Re face-specific citrate synthase is synthesized by Dehalococcoides strain CBDB1 and that this function is encoded by the gene cbdbA1708 (NCBI accession number CAI83711), previously annotated as encoding homocitrate synthase. Gene cbdbA1708 was heterologously expressed in Escherichia coli, and the recombinant enzyme was purified. The enzyme catalyzed the condensation of oxaloacetate and acetyl coenzyme A (acetyl-CoA) to citrate. The protein did not have homocitrate synthase activity and was inhibited by citrate, and Mn2+ was needed for full activity. The stereospecificity of the heterologously expressed citrate synthase was determined by electrospray ionization liquid chromatography-mass spectrometry (ESI LC/MS). Citrate was synthesized from [2-(13)C]acetyl-CoA and oxaloacetate by the Dehalococcoides recombinant citrate synthase and then converted to acetate and malate by commercial citrate lyase plus malate dehydrogenase. The formation of unlabeled acetate and 13C-labeled malate proved the Re face-specific activity of the enzyme. Shotgun proteome analyses of cell extracts of strain CBDB1 demonstrated that cbdbA1708 is expressed in strain CBDB1.  相似文献   

19.
Citrate specimens derived from chiral acetates were converted to the CoA derivatives. These were reconverted with citrate synthase to citrates under conditions of either predominating hydrolytic burst or predominating steady-state period. The stereochemical purity of substrates and products was determined. Reversal of the synthase condensation step occurs under both conditions but is markedly increased during the steady-state period. The results indicate that citryl-CoA-derived acetyl-CoA and oxaloacetate create the steady-state conditions. The hydrolase state and the ligase/lyase state of the synthase predominate under burst and steady-state conditions, respectively. This result indicates a conversion of the hydrolase state into the ligase/lyase state during the transition.  相似文献   

20.
The biosynthesis of the 2'-(5"-phosphoribosyl)-3'-dephospho-coenzyme A (CoA) prosthetic group of citrate lyase (EC 4.1.3.6), a key enzyme of citrate fermentation, proceeds via the initial formation of the precursor 2'-(5"-triphosphoribosyl)-3'-dephospho-CoA and subsequent transfer to apo-citrate lyase with removal of pyrophosphate. In Escherichia coli, the two steps are catalyzed by CitG and CitX, respectively, and the corresponding genes are part of the citrate lyase gene cluster, citCDEFXG. In the homologous citCDEFG operon of Klebsiella pneumoniae, citX is missing. A search for K. pneumoniae citX led to the identification of a second genome region involved in citrate fermentation which comprised the citWX genes and the divergent citYZ genes. The citX gene was confirmed to encode holo-citrate lyase synthase, whereas citW was shown to encode a citrate carrier, the third one identified in this species. The citYZ genes were found to encode a two-component system consisting of the sensor kinase CitY and the response regulator CitZ. Remarkably, both proteins showed >or=40% sequence identity to the citrate-sensing CitA-CitB two-component system, which is essential for the induction of the citrate fermentation genes in K. pneumoniae. A citZ insertion mutant was able to grow anaerobically with citrate, indicating that CitZ is not essential for expression of citrate fermentation genes. CitX synthesis was induced to a basal level under anaerobic conditions, independent of citrate, CitB, and CitZ, and to maximal levels during anaerobic growth with citrate as the sole carbon source. Similar to the other citrate fermentation enzymes, CitX synthesis was apparently subject to catabolite repression.  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