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 31 毫秒
1.
The production of poly--hydroxyalkanoate (PHA) in Syntrophomonas wolfei grown in pure culture or in coculture with Methanospirillum hungatei was studied. PHA was produced by S. wolfei during the exponential phase of growth under both of these cultural conditions. S. wolfei in pure culture also produced PHA in stationary phase when the medium was supplemented with high concentrations of the substrate crotonate. In S. wolfei, PHA levels decreased after growth stopped and most of the substrate was depleted. Altering the C to N ratio of the medium did not affect the amount of PHA made per mg of protein. The incorporation of labeled butyrate but not acetate into PHA during the early stages of growth of S. wolfei and the fact that some of the PHA in a pure culture of S. wolfei grown with trans-2-pentenoate was a polymer of 5-carbon monomer units indicated that a pathway exists for the synthesis of PHA without degradation of the substrate to acetyl-CoA. During the later stages of growth, PHA was made by a pathway which was in equilibrium with the acetate pool. These data indicate that PHA served as a carbon/energy reserve material in S. wolfei, but the synthesis of this polymer was regulated in a manner different from the known pathways.  相似文献   

2.
A Synthrophomonas wolfei-Methanospirillum hungatei coculture was adapted to catabolize crotonate. S. wolfei was then isolated in axenic culture using agar spread plates and roll tubes with crotonate as the sole energy source. S. wolfei catabolized crotonate via a disproportionation mechanism similar to that of some Clostridium species. Growth on crotonate was very slow (specific growth rate of 0.029 h–1) but the conversion of energy into cell material was very efficient with cell yields of 14.6 g (dry wt.) per mol of crotonate. S. wolfei alone did not catabolize butyrate, but butyrate was stoichiometrically degraded to acetate and presumably methane when S. wolfei was reassociated with M. hungatei. S. wolfei-M. hungatei cocultures accumulated some butyrate during growth on crotonate indicating that protons were not the sole electron acceptors used for crotonate oxidation by the coculture.  相似文献   

3.
Cell-free extracts of Syntrophomonas wolfei subsp. wolfei synthesized d-(-)-3-hydroxybutyryl-coenzyme A (CoA) (the stereoisomer required for the synthesis of poly--hydroxyalkanoate) from acetoacetyl-CoA, but not crotonyl-CoA, and NAD(P)H. Ammonium sulfate fractionation and ion exchange chromatography separated an acetoacetyl-CoA reductase activity that formed d-(-)-3-hydroxybutyryl-CoA from the -oxidation enzyme activity, l-(+)-3-hydroxyacyl-CoA dehydrogenase. The former activity was further purified by hydroxylapatite and affinity chromatography. The most pure acetoacetyl-CoA reductase preparations formed d-(-)-3-hydroxybutyryl-CoA from acetoacetyl-CoA and had high specific activities using either NADH or NADPH as the electron donor. Thus, S. wolfei makes d-(-)-3-hydroxybutyryl-CoA by an acetoacetyl-CoA reductase rather than by a d-isomer specific enoyl-CoA hydratase and the reducing equivalents required for PHA synthesis from acetoacetyl-CoA can be supplied from the NADH made during -oxidation.  相似文献   

4.
The membrane phospholipid fatty acids (PLFAs) from several cocultures and a pure culture of Syntrophomonas wolfei were determined by capillary column gas chromatography. Cocultures of S. wolfei with a Desulfovibrio sp. contained PLFAs from both organisms, whereas PLFAs from a coculture with Methanospirillum hungatei contained very little biomass to analyze. The pure culture of S. wolfei grown on crotonate provided the best material for analysis of the PLFAs. The predominant PLFAs of S. wolfei were the monounsaturated 16:1ω7c and 16:1ω9c and the saturated 16:0 and 14:0. A low concentration of the diunsaturated 18:2ω6 was detected. The PLFA analysis provides additional information for consideration in the determination of the profile of PLFAs obtained from anaerobic environments. In addition, this information may aid in the understanding of the physiology and phylogeny of S. wolfei and other syntrophic bacteria.  相似文献   

5.
A strictly anaerobic, benzoate-degrading bacterium,Syntrophus buswellii strain GA, was able to degrade benzoate or 3-phenylpropionate to acetate, CO2 and H2 if the hydrogen partial pressure was sufficiently low. The hydrogen was removed in syntrophic coculture byMethanospirillum hungatei or byDesulfovibrio sp. through interspecies hydrogen transfer or in pure culture by the use of crotonate as reducible cosubstrate. Alternatively,S. buswellii strain GA could grow in pure culture with crotonate. Activities of seven catabolic enzymes were measured in crude cell extracts ofS. buswellii strain GA grown with various substrates and of crotonate-grownS. buswellii strain DSM 2612A. Benzoate, 3-phenylpropionate and crotonate were activated by CoA ligases. Glutaryl-CoA dehydrogenase was found to be involved in the degradation of aromatic compounds and enzymes catalysing -oxidation were involved in the reaction sequence from crotonyl-CoA to acetate. Ac-type cytochrome was present in the cytoplasm, whereasb-type cytochromes were associated with the membranes of bothS. buswelli strains grown on crotonate. These indicated the presence of an electron-transport system. A high growth yield of crotonate-grownS. buswellii strain GA might be explained by electron-transport phosphorylation in addition to substrate-level phosphorylation.  相似文献   

6.
Syntrophomonas wolfei and Syntrophus buswellii were grown with butyrate or benzoate in a defined binary coculture with Methanospirillum hungatei. Both strains also grew independent of the partner bacteria with crotonate as substrate. Localization of enzymes involved in butyrate oxidation by S. wolfei revealed that ATP synthase, hydrogenase, and butyryl-CoA dehydrogenase were at least partially membrane-associated whereas 3-hydroxybutyryl-CoA dehydrogenase and crotonase were entirely cytoplasmic. Inhibition experiments with copper chloride indicated that hydrogenase faced the outer surface of the cytoplasmic membrane. Suspensions of butyrate-or benzoate-grown cells of either strain accumulated hydrogen during oxidation of butyrate or benzoate to a low concentration that was thermodynamically in equilibrium with calculated reaction energetics. The protonophore carbonylcyanide m-chlorophenyl-hydrazone (CCCP) and the proton-translocating ATPase inhibitor N,Ndicyclohexylcarbodiimide (DCCD) both specifically inhibited hydrogen formation from butyrate or benzoate at low concentrations, whereas hydrogen formation from crotonate was not affected. A menaquinone was extracted from cells of S. wolfei and S. buswellii grown syntrophically in a binary methanogenic culture. The results indicate that a proton-potential-driven process is involved in hydrogen release from butyrate or benzoate oxidation.Abbreviations BES Bromoethanesulfonate - CCCP Carbonyl cyanide-m-chlorophenyl-hydrazone - DCCD N,Ndicyclohexylcarbodiimide - DCPIP Dichlorophenol indophenol - PMS Phenazine methosulfate  相似文献   

7.
The effects of organic acid anions on the growth of Syntrophomonas wolfei was determined by varying the initial concentration of the acid anion in the medium. The addition of 15 mM acetate decreased the growth rate of a butyrate-catabolizing coculture containing Methanospirillum hungatei from 0.0085 to 0.0029 per hour. Higher initial acetate concentrations decreased the butyrate degradation rate and the yield of cells of S. wolfei per butyrate degraded. Inhibition was not due to the counter ion or the effect of acetate on the methanogen. Initial acetate concentrations above 25 mM inhibited crotonate-using pure cultures and cocultures of S. wolfei. Benzoate and lactate inhibited the growth of S. wolfei on crotonate in pure culture and coculture. Lactate was an effective inhibitor of S. wolfei cultures at concentrations greater than 10 mM. High concentrations of acetate and lactate altered the electron flow in crotonate-catabolizing cocultures, resulting in the formation of less methane and more butyrate and caproate. The inclusion of the acetate-using methanogen, Methanosarcina barkeri, in a methanogenic butyrate-catabolizing coculture increased both the yield of S. wolfei cells per butyrate degraded and the efficacy of butyrate degradation. Butyrate degradation by acetate-inhibited cocultures occurred only after the addition of Methanosarcina barkeri. These results showed that the metabolism of S. wolfei was inhibited by high levels of organic acid anions. The activity of acetate-using methanogens is important for the syntrophic degradation of fatty acids when high levels of acetate are present.  相似文献   

8.
The metabolism of benzoate, cyclohex-1-ene carboxylate, and cyclohexane carboxylate by "Syntrophus aciditrophicus" in cocultures with hydrogen-using microorganisms was studied. Cyclohexane carboxylate, cyclohex-1-ene carboxylate, pimelate, and glutarate (or their coenzyme A [CoA] derivatives) transiently accumulated during growth with benzoate. Identification was based on comparison of retention times and mass spectra of trimethylsilyl derivatives to the retention times and mass spectra of authentic chemical standards. (13)C nuclear magnetic resonance spectroscopy confirmed that cyclohexane carboxylate and cyclohex-1-ene carboxylate were produced from [ring-(13)C(6)]benzoate. None of the metabolites mentioned above was detected in non-substrate-amended or heat-killed controls. Cyclohexane carboxylic acid accumulated to a concentration of 260 microM, accounting for about 18% of the initial benzoate added. This compound was not detected in culture extracts of Rhodopseudomonas palustris grown phototrophically or Thauera aromatica grown under nitrate-reducing conditions. Cocultures of "S. aciditrophicus" and Methanospirillum hungatei readily metabolized cyclohexane carboxylate and cyclohex-1-ene carboxylate at a rate slightly faster than the rate of benzoate metabolism. In addition to cyclohexane carboxylate, pimelate, and glutarate, 2-hydroxycyclohexane carboxylate was detected in trace amounts in cocultures grown with cyclohex-1-ene carboxylate. Cyclohex-1-ene carboxylate, pimelate, and glutarate were detected in cocultures grown with cyclohexane carboxylate at levels similar to those found in benzoate-grown cocultures. Cell extracts of "S. aciditrophicus" grown in a coculture with Desulfovibrio sp. strain G11 with benzoate or in a pure culture with crotonate contained the following enzyme activities: an ATP-dependent benzoyl-CoA ligase, cyclohex-1-ene carboxyl-CoA hydratase, and 2-hydroxycyclohexane carboxyl-CoA dehydrogenase, as well as pimelyl-CoA dehydrogenase, glutaryl-CoA dehydrogenase, and the enzymes required for conversion of crotonyl-CoA to acetate. 2-Ketocyclohexane carboxyl-CoA hydrolase activity was detected in cell extracts of "S. aciditrophicus"-Desulfovibrio sp. strain G11 benzoate-grown cocultures but not in crotonate-grown pure cultures of "S. aciditrophicus". These results are consistent with the hypothesis that ring reduction during syntrophic benzoate metabolism involves a four- or six-electron reduction step and that once cyclohex-1-ene carboxyl-CoA is made, it is metabolized in a manner similar to that in R. palustris.  相似文献   

9.
The metabolism of benzoate, cyclohex-1-ene carboxylate, and cyclohexane carboxylate by “Syntrophus aciditrophicus” in cocultures with hydrogen-using microorganisms was studied. Cyclohexane carboxylate, cyclohex-1-ene carboxylate, pimelate, and glutarate (or their coenzyme A [CoA] derivatives) transiently accumulated during growth with benzoate. Identification was based on comparison of retention times and mass spectra of trimethylsilyl derivatives to the retention times and mass spectra of authentic chemical standards. 13C nuclear magnetic resonance spectroscopy confirmed that cyclohexane carboxylate and cyclohex-1-ene carboxylate were produced from [ring-13C6]benzoate. None of the metabolites mentioned above was detected in non-substrate-amended or heat-killed controls. Cyclohexane carboxylic acid accumulated to a concentration of 260 μM, accounting for about 18% of the initial benzoate added. This compound was not detected in culture extracts of Rhodopseudomonas palustris grown phototrophically or Thauera aromatica grown under nitrate-reducing conditions. Cocultures of “S. aciditrophicus” and Methanospirillum hungatei readily metabolized cyclohexane carboxylate and cyclohex-1-ene carboxylate at a rate slightly faster than the rate of benzoate metabolism. In addition to cyclohexane carboxylate, pimelate, and glutarate, 2-hydroxycyclohexane carboxylate was detected in trace amounts in cocultures grown with cyclohex-1-ene carboxylate. Cyclohex-1-ene carboxylate, pimelate, and glutarate were detected in cocultures grown with cyclohexane carboxylate at levels similar to those found in benzoate-grown cocultures. Cell extracts of “S. aciditrophicus” grown in a coculture with Desulfovibrio sp. strain G11 with benzoate or in a pure culture with crotonate contained the following enzyme activities: an ATP-dependent benzoyl-CoA ligase, cyclohex-1-ene carboxyl-CoA hydratase, and 2-hydroxycyclohexane carboxyl-CoA dehydrogenase, as well as pimelyl-CoA dehydrogenase, glutaryl-CoA dehydrogenase, and the enzymes required for conversion of crotonyl-CoA to acetate. 2-Ketocyclohexane carboxyl-CoA hydrolase activity was detected in cell extracts of “S. aciditrophicus”-Desulfovibrio sp. strain G11 benzoate-grown cocultures but not in crotonate-grown pure cultures of “S. aciditrophicus”. These results are consistent with the hypothesis that ring reduction during syntrophic benzoate metabolism involves a four- or six-electron reduction step and that once cyclohex-1-ene carboxyl-CoA is made, it is metabolized in a manner similar to that in R. palustris.  相似文献   

10.
The strictly anaerobic Syntrophus aciditrophicus is a fermenting deltaproteobacterium that is able to degrade benzoate or crotonate in the presence and in the absence of a hydrogen-consuming partner. During growth in pure culture, both substrates are dismutated to acetate and cyclohexane carboxylate. In this work, the unknown enzymes involved in the late steps of cyclohexane carboxylate formation were studied. Using enzyme assays monitoring the oxidative direction, a cyclohex-1-ene-1-carboxyl-CoA (Ch1CoA)-forming cyclohexanecarboxyl-CoA (ChCoA) dehydrogenase was purified and characterized from S. aciditrophicus and after heterologous expression of its gene in Escherichia coli. In addition, a cyclohexa-1,5-diene-1-carboxyl-CoA (Ch1,5CoA)-forming Ch1CoA dehydrogenase was characterized after purification of the heterologously expressed gene. Both enzymes had a native molecular mass of 150 kDa and were composed of a single, 40- to 45-kDa subunit; both contained flavin adenine dinucleotide (FAD) as a cofactor. While the ChCoA dehydrogenase was competitively inhibited by Ch1CoA in the oxidative direction, Ch1CoA dehydrogenase further converted the product Ch1,5CoA to benzoyl-CoA. The results obtained suggest that Ch1,5CoA is a common intermediate in benzoate and crotonate fermentation that serves as an electron-accepting substrate for the two consecutively operating acyl-CoA dehydrogenases characterized in this work. In the case of benzoate fermentation, Ch1,5CoA is formed by a class II benzoyl-CoA reductase; in the case of crotonate fermentation, Ch1,5CoA is formed by reversing the reactions of the benzoyl-CoA degradation pathway that are also employed during the oxidative (degradative) branch of benzoate fermentation.  相似文献   

11.
Whole cells of Desulfobulbus propionicus fermented [1-13C]ethanol to [2-13C] and [3-13C]propionate and [1-13C]-acetate, which indicates the involvement of a randomizing pathway in the formation of propionate. Cell-free extracts prepared from cells grown on lactate (without sulfate) contained high activities of methylmalonyl-CoA: pyruvate transacetylase, acetase kinase and reasonably high activities of NAD(P)-independent L(+)-lactate dehydrogenase NAD(P)-independent pyruvate dehydrogenase, phosphotransacetylase, acetate kinase and reasonably high activity of NAD(P)-independent L(+)-lactate dehydrogenase, fumarate reductase and succinate dehydrogenase. Cell-free extracts catalyzed the conversion of succinate to propionate in the presence of pyruvate, CoA and ATP and the oxaloacetate-dependent conversion of propionate to succinate. After growth on lactate or propionate in the presence of sulfate similar enzyme levels were found except for fumarate reductase which was considerably lower. Fermentative growth on lactate led to higher cytochrome b contents than growth with sulfate as electron acceptor.The labeling studies and the enzyme measurements demonstrate that in Desulfobulbus propionate is formed via a succinate pathway involving a transcarboxylase like in Propionibacterium. The same pathway may be used for the degradation of propionate to acetate in the presence of sulfate.Abbreviations DCPIP 2,6-dichlorophenolindophenol - PEP phosphoenolpyruvate  相似文献   

12.
Cell-free extracts of Pseudomonas sp. strains KB 740 and K 172 both contained high levels of glutaryl-CoA dehydrogenase when grown anaerobically on benzoate or other aromatic compounds and with nitrate as electron acceptor. These aromatic compounds have in common benzoyl-CoA as the central aromatic intermediate of anerobic metabolism. The enzymatic activity was almost absent in cells grown aerobically on benzoate regardless whether nitrate was present. Glutaryl-CoA dehydrogenase activity was also detected in cell-free extracts of Rhodopseudomonas, Rhodomicrobium and Rhodocyclus after phototrophic growth on benzoate. Parallel to the induction of glutaryl-CoA dehydrogenase as measured with ferricenium ion as electron acceptor, an about equally high glutaconyl-CoA decarboxylase activity was detected in cell-free extracts. The latter activity was measured with the NAD-dependent assay, as described for the biotin-containing sodium ion pump glutaconyl-CoA decarboxylase from glutamate fermenting bacteria. Glutaryl-CoA dehydrogenase was purified to homogeneity from both Pseudomonas strains. The enzymes catalyse the decarboxylation of glutaconyl-CoA at about the same rate as the oxidative decarboxylation of glutaryl-CoA. The green enzymes are homotetramers (m=170 kDa) and contain 1 mol FAD per subunit. No inhibition was observed with avidin indicating the absence of biotin. The N-terminal sequences of the enzymes from both strains are similar (65%).  相似文献   

13.
Cyclohexane carboxylate supported relatively rapid growth (doubling times 7–8 h) of Rhodopseudomonas palustris under oxic or photosynthetic conditions, but did not serve as a substrate for either of the known aromatic CoA ligases. A CoA ligase that thioesterifies cyclohexane carboxylate was partially purified and did not cross react immunologically with the two CoA ligases purified previously from this bacterium. Crude extracts of R. palustris cells grown with a range of aromatic or alicyclic acids contained a dehydrogenase that reacted with cyclohexane carboxyl-CoA or cyclohex-1-ene carboxyl-CoA, using 2,6-dichlorophenolindophenol or ferricenium ion as electron carrier. This activity was not detected in extracts of adipate-, glutamate-, or succinate-grown cells. No oxidation or reduction of nonesterified cyclohexane carboxylate or cyclohexene carbocylate was detected in extracts of cells grown with aromatic or aliphatic substrates, neither aerobically nor anaerobically. A constitutively expressed thioesterase that hydrolyzed cyclohexane carboxyl-CoA and also some alicyclic and aliphatic CoA derivatives was purified and characterized. The enzyme had little or no activity on benzoyl-CoA or 4-hydroxybenzoyl-CoA. The presence of a thioesterase that effectively hydrolyzes cyclohexane carboxyl-CoA suggests that transient production of cyclohexane carboxylate is a physiological response to temporary excess of reductant during metabolism of aromatic compounds. Received: 22 May 1995 / Accepted: 13 September 1995  相似文献   

14.
Syntrophy is essential for the efficient conversion of organic carbon to methane in natural and constructed environments, but little is known about the enzymes involved in syntrophic carbon and electron flow. Syntrophus aciditrophicus strain SB syntrophically degrades benzoate and cyclohexane-1-carboxylate and catalyses the novel synthesis of benzoate and cyclohexane-1-carboxylate from crotonate. We used proteomic, biochemical and metabolomic approaches to determine what enzymes are used for fatty, aromatic and alicyclic acid degradation versus for benzoate and cyclohexane-1-carboxylate synthesis. Enzymes involved in the metabolism of cyclohex-1,5-diene carboxyl-CoA to acetyl-CoA were in high abundance in S. aciditrophicus cells grown in pure culture on crotonate and in coculture with Methanospirillum hungatei on crotonate, benzoate or cyclohexane-1-carboxylate. Incorporation of 13C-atoms from 1-[13C]-acetate into crotonate, benzoate and cyclohexane-1-carboxylate during growth on these different substrates showed that the pathways are reversible. A protein conduit for syntrophic reverse electron transfer from acyl-CoA intermediates to formate was detected. Ligases and membrane-bound pyrophosphatases make pyrophosphate needed for the synthesis of ATP by an acetyl-CoA synthetase. Syntrophus aciditrophicus, thus, uses a core set of enzymes that operates close to thermodynamic equilibrium to conserve energy in a novel and highly efficient manner.  相似文献   

15.
A stable, syntrophic benzoate-degrading bacterial consortium was enriched from sewage sludge. It oxidized benzoate or 3-phenylpropionate to acetate, H2 and CO2. As hydrogen scavengers Methanospirillum hungatei and Desulfovibrio sp. were present. The benzoate-degrading bacteria of this syntrophic culture and of Syntrophus buswelli were able to grow with benzoate/crotonate or crotonate alone in the absence of a hydrogen-utilizing partner organism. If crotonate was the only substrate, acetate and butyrate were produced, while during growth on benzoate or 3-phenylpropionate crotonate served as a reducible co-substrate and was exclusively converted to butyrate. In the presence of crotonate interspecies hydrogen transfer was not necessary as a hydrogen sink. The benzoate degrader was isolated as a pure culture with crotonate as the only carbon source. The pure culture could also grow with benzoate/crotonate or 3-phenylpropionate/crotonate. The effect of high concentrations of crotonate and of acetate or butyrate on growth of the benzoate degrader was investigated. The benzoate degrader was compared with S. buswellii for its morphology, physiology and DNA base composition. Except for the fact that S. buswellii was also able to grow on cinnamate, no differences between the two organisms were detected. The isolate is named S. buswelli, strain GA.  相似文献   

16.
We calculated the potential H2 and formate diffusion between microbes and found that at H2 concentrations commonly found in nature, H2 could not diffuse rapidly enough to dispersed methanogenic cells to account for the rate of methane synthesis but formate could. Our calculations were based on individual organisms dispersed in the medium, as supported by microscopic observations of butyrate-degrading cocultures. We isolated an axenic culture of Syntrophomonas wolfei and cultivated it on butyrate in syntrophic coculture with Methanobacterium formicicum; during growth the H2 concentration was 63 nM (10.6 Pa). S. wolfei contained formate dehydrogenase activity (as does M. formicicum), which would allow interspecies formate transfer in that coculture. Thus, interspecies formate transfer may be the predominant mechanism of syntrophy. Our diffusion calculations also indicated that H2 concentration at the cell surface of H2-consuming organisms was low but increased to approximately the bulk-fluid concentration at a distance of about 10 μm from the surface. Thus, routine estimation of kinetic parameters would greatly overestimate the Km for H2 or formate.  相似文献   

17.
In syntrophic conversion of butyrate to methane and CO2, butyrate is oxidized to acetate by secondary fermenting bacteria such as Syntrophomonas wolfei in close cooperation with methanogenic partner organisms, e.g., Methanospirillum hungatei. This process involves an energetically unfavourable shift of electrons from the level of butyryl-CoA oxidation to the substantially lower redox potential of proton and/or CO2 reduction, in order to transfer these electrons to the methanogenic partner via hydrogen and/or formate.In the present study, all prominent membrane-bound and soluble proteins expressed in S. wolfei specifically during syntrophic growth with butyrate, in comparison to pure-culture growth with crotonate, were examined by one- and two-dimensional gel electrophoresis, and identified by peptide fingerprinting-mass spectrometry. A membrane-bound, externally oriented, quinone-linked formate dehydrogenase complex was expressed at high level specifically during syntrophic butyrate oxidation, comprising a selenocystein-linked catalytic subunit with a membrane-translocation pathway signal (TAT), a membrane-bound iron-sulfur subunit, and a membrane-bound cytochrome. Soluble hydrogenases were expressed at high levels specifically during growth with crotonate. The results were confirmed by native protein gel electrophoresis, by formate dehydrogenase and hydrogenase-activity staining, and by analysis of formate dehydrogenase and hydrogenase activities in intact cells and cell extracts. Furthermore, constitutive expression of a membrane-bound, internally oriented iron-sulfur oxidoreductase (DUF224) was confirmed, together with expression of soluble electron-transfer flavoproteins (EtfAB) and two previously identified butyryl-CoA dehydrogenases.Our findings allow to depict an electron flow scheme for syntrophic butyrate oxidation in S. wolfei. Electrons derived from butyryl-CoA are transferred through a membrane-bound EtfAB:quinone oxidoreductase (DUF224) to a menaquinone cycle and further via a b-type cytochrome to an externally oriented formate dehydrogenase. Hence, an ATP hydrolysis-driven proton-motive force across the cytoplasmatic membrane would provide the energy input for the electron potential shift necessary for formate formation.  相似文献   

18.
Specific enzymes of ammonium assimilation were measured in cell-free extracts ofNocardia asteroides grown in a synthetic medium with glutamate as the nitrogen source. Cell-free extracts had active glutamine synthetase (GS) and glutamate synthase (GOGAT) and alanine dehydrogenase (ADH) but glutamate dehydrogenase (GDH) could not be detected in the enzyme preparation. This shows that GS/GOGAT is the major pathway of ammonium assimilation inN. asteroides.  相似文献   

19.
Resting cells of Clostridium sticklandii took up thymine or uracil, when grown in a medium containing 40 mM serine and 20 mM thymine or uracil. The uptake was much lower, when the cells had been grown in a complex medium. Cell-free extracts from cells grown in the complex medium reduced the two bases to the dihydro compounds and decomposed dihydrothymine to -ureidoisobutyrate, as indicated by thin-layer chromatography. Uptake and degradation were stimulated by both NADH and NADPH. Further breakdown did not occur, as 14CO2 was not evolved from C-2-labelled thymine or uracil. The rates of pyrimidine uptake and breakdown of C. sticklandii were lower than those reported for C. sporogenes (Hilton et al., 1975).  相似文献   

20.
The present study examined the aerobic metabolism of trimethylamine in Pseudomonas putida A ATCC 12633 grown on tetradecyltrimethylammonium bromide or trimethylamine. In both conditions, the trimethylamine was used as a nitrogen source and also accumulated in the cell, slowing the bacterial growth. Decreased bacterial growth was counteracted by the addition of AlCl3. Cell-free extracts prepared from cells grown aerobically on tetradecyltrimethylammonium bromide exhibited trimethylamine monooxygenase activity that produced trimethylamine N-oxide and trimethylamine N-oxide demethylase activity that produced dimethylamine. Cell-free extracts from cells grown on trimethylamine exhibited trimethylamine dehydrogenase activity that produced dimethylamine, which was oxidized to methanal and methylamine by dimethylamine dehydrogenase. These results show that this bacterial strain uses two enzymes to initiate the oxidation of trimethylamine in aerobic conditions. The apparent Km for trimethylamine was 0.7 mM for trimethylamine monooxygenase and 4.0 mM for trimethylamine dehydrogenase, but both enzymes maintain similar catalytic efficiency (0.5 and 0.4, respectively). Trimethylamine dehydrogenase was inhibited by trimethylamine from 1 mM. Therefore, the accumulation of trimethylamine inside Pseudomonas putida A ATCC 12633 grown on tetradecyltrimethylammonium bromide or trimethylamine may be due to the low catalytic efficiency of trimethylamine monooxygenase and trimethylamine dehydrogenase.  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