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.
Two strains of new strictly anaerobic, gramnegative bacteria were enriched and isolated from a freshwater (strain WoG13) and a saltwater (strain CuG11) anoxic sediment with glutarate as sole energy source. Strain WoG13 formed spores whereas strain CuG11 did not. Both strains were rod-shaped, motile bacteria growing in carbonate-buffered, sulfide-reduced mineral medium supplemented with 2% of rumen fluid. Both strains fermented glutarate to butyrate, isobutyrate, CO2, and small amounts of acetate. With methylsuccinate, the same products were formed, and succinate was fermented to propionate and CO2. No sugars, amino acids or other organic acids were used as substrates. Molar growth yields (Ys) were very small (0.5–0.9 g cell dry mass/mol dicarboxylate). Cells of strain WoG13 contained no cytochromes, and the DNA base ratio was 49.0±1.4 mol% guanine-plus-cytosine. Enzyme activities involved in glutarate degradation could bedemonstrated in cell-free extracts of strain WoG13. A pathway of glutarate fermentation via decarboxylation of glutaconyl-CoA to crotonyl-CoA is suggested which forms butyrate and partly isobutyrate by subsequent isomerization.  相似文献   

2.
Propionate metabolism was studied in ovine hepatocytes. The main products of metabolism were CO2, glucose, L-lactate and pyruvate. The fatty acids, butyrate and palmitate inhibited propionate oxidation; butyrate inhibited but palmitate slightly stimulated gluconeogenesis from propionate. Butyrate and palmitate also inhibited lactate and pyruvate production from both endogenous substrates and from propionate.  相似文献   

3.
Spirochete MA-2, which is anaerobic, ferments glucose, forming acetate as a major product. The spirochete also ferments (but does not utilize as growth substrates) small amounts of l-leucine, l-isoleucine, and l-valine, forming the branched-chain fatty acids isovalerate, 2-methylbutyrate, and isobutyrate, respectively, as end products. Energy generated through the fermentation of these amino acids is utilized to prolong cell survival under conditions of growth substrate starvation. A branched-chain fatty acid kinase and two acetate kinase isozymes were resolved from spirochete MA-2 cell extracts. Kinase activity was followed by measuring the formation of acyl phosphate from fatty acid and ATP. The branched-chain fatty acid kinase was active with isobutyrate, 2-methylbutyrate, isovalerate, butyrate, valerate, or propionate as a substrate but not with acetate as a substrate. The acetate kinase isozymes were active with acetate and propionate as substrates but not with longer-chain fatty acids as substrates. The acetate kinase isozymes and the branched-chain fatty acid kinase differed in nucleoside triphosphate and cation specificities. Each acetate kinase isozyme had an apparent molecular weight of approximately 125,000, whereas the branched-chain fatty acid kinase had a molecular weight of approximately 76,000. These results show that spirochete MA-2 synthesizes a branched-chain fatty acid kinase specific for leucine, isoleucine, and valine fermentation. It is likely that a phosphate branched-chain amino acids is also synthesized by spirochete MA-2. Thus, in spirochete MA-2, physiological mechanisms have evolved which serve specifically to generate maintenance energy from branched-chain amino acids.  相似文献   

4.
Three strains of new strictly anaerobic, Grampositive, non-sporeforming bacteria were isolated from various anoxic sediment samples with putrescine as sole carbon and energy source. Optimal growth in carbonate-buffered defined medium occurred at 37°C at pH 7.2–7.6. The DNA base ratio of strain NorPut1 was 29.6±1 mol% guanine plus cytosine. In addition to a surface layer and the peptidoglycan layer, the cell wall contained a second innermost layer with a periodic arrangement of subunits. All strains fermented putrescine to acetate, butyrate, and molecular hydrogen; the latter originated from both oxidative putrescine deamination and 4-aminobutyraldehyde oxidation. In defined mixed cultures with methanogens or homoacetogenic bacteria, methane or additional acetate were formed due to interspecies hydrogen transfer. Also 4-aminobutyrate and 4-hydroxybutyrate were fermented to acetate and butyrate, but no hydrogen was released from these substrates. No sugars, organic acids, other primary amines or amino acids were used as substrates. Neither sulfate, thiosulfate, sulfur, nitrate nor fumarate was reduced. Most of the enzymes involved in putrescine degradation could be demonstrated in cell-free extracts. A pathway of putrescine fermentation via 4-aminobutyrate and crotonyl-CoA with subsequent dismutation to acetate and butyrate is suggested.  相似文献   

5.
Anaerobic bacteria from the orders Clostridiales and Fusobacteriales are able to ferment glutamate by at least five different pathways, most of which contain enzymes with radicals in their catalytic pathways. The first two pathways proceed to ammonia, acetate and pyruvate via the coenzyme B12-dependent glutamate mutase, which catalyses the re-arrangement of the linear carbon skeleton to that of the branched-chain amino acid (2S,3S)-3-methylaspartate. Pyruvate then disproportionates either to CO2 and butyrate or to CO2, acetate and propionate. In the third pathway, glutamate again is converted to ammonia, CO2, acetate and butyrate. The key intermediate is (R)-2-hydroxyglutaryl-CoA, which is dehydrated to glutaconyl-CoA, followed by decarboxylation to crotonyl-CoA. The unusual dehydratase, containing an iron-sulfur cluster, is activated by an ATP-dependent one-electron reduction. The remaining two pathways require more then one organism for the complete catabolism of glutamate to short chain fatty acids. Decarboxylation of glutamate leads to 4-aminobutyrate, which is fermented by a second organism via the fourth pathway to acetate and butyrate, again mediated by an unusual dehydratase which catalyses the reversible dehydration of 4-hydroxybutyryl-CoA to crotonyl-CoA. The fifth pathway is the only one without decarboxylation, since the gamma-carboxylate of glutamate is reduced to the amino group of delta-aminovalerate, which then is fermented to acetate, propionate and valerate. The pathway involves the oxidative dehydration of 5-hydroxyvaleryl-CoA to 2,4-pentadienoyl-CoA followed by reduction to 3-pentenoyl-CoA and isomerisation to 2-pentenoyl-CoA.  相似文献   

6.
Summary The effect of monensin and 2-bromoethanesulfonic acid (BESA) on methane production from cattle manure and on volatile fatty acids metabolism was tested. At 10 days retention time 0.81 biogas per liter cattle manure and day were produced. Methanogenesis was inhibited 20% by 3 mM BESA per liter and 45% by 2–5 mg monensin per liter. When the digestion was inhibited with either of the both drugs, the acetate pool increased drastically. Like in untreated fermentations the propionate pool increased in BESA-inhibited fermentations for several hours after substrate addition. After 24 h however it did not decrease to the low level reached in non-inhibited fermentations. When monensin was the inhibitor, the propionate pool did not change for 15 h, but then decreased with the same rate as in the control experiment. Adaptation processes or detoxification may be responsible for the delayed degradation.The degradation of low concentrations of buty-rate to acetate and the turn over rates of the butyrate pool are almost identical in cattle manure containing BESA, monensin, or no inhibitor. The turn over of 14C-acetate from butyrate degradation is delayed in BESA and monensin inhibited fermentations.From the data presented it can be concluded, that BESA mainly inhibits the methanogens, while monensin seems to inhibit both, methanogenic and nonmethanogenic organisms. However, a fast adaptation to or detoxification of the antibiotic seems to occur.  相似文献   

7.
High amylose maize starch (HAS) is not digested in the small intestine and most of it reaches the large intestine. In the large intestine, HAS is fermented by intestinal bacteria, resulting in production of short-chain fatty acids (SCFA), particularly butyrate. Clostridium butyricum can utilize HAS and produce butyrate and acetate. It has been proposed that butyrate inhibits carcinogenesis in the colon. In this study, we examined the inhibitory effects of HAS and C. butyricum strain MIYAIRI588 (CBM588) on azoxymethane-induced aberrant crypt foci (ACF) formation in rats. In the group of rats administered only CBM588 spores, the concentration of butyrate in the cecum increased, but there was no decrease in the number of ACF. In the group of rats fed an HAS diet, a decrease in the number of ACF was observed, and in the group of rats administered HAS and CBM588, the number of ACF decreased significantly. In these two groups, the concentrations of acetate and propionate in intestinal contents significantly increased, but the concentration of butyrate did not change. It was found that the beta-glucuronidase activity level of colonic contents decreased significantly in the two groups of rats fed HAS. This study showed that HAS and CBM588 changed the metabolism of colonic microbiota and decreased the level of beta-glucuronidase activity, phenomena that may play a role in the inhibition of ACF formation in the rat colon.  相似文献   

8.
In vitro ruminal fermentation of organic acids common in forage.   总被引:7,自引:2,他引:5       下载免费PDF全文
Mixed rumen bacteria from cows fed either timothy hay or a 60% concentrate were incubated with 7.5 mM citrate, trans-aconitate, malate, malonate, quinate, and shikimate. Citrate, trans-aconitate, and malate were fermented at faster rates than malonate, quinate, and shikimate. Acetate was the primary fermentation product for all six acids. Quinate and shikimate fermentations gave rist to butyrate, whereas malate and malonate produced significant amounts of propionic acid. High-pressure liquid chromatography of fermentation products from trans-aconitate incubations revealed a compound that was subsequently identified as tricarballylate. As much as 40% of the trans-aconitate acid was converted to tricarballylate, and tricarballylate was fermented slowly. The slow rate of tricarballylate metabolism by mixed rumen bacteria and its potential as a magnesium chelator suggest that tricarballylate formation could be an important factor in the hypomagnesemia that leads to grass tetany.  相似文献   

9.
Short-chain fatty acids (SCFAs), primarily acetate, propionate, and butyrate, are metabolites formed by gut microbiota from complex dietary carbohydrates. Butyrate and acetate were reported to protect against diet-induced obesity without causing hypophagia, while propionate was shown to reduce food intake. However, the underlying mechanisms for these effects are unclear. It was suggested that SCFAs may regulate gut hormones via their endogenous receptors Free fatty acid receptors 2 (FFAR2) and 3 (FFAR3), but direct evidence is lacking. We examined the effects of SCFA administration in mice, and show that butyrate, propionate, and acetate all protected against diet-induced obesity and insulin resistance. Butyrate and propionate, but not acetate, induce gut hormones and reduce food intake. As FFAR3 is the common receptor activated by butyrate and propionate, we examined these effects in FFAR3-deficient mice. The effects of butyrate and propionate on body weight and food intake are independent of FFAR3. In addition, FFAR3 plays a minor role in butyrate stimulation of Glucagon-like peptide-1, and is not required for butyrate- and propionate-dependent induction of Glucose-dependent insulinotropic peptide. Finally, FFAR3-deficient mice show normal body weight and glucose homeostasis. Stimulation of gut hormones and food intake inhibition by butyrate and propionate may represent a novel mechanism by which gut microbiota regulates host metabolism. These effects are largely intact in FFAR3-deficient mice, indicating additional mediators are required for these beneficial effects.  相似文献   

10.
The human gut microbiota ferments dietary non‐digestible carbohydrates into short‐chain fatty acids (SCFA). These microbial products are utilized by the host and propionate and butyrate in particular exert a range of health‐promoting functions. Here an overview of the metabolic pathways utilized by gut microbes to produce these two SCFA from dietary carbohydrates and from amino acids resulting from protein breakdown is provided. This overview emphasizes the important role played by cross‐feeding of intermediary metabolites (in particular lactate, succinate and 1,2‐propanediol) between different gut bacteria. The ecophysiology, including growth requirements and responses to environmental factors, of major propionate and butyrate producing bacteria are discussed in relation to dietary modulation of these metabolites. A detailed understanding of SCFA metabolism by the gut microbiota is necessary to underpin effective strategies to optimize SCFA supply to the host.  相似文献   

11.
Abstract Batch culture incubations were used to investigate the effects of pH (6.8 or 5.5) and carbohydrate (starch) availability on dissimilatory aromatic amino acid metabolism in human fecal bacteria. During growth on peptide mixtures, tyrosine and phenylalanine fermentations occurred optimally at pH 6.8, while individual metabolic reactions were inhibited by up to 80% in the presence of 10 g l−1 starch. Tryptophan metabolites were not detected in these experiments. When free amino acids replaced peptides, phenol production was increased during carbohydrate fermentation, although formation of p-cresol, another tyrosine metabolite was strongly inhibited. Phenylpropionate, which is produced from phenylalanine, was unaffected by starch. Tryptophan was fermented in these studies, although indole production was reduced in the starch fermentors. The importance of different fermentation substrates (casein, peptide mixtures, free amino acids) on aromatic amino acid metabolism was investigated in incubations of material taken from the proximal bowel. The phenylalanine metabolites, phenylacetate and phenylpropionate, were the principal phenolic compounds formed from all three substrates. Phenol was the major tyrosine metabolite produced in casein and peptide fermentations, while hydroxyphenylpropionate was a more important tyrosine product from free amino acids. Indole was the sole product of tryptophan metabolism, but was formed only from the free amino acid. Bacterial metabolism of individual phenolic and indolic compounds was also investigated. Phenol, p-cresol, phenylacetate, phenylpropionate, 4-ethylphenol, indole, indoleacetate, and indolepropionate were not metabolized by colonic bacteria. However, hydroxyphenylacetate was hydrolyzed to p-cresol, while hydroxyphenylpropionate was transformed into phenylpropionate. Indolepyruvate was either converted to indoleacetate or metabolized into indole. Indolepropionate, and to a lesser degree indoleacetate were produced from indolelactate. These data show that human colonic anaerobes are able to extensively degrade either free or peptide-bound aromatic amino acids, with the concomitant formation of toxic metabolic products. These processes are controlled to a significant degree by environmental factors such as pH and carbohydrate availability, and this ultimately influences the types and amounts of fermentation products that can be formed in different regions of the large bowel. Received: 25 January 1996; Accepted: 8 May 1996  相似文献   

12.
Propionate is an important intermediate in the anaerobic degradation of complex organic matter to methane and carbon dioxide. The metabolism of propionate-forming and propionate-degrading bacteria is reviewed here. Propionate is formed during fermentation of polysaccharides, proteins and fats. The study of the fate of 13C-labelled compounds by nuclear magnetic resonance (NMR) spectroscopy has contributed together with other techniques to the present knowledge of the metabolic routes which lead to propionate formation from these substrates. Since propionate oxidation under methanogenic conditions is thermodynamically difficult, propionate often accumulates when the rates of its formation and degradation are unbalanced. Bacteria which are able to degrade propionate to the methanogenic substrates acetate and hydrogen can only perform this reaction when the methanogens consume acetate and hydrogen efficiently. As a consequence, propionate can only be degraded by obligatory syntrophic consortia of microorganisms. NMR techniques were used to study the degradation of propionate by defined and less defined cultures of these syntrophic consortia. Different types of side-reactions were reported, like the reductive carboxylation to butyrate and the reductive acetylation to higher fatty acids.  相似文献   

13.
Intestinal inflammation causes metabolic disorders. The purpose of this study was to determine the effect of dietary supplementation with lactosucrose (LS) on the serum metabolome and intestinal luminal content of fatty acids in colitic rats. Colitis was induced in rats using trinitrobenzene sulfonic acid. Subsequently, rats received intragastric administration of either 250 mg LS/kg body weight or saline (the control group) every day for 5 weeks. Short-chain fatty acids in the intestinal lumen, blood profile, and metabolites in serum were measured, respectively, using gas chromatography, biochemistry analyzer, and nuclear magnetic resonance-based metabolomics combined with multivariate statistics. Metabolic effects of LS included: (1) decreases in concentrations of branched-chain amino acids (isoleucine and valine), alanine, citric acid, trimethylamine oxide and taurine, and the abundance of aspartate aminotransferase in serum; (2) increases in concentrations of glucose metabolites (including succinate) in serum; and (3) altered concentrations of butyrate in the cecal content and of butyrate and acetate in the colon content. The results indicate that LS supplementation to colitic rats affects whole-body metabolism of amino acids and release of aspartate aminotransferase and alkaline phosphatase from tissues into the blood circulation, and enhances the production of short-chain fatty acids in the intestinal lumen.  相似文献   

14.
1. Free glutamic acid, aspartic acid, glutamic acid from glutamine and, in some instances, the glutamic acid from glutathione and the aspartic acid from N-acetyl-aspartic acid were isolated from the brains of sheep and assayed for radioactivity after intravenous injection of [2-14C]glucose, [1-14C]acetate, [1-14C]butyrate or [2-14C]propionate. These brain components were also isolated and analysed from rats that had been given [2-14C]propionate. The results indicate that, as in rat brain, glucose is by far the best precursor of the free amino acids of sheep brain. 2. Degradation of the glutamate of brain yielded labelling patterns consistent with the proposal that the major route of pyruvate metabolism in brain is via acetyl-CoA, and that the short-chain fatty acids enter the brain without prior metabolism by other tissue and are metabolized in brain via the tricarboxylic acid cycle. 3. When labelled glucose was used as a precursor, glutamate always had a higher specific activity than glutamine; when labelled fatty acids were used, the reverse was true. These findings add support and complexity to the concept of the metabolic `compartmentation' of the free amino acids of brain. 4. The results from experiments with labelled propionate strongly suggest that brain metabolizes propionate via succinate and that this metabolic route may be a limited but important source of dicarboxylic acids in the brain.  相似文献   

15.
The fate of representative fermentation products (acetate, propionate, butyrate, lactate, and ethanol) in hot spring cyanobacterial mats was investigated. The major fate during incubations in the light was photoassimilation by filamentous bacteria resembling Chloroflexus aurantiacus. Some metabolism of all compounds occurred under dark aerobic conditions. Under dark anaerobic conditions, only lactate was oxidized extensively to carbon dioxide. Extended preincubation under dark anaerobic conditions did not enhance anaerobic catabolism of acetate, propionate, or ethanol. Acetogenesis of butyrate was suggested by the hydrogen sensitivity of butyrate conversion to acetate and by the enrichment of butyrate-degrading acetogenic bacteria. Accumulation of fermentation products which were not catabolized under dark anaerobic conditions revealed their importance. Acetate and propionate were the major fermentation products which accumulated in samples collected at temperatures ranging from 50 to 70°C. Other organic acids and alcohols accumulated to a much lesser extent. Fermentation occurred mainly in the top 4 mm of the mat. Exposure to light decreased the accumulation of acetate and presumably of other fermentation products. The importance of interspecies hydrogen transfer was investigated by comparing fermentation product accumulation at a 65°C site, with naturally high hydrogen levels, and a 55°C site, where active methanogenesis prevented significant hydrogen accumulation. There was a greater relative accumulation of reduced products, notably ethanol, in the 65°C mat.  相似文献   

16.
1. Interactions in the rates of consumption of acetate, propionate and butyrate in sheep liver mitochondria were examined in the presence and absence of l-malate and alpha-oxoglutarate. 2. Acetate was not consumed in absence of ancillary substrate but utilization of acetate (7.2nmol/min per mg of protein) occurred in the presence of alpha-oxoglutarate. This consumption was abolished by propionate or butyrate but the presence of acetate did not affect consumption of propionate or butyrate. 3. Propionate consumption (10.1nmol/min per mg of protein) was unaffected by malate but was stimulated by 63% by butyrate or by 180% by alpha-oxoglutarate. 4. Butyrate consumption (3.3nmol/min per mg of protein) was stimulated by 117% by malate, by 151% by propionate and by 310% by alpha-oxoglutarate. 5. In the absence of ancillary substrates the maximum rate of total volatile fatty acid utilization (24.7nmol/min per mg of protein) occurred with a mixture of propionate and butyrate. When both propionate and butyrate were present total consumption was not affected by malate but was stimulated by 24% by alpha-oxoglutarate. With alpha-oxoglutarate present, propionate and butyrate each decreased the other's consumption by about 26%, but the total utilization was the greatest observed. 6. The inhibition of acetate consumption by propionate or butyrate is unexplained, but the remaining effects are consistent with an interaction of propionate and butyrate through oxaloacetate together with a general limitation imposed by a need for GTP to rephosphorylate AMP formed during activation of the volatile fatty acids.  相似文献   

17.
1. The total capacities of homogenates of bovine liver and rumen epithelium to activate acetate, propionate and butyrate were determined. 2. Activating capacities were assayed by measuring the rate of formation of the corresponding CoA esters. The methods used for determining the concentrations of the CoA esters allowed the CoA esters of acetate, propionate and butyrate to be distinguished. It was thus possible to investigate the effect of the presence of a second volatile fatty acid on the rate at which a given volatile fatty acid was activated. 3. The propionate-activating capacity in rumen epithelium was decreased by about 87% in the presence of butyrate, the acetate-activating capacity in liver was decreased by about 55% in the presence of either propionate or butyrate, and the butyrate-activating capacity in liver was decreased by about 40-50% in the presence of propionate. 4. All three activating capacities in liver appeared to be located in the mitochondrial matrix and membrane. The three activating capacities had similar locations to each other in rumen epithelium as well, although in this case activity was more evenly divided between the mitochondria and the cytoplasm. 5. The relative activating capacities towards the volatile fatty acids in the two tissues, together with the ability of one volatile fatty acid to inhibit the activation of another volatile fatty acid, appear to ensure that butyrate is mainly metabolized in the rumen epithelium and that propionate is metabolized in the liver.  相似文献   

18.
Viable counts of sulphate-reducing bacteria, able to use a range of different growth substrates were determined in sediments from two Sea Lochs (Etive and Eil) and an estuarine site (Tay), in Scotland. The composition of the sulphate-reducing bacterial population, in terms of substrate utilization, broadly corresponded to the in situ substrates for sulphate reduction and concentration of substrates at each site. Addition of acetate, lactate, propionate, butyrate, hydrogen and glutamate/serine (20 mM) to replicate slurries from each site resulted in stimulation of the corresponding population of sulphate-reducing bacteria and the in situ rates of sulphate reduction. The metabolism of the added substrates and changes in bacterial phospholipid fatty acids (PLFA) were quantified. With the exception of acetate and hydrogen, added substrates were incompletely oxidised, producing a mixture of further substrates, which predominantly were sequentially oxidised, and resulted in the stimulation of a mixed population of sulphate-reducing bacteria. There were significant changes in the PLFA of slurries with added substrate compared to controls. Acetate was completely removed at all sites and the small increase in even chain PLFA together with the absence of stimulation of any other biomarker, indicated that acetate was oxidised by sulphate-reducing bacteria distinctly different from those using other substrates. A biomarker for Desulfobacter, 10 Methyl 16:0, was not stimulated in any of the acetate slurries or in slurries where acetate was produced. Biomarkers for the propionate utilizing Desulfobulbus sp (17:1w6, 15:1w6) were always stimulated in propionate slurries and also in lactate slurries, where partial lactate fermentation produced propionate and acetate. In lactate and glutamate / serine slurries from the Tay estuary and lactate and hydrogen slurries from Loch Etive the biomarker for Desulfovibrio sp (i17:1w7) as well as those for Desulfobulbus were stimulated. This provides direct evidence for the significance of Desulfovibrio sp. within sediment slurries and demonstrates the competitive interaction between members of this genus and Desulfobulbus sp. for lactate, hydrogen and amino acid metabolism. At the estuarine site, sulphate reduction was limited at higher sulphate concentrations (about 3.5 mM) than the Sea Loch sites (<2 mM) and this had a significant effect on propionate and butyrate metabolism, as well as on methane production. These results demonstrate that although the sulphate-reducing bacterial population at each site could metabolise identical substrates, the types of sulphate-reducing bacteria involved and their sulphate thresholds were characteristically different.  相似文献   

19.
Two Pathways of Glutamate Fermentation by Anaerobic Bacteria   总被引:12,自引:6,他引:6  
Two pathways are involved in the fermentation of glutamate to acetate, butyrate, carbon dioxide, and ammonia-the methylaspartate and the hydroxyglutarate pathways which are used by Clostridium tetanomorphum and Peptococcus aerogenes, respectively. Although these pathways give rise to the same products, they are easily distinguished by different labeling patterns of the butyrate when [4-(14)C]glutamate is used as substrate. Schmidt degradation of the radioactive butyrate from C. tetanomorphum yielded equally labeled propionate and carbon dioxide, whereas nearly all the radioactivity of the butyrate from P. aerogenes was recovered in the corresponding propionate. This procedure was used as a test for the pathway of glutamate fermentation by 15 strains (9 species) of anaerobic bacteria. The labeling patterns of the butyrate indicate that glutamate is fermented via the methylaspartate pathway by C. tetani, C. cochlearium, and C. saccarobutyricum, and via the hydroxyglutarate pathway by Acidaminococcus fermentans, C. microsporum, Fusobacterium nucleatum, and F. fusiformis. Enzymes specific for each pathway were assayed in crude extracts of the above organisms. 3-Methylaspartase was found only in clostridia which use the methylaspartate pathway, including Clostridium SB4 and C. sticklandii, which probably degrade glutamate to acetate and carbon dioxide by using a second amino acid as hydrogen acceptor. High levels of 2-hydroxyglutarate dehydrogenase were found exclusively in organisms that use the hydroxyglutarate pathway. The data indicate that only two pathways are involved in the fermentation of glutamate by the bacteria analyzed. The methylaspartate pathway appears to be used only by species of Clostridium, whereas the hydroxyglutarate pathway is used by representatives of several genera.  相似文献   

20.
内蒙古自治区二连盆地、海拉尔盆地是我国重要的煤层气产区,其中生物成因煤层气是煤层气的重要来源,但复杂物质转化产甲烷相关微生物群落结构及功能尚不清楚。【目的】研究煤层水中的微生物代谢挥发性脂肪酸产甲烷的生理特征及群落特征。【方法】以内蒙古自治区二连盆地和海拉尔盆地的四口煤层气井水作为接种物,分别添加乙酸钠、丙酸钠和丁酸钠厌氧培养;定期监测挥发性脂肪酸降解过程中甲烷和底物的变化趋势,应用高通量测序技术,分析原始煤层气井水及稳定期产甲烷菌液的微生物群落结构。【结果】除海拉尔盆地H303煤层气井微生物不能代谢丙酸外,其他样品均具备代谢乙酸、丙酸和丁酸产生甲烷的能力,其生理生态参数存在显著差异,产甲烷延滞期依次是乙酸丁酸丙酸;最大比产甲烷速率和底物转化效率依次是丙酸乙酸丁酸。富集培养后,古菌群落结构与煤层气井水的来源显著相关,二连盆地优势古菌为氢营养型产甲烷古菌Methanocalculus (相对丰度13.5%–63.4%)和复合营养型产甲烷古菌Methanosarcina (7.9%–51.3%),海拉尔盆地的优势古菌为氢营养型产甲烷古菌Methanobacterium(24.3%–57.4%)和复合营养型产甲烷古菌Methanosarcina(29.6%–66.5%);细菌群落则与底物类型显著相关,硫酸盐还原菌Desulfovibrio(12.0%–41.0%)、互营丙酸氧化菌Syntrophobacter(39.6%–75.5%)和互营丁酸菌Syntrophomonas(8.5%–21.9%)分别在乙酸钠、丙酸钠和丁酸钠处理组显著富集。【结论】煤层气井水微生物可降解挥发性脂肪酸(乙酸、丙酸和丁酸)并具有产甲烷潜力;乙酸可能被古菌直接代谢产甲烷,而丙酸和丁酸通过互营细菌和产甲烷古菌代谢产甲烷。Desulfovibrio、Syntrophobacter和Syntrophomonas分别在乙酸、丙酸和丁酸代谢过程中发挥了重要作用。这些结果为煤层气生物强化开采提供了一定的微生物资源基础。  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