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 15 毫秒
1.
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   

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.
Extracts of Methanosarcina barkeri reduced methanol and CO2 to CH4 in the presence of H2 and converted methanol stoichiometrically into CH4 and CO2 in the absence of H2. In dialyzed cell-free extracts these reactions were stimulated by 2-mercaptoethanesulfonic acid (coenzyme M) and some derivatives (acetyl and formylcoenzyme M and the oxidized form of coenzyme M), which could be converted to coenzyme M by enzyme systems present in the extracts. Methylcoenzyme M could not be used in these systems.  相似文献   

4.
The nickel enzyme methyl-coenzyme M reductase (MCR) catalyzes the terminal step of methane formation in the energy metabolism of all methanogenic archaea. In this reaction methyl-coenzyme M and coenzyme B are converted to methane and the heterodisulfide of coenzyme M and coenzyme B. The crystal structures of methyl-coenzyme M reductase from Methanosarcina barkeri (growth temperature optimum, 37 degrees C) and Methanopyrus kandleri (growth temperature optimum, 98 degrees C) were determined and compared with the known structure of MCR from Methanobacterium thermoautotrophicum (growth temperature optimum, 65 degrees C). The active sites of MCR from M. barkeri and M. kandleri were almost identical to that of M. thermoautotrophicum and predominantly occupied by coenzyme M and coenzyme B. The electron density at 1.6 A resolution of the M. barkeri enzyme revealed that four of the five modified amino acid residues of MCR from M. thermoautotrophicum, namely a thiopeptide, an S-methylcysteine, a 1-N-methylhistidine and a 5-methylarginine were also present. Analysis of the environment of the unusual amino acid residues near the active site indicates that some of the modifications may be required for the enzyme to be catalytically effective. In M. thermoautotrophicum and M. kandleri high temperature adaptation is coupled with increasing intracellular concentrations of lyotropic salts. This was reflected in a higher fraction of glutamate residues at the protein surface of the thermophilic enzymes adapted to high intracellular salt concentrations.  相似文献   

5.
The mechanism of ammonia assimilation in Methanosarcina barkeri and Methanobacterium thermoautotrophicum was documented by analysis of enzyme activities, 13NH3 incorporation studies, and comparison of growth and enzyme activity levels in continuous culture. Glutamate accounted for 65 and 52% of the total amino acids in the soluble pools of M. barkeri and M. thermoautotrophicum. Both organisms contained significant activities of glutamine synthetase, glutamate synthase, glutamate oxaloacetate transaminase, and glutamate pyruvate transaminase. Hydrogen-reduced deazaflavin-factor 420 or flavin mononucleotide but not NAD, NADP, or ferredoxin was used as the electron donor for glutamate synthase in M. barkeri. Glutamate dehydrogenase activity was not detected in either organism, but alanine dehydrogenase activity was present in M. thermoautotrophicum. The in vivo activity of the glutamine synthetase was verified in M. thermoautotrophicum by analysis of 13NH3 incorporation into glutamine, glutamate, and alanine. Alanine dehydrogenase and glutamine synthetase activity varied in response to [NH4+] when M. thermoautotrophicum was cultured in a chemostat with cysteine as the sulfur source. Alanine dehydrogenase activity and growth yield (grams of cells/mole of methane) were highest when the organism was cultured with excess ammonia, whereas growth yield was lower and glutamine synthetase was maximal when ammonia was limiting.  相似文献   

6.
Trimethylamine dehydrogenase from a facultative methylotroph contains 4 g atoms each of Fe and S and an unknown, covalently bound, yellow coenzyme. The absorbance of the enzyme in the visible range (λmax=445 nm) is extensively bleached by dithionite. Reduction by substrate causes less extensive bleaching and the appearance of a three banded spectrum which may be representative of a free radical form. Denaturation liberates the FeS center(s) but not the organic coenzyme. The latter is covalently linked to the protein via an amino acid residue and is solubilized on proteolytic digestion in the form of the peptide. The coenzyme-peptide has been purified to a constant ratio of amino acid to coenzyme. The oxidized and reduced forms show maximal absorbance at 437 nm and 380 nm respectively. Based on dithionite titrations its molar absorbance at 437 nm is 12,300 in the oxidized and 4000 in the dithionite reduced form. The cofactor is very labile to photolysis giving rise to several products the predominant one of which shows fluorescence excitation and emission maxima at 394 and 500 nm, respectively. After cleavage of the hydrolyzable amino acids in HCl, the compound consumed 3 moles of periodate. Digestion with aminopeptidase M yields a compound with a single amino acid and ~1 mole of organic P present. Acid phosphatase, but not nucleotide pyrophosphatase affects its mobility. These findings suggest that the coenzyme-peptide is isolated in the form of a mononucleotide, containing a 5-carbon alcohol. The physical and chemical properties of the compound do not agree with those of known flavin or pyridoxine derivatives but are not incompatible with a covalently linked pteridine (lumazine) derivative, although no proof for such a structure is so far available.  相似文献   

7.
N 5 N 10-Methylenetetrahydromethanopterin reductase was purified 13-fold to apparent homogeneity from methanol grown Methanosarcina barkeri . The colourless enzyme was found to be composed of four identical subunits of apparent molecular mass 36 kDa. It catalysed the reduction of methylenetetrahydromethanopterin ( K m=15 μM) to methyltetrahydromethanopterin with reduced coenzyme F420 ( K m=12 μM) at a specific rate ( V max) of 2200 μmol min−1· mg protein−1 ( K cat=1320 s−1). With respect to coenzyme specificity, molecular properties and catalytic mechanism the enzyme was found to be similar to CH2=H4MPT reductase of Methanobacterium thermoautotrophicum which phylogenetically is only distantly related to M. barkeri .  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.
We used (13)C-labeled methane to document the extent of trace methane oxidation by Archaeoglobus fulgidus, Archaeoglobus lithotrophicus, Archaeoglobus profundus, Methanobacterium thermoautotrophicum, Methanosarcina barkeri and Methanosarcina acetivorans. The results indicate trace methane oxidation during growth varied among different species and among methanogen cultures grown on different substrates. The extent of trace methane oxidation by Mb. thermoautotrophicum (0.05 +/- 0.04%, +/- 2 standard deviations of the methane produced during growth) was less than that by M. barkeri (0.15 +/- 0.04%), grown under similar conditions with H(2) and CO(2). Methanosarcina acetivorans oxidized more methane during growth on trimethylamine (0.36 +/- 0.05%) than during growth on methanol (0.07 +/- 0.03%). This may indicate that, in M. acetivorans, either a methyltransferase related to growth on trimethylamine plays a role in methane oxidation, or that methanol is an intermediate of methane oxidation. Addition of possible electron acceptors (O(2), NO(3) (-), SO(4) (2-), SO(3) (2-)) or H(2) to the headspace did not substantially enhance or diminish methane oxidation in M. acetivorans cultures. Separate growth experiments with FAD and NAD(+) showed that inclusion of these electron carriers also did not enhance methane oxidation. Our results suggest trace methane oxidized during methanogenesis cannot be coupled to the reduction of these electron acceptors in pure cultures, and that the mechanism by which methane is oxidized in methanogens is independent of H(2) concentration. In contrast to the methanogens, species of the sulfate-reducing genus Archaeoglobus did not significantly oxidize methane during growth (oxidizing 0.003 +/- 0.01% of the methane provided to A. fulgidus, 0.002 +/- 0.009% to A. lithotrophicus and 0.003 +/- 0.02% to A. profundus). Lack of observable methane oxidation in the three Archaeoglobus species examined may indicate that methyl-coenzyme M reductase, which is not present in this genus, is required for the anaerobic oxidation of methane, consistent with the "reverse methanogenesis" hypothesis.  相似文献   

10.
11.
A fraction previously isolated from acid-treated supernatant fraction of Methanobacterium thermoautotrophicum by DEAE-Sephadex chromatography [Sauer, Mahadevan & Erfle (1984) Biochem. J. 221, 61-97] which was absolutely required for methane synthesis, has been separated into two compounds, tetrahydromethanopterin (H4MPT) and an as-yet-unidentified cofactor we call 'cytoplasmic cofactor'. H4MPT was identified by its u.v. spectrum and by 13C- and 1H-n.m.r. spectroscopy. The reduction of 2-(methylthio)ethanesulphonic acid (CH3-S-CoM) to methane by the membrane fraction from M. thermoautotrophicum was completely dependent on the addition of cytoplasmic cofactor. Methane synthesis from CO2, however, was only partially dependent on cofactor addition, and 57% of the original activity was retained in its absence. The kinetics of 14C labelling were consistent with the scheme methyl-H4MPT----CH3-S-CoM----methane, as has been proposed. This is the first time that direct experimental evidence has been presented to show that the proposed methyl transfer from H4MPT to coenzyme M (HS-CoM) actually occurs.  相似文献   

12.
Intact membrane vesicles are required to synthesize methane from CO2 and H2 by disrupted preparations of Methanobacterium thermoautotrophicum cells. When membrane vesicles were removed by high-speed centrifugation at 226 600 g, the remaining supernatant fraction no longer synthesized methane. Alternatively, if vesicle structure was disrupted by passage through a Ribi cell fractionator at very high pressures (345 MPa), the bacterial cell extract, with all the particulate fraction in it, did not synthesize methane. Methyl-coenzyme M, a new coenzyme first described by McBride & Wolfe [(1971) Biochemistry 10, 2317--2324], was shown to stimulate methane production from CO2 and H2, as previously reported, but the methyl group of the coenzyme did not appear to be a precursor of methane in this reaction. No methyl-coenzyme M reductase activity was detected in the cytoplasmic fraction of M. thermoautotrophicum cells.  相似文献   

13.
The novel compound 2,3-cyclopyrophosphoglycerate (CPP) is the major small molecule carbon pool in Methanobacterium thermoautotrophicum. High-field 13C NMR 13CO2 pulse/unenriched CO2 chase experiments have shown that the labeled CPP rapidly loses its 13C to an insoluble pool, while the CPP steady-state concentration is maintained (as monitored by 31P NMR spectroscopy). The biosynthesis of CPP from CO2, acetyl coenzyme A, and pyruvate as precursors has been established by a 13C NMR study of ethanol extracts of Mb. thermoautotrophicum fed with 13CO2, [1-13C]- and [2-13C]acetate, and [1-13C]pyruvate. That CPP is a post-phosphoenolpyruvate metabolite has been confirmed by in vitro experiments with cell extracts. A role for CPP in carbohydrate metabolism was established when [1-13C]glucose fed to cells resulted in the formation of [3-13C]CPP exclusively. Possible functions of CPP within the cell are discussed.  相似文献   

14.
Buffer-soluble extracts of acetate-grown Methanosarcina barkeri catalyzed methanogenesis from acetate in the presence of hydrogen and ATP. The rates of methane formation from either acetate plus ATP, or acetylphosphate without ATP added, were approximately doubled by the addition of coenzyme A (CoA). In vitro methyl group transfer from [2-14C]acetate to form [14CH3]methyl coenzyme M (14CH3S-CoM) was monitored by causing the accumulation of 14CH3S-CoM (14CH3-SCH2CH2SO3-) in the presence of 2-bromoethanesulfonate. The rate of 14CH3S-CoM formation was increased 2.5-fold by 0.2 mM CoA.  相似文献   

15.
Methanofuran (4-[N-(4,5,7-tricarboxyheptanoyl-gamma-L-glutamyl)-gamma-L- glutamyl)-p-(beta-aminoethyl)phenoxymethyl]-2-(aminomethyl)furan is a coenzyme involved in methanogenesis. The N-formyl derivative is an intermediate in the reduction of CO2 to CH4 and the disproportionation of methanol to CO2 and CH4. Formylmethanofuran dehydrogenase and formylmethanofuran:tetrahydromethanopterin formyltransferase are the enzymes catalyzing its conversions. We report here that the two enzymes from Methanosarcina barkeri and the formyltransferase from Methanobacterium thermoautotrophicum can also use N-furfurylformamide as a pseudo-substrate albeit with higher apparent Km and lower apparent Vmax values. N-Methylformamide, formamide, and formate were not converted indicating that the furfurylamine moiety of methanofuran is the minimum structure required for the correct binding of the coenzyme.  相似文献   

16.
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   

17.
The 5,10-methenyltetrahydromethanopterin cyclohydrolase from Methanosarcina barkeri was purified 313-fold to a specific activity of 470 mumol min-1 mg-1 at 37 degrees C and pH 7.8. At this stage, the enzyme was pure as judged from polyacrylamide gel electrophoresis. The monofunctional enzyme was oxygen stable, but the presence of a detergent proved to be essential for its stability. Like the cyclohydrolase purified from Methanobacterium thermoautotrophicum (A. A. Dimarco, M. I. Donnelly, and R. S. Wolfe, J. Bacteriol. 168:1372-1377, 1986), the protein showed an apparent Mr of 82,000, and it is composed of two identical subunits as was concluded from nondenaturating and denaturating polyacrylamide gel electrophoresis. The enzymes from M. thermoautotrophicum and M. barkeri markedly differ with respect to the hydrolysis product of 5,10-methenyltetrahydromethanopterin: 5-formyl- and 10-formyltetrahydromethanopterin, respectively. The apparent Km for 5,10-methenyltetrahydromethanopterin was 0.57 mM at 37 degrees C and pH 7.8.  相似文献   

18.
Among 17 purine and ATP derivatives tested, 3 were found to totally inhibit the methyl coenzyme M methylreductase system of Methanobacterium thermoautotrophicum at a concentration of 5 mM; these derivatives were 8-azido-ATP, alpha, beta-thio-ADP and 2',3'-dialdehyde-ATP. 2',3'-Dialdehyde-ATP specifically and irreversibly bound to component A3 of the methylreductase system during ATP activation of the system.  相似文献   

19.
The stimulation of carbon dioxide reduction to methane by addition of 2-(methylthio)ethanesulfonate (CH3-S-CoM) to cell extracts of Methanobacterium thermoautotrophicum was investigated. Similar stimulation of CO2 reduction by CH3-S-CoM was found for cell extracts of Methanobacterium bryantii and Methanospirillum hungatei. The CH3-S-CoM requirement could be met by the methanogenic precursors formaldehyde, serine, or pyruvate, or by 2-(ethylthio)ethanesulfonate (CH3CH2-S-CoM), but not by other coenzyme M derivatives. Efficient reduction of CO2 to CH4 was favored by low concentrations of CH3-S-CoM and high concentrations of CO2. Sulfhydryl compounds were identified as effective inhibitors of CO2 reduction. Both an allosteric model and a free-radical model for the mechanism of CO2 activation and reduction are discussed.  相似文献   

20.
J Breitung  R K Thauer 《FEBS letters》1990,275(1-2):226-230
Formylmethanofuran: tetrahydromethanopterin formyltransferase was purified from methanol grown Methanosarcina barkeri to apparent homogeneity and characterized with respect to its molecular and kinetic properties. The enzyme was found to be very similar to the formyltransferase from H2/CO2 grown Methanobacterium thermoautotrophicum. It also catalyzed the formation of N5-formyltetrahydromethanopterin rather than of N10-formyltetrahydromethanopterin from formylmethanofuran and tetrahydromethanopterin.  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