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 transfer of the methyl group of acetate to coenzyme M (2-mercaptoethanesulfonic acid; HS-CoM) during the metabolism of acetate to methane was investigated in cultures of Methanosarcina strain TM-1. The organism metabolized CD3COO- to 83% CD3H and 17% CD2H2 and produced no CDH3 or CH4. The isotopic composition of coenzyme M in cells grown on CD3COO- was analyzed with a novel gas chromatography-mass spectrometry technique. The cells contained CD3-D-CoM and CD2H-S-CoM) in a proportion similar to that of CD3H to CD2H2. These results, in conjunction with a report (J.K. Nelson and J.G. Ferry, J. Bacteriol. 160:526-532, 1984) that extracts of acetate-grown strain TM-1 contain high levels of CH3-S-CoM methylreductase, indicate that CH3-S-CoM is an intermediate in the metabolism of acetate to methane in this organism.  相似文献   

2.
Cell extracts from acetate-grown Methanosarcina thermophila contained CO-oxidizing:H2-evolving activity 16-fold greater than extracts from methanol-grown cells. Following fractionation of cell extracts into soluble and membrane components, CO-dependent H2 evolution and CO-dependent methyl-coenzyme M methylreductase activities were only present in the soluble fraction, but addition of the membrane fraction enhanced both activities. A b-type cytochrome(s), present in the membrane fraction, was linked to a membrane-bound hydrogenase. CO-oxidizing:H2-evolving activity was reconstituted with: (i) CO dehydrogenase complex, (ii) a ferredoxin, and (iii) purified membranes with associated hydrogenase. The ferredoxin was a direct electron acceptor for the CO dehydrogenase complex. The ferredoxin also coupled CO oxidation by CO dehydrogenase complex to metronidazole reduction.  相似文献   

3.
Methanosarcina sp. strain TM-1 and Methanosarcina acetivorans produced and consumed H(2) to maintain H(2) partial pressures of 16 to 92 Pa in closed cultures during growth on acetate. Strain TM-1 produced H(2) continuously when H(2) was continuously removed from the culture. The potential physiological significance of H(2) in acetate metabolism to methane is discussed.  相似文献   

4.
Cell lysates of acetate-grown Methanosarcina barkeri 227 were found to cleave acetate to CH4 and CO2. The aceticlastic reaction was identified by using radioactive methyl-labeled acetate. Cell lysates decarboxylated acetate in a nitrogen atmosphere, conserving the methyl group in methane. The rate of methanogenesis from acetate in the cell lysates was comparable to that observed with whole cells. Aceticlastic activity was found in the particulate fraction seperate from methylcoenzyme M methylreductase activity, which occurs in the soluble fraction. Pronase treatment eliminated methylcoenzyme M methylreductase activity in lysates and stimulated aceticlastic activity, indicating the aceticlastic activity was not derived from unbroken cells, which are unaffected by proteolytic treatment.  相似文献   

5.
The cell extract protein content of acetate- and methanol-grown Methanosarcina thermophila TM-1 was examined by two-dimensional polyacrylamide gel electrophoresis. More than 100 mutually exclusive spots were present in acetate- and methanol-grown cells. Spots corresponding to acetate kinase, phosphotransacetylase, and the five subunits of the carbon monoxide dehydrogenase complex were identified in acetate-grown cells. Activities of formylmethanofuran dehydrogenase, formylmethanofuran:tetrahydromethanopterin formyltransferase, 5,10-methenyltetrahydromethanopterin cyclohydrolase, methylene tetrahydromethanopterin:coenzyme F420 oxidoreductase, formate dehydrogenase, and carbonic anhydrase were examined in acetate- and methanol-grown Methanosarcina thermophila. Levels of formyltransferase in either acetate- or methanol-grown Methanosarcina thermophila were approximately half the levels detected in H2-CO2-grown Methanobacterium thermoautotrophicum. All other enzyme activities were significantly lower in acetate- and methanol-grown Methanosarcina thermophila.  相似文献   

6.
During the methanogenic fermentation of acetate by Methanosarcina thermophila, the CO dehydrogenase complex cleaves acetyl coenzyme A and oxidizes the carbonyl group (or CO) to CO2, followed by electron transfer to coenzyme M (CoM)-S-S-coenzyme B (CoB) and reduction of this heterodisulfide to HS-CoM and HS-CoB (A. P. Clements, R. H. White, and J. G. Ferry, Arch. Microbiol. 159:296-300, 1993). The majority of heterodisulfide reductase activity was present in the soluble protein fraction after French pressure cell lysis. A CO:CoM-S-S-CoB oxidoreductase system from acetate-grown cells was reconstituted with purified CO dehydrogenase enzyme complex, ferredoxin, membranes, and partially purified heterodisulfide reductase. Coenzyme F420 (F420) was not required, and CO:F420 oxidoreductase activity was not detected in cell extracts. The membranes contained cytochrome b that was reduced with CO and oxidized with CoM-S-S-CoB. The results suggest that a novel CoM-S-S-CoB reducing system operates during acetate conversion to CH4 and CO2. In this system, ferredoxin transfers electrons from the CO dehydrogenase complex to membrane-bound electron carriers, including cytochrome b, that are required for electron transfer to the heterodisulfide reductase. The cytochrome b was purified from solubilized membrane proteins in a complex with six other polypeptides. The cytochrome was not reduced when the complex was incubated with H2 or CO, and H2 uptake hydrogenase activity was not detected; however, the addition of CO dehydrogenase enzyme complex and ferredoxin enabled the CO-dependent reduction of cytochrome b.  相似文献   

7.
Biochemistry of methanogenesis.   总被引:9,自引:0,他引:9  
Methane is a product of the energy-yielding pathways of the largest and most phylogenetically diverse group in the Archaea. These organisms have evolved three pathways that entail a novel and remarkable biochemistry. All of the pathways have in common a reduction of the methyl group of methyl-coenzyme M (CH3-S-CoM) to CH4. Seminal studies on the CO2-reduction pathway have revealed new cofactors and enzymes that catalyze the reduction of CO2 to the methyl level (CH3-S-CoM) with electrons from H2 or formate. Most of the methane produced in nature originates from the methyl group of acetate. CO dehydrogenase is a key enzyme catalyzing the decarbonylation of acetyl-CoA; the resulting methyl group is transferred to CH3-S-CoM, followed by reduction to methane using electrons derived from oxidation of the carbonyl group to CO2 by the CO dehydrogenase. Some organisms transfer the methyl group of methanol and methylamines to CH3-S-CoM; electrons for reduction of CH3-S-CoM to CH4 are provided by the oxidation of methyl groups to CO2.  相似文献   

8.
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   

9.
Methanogenic archaea conserve energy for growth by reducing some one- and two-carbon compounds to methane and concomitantly generating an ion motive force. Growth of Methanosarcina acetivorans on carbon monoxide (CO) is peculiar as it involves formation of, besides methane, formate, acetate and methylated thiols. It has been argued that methane formation is partially inhibited under carboxidotrophic conditions and that the other products result from either detoxification of CO or from bypassing methanogenesis with other pathways for energy conservation. To gain a deeper understanding of the CO-dependent physiology of M. acetivorans we analyzed metabolite formation in resting cells. The initial rates of methane, acetate, formate, and dimethylsulfide formation increased differentially with increasing CO concentrations but were maximal already at the same moderate CO partial pressure. Strikingly, further increase of the amount of CO was not inhibitory. The maximal rate of methane formation from CO was approximately fivefold lower than that from methanol, consistent with the previously observed significant downregulation of the energy converting sodium-dependent methyltransferase. The rate of dimethylsulfide formation from CO was only 1–2% of that of methane formation under any conditions tested. Implications of the data presented for previously proposed pathways of CO utilization are discussed.  相似文献   

10.
Methanosarcina barkeri strain 227 maintained on an acetate medium for 2 years was found to possess hydrogenase, methylcoenzyme M methylreductase, coenzyme F420, and coenzyme M. The levels of these constituents in acetate-grown cells were similar to those found in cells of the same strain grown on methanol or hydrogen and carbon dioxide.  相似文献   

11.
Methanosarcina barkeri strain 227 maintained on an acetate medium for 2 years was found to possess hydrogenase, methylcoenzyme M methylreductase, coenzyme F420, and coenzyme M. The levels of these constituents in acetate-grown cells were similar to those found in cells of the same strain grown on methanol or hydrogen and carbon dioxide.  相似文献   

12.
Despite its toxicity for the majority of living matter on our planet, numerous microorganisms, both aerobic and anaerobic, can use carbon monoxide (CO) as a source of carbon and/or energy for growth. The capacity to employ carboxidotrophic energy metabolism anaerobically is found in phylogenetically diverse members of the Bacteria and the Archaea. The oxidation of CO is coupled to numerous respiratory processes, such as desulfurication, hydrogenogenesis, acetogenesis, and methanogenesis. Although as diverse as the organisms capable of it, any CO-dependent energy metabolism known depends on the presence of carbon monoxide dehydrogenase. This review summarizes recent insights into the CO-dependent physiology of anaerobic microorganisms with a focus on methanogenic archaea. Carboxidotrophic growth of Methanosarcina acetivorans, thought to strictly rely on the process of methanogenesis, also involves formation of methylated thiols, formate, and even acetogenesis, and, thus, exemplifies how the beneficial redox properties of CO can be exploited in unexpected ways by anaerobic microorganisms.  相似文献   

13.
A thermophilic strain of Methanosarcina, designated Methanosarcina strain TM-1, was isolated from a laboratory-scale 55 degrees C anaerobic sludge digestor by the Hungate roll-tube technique. Penicillin and d-cycloserine, inhibitors of peptidoglycan synthesis, were used as selective agents to eliminate contaminating non-methanogens. Methanosarcina strain TM-1 had a temperature optimum for methanogenesis near 50 degrees C and grew at 55 degrees C but not at 60 degrees C. Substrates used for methanogenesis and growth by Methanosarcina strain TM-1 were acetate (12-h doubling time), methanol (7- to 10-h doubling time), methanol-acetate mixtures (5-h doubling time), methylamine, and trimethylamine. When radioactively labeled acetate was the sole methanogenic substrate added to the growth medium, it was predominantly split to methane and carbon dioxide. When methanol was also present in the medium, the metabolism of acetate shifted to its oxidation and incorporation into cell material. Electrons derived from acetate oxidation apparently were used to reduce methanol. H(2)-CO(2) was not used for growth and methanogenesis by Methanosarcina strain TM-1. When presented with both H(2)-CO(2) and methanol, Methanosarcina strain TM-1 was capable of limited hydrogen metabolism during growth on methanol, but hydrogen metabolism ceased once the methanol was depleted. Methanosarcina strain TM-1 required a growth factor (or growth factors) present in the supernatant of anaerobic digestor sludge. Growth factor requirements and the inability to use H(2)-CO(2) are characteristics not found in other described Methanosarcina strains. The high numbers of Methanosarcina-like clumps in sludges from thermophilic digestors and the fast generation times reported here for Methanosarcina TM-1 indicate that Methanosarcina may play an important role in thermophilic methanogenesis.  相似文献   

14.
Fast protein liquid chromatography of cell extract from methanol- or acetate-grown Methanosarcina thermophila resolved two peaks of CO dehydrogenase activity. The activity of one of the CO dehydrogenases was sixfold greater in acetate-grown compared with methanol-grown cells. This CO dehydrogenase was purified to apparent homogeneity (70 mumol of methyl viologen reduced per min per mg of protein) and made up greater than 10% of the cellular protein of acetate-grown cells. The native enzyme (Mr 250,000) formed aggregates with an Mr of approximately 1,000,000. The enzyme contained five subunits (Mrs 89,000, 71,000, 60,000, 58,000, and 19,000), suggesting a multifunctional enzyme complex. Nickel, iron, cobalt, zinc, inorganic sulfide, and a corrinoid were present in the complex. The UV-visible spectrum suggested the presence of iron-sulfur centers. The electron paramagnetic resonance spectrum contained g values of 2.073, 2.049, and 2.028; these features were broadened in enzyme that was purified from cells grown in the presence of medium enriched with 61Ni, indicating the involvement of this metal in the spectrum. The pattern of potassium cyanide inhibition indicated that cyanide binds at or near the CO binding site. The properties of the enzyme imply an involvement in the dissimilation of acetate to methane, possibly by cleavage of acetate or activated acetate.  相似文献   

15.
When grown in the absence of added sulfate, cocultures of Desulfovibrio desulfuricans or Desulfovibrio vulgaris with Methanobrevibacter smithii (Methanobacterium ruminantium), which uses H(2) and CO(2) for methanogenesis, degraded lactate, with the production of acetate and CH(4). When D. desulfuricans or D. vulgaris was grown in the absence of added sulfate in coculture with Methanosarcina barkeri (type strain), which uses both H(2)-CO(2) and acetate for methanogenesis, lactate was stoichiometrically degraded to CH(4) and presumably to CO(2). During the first 12 days of incubation of the D. desulfuricans-M. barkeri coculture, lactate was completely degraded, with almost stoichiometric production of acetate and CH(4). Later, acetate was degraded to CH(4) and presumably to CO(2). In experiments in which 20 mM acetate and 0 to 20 mM lactate were added to D. desulfuricans-M. barkeri cocultures, no detectable degradation of acetate occurred until the lactate was catabolized. The ultimate rate of acetate utilization for methanogenesis was greater for those cocultures receiving the highest levels of lactate. A small amount of H(2) was detected in cocultures which contained D. desulfuricans and M. barkeri until after all lactate was degraded. The addition of H(2), but not of lactate, to the growth medium inhibited acetate degradation by pure cultures of M. barkeri. Pure cultures of M. barkeri produced CH(4) from acetate at a rate equivalent to that observed for cocultures containing M. barkeri. Inocula of M. barkeri grown with H(2)-CO(2) as the methanogenic substrate produced CH(4) from acetate at a rate equivalent to that observed for acetate-grown inocula when grown in a rumen fluid-vitamin-based medium but not when grown in a yeast extract-based medium. The results suggest that H(2) produced by the Desulfovibrio species during growth with lactate inhibited acetate degradation by M. barkeri.  相似文献   

16.
Methyl coenzyme M methylreductase from acetate-grown Methanosarcina thermophila TM-1 was purified 16-fold from a cell extract to apparent homogeneity as determined by native polyacrylamide gel electrophoresis. Ninety-four percent of the methylreductase activity was recovered in the soluble fraction of cell extracts. The estimated native molecular weight of the enzyme was between 132,000 (standard deviation [SD], 1,200) and 141,000 (SD, 1,200). Denaturing polyacrylamide gel electrophoresis revealed three protein bands corresponding to molecular weights of 69,000 (SD, 1,200), 42,000 (SD, 1,200), and 33,000 (SD, 1,200) and indicated a subunit configuration of alpha 1 beta 1 gamma 1. As isolated, the enzyme was inactive but could be reductively reactivated with titanium (III) citrate or reduced ferredoxin. ATP stimulated enzyme reactivation and was postulated to be involved in a conformational change of the inactive enzyme from an unready state to a ready state that could be reductively reactivated. The temperature and pH optima for enzyme activity were 60 degrees C and between 6.5 and 7.0, respectively. The active enzyme contained 1 mol of coenzyme F430 per mol of enzyme (Mr, 144,000). The Kms for 2-(methylthio)ethane-sulfonate and 7-mercaptoheptanoylthreonine phosphate were 3.3 mM and 59 microM, respectively.  相似文献   

17.
A mutation in the mch gene, encoding the enzyme 5,10-methenyl tetrahydromethanopterin (H(4)MPT) cyclohydrolase, was constructed in vitro and recombined onto the chromosome of the methanogenic archaeon Methanosarcina barkeri. The resulting mutant does not grow in media using H(2)/CO(2), methanol, or acetate as carbon and energy sources, but does grow in media with methanol/H(2)/CO(2), demonstrating its ability to utilize H(2) as a source of electrons for reduction of methyl groups. Cell suspension experiments showed that methanogenesis from methanol or from H(2)/CO(2) is blocked in the mutant, explaining the lack of growth on these substrates. The corresponding mutation in Methanosarcina acetivorans C2A, which cannot grow on H(2)/CO(2), could not be made in wild-type strains, but could be made in strains carrying a second copy of mch, suggesting that M. acetivorans is incapable of methyl group reduction using H(2). M. acetivorans mch mutants could also be constructed in strains carrying the M. barkeri ech hydrogenase operon, suggesting that the block in the methyl reduction pathway is at the level of H(2) oxidation. Interestingly, the ech-dependent methyl reduction pathway of M. acetivorans involves an electron transport chain distinct from that used by M. barkeri, because M. barkeri ech mutants remain capable of H(2)-dependent methyl reduction.  相似文献   

18.
19.
Carbon monoxide-dependent production of H2, CO2, and CH4 was detected in crude cell extracts of acetate-grown Methanosarcina barkeri. This metabolic transformation was associated with an active methyl viologen-linked CO dehydrogenase activity (5 to 10 U/mg of protein). Carbon monoxide dehydrogenase activity was inhibited 85% by 10 microM KCN and was rapidly inactivated by O2. The enzyme was nearly homogeneous after 20-fold purification, indicating that a significant proportion of soluble cell protein was CO dehydrogenase (ca. 5%). The native purified enzyme displayed a molecular weight of 232,000 and a two-subunit composition of 92,000 and 18,000 daltons. The enzyme was shown to contain nickel by isolation of radioactive CO dehydrogenase from cells grown in 63Ni. Analysis of enzyme kinetic properties revealed an apparent Km of 5 mM for CO and a Vmax of 1,300 U/mg of protein. The spectral properties of the enzyme were similar to those published for CO dehydrogenase from acetogenic anaerobes. The physiological functions of the enzyme are discussed.  相似文献   

20.
The genome of Methanosarcina acetivorans encodes three homologs, initially annotated as hypothetical fused corrinoid/methyl transfer proteins, which are highly elevated in CO-grown cells versus cells grown with alternate substrates. Based only on phenotypic analyses of deletion mutants, it was previously concluded that the homologs are strictly dimethylsulfide:coenzyme M (CoM) methyltransferases not involved in the metabolism of CO (E. Oelgeschlager and M. Rother, Mol. Microbiol. 72:1260 -1272, 2009). The homolog encoded by MA4383 (here designated CmtA) was reexamined via biochemical characterization of the protein overproduced in Escherichia coli. Purified CmtA reconstituted with methylcob(III)alamin contained a molar ratio of cobalt to protein of 1.0 ± 0.2. The UV-visible spectrum was typical of methylated corrinoid-containing proteins, with absorbance maxima at 370 and 420 nm and a band of broad absorbance between 450 and 600 nm with maxima at 525, 490, and 550 nm. CmtA reconstituted with aquocobalamin showed methyl-tetrahydromethanopterin:CoM (CH(3)-THMPT:HS-CoM) methyltransferase activity (0.31 μmol/min/mg) with apparent K(m) values of 135 μM for CH(3)-THMPT and 277 μM for HS-CoM. The ratio of CH(3)-THMPT:HS-CoM methyltransferase activity in the soluble versus membrane cellular fractions was 15-fold greater in CO-grown versus methanol-grown cells. A mutant strain deleted for the CmtA gene showed lower growth rates and final yields when cultured with growth-limiting partial pressures of CO, demonstrating a role for CmtA during growth with this substrate. The results establish that CmtA is a soluble CH(3)-THSPT:HS-CoM methyltransferase postulated to supplement the membrane-bound CH(3)-THMPT:HS-CoM methyltransferase during CO-dependent growth of M. acetivorans. Thus, we propose that the name of the enzyme encoded by MA4384 be CmtA (for cytoplasmic methyltransferase).  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