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 599 毫秒
1.
Intracellular levels of coenzyme A (CoA) and its derivatives involved in the metabolic pathways for Clostridium acetobutylicum ATCC 824 were analyzed by using reverse-phase high-performance liquid chromatography (HPLC). During the shift from the acidogenic to the solventogenic or stationary growth phase, the concentration of butyryl-CoA increased rapidly and the concentrations of free CoA and acetyl-CoA decreased. These changes were accompanied by a rapid increase of the solvent pathway enzyme activity and a decrease of the acid pathway enzyme activity. Assays with several non-solvent-producing mutant strains were also carried out. Upon entry of the mutant strains to the stationary phase, the butyryl-CoA concentrations for these mutant strains were comparable to those for the wild type even though the mutants were deficient in solvent-producing enzymes. Levels of acetoacetyl-CoA, β-hydroxy-butyryl-CoA, and crotonyl-CoA compounds in both wild-type and mutant extracts were below HPLC detection thresholds (<21 μM).  相似文献   

2.
The properties and developmental change in the activity of cytosolic 3-hydroxy-3-methyl glutaryl coenzyme A (HMG-CoA) synthase in brain was examined and whether or not HMG-CoA lyase is present in cytosol and mitochondria from brain was determined. Although mitochondrial fractions contained significant HMG-CoA lyase activity, the enzyme activity was not detected in brain cytosol. The synthase activity was present in both mitochondrial and cytosolic fraction. The HMG-CoA synthesis by brain cytosol was optimal at pH 8.0 and did not require Mg2+ or exogenous acetoacetyl CoA. This indicates that brain cytosol can synthesize sufficient quantity of acetoacetyl CoA from acetyl CoA to be utilized for HMG-CoA synthesis. Our results also showed that the specific activity (nmol acetyl CoA incorporated/mg protein) of HMG-CoA synthase in brain cytosol was high (between 2–11 days of postnatal age) when the cholesterol content of brain is increasing rapidly, and the activity declined slowly thereafter. This suggests that in brain, cytosolic enzyme HMG-CoA synthase plays a role in the regulation of cholesterol synthesis.  相似文献   

3.
Histone deacetylases (HDACs) function in a wide range of molecular processes, including gene expression, and are of significant interest as therapeutic targets. Although their native complexes, subcellular localization, and recruitment mechanisms to chromatin have been extensively studied, much less is known about whether the enzymatic activity of non-sirtuin HDACs can be regulated by natural metabolites. Here, we show that several coenzyme A (CoA) derivatives, such as acetyl-CoA, butyryl-CoA, HMG-CoA, and malonyl-CoA, as well as NADPH but not NADP(+), NADH, or NAD(+), act as allosteric activators of recombinant HDAC1 and HDAC2 in vitro following a mixed activation kinetic. In contrast, free CoA, like unconjugated butyrate, inhibits HDAC activity in vitro. Analysis of a large number of engineered HDAC1 mutants suggests that the HDAC activity can potentially be decoupled from "activatability" by the CoA derivatives. In vivo, pharmacological inhibition of glucose-6-phosphate dehydrogenase (G6PD) to decrease NADPH levels led to significant increases in global levels of histone H3 and H4 acetylation. The similarity in structures of the identified metabolites and the exquisite selectivity of NADPH over NADP(+), NADH, and NAD(+) as an HDAC activator reveal a previously unrecognized biochemical feature of the HDAC proteins with important consequences for regulation of histone acetylation as well as the development of more specific and potent HDAC inhibitors.  相似文献   

4.
A new coenzyme A (CoA)-transferase from the anaerobe Clostridium aminobutyricum catalyzing the formation of 4-hydroxybutyryl-CoA from 4-hydroxybutyrate and acetyl-CoA is described. The enzyme was purified to homogeneity by standard techniques, including fast protein liquid chromatography under aerobic conditions. Its molecular mass was determined to be 110 kDa, and that of the only subunit was determined to be 54 kDa, indicating a homodimeric structure. Besides acetate and acetyl-CoA, the following substrates were detected (in order of decreasing kcat/Km): 4-hydroxybutyryl-CoA, butyryl-CoA and propionyl-CoA, vinyl-acetyl-CoA (3-butenoyl-CoA), and 5-hydroxyvaleryl-CoA. In an indirect assay the corresponding acids were also found to be substrates; however, DL-lactate, DL-2-hydroxybutyrate, DL-3-hydroxybutyrate, crotonate, and various dicarboxylates were not.  相似文献   

5.
A new coenzyme A (CoA)-transferase from the anaerobe Clostridium aminobutyricum catalyzing the formation of 4-hydroxybutyryl-CoA from 4-hydroxybutyrate and acetyl-CoA is described. The enzyme was purified to homogeneity by standard techniques, including fast protein liquid chromatography under aerobic conditions. Its molecular mass was determined to be 110 kDa, and that of the only subunit was determined to be 54 kDa, indicating a homodimeric structure. Besides acetate and acetyl-CoA, the following substrates were detected (in order of decreasing kcat/Km): 4-hydroxybutyryl-CoA, butyryl-CoA and propionyl-CoA, vinyl-acetyl-CoA (3-butenoyl-CoA), and 5-hydroxyvaleryl-CoA. In an indirect assay the corresponding acids were also found to be substrates; however, DL-lactate, DL-2-hydroxybutyrate, DL-3-hydroxybutyrate, crotonate, and various dicarboxylates were not.  相似文献   

6.
Clostridium aminobutyricum ferments 4-aminobutyrate (γ-aminobutyrate, GABA) to ammonia, acetate and butyrate via 4-hydroxybutyrate that is activated to the CoA-thioester catalyzed by 4-hydroxybutyrate CoA-transferase. Then, 4-hydroxybutyryl-CoA is dehydrated to crotonyl-CoA, which disproportionates to butyryl-CoA and acetyl-CoA. Cocrystallization of the CoA-transferase with the alternate substrate butyryl-CoA yielded crystals with non-covalently bound CoA and two water molecules at the active site. Most likely, butyryl-CoA reacted with the active site Glu238 to CoA and the mixed anhydride, which slowly hydrolyzed during crystallization. The structure of the CoA is similar but less stretched than that of the CoA-moiety of the covalent enzyme-CoA-thioester in 4-hydroxybutyrate CoA-transferase from Shewanella oneidensis. In contrast to the structures of the apo-enzyme and enzyme-CoA-thioester, the structure described here has a closed conformation, probably caused by a flip of the active site loop (residues 215–219). During turnover, the closed conformation may protect the anhydride intermediate from hydrolysis and CoA from dissociation from the enzyme. Hence, one catalytic cycle changes conformation of the enzyme four times: free enzyme—open conformation, CoA+ anhydride 1—closed, enzyme-CoA-thioester—open, CoA + anhydride-2—closed, free enzyme—open.  相似文献   

7.
The activation of pyruvate dehydrogenasea kinase activity by CoA esters has been further characterized. Half-maximal activation of kinase activity was achieved with about 1.0 microM acetyl-CoA after a 20-s preincubation in the presence of NADH. More than 80% of the acetyl-CoA was consumed during this period in acetylating sites in the pyruvate dehydrogenase complex as a result of the transacetylation reaction proceeding to equilibrium. At 1.0 microM acetyl-CoA, this resulted in more than a 4-fold higher level of CoA than residual acetyl-CoA. Activation of kinase activity could result either from acetylation of specific sites in the complex or tight binding of acetyl-CoA. Removal of CoA enhanced both acetylation and activation, suggesting acetylation mediates activation. For allosteric binding of acetyl-CoA to elicit activation, an activation constant, Ka, less than 50 nM would be required. To further distinguish between those mechanisms, the effects of other CoA esters as well as the reactivity of most of the effective CoA esters were characterized. Several short-chain CoA esters enhanced kinase activity including (in decreasing order of effectiveness) malonyl-CoA, acetoacetyl-CoA, propionyl-CoA, and methylmalonyl-CoA. Butyryl-CoA inhibited kinase activity as did high concentrations of long-chain acyl-CoAs. Inhibition by long-chain acyl-CoAs may result, in part, from detergent-like properties of those esters. Malonyl-CoA, propionyl-CoA, butyryl-CoA, and methylmalonyl-CoA, obtained with radiolabeled acyl groups, were shown to acylate sites in the complex. Propionyl-CoA and butyryl-CoA were tested, in competition with acetyl-CoA or pyruvate, as alternative substrates for acylation of sites in the complex and as competitive effectors of kinase activity. Propionyl-CoA alone rapidly acylated sites in the complex at low concentrations, and low concentrations of propionyl-CoA were effective in activating kinase activity although only a relatively small activation was observed. When an equivalent level (20 microM) of acetyl-CoA and propionyl-CoA was used, marked activation of kinase activity due to a dominant effect of acetyl-CoA was associated with acetylation of a major portion of sites in the complex and with a small portion undergoing acylation with propionyl-CoA. Those results were rapidly achieved in a manner independent of the order of addition of the two CoA esters. That indicates that tight slowly reversible binding of acetyl-CoA is not involved in kinase activation. High levels of propionyl-CoA greatly reduced acetylation by acetyl-CoA and nearly prevented activation of kinase activity by acetyl-CoA.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

8.
Acetyl-CoA synthase (ACS ACS/CODH CODH/ACS) from Moorella thermoacetica catalyzes the synthesis of acetyl-CoA from CO, CoA, and a methyl group of a corrinoid-iron-sulfur protein (CoFeSP). A time lag prior to the onset of acetyl-CoA production, varying from 4 to 20 min, was observed in assay solutions lacking the low-potential electron-transfer agent methyl viologen (MV). No lag was observed when MV was included in the assay. The length of the lag depended on the concentrations of CO and ACS, with shorter lags found for higher [ACS] and sub-saturating [CO]. Lag length also depended on CoFeSP. Rate profiles of acetyl-CoA synthesis, including the lag phase, were numerically simulated assuming an autocatalytic mechanism. A similar reaction profile was monitored by UV-vis spectrophotometry, allowing the redox status of the CoFeSP to be evaluated during this process. At early stages in the lag phase, Co2+FeSP reduced to Co+FeSP, and this was rapidly methylated to afford CH3-Co3+FeSP. During steady-state synthesis of acetyl-CoA, CoFeSP was predominately in the CH3-Co3+FeSP state. As the synthesis rate declined and eventually ceased, the Co+FeSP state predominated. Three activation reductive reactions may be involved, including reduction of the A- and C-clusters within ACS and the reduction of the cobamide of CoFeSP. The B-, C-, and D-clusters in the subunit appear to be electronically isolated from the A-cluster in the connected subunit, consistent with the ~70 Å distance separating these clusters, suggesting the need for an in vivo reductant that activates ACS and/or CoFeSP.Abbreviations ACS acetyl-CoA synthase, also known as CODH (carbon monoxide dehydrogenase) or CODH/ACS or ACS/CODH - CH3-Co3+FeSP, Co2+FeSP, and Co+FeSP corrinoid-iron-sulfur protein with the cobalamin in the methylated 3+, unmethylated 2+, and unmethylated 1+ states - CoA coenzyme A - DTT dithiothreitol - H-THF or THF tetrahydrofolic acid or tetrahydrofolate - MT methyl transferase - MV methyl viologen  相似文献   

9.
Cell extracts of butyrate-forming clostridia have been shown to catalyze acetyl-coenzyme A (acetyl-CoA)- and ferredoxin-dependent formation of H2 from NADH. It has been proposed that these bacteria contain an NADH:ferredoxin oxidoreductase which is allosterically regulated by acetyl-CoA. We report here that ferredoxin reduction with NADH in cell extracts from Clostridium kluyveri is catalyzed by the butyryl-CoA dehydrogenase/Etf complex and that the acetyl-CoA dependence previously observed is due to the fact that the cell extracts catalyze the reduction of acetyl-CoA with NADH via crotonyl-CoA to butyryl-CoA. The cytoplasmic butyryl-CoA dehydrogenase complex was purified and is shown to couple the endergonic reduction of ferredoxin (E0′ = −410 mV) with NADH (E0′ = −320 mV) to the exergonic reduction of crotonyl-CoA to butyryl-CoA (E0′ = −10 mV) with NADH. The stoichiometry of the fully coupled reaction is extrapolated to be as follows: 2 NADH + 1 oxidized ferredoxin + 1 crotonyl-CoA = 2 NAD+ + 1 ferredoxin reduced by two electrons + 1 butyryl-CoA. The implications of this finding for the energy metabolism of butyrate-forming anaerobes are discussed in the accompanying paper.  相似文献   

10.
A high-performance liquid chromatographic method has been developed to measure short-chain CoA compounds in freeze-clamped liver. Seventeen CoA compounds can be quantitated in 37 min using a 3-micron octadecylsilica column (4.6 mm X 7.5 cm). The chromatographic separation of CoA compounds is conducted with a gradient system of sodium phosphate and acetonitrile. The large amount of uv-absorbing, non-CoA material present in liver extracts is eluted earlier than the CoA compounds when the phosphate concentration is 0.2 M. The CoA compounds that can be resolved by this method include acetoacetyl-CoA, acetyl-CoA, butyryl-CoA, CoASH, crotonyl-CoA, dephospho-CoA, glutathione-CoA, 3-hydroxy-3-methylglutaryl-CoA, isobutyryl-CoA, isovaleryl-CoA, malonyl-CoA, 3-methylcrotonyl-CoA, methylmalonyl-CoA, oxidized-CoA, propionyl CoA, succinyl-CoA, and valeryl-CoA. Comparisons at pH 3 and 6 showed that the stability of the CoA compounds is much greater when perchloric acid extracts of rat liver are adjusted to pH 3. Recovery of CoA standards added in tissue extracts ranged from 83 to 107%. The method is linear over the range of 12 to 700 pmol, and this sensitivity allows acetyl-CoA content to be determined in extracts of as little as 0.1 mg of liver. The values for CoA compounds obtained for freeze-clamped liver from starved rats include (units are nmol/g wet weight +/- SE) malonyl-CoA, 1.50 +/- 0.14; glutathione-CoA, 6.57 +/- 1.72; CoASH, 56.06 +/- 2.90; methylmalonyl-CoA, 4.60 +/- 1.27; succinyl-CoA, 13.52 +/- 0.76; 3-hydroxy-3-methylglutaryl-CoA, 7.06 +/- 0.89; and acetyl-CoA, 100.5 +/- 6.4.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

11.
Abstract: The pathway from acetyl-CoA to butyryl-CoA serves as a major carbon metabolism channel in Clostridium acetobutylicum and other butyrate-forming clostridia, and the steps are similar to those involved in fatty acid metabolism. Recent findings are discussed, reviewing the isolation and characterization of the enzymes of the pathway, and the analyses of metabolic intermediate levels and possible points of regulation of enzyme activity by CoA compounds. DNA analyses have identified the genes for two thiolase proteins, and an apparent operon encoding five proteins involved in the conversion of acetoacetyl-CoA to butyryl-CoA. These five proteins are β-hydroxybutyryl-CoA dehydrogenase, crotonase, butyryl-CoA dehydrogenase and the α and β subunits of an electron transfer flavoprotein.  相似文献   

12.
The effects of carnitine on the metabolism of palmitoylcarnitine were studied by using isolated rat liver mitochondria. Particular attention was given to carnitine acyltransferase-mediated interactions between carnitine and the mitochondrial CoA pool. Carnitine concentrations less than 1.25mm resulted in an increased production of acetylcarnitine during palmitoylcarnitine oxidation. Despite this shunting of C2 units to acetylcarnitine formation, no change was observed in the rate of oxygen consumption or major product formation (citrate or acetoacetate). Further, no changes were observed in the mitochondrial content of acetyl-CoA, total acid-soluble CoA or acid-insoluble acyl-CoA. These observations support the concept, based on studies in vivo, that the carnitine/acylcarnitine pool is metabolically sluggish and the acyl-group flux low as compared with the CoA/acyl-CoA pool. Acid-insoluble acyl-CoA content was decreased and CoA content increased at carnitine concentrations greater than 1.25mm. When [14C]carnitine was used in the incubations, it was demonstrated that this resulted from acid-insoluble acylcarnitine formation from intramitochondrial acid-insoluble acyl-CoA mediated by carnitine palmitoyltransferase B. Again, the higher carnitine concentrations resulted in no changes in the rates of oxygen consumption or major product formation. The above effects of carnitine were observed whether citrate or acetoacetate was the major product of oxidation. In contrast, an increase in acetyl-CoA concentration was observed at high carnitine concentrations only when acetoacetate was the product. Since the rate of acetoacetate production was not changed, these higher acetyl-CoA concentrations suggest that a new steady state had been established to maintain acetoacetate-production rates. Since there was no change in acetyl-CoA concentration when citrate was the major product, a change in the activity of the pathway utilizing acetyl-CoA for ketone-body synthesis and the potential regulation of this pathway must be considered.  相似文献   

13.
Long-chain acyl-CoA hydrolase (EC 3.1.2.2.) has been partially purified from the 100,000 × g supernatant fraction of rat brain tissue. The purification procedure included chromatography on gel filtration media, DEAE-cellulose, CM-cellulose, and hydroxyapatite. The partially purified enzyme had a specific activity of 7.1 mol/min-mg, and when analyzed by polyacrylamide gel electrophoresis, revealed one major and three minor bands of protein in the presence of dodecyl sulfate and two major bands of protein in the absence of dodecyl sulfate. The enzyme had a molecular weight of 65,000 and showed no evidence of aggregated or dissociated forms. The highest catalytic activity was exhibited with palmitoyl-CoA and oleoyl-CoA as substrates. Lower activity was found with decanoyl-CoA as the substrate and little or no activity was found with acetyl-CoA, malonyl-CoA, butyryl-CoA, or acetoacetyl-CoA. The enzyme was inhibited by CoA, various metal ions, including Mn2+, Mg2+ and Ca2+, and by bovine serum albumin. Heating the enzyme produced a loss of activity which corresponded to a first-order kinetic process, the rate of which was independent of the choice of substrate used to measure enzyme activity. This finding supports the idea that the purification procedure yields a single species of long-chain acyl-CoA hydrolase.  相似文献   

14.
15.
The disease process of ulcerative colitis (UC) is associated with a block in -oxidation of short chain fatty acid in colonic epithelial cells which can be reproduced by exposure of cells to sulfides. The aim of the current work was to assess the level in the -oxidation pathway at which sulfides might be inhibitory in human colonocytes. Isolated human colonocytes from cases without colitis (n = 12) were exposed to sulfide (1.5 mM) in the presence or absence of exogenous CoA and ATP. Short chain acyl-CoA esters were measured by a high performance liquid chromatographic assay. 14CO2 generation was measured from [1-14C]butyrate and [6-14C]glucose. 14CO2 from butyrate was significantly reduced (p < 0.001) by sulfide. When colonocytes were incubated with hydrogen sulfide in the presence of CoA and ATP, butyryl-CoA concentration was increased (p < 0.01), while crotonyl-CoA (p < 0.01) and acetyl-CoA (p < 0.01) concentrations were decreased. These results show that sulfides inhibit short chain acyl-CoA dehydrogenase. As oxidation of n-butyrate governs the epithelial barrier function of colonocytes the functional activity of short chain acyl-CoA dehydrogenase may be critical in maintaining colonic mucosal integrity. Maintaining the functional activity of dehydrogenases could be an important determinant in the expression of ulcerative colitis.  相似文献   

16.
Glucose metabolism and the mechanisms of NADH oxidation by Treponema hyodysenteriae were studied. Under an N2 atmosphere, washed cell suspensions of the spirochete consumed glucose and produced acetate, butyrate, H2, and CO2. Approximately twice as much H2 as CO2 was produced. Determinations of radioactivity in products of [14C]glucose and [14C]pyruvate metabolism and analyses of enzyme activities in cell lysates revealed that glucose was catabolized to pyruvate via the Embden-Meyerhof-Parnas pathway. The results of pyruvate exchange reactions with NaH14CO3 and Na14COOH demonstrated that pyruvate was converted to acetyl coenzyme A (acetyl-CoA), H2, and CO2 by a clostridium-type phosphoroclastic mechanism. NADH:ferredoxin oxidoreductase and hydrogenase activities were present in cell lysates and produced H2 from NADH oxidation. Phosphotransacetylase and acetate kinase catalyzed the formation of acetate from acetyl-CoA. Butyrate was formed from acetyl-CoA via a pathway that involved 3-hydroxybutyryl-coenzyme A (CoA) dehydrogenase, butyryl-CoA dehydrogenase, and butyryl-CoA transferase. T. hyodysenteriae cell suspensions generated less H2 and butyrate under 10% O2-90% N2 than under 100% N2. Cell lysates contained NADH oxidase, NADH peroxidase, and superoxide dismutase activities. These findings indicated there are three major mechanisms that T. hyodysenteriae cells use to recycle NADH generated from the Embden-Meyerhof-Parnas pathway--enzymes in the pathway from acetyl-CoA to butyrate, NADH:ferredoxin oxidoreductase, and NADH oxidase. Versatility in methods of NADH oxidation and an ability to metabolize oxygen could benefit T. hyodysenteriae cells in the colonization of tissues of the swine large bowel.  相似文献   

17.
Complete arrest of maternal-fetal blood supply for up to 30 min caused a time-dependent increase in the endogenous levels of lipid peroxides (LPO) in fetal brain and liver extracts and fetal blood and amniotic fluids as indicated by the appearance of thiobarbituric acid reactive substances (TBARS), A steady increase of TBARS from 48.0 ± 2.2 pmol/g wet weight to 75.0 ± 5.6 pmol/g wet weight up to 30 min restriction was noticed in the fetal brain. The fetal liver TBARS values increased by approximately 69% after 5 min restriction and remained high, above the control level, for 30 min. After two days reperfusion following 30 min restriction, the TBARS levels in the fetal brain were 1.8 fold higher above the control, while those of the liver returned to control values. The levels of the lipid-soluble antioxidant -tocopherol were reduced by about 40% and 50% in the placenta and brain tissues after 5 min restriction, respectively. Slices of fetal brain incubated at 37°C in DMEM under oxygen in the presence of 50 M Fe2+ were able to generate LPO in a time- and tissue concentration-dependent manner. After 15 min incubation, about 6.3 fold increase in total TBARS levels could be measured in the presence of 50 M Fe2+, most of which was released in the medium. The iron chelator desferrioxamine (25 M) and the antioxidant -tocopherol (10 M) added to the incubation medium, each inhibited by about 88% TBARS production. After 20 min episode of ischemia, fetal brain slices released into the medium 138.5 ± 9.8 nmol/15 min/mg DNA compared to 75.9 ± 4.5 nmol/15 min/mg DNA released by the sham preparations. After 3 h reperfusion, brain slices from fetuses exposed to 20 min ischemia continued to produce TBARS above control levels, whereas those of brief ischemia (5 min) returned to control levels. The data indicate that the limited resistance of the fetal brain to brief, rather than prolonged, periods of ischemia, is likely due to a lack of free FA for LPO generation, rather than the levels of tissue lipid antioxidants.  相似文献   

18.
The effects of varying concentrations and types of volatile anesthetics on neurochemical sequelae of brain ischemia were evaluated in the rat. Rats were assigned to treatment defined by a 3×3 design (anesthetic type and dose) with 5 rats/cell. Each group received halothane, enflurane, or isoflurane 0.5, 1.0, or 2.0 MAC (minimal alevolar concentration). This was followed by preischemic plasma glucose sampling, 5 min hypotension (30 mmHg) and 5 min decapitation cerebral ischemia. Preischemia plasma glucose increased with increasing anesthetic concentration and was highest in the isoflurane groups, varying from a low (±SD) of 7.19±1.79 mol/ml in the 0.5 MAC halothane group to a high of 12.68±3.65 mol/ml in the 2.0 MAC isoflurane group. End-ischemic brain lactate correlated with preischemic plasma glucose (r=0.5, =0.5). We conclude that increasing concentration of volatile anesthesia with iv phenylephrine blood pressure support produces higher levels of plasma glucose and brain lactate with cerebral ischemia.  相似文献   

19.
Glucose metabolism and the mechanisms of NADH oxidation by Treponema hyodysenteriae were studied. Under an N2 atmosphere, washed cell suspensions of the spirochete consumed glucose and produced acetate, butyrate, H2, and CO2. Approximately twice as much H2 as CO2 was produced. Determinations of radioactivity in products of [14C]glucose and [14C]pyruvate metabolism and analyses of enzyme activities in cell lysates revealed that glucose was catabolized to pyruvate via the Embden-Meyerhof-Parnas pathway. The results of pyruvate exchange reactions with NaH14CO3 and Na14COOH demonstrated that pyruvate was converted to acetyl coenzyme A (acetyl-CoA), H2, and CO2 by a clostridium-type phosphoroclastic mechanism. NADH:ferredoxin oxidoreductase and hydrogenase activities were present in cell lysates and produced H2 from NADH oxidation. Phosphotransacetylase and acetate kinase catalyzed the formation of acetate from acetyl-CoA. Butyrate was formed from acetyl-CoA via a pathway that involved 3-hydroxybutyryl-coenzyme A (CoA) dehydrogenase, butyryl-CoA dehydrogenase, and butyryl-CoA transferase. T. hyodysenteriae cell suspensions generated less H2 and butyrate under 10% O2-90% N2 than under 100% N2. Cell lysates contained NADH oxidase, NADH peroxidase, and superoxide dismutase activities. These findings indicated there are three major mechanisms that T. hyodysenteriae cells use to recycle NADH generated from the Embden-Meyerhof-Parnas pathway--enzymes in the pathway from acetyl-CoA to butyrate, NADH:ferredoxin oxidoreductase, and NADH oxidase. Versatility in methods of NADH oxidation and an ability to metabolize oxygen could benefit T. hyodysenteriae cells in the colonization of tissues of the swine large bowel.  相似文献   

20.
The butyrogenic genes from Clostridium difficile DSM 1296T have been cloned and expressed in Escherichia coli. The enzymes acetyl-coenzyme A (CoA) C-acetyltransferase, 3-hydroxybutyryl-CoA dehydrogenase, crotonase, phosphate butyryltransferase, and butyrate kinase and the butyryl-CoA dehydrogenase complex composed of the dehydrogenase and two electron-transferring flavoprotein subunits were individually produced in E. coli and kinetically characterized in vitro. While most of these enzymes were measured using well-established test systems, novel methods to determine butyrate kinase and butyryl-CoA dehydrogenase activities with respect to physiological function were developed. Subsequently, the individual genes were combined to form a single plasmid-encoded operon in a plasmid vector, which was successfully used to confer butyrate-forming capability to the host. In vitro and in vivo studies demonstrated that C. difficile possesses a bifurcating butyryl-CoA dehydrogenase which catalyzes the NADH-dependent reduction of ferredoxin coupled to the reduction of crotonyl-CoA also by NADH. Since the reoxidation of ferredoxin by a membrane-bound ferredoxin:NAD+-oxidoreductase enables electron transport phosphorylation, additional ATP is formed. The butyryl-CoA dehydrogenase from C. difficile is oxygen stable and apparently uses oxygen as a co-oxidant of NADH in the presence of air. These properties suggest that this enzyme complex might be well suited to provide butyryl-CoA for solventogenesis in recombinant strains. The central role of bifurcating butyryl-CoA dehydrogenases and membrane-bound ferredoxin:NAD oxidoreductases (Rhodobacter nitrogen fixation [RNF]), which affect the energy yield of butyrate fermentation in the clostridial metabolism, is discussed.  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