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.
Carbon isotope effects associated with aceticlastic methanogenesis   总被引:1,自引:0,他引:1  
The carbon isotope effects associated with synthesis of methane from acetate have been determined for Methanosarcina barkeri 227 and for methanogenic archaea in sediments of Wintergreen Lake, Michigan. At 37 degrees C, the 13C isotope effect for the reaction acetate (methyl carbon) --> methane, as measured in replicate experiments with M. barkeri, was - 21.3% +/- 0.3%. The isotope effect at the carboxyl portion of acetate was essentially equal, indicating participation of both positions in the rate-determining step, as expected for reactions catalyzed by carbon monoxide dehydrogenase. A similar isotope effect, - 19.2% +/- 0.3% was found for this reaction in the natural community (temperature = 20 degrees C). Given these observations, it has been possible to model the flow of carbon to methane within lake sediment communities and to account for carbon isotope compositions of evolving methane. Extension of the model allows interpretation of seasonal fluctuations in 13C contents of methane in other systems.  相似文献   

2.
Influence of corrinoid antagonists on methanogen metabolism.   总被引:11,自引:11,他引:11       下载免费PDF全文
Iodopropane inhibited cell growth and methane production when Methanobacterium thermoautotrophicum, Methanobacterium formicicum, and Methanosarcina barkeri were cultured on H2-CO2. Iodopropane (40 microM) inhibited methanogenesis (30%) and growth (80%) when M. barkeri was cultured mixotrophically on H2-CO2-methanol. The addition of acetate to the medium prevented the observed iodopropane-dependent inhibition of growth. The concentrations of iodopropane that caused 50% inhibition of growth of M. barkeri on either H2-CO2, H2-CO2-methanol, methanol, and acetate were 112 +/- 6, 24 +/- 2, 63 +/- 11, and 4 +/- 1 microM, respectively. Acetate prevented the iodopropane-dependent inhibition of one-carbon metabolism. Cultivation of M. barkeri on H2-CO2-methanol in bright light also inhibited growth and methanogenesis to a greater extent in the absence than in the presence of acetate in the medium. Acetate was the only organic compound examined that prevented iodopropane-dependent inhibition of one-carbon metabolism in M. barkeri. The effect of iodopropane and acetate on the metabolic fates of methanol and carbon dioxide was determined with 14C tracers when M. barkeri was grown mixotrophically on H2-CO2-methanol. The addition of iodopropane decreased the contribution of methanol to methane and cell carbon while increasing the contribution of CO2 to cell carbon. Regardless of iodopropane, acetate addition decreased the contribution of methanol and CO2 to cell carbon without decreasing their contribution to methane. The corrinoid antagonists, light and iodopropane, appeared most specific for methanogen metabolic reactions involved in acetate synthesis from one-carbon compounds and acetate catabolism.  相似文献   

3.
In the pyrimidine biosynthetic pathway, N-carbamyl-L-aspartate (CA-asp) is converted to L-dihydroorotate (DHO) by dihydroorotase (DHOase). The mechanism of this important reaction was probed using primary and secondary 15N and 13C isotope effects on the ring opening of DHO using isotope ratio mass spectrometry (IRMS). The reaction was performed at three different temperatures (25, 37, and 45 degrees C for hamster DHOase; 37, 50, and 60 degrees C for Bacillus caldolyticus), and the product CA-asp was purified for analysis. The primary and secondary kinetic isotope effects for the ring opening of the DHO were determined from analysis of the N and C of the carbamyl group after hydrolysis. In addition, the beta-carboxyl of the residual aspartate was liberated enzymatically by transamination to oxaloacetate with aspartate aminotransferase and then decarboxylation with oxaloacetate decarboxylase. The 13C/12C ratio from the released CO2 was determined by IRMS, yielding a second primary isotope effect. The primary and secondary isotope effects for the reaction catalyzed by DHOase showed little variation between enzymes or temperatures, the primary 13C and 15N isotope effects being approximately 1% on average, while the secondary 13C isotope effect is negligible or very slightly normal (>1.0000). These data indicate that the chemistry is at least partially rate-limiting while the secondary isotope effects suggest that the transition state may have lost some bending and torsional modes leading to a slight lessening of bond stiffness at the carbonyl carbon of the amide of CA-asp. The equilibrium isotope effects for DHO --> CA-asp have also been measured (secondary 13K(eq) = 1.0028 +/- 0.0002, primary 13K(eq) = 1.0053 +/- 0.0003, primary 15K(eq) = 1.0027 +/- 0.0003). Using these equilibrium isotope effects, the kinetic isotope effects for the physiological reaction (CA-asp --> DHO) have been calculated. These values indicate that the carbon of the amide group is more stiffly bonded in DHO while the slightly lesser, but still normal, values of the primary kinetic isotope effect show that the chemistry remains at least partially rate-limiting for the physiological reaction. It appears that the ring opening and closing is the slow step of the reaction.  相似文献   

4.
Methane formation and methane oxidation by methanogenic bacteria.   总被引:25,自引:1,他引:24       下载免费PDF全文
Methanogenic bacteria were found to form and oxidize methane at the same time. As compared to the quantity of methane formed, the amount of methane simultaneously oxidized varied between 0.3 and 0.001%, depending on the strain used. All the nine tested strains of methane producers (Methanobacterium ruminantium, Methanobacterium strain M.o.H., M. formicicum, M. thermoautotrophicum, M. arbophilicum, Methanobacterium strain AZ, Methanosarcina barkeri, Methanospirillum hungatii, and the "acetate organism") reoxidized methane to carbon dioxide. In addition, they assimilated a small part of the methane supplied into cell material. Methanol and acetate also occurred as oxidation products in M. barkeri cultures. Acetate was also formed by the "acetate organism," a methane bacterium unable to use methanogenic substrates other than acetate. Methane was the precursor of the methyl group of the acetate synthesized in the course of methane oxidation. Methane formation and its oxidation were inhibited equally by 2-bromoethanesulfonic acid. Short-term labeling experiments with M. thermoautotrophicum and M. hungatii clearly suggest that the pathway of methane oxidation is not identical with a simple back reaction of the methane formation process.  相似文献   

5.
A method has been developed for the positional 13C isotope analysis of pyruvate and acetate by stepwise quantitative degradation. On its base, the kinetic isotope effects on the pyruvate dehydrogenase reaction (enzymes from Escherichia coli and Saccharomyces cerevisiae) for both of the carbon atoms involved in the bond scission (double isotope effect determination) and on C-3 of pyruvate have been determined. The experimental k12/k13 values with the enzyme from E. coli on C-1 and C-2 of pyruvate are 1.0093 +/- 0.0007 and 1.0213 +/- 0.0017, respectively, and, with the enzyme from S. cerevisiae, the values are 1.0238 +/- 0.0013 and 1.0254 +/- 0.0016, respectively. A secondary isotope effect of 1.0031 +/- 0.0009 on C-3 (CH3-group) was found with both enzymes. The size of the isotope on C-1 indicates that decarboxylation is more rate-determining with the yeast enzyme than with the enzyme from E. coli, although it is not the entirely rate-limiting step in the overall reaction sequence. Assuming appropriate values for the intrinsic isotope effect on the decarboxylation step (k3) and the equilibrium isotope effect on the reversible substrate binding (k1, k2), one can calculate values for the partitioning factor R (k3/k2: E. coli enzyme 4.67, S. cerevisiae enzyme 1.14) and the intrinsic isotope effects related to the carbonyl-C (k1/k'1 = 1.019; k3/k'3 = 1.033). The isotope fractionation at C-2 of pyruvate gives strong evidence that the well known relative carbon-13 depletion in lipids from biological material is mainly caused by the isotope effect on the pyruvate dehydrogenase reaction. In addition, our results indicate an alternating 13C abundance in fatty acids, that has already been verified in some cases.  相似文献   

6.
Kinetics of methanogenesis from acetate was studied using pure cultures of Methanosarcina barkeri and Methanosarcina mazei. Methane formation was found to be associated with cell growth. Nearly equimolar methane was produced from acetate during the methanogenic growth, and about 1.94 g of cells were formed from each mole of acetate consumed. Cell growth can be estimated from methane production. Significant substrate inhibition was found when acetate concentration was higher than 0.12 M. Among the three methanogenic strains studied, M. mazei strain S6 had the highest specific growth rate at all acetate concentrations studied and was least sensitive to environmental factors investigated (e.g., acetate concentration). The maximum specific growth rate found for strain S6 was 0.022 hr(-1) at acetic acid concentration around 7 g/L. The other two strains studied were M. barkeri strain 227 and strain MS. Growth of M. barkeri was completely inhibited at sodium acetate concentrations higher than 0.24 M. The maximum specific growth rate found for strains 227 and MS was 0.019 and 0.021 h(-1) at acetic acid concentrations of 3.6 and 6.8 g/L, respectively. A kinetic model with substrate inhibition was developed and can be used to simulate the methane formation from M. mazei strain S6 grown on acetate at 35 degrees C, pH 7.  相似文献   

7.
Chloramphenicol produced by cultures of Streptomyces species 3022a supplemented with sodium [1,2-13C]acetate was labelled with 13C exclusively in the dichloromethine (2.6 +/- 0.1%) and carbonyl (0.59 +/- 0.05% carbon atoms. Satellite signals from 13C-13C coupling between covalently bonded 13C-enriched carbon atoms were too intense to be attributed to random combination of labelled atoms at the average enrichments measured, but their intensity relative to those of the signals for uncoupled 13C atoms indicated that most of the precursor had been incorporated after 13C-13C bond fission. Since [2,3-13c]succinic acid enriched only the carbonyl carbon atom of chloramphenicol, these results suggest that neither acetate nor a Krebs cycle intermediate is a direct precursor of the dichloroacetyl group. Cultures supplemented with [2-3h]-or [2h2]-dichloroacetic acid incorporated negligible amounts of isotope into the antibiotic; on this evidence, the free acid is not an intermediate in chloramphenicol biosynthesis and the acylation step may precede chlorination.  相似文献   

8.
Although aldolase-catalyzed condensations proceed by stepwise mechanisms via the intermediacy of nucleophilic enol(ate)s or enamines, the mechanisms of those enzymes that catalyze Claisen-type condensations are unclear. The reaction pathway followed by an enzyme from this second group, malate synthase, has been studied by the double-isotope fractionation method to determine whether the reaction is stepwise or concerted. In agreement with earlier work, a deuterium kinetic isotope effect D(V/K) of 1.3 +/- 0.1 has been found when [2H3]acetyl-CoA is the substrate. The 13C isotope effect at the aldehydic carbon of glyoxylate has also been measured. For this determination, the malate product (containing the carbon of interest at C-2) was quantitatively transformed into a new sample of malate having the carbon of interest at C-4. This material was decarboxylated by malic enzyme to produce the appropriate CO2 for isotope ratio mass spectrometric analysis. The 13C isotope effect with [1H3]acetyl-CoA [that is, 13(V/K)H] is 1.0037 +/- 0.0004. By use of the known values of the intermolecular and intramolecular deuterium effects and of 13(V/K)H, the value of the 13C isotope effect when deuteriated [2H3]acetyl-CoA is the substrate [that is, 13(V/K)D] can be predicted for three possible mechanisms. If 13(V/K)H is a kinetic isotope effect and the reaction is concerted, the value of the 13C effect on deuteriation of acetyl-CoA will rise to 1.011; if 13(V/K)H is a kinetic isotope effect and the reaction is stepwise, the value of the 13C effect will fall to 1.0025; and if the 13C effect is an equilibrium isotope effect deriving from glyoxylate dehydration, the reaction is necessarily stepwise, and the value of 13(V/K)D will be 1.0037, unchanged from that of 13(V/K)H. Experimentally, the value of 13(V/K)D is 1.0037 +/- 0.0007, which requires that malate synthase follow a stepwise path. It is therefore clear that the two salient characteristics of enzymes that catalyze Claisen-like condensations, namely, the absence of enzyme-catalyzed proton exchange with solvent and the inversion of the configuration at the nucleophilic center, which had been suggestive of a concerted pathway, are not mechanistically diagnostic.  相似文献   

9.
Cell suspensions of Methanosarcina barkeri, grown on acetate, catalyzed the conversion of carbon monoxide and H2O to CO2 and H2 in stoichiometric amounts when methane formation was inhibited by bromoethanesulfonate. The specific activity was 80-120 nmol min-1 mg protein-1 at 5% CO in the gas phase. CO oxidation was coupled with the phosphorylation of ADP as indicated by a rapid increase of the intracellular ATP level upon start of the reaction. At least 0.1 mol ATP was formed/mol CO consumed. The onset of CO oxidation was also accompanied by an increase of the proton motive force (delta p) from 100 mV to 150 mV (inside negative). Addition of the uncoupler tetrachlorosalicylanilide to CO-metabolizing cells led to a rapid decrease of the ATP level and of delta p, and to an increase of the CO oxidation rate up to 70%. In the presence of the proton-translocating ATPase inhibitor N,N'-dicyclohexylcarbodiimide the phosphorylation of ADP was inhibited and CO oxidation slowed down, whereas delta p was almost unaffected. Inhibition of CO oxidation under these conditions was relieved by the addition of the protonophore tetrachlorosalicylanilide. The results indicate that in acetate-grown M. barkeri the free-energy change associated with the formation of CO2 and H2 from CO and H2O (delta G degrees = -20 kJ/mol) can be used to drive the phosphorylation of ADP and that the coupling proceeds via a chemiosmotic mechanism. A possible role of the carbon monoxide oxidation reaction as an energy-conserving site in acetate fermentation to CH4 and CO2 is discussed.  相似文献   

10.
The pathway of acetate catabolism in Methanosarcina barkeri strain MS was studied by using a recently developed assay for methanogenesis from acetate by soluble enzymes in cell extracts. Extracts incubated with [2-14C]acetate, hydrogen, and ATP formed 14CH4 and [14C]methyl coenzyme M as products. The apparent Km for acetate conversion to methane was 5 mM. In the presence of excess acetate, both the rate and duration of methane production was dependent on ATP. Acetyl phosphate replaced the cell extract methanogenic requirement for both acetate and ATP (the Km for ATP was 2 mM). Low concentrations of bromoethanesulfonic acid and cyanide, inhibitors of methylreductase and carbon monoxide dehydrogenase, respectively, greatly reduced the rate of methanogenesis. Precipitation of CO dehydrogenase in cell extracts by antibodies raised to 95% purified enzyme inhibited both CO dehydrogenase and acetate-to-methane conversion activity. The data are consistent with a model of acetate catabolism in which methylreductase, methyl coenzyme M, CO dehydrogenase, and acetate-activating enzymes are components. These results are discussed in relation to acetate uptake and rate-limiting transformation mechanisms in methane formation.  相似文献   

11.
We compared the metabolism of methanol and acetate when Methanosarcina barkeri was grown in the presence and absence of Desulfovibrio vulgaris. The sulfate reducer was not able to utilize methanol or acetate as the electron donor for energy metabolism in pure culture, but was able to grow in coculture. Pure cultures of M. barkeri produced up to 10 mumol of H(2) per liter in the culture headspace during growth on acetate or methanol. In coculture with D. vulgaris, the gaseous H(2) concentration was 相似文献   

12.
During alcohol fermentation, the carbon isotope composition of ethyl alcohol produced depended on the substrate used and was characterized by the value of delta 13C equal to -24.7 +/- 0.8/1000 (wheat grain), -22 +/- 0.1/1000 (rye grain), -22 +/- 0.5/1000 (products of wood hydrolysis), -15.3 +/- 0.3/1000 (maize grain) and -10 +/- 0.1/1000 (sugar cane). The isotope composition of carbon of ethyl alcohol obtained during catalytic hydroxylation of ethylene has a delta 13C of -30.6 +/- 0.3/1000. The possibility of quantitative determination of specific components in mixtures of ethanol samples with various isotope compositions (chemical synthesis and alcohol fermentation of raw material from C3- or C4-plants) was shown.  相似文献   

13.
Acetate assimilation pathway of Methanosarcina barkeri.   总被引:18,自引:11,他引:7       下载免费PDF全文
The pathway of acetate assimilation in Methanosarcina barkeri was determined from analysis of the position of label in alanine, aspartate, and glutamate formed in cells grown in the presence of [14C]acetate and by measurement of enzyme activities in cell extracts. The specific radioactivity of glutamate from cells grown on [1-14C]- or [2-14C]acetate was approximately twice that of aspartate. The methyl and carboxyl carbons of acetate were incorporated into aspartate and glutamate to similar extents. Degradation studies revealed that acetate was not significantly incorporated into the C1 of alanine, C1 or C4 of aspartate, or C1 of glutamate. The C5 of glutamate, however, was partially derived from the carboxyl carbon of acetate. Cell extracts were found to contain the following enzyme activities, in nanomoles per minute per milligram of protein at 37 degrees C: F420-linked pyruvate synthase, 170; citrate synthase, 0.7; aconitase, 55; oxidized nicotinamide adenine dinucleotide phosphate-linked isocitrate dehydrogenase, 75; and oxidized nicotinamide adenine dinucleotide-linked malate dehydrogenase, 76. The results indicate that M. barkeri assimilates acetate into alanine and aspartate via pyruvate and oxaloacetate and into glutamate via citrate, isocitrate, and alpha-ketoglutarate. The data reveal differences in the metabolism of M. barkeri and Methanobacterium thermoautotrophicum and similarities in the assimilation of acetate between M. barkeri and other anaerobic bacteria, such as Clostridium kluyveri.  相似文献   

14.
Methane formation from acetate by resting cells of Methanosarcina barkeri was accompanied by an increase in the intracellular ATP content from 0.9 to 4.0 nmol/mg of protein. Correspondingly, the proton motive force increased to a steady-state level of -120 mV. The transmembrane pH gradient however, was reversed under these conditions and amounted to +20 mV. The addition of the protonophore 3,5,3',4'-tetrachlorosalicylanilide led to a drastic decrease in the proton motive force and in the intracellular ATP content and to an inhibition of methane formation. The ATPase inhibitor N,N'-dicyclohexylcarbodiimide stopped methanogenesis, and the intracellular ATP content decreased. The proton motive force decreased also under these conditions, indicating that the proton motive force could not be generated from acetate without ATP. The overall process of methane formation from acetate was dependent on the presence of sodium ions; upon addition of acetate to cell suspensions of M. barkeri, a transmembrane Na+ gradient in the range of 4:1 (Na+ out/Na+ in) was established. Possible sites of involvement of the Na+ gradient in the conversion of acetate to methane and carbon dioxide are discussed. Na+ is not involved in the CO dehydrogenase reaction.  相似文献   

15.
C A Roeske  M H O'Leary 《Biochemistry》1985,24(7):1603-1607
The carbon isotope effect at CO2 has been measured in the carboxylation of ribulose 1,5-bisphosphate by the ribulosebisphosphate carboxylase from Rhodospirillum rubrum. The isotope effect is obtained by comparing the isotopic composition of carbon 1 of the 3-phosphoglyceric acid formed in the reaction with that of the carbon dioxide source. A correction is made for carbon 1 of 3-phosphoglyceric acid which arises from carbon 3 of the starting ribulose bisphosphate. The isotope effect is k12/k13 = 1.0178 +/- 0.0008 at 25 degrees C, pH 7.8. This value is smaller than the corresponding value for the spinach enzyme. It appears that substrate addition with the R. rubrum enzyme is principally ordered, with ribulose bisphosphate binding first, whereas substrate addition is random with the spinach enzyme. The carboxylation step is partially rate limiting with both enzymes.  相似文献   

16.
Yang J  Schenkman S  Horenstein BA 《Biochemistry》2000,39(19):5902-5910
Trypanosoma cruzi trans-sialidase catalyzes a novel reaction that involves the transfer of sialic acid between host and parasite glycoconjugates. In this paper, we report kinetic isotope effect studies on recombinant trans-sialidase. beta-Dideuterium and primary 13C isotope effects were measured for a good substrate, sialyl-lactose, and a slow substrate, sialyl-galactose, in both acid-catalyzed solvolysis and enzymatic transfer reactions. The beta-dideuterium isotope effect for sialyl-lactose in the acid hydrolysis reaction was 1.113 +/- 0.012. The primary 13C isotope effects for hydrolysis of sialyl-lactose and sialyl-galactose were 1. 016 +/- 0.011 and 1.015 +/- 0.008, respectively. In the enzymatic transfer reactions, the beta-dideuterium and primary 13C effects for sialyl-galactose were 1.060 +/- 0.008 and 1.032 +/- 0.008, respectively. The isotope effects for hydrolysis describe a dissociative SN1-like mechanism, and these data are contrasted by the data for the enzyme-catalyzed reaction. The enzymatic deuterium isotope effects are lower by a factor of 2, but the primary carbon isotope effects are higher by a factor of 2. This pattern describes a mechanism involving nucleophilic participation in the rate-determining transition state.  相似文献   

17.
A study of anaerobic sediments below cyanobacterial mats of a low-salinity meltwater pond called Orange Pond on the McMurdo Ice Shelf at temperatures simulating those in the summer season (<5 degrees C) revealed that both sulfate reduction and methane production were important terminal anaerobic processes. Addition of [2-(14)C]acetate to sediment samples resulted in the passage of label mainly to CO(2). Acetate addition (0 to 27 mM) had little effect on methanogenesis (a 1.1-fold increase), and while the rate of acetate dissimilation was greater than the rate of methane production (6.4 nmol cm(-3) h(-1) compared to 2.5 to 6 nmol cm(-3) h(-1)), the portion of methane production attributed to acetate cleavage was <2%. Substantial increases in the methane production rate were observed with H(2) (2.4-fold), and H(2) uptake was totally accounted for by methane production under physiological conditions. Formate also stimulated methane production (twofold), presumably through H(2) release mediated through hydrogen lyase. Addition of sulfate up to 50-fold the natural levels in the sediment (interstitial concentration, approximately 0.3 mM) did not substantially inhibit methanogenesis, but the process was inhibited by 50-fold chloride (36 mM). No net rate of methane oxidation was observed when sediments were incubated anaerobically, and denitrification rates were substantially lower than rates for sulfate reduction and methanogenesis. The results indicate that carbon flow from acetate is coupled mainly to sulfate reduction and that methane is largely generated from H(2) and CO(2) where chloride, but not sulfate, has a modulating role. Rates of methanogenesis at in situ temperatures were four- to fivefold less than maximal rates found at 20 degrees C.  相似文献   

18.
Summary Organic waste is converted in a two-stage process to methane and carbon dioxide by mixed cultures of microorganisms. Acetate, a product of acidogenic and acetogenic bacteria and the main substrate for methanogenic bacteria, is an important intermediate of the anaerobic degradation process, which results in the generation of methane. It was shown by labelling experiments using (U-14C) acetate that as much as 65%–96% of the total methane produced came from the acetate. The first order utilization rate for acetate in the methanogenic stages of a two-stage digestion process was between 0.17 h-1 and 0.5 h-1. The kinetics as well as the mass flow and yields of acetate and the methyl group of acetate were determined by pulse-labelling experiments with (U-14C) acetate and (2-14C) acetate without a significant rise of the total concentrations. Up to 58% of the acetate carbon was transformed to methane, and about 30% to carbon dioxide; only 4%–15% was incorporated into the biomass. There are at least two parallel degradation mechanisms in the metabolic transformation of acetate to methane: acetate is cleaved either to form methane and carbon dioxide or to form hydrogen and carbon dioxide, which can be transformed by an additional reaction to methane. Labelling experiments with (2-14C) acetate show that both mechanisms took place at similar order.  相似文献   

19.
Carbon isotope fractionation during aerobic mineralization of 1, 2-dichloroethane (1,2-DCA) by Xanthobacter autotrophicus GJ10 was investigated. A strong enrichment of (13)C in residual 1,2-DCA was observed, with a mean fractionation factor alpha +/- standard deviation of 0.968 +/- 0.0013 to 0.973 +/- 0.0015. In addition, a large carbon isotope fractionation between biomass and inorganic carbon occurred. A mechanistic model that links the fractionation factor alpha to the rate constants of the first catabolic enzyme was developed. Based on the model, it was concluded that the strong enrichment of (13)C in 1,2-DCA arises because the first irreversible step of the initial enzymatic transformation of 1,2-DCA consists of an S(N)2 nucleophilic substitution. S(N)2 reactions are accompanied by a large kinetic isotope effect. The substantial carbon isotope fractionation between biomass and inorganic carbon could be explained by the kinetic isotope effect associated with the initial 1,2-DCA transformation and by the metabolic pathway of 1,2-DCA degradation. Carbon isotope fractionation during 1,2-DCA mineralization leads to 1,2-DCA, inorganic carbon, and biomass with characteristic carbon isotope compositions, which may be used to trace the process in contaminated environments.  相似文献   

20.
In abandoned coal mines, methanogenic archaea are responsible for the production of substantial amounts of methane. The present study aimed to directly unravel the active methanogens mediating methane release as well as active bacteria potentially involved in the trophic network. Therefore, the stable-isotope-labeled precursors of methane, [(13)C]acetate and H(2)-(13)CO(2), were fed to liquid cultures from hard coal and mine timber from a coal mine in Germany. Guided by methane production rates, samples for DNA stable-isotope probing (SIP) with subsequent quantitative PCR and denaturing gradient gel electrophoretic (DGGE) analyses were taken over 6 months. Surprisingly, the formation of [(13)C]methane was linked to acetoclastic methanogenesis in both the [(13)C]acetate- and the H(2)-(13)CO(2)-amended cultures of coal and timber. H(2)-(13)CO(2) was used mainly by acetogens related to Pelobacter acetylenicus and Clostridium species. Active methanogens, closely affiliated with Methanosarcina barkeri, utilized the readily available acetate rather than the thermodynamically more favorable hydrogen. Thus, the methanogenic microbial community appears to be highly adapted to the low-H(2) conditions found in coal mines.  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