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 281 毫秒
1.
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   

2.
A new enzyme, tetrahydromethanopterin methyltransferase, which catalyzes the transfer of methyl groups from methyl-tetrahydromethanopterin to 2-mercaptoethane-sulfonate, has been found in the methane synthesizing complex of Methanobacterium thermoautotrophicum. The enzyme is oxygen sensitive and has a well defined pH optimum at pH 6.7. There was no methyl group transfer when the enzyme was heated to 100 degrees for 5 min. The product of the forward reaction, methyl-CoM, was positively identified by TLC and high voltage paper electrophoresis. The demethylation of methyl-CoM, in the absence of methane synthesis, was dependent on the addition of H4MPT which suggests that the enzyme reaction is reversible.  相似文献   

3.
Methanogenesis and methane oxidation are the major biological processes affecting the global cycling of the powerful greenhouse gas methane. To carry out the two alternative bioconversions, Nature has cleverly recycled key reactions for the C1 transfers between the oxidation levels of formaldehyde and formate, and these involve analogous enzyme systems and common specialized cofactors, methanopterin and methanofuran. Until recently, the distribution of these functions has been limited to methanogenic archaea and methylotrophic proteobacteria, and their evolutionary history remained obscure. Single interdomain lateral transfer of the respective genes has been suggested to play a role. Here we show that genes for C1 transfer reactions linked to methanopterin and methanofuran are also present in diverse representatives of the enigmatic bacterial clade, the Planctomycetes. Phylogenetic analysis places the planctomycete sequences as distantly from their archaeal counterparts as from their proteobacterial counterparts, suggesting novel scenarios for the evolution of the C1 transfer functions in both methanogens and methylotrophs. This finding suggests a possible role for Planctomycetes in the evolution of the methane cycle on Earth.  相似文献   

4.
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   

5.
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   

6.
The 13C NMR signals of methanofuran were assigned by two-dimensional 1H and 13C NMR experiments. On this basis, the incorporation of 13C-labeled acetate and pyruvate into methanofuran by growing cells of Methanobacterium thermoautotrophicum was analyzed by one- and two-dimensional 13C NMR experiments. The data were analyzed by a retrobiosynthetic approach based on a comparison of labeling patterns in a variety of metabolites. The data show that the furan ring is formed by condensation of two molecules from the pyruvate/triose pool. The tetracarbocylic acid moiety is assembled from ketoglutarate, two molecules of acetyl CoA, and one molecule of carbon dioxide.  相似文献   

7.
Abstract During the first stage of the preparation of mushroom compost oxygen is believed to be readily available. However we measured methane in the evoking air above the compost piles and were able to isolate thermophilic methanogenic bacteria from this compost. The isolates grow only on H2 and CO2 as energy and carbon source and do not require complex factors for growth. On the basis of nutritional and morphological characteristics these methanogens were identified as strains of Methanobacterium thermoautotrophicum .  相似文献   

8.
Formylmethanofuran: tetrahydromethanopterin formyltransferase was purified to electrophoretic homogeneity from cells of Methanobacterium thermoautotrophicum. The enzyme is a tetramer of similar or identical subunits (Mr = 41,000). The equilibrium favors transfer of the formyl group to tetrahydromethanopterin (H4MPT) at physiological pH. The product of formyl transfer by the purified enzyme was shown by a number of criteria to be 5-formyl-H4MPT, as opposed to 10-formyl-H4MPT or 5,10-methenyl-H4MPT. Reconstitution of a portion of the methanogenic C1 cycle was effected by combining purified formyltransferase, methenyl-H4MPT cyclohydrolase, formylmethanofuran, and H4MPT to give methenyl-H4MPT. Additional reconstitution experiments established that the formyltransferase is an essential enzyme for the conversion of carbon dioxide to methane. In conjunction with previously published data (Donnelly, M.I., Escalante-Semerena, J.C., Rinehart, K. L., Jr., and Wolfe, R.S. (1985) Arch. Biochem. Biophys. 242, 430-439), these data substantiate the role of 5-formyl-H4MPT as an intermediate of methanogenesis.  相似文献   

9.
Methanofuran is the first coenzyme in the methanogenic pathway used by the archaeon Methanocaldococcus jannaschii, as well as other methanogens, to reduce CO2 to methane. The details of the pathway for the biosynthesis of methanofuran and the responsible genes have yet to be established. A clear structural element in all known methanofurans is tyramine, likely produced by the decarboxylation of L-tyrosine. We show here that the mfnA gene at M. jannaschii locus MJ0050 encodes a thermostable pyridoxal phosphate-dependent L-tyrosine decarboxylase that specifically produces tyramine. Homologs of this gene are widely distributed among euryarchaea but are not specifically related to known bacterial or plant tyrosine decarboxylases.  相似文献   

10.
The requirement of ATP for the methyl coenzyme M methylreductase in extracts of Methanobacterium thermoautotrophicum was found to be catalytic; for each mol of ATP added, 15 mol of methane was produced from methyl coenzyme M [2-(methylthio)ethanesulfonic acid]. Other nucleotide triphosphates partially replaced ATP in activation of the reductase. All components of the reaction were found in the supernatant fraction of cell extracts after centrifugation at 100,000 X g for 1 h; optimal reaction rates occurred at 65 degrees C, at a pH range of 5.6 to 6.0, and at concentrations of ATP and MgCl2 of 1 mM and 40 mM, respectively. Chloral hydrate, chloroform, nitrite, 2,4-dinitrophenol, and viologen dyes (compounds known to inhibit methanogenesis from a variety of substrates) were found to inhibit the conversion of methyl coenzyme M to methane. Methyl coenzyme M methylreductase was shown to be present in a variety of methanogens.  相似文献   

11.
Rao AV  Shaha C 《FEBS letters》2002,523(1-3):133-137
Methylobacterium extorquens AM1 possesses a formyltransferase (Ftr) complex that is essential for growth in the presence of methanol and involved in formaldehyde oxidation to CO(2). One of the subunits of the complex carries the catalytic site for transfer of the formyl group from tetrahydromethanopterin to methanofuran (MFR). We now found via nuclear magnetic resonance-based studies that the Ftr complex also catalyzes the hydrolysis of formyl-MFR and generates formate. The enzyme was therefore renamed Ftr/hydrolase complex (Fhc). FhcA shares a sequence pattern with amidohydrolases and is assumed to be the catalytic site where the hydrolysis takes place.  相似文献   

12.
Formylmethanofuran:tetrahydromethanopterin formyltransferase is an essential enzyme in the one-carbon metabolism of methanogenic and sulfate-reducing archaea and of methylotrophic bacteria. The enzyme, which is devoid of a prosthetic group, catalyzes the reversible formyl transfer between the two substrates coenzyme methanofuran and coenzyme tetrahydromethanopterin (H4MPT) in a ternary complex catalytic mechanism. The structure of the formyltransferase without its coenzymes has been determined earlier. We report here the structure of the enzyme in complex with both coenzymes at a resolution of 2.0 A. Methanofuran, characterized for the first time in an enzyme structure, is embedded in an elongated cleft at the homodimer interface and fixed by multiple hydrophobic interactions. In contrast, tetrahydromethanopterin is only weakly bound in a shallow and wide cleft that provides two binding sites. It is assumed that the binding of the bulky coenzymes induces conformational changes of the polypeptide in the range of 3A that close the H4MPT binding cleft and position the reactive groups of both substrates optimally for the reaction. The key residue for substrate binding and catalysis is the strictly conserved Glu245. Glu245, embedded in a hydrophobic region and completely buried upon tetrahydromethanopterin binding, is presumably protonated prior to the reaction and is thus able to stabilize the tetrahedral oxyanion intermediate generated by the nucleophilic attack of the N5 atom of tetrahydromethanopterin onto the formyl carbon atom of formylmethanofuran.  相似文献   

13.
The levels of seven water-soluble vitamins in Methanobacterium thermoautotrophicum, Methanococcus voltae, Escherichia coli, Bacillus subtilis, Pseudomonas fluorescens, and Bacteroides thetaiotaomicron were compared by using a vitamin-requiring Leuconostoc strain. Both methanogens contained levels of folic acid and pantothenic acid which were approximately two orders of magnitude lower than levels in the nonmethanogens. Methanobacterium thermoautotrophicum contained levels of thiamine, biotin, nicotinic acid, and pyridoxine which were approximately one order of magnitude lower than levels in the nonmethanogens. The thiamine level in Methanococcus voltae was approximately one order of magnitude lower than levels in the nonmethanogens. Only the levels of riboflavin (and nicotinic acid and pyridoxine in Methanococcus voltae) were approximately equal in the methanogens and nonmethanogens. Folic acid may have been present in extracts of methanogens merely as a precursor, by-product, or hydrolysis product of methanopterin.  相似文献   

14.
The structural genes encoding the four largest subunits of RNA polymerase, A, B', B", and C, were physically mapped in Methanobacterium thermoautotrophicum Winter. The genes formed a cluster in the order B", B', A, C and had a common orientation. DNA hybridization experiments yielded different degrees of homology between RNA polymerase gene sequences of different species of Methanobacterium and Methanococcus voltae. No homology was detectable between Methanobacterium thermoautotrophicum and Methanosarcina barkeri. From Southern hybridization experiments in which probes of the four genes from Methanobacterium thermoautotrophicum Winter and restriction digests of the genomic DNAs of the different methanogens were used, a common gene order of the RNA polymerase genes could be deduced.  相似文献   

15.
Methanococcus deltae (Delta)LH was grown on H(inf2)-CO(inf2) in the presence of various concentrations of furfural. Furfural at higher concentrations, namely, 20 and 25 mM, inhibited growth of this organism. At concentration of 5 and 10 mM, no inhibition of growth was observed. The other methanogens in this study were not inhibited by 10 mM furfural. Among the methanogens tested, M. deltae was capable of transforming furfural, whereas Methanobacterium thermoautotrophicum Marburg, Methanosarcina barkeri 227, Methanococcus thermolithotrophicus, and Methanobrevibacter ruminantium lacked this capability. One hundred percent removal of furfural was observed within 48 h of incubation in M. deltae cultures. The end product observed during furfural metabolism was furfuryl alcohol. An almost stoichiometric amount of furfuryl alcohol was produced by M. deltae. This transformation is likely to be of value in the detoxification of furfural and in its ultimate conversion to methane and CO(inf2) by anaerobic digestion.  相似文献   

16.
The formylmethanofuran:tetrahydromethanopterin formyltransferase (FTR) from Methanobacterium thermoautotrophicum delta H was cloned and its sequence was determined. The clone was contained on a 4.8-kilobase BamHI fragment of M. thermoautotrophicum DNA ligated into pBR329. When this fragment was subcloned into the phagemid pTZ18R, a functional enzyme was synthesized under control of the lac promoter. Sequence analysis revealed the presence of a ribosome binding site and a possible terminator structure. The absence of an identifiable promoter lends credibility to the open reading frame which is present 5' to ftr. The ftr gene encodes an acidic protein with a calculated molecular weight of 31,401. The sequence of FTR does not appear to be homologous to any other sequenced proteins, including proteins which use pterin substrates.  相似文献   

17.
《FEBS letters》1985,181(2):303-307
Methanobacterium thermoautotrophicum, a methane forming archaebacterium, grows autotrophically by synthesizing activated acetic acid from 2 CO2. It is demonstrated in vitro that the methyl group of acetate is derived from methenyl tetrahydromethanopterin, which is known to be a one-carbon carrying coenzyme in CO2 reduction to methane. The direct acetate precursors are suggested to be methyl tetrahydromethanopterin (“activated methanol”) and “activated carbon monoxide”.  相似文献   

18.
The ftr gene encoding formylmethanofuran: tetrahydromethanopterin formyltransferase (Ftr) from Methanosarcina barkeri was cloned, sequenced, and functionally expressed in Escherichia coli. The overproduced enzyme was purified eightfold to apparent homogeneity, and its catalytic properties were determined. The primary structure and the hydropathic character of the formyltransferase from Methanosarcina barkeri were compared with those of the enzymes from Methanobacterium thermoautotrophicum, Methanothermus fervidus, and Methanopyrus kandleri. The amino acid sequence of the enzyme from Methanosarcina barkeri was 64%, 61%, and 59% identical to that of the enzyme from Methanobacterium thermoautotrophicum, Methanothermus fervidus, and Methanopyrus kandleri, respectively. A negative correlation between the hydrophobicity of the enzymes and both the growth temperature optimum and the intracellular salt concentration of the four organisms was observed. The hydrophobicity of amino acid composition was +21.6 for the enzyme from Methanosarcina barkeri (growth temperature optimum 37° C, intracellular salt concentration ≈ 0.3 M), +9.9 for the enzyme from Methanobacterium thermoautotrophicum (65°C, ≈ 0.7 M), –20.8 for the enzyme from Methanothermus fervidus (83° C, ≈ 1.0 M) and –31.4 for the enzyme from Methanopyrus kandleri (98° C, > 1.1 M). Generally, a positive correlation between hydrophobicity and thermophilicity of enzymes and a negative correlation between hydrophobicity and halophilicity of enzymes are observed. The findings therefore indicate that the hydropathic character of the formyltransferases compared is mainly determined by the intracellular salt concentration rather than by temperature. Sequence similarities between the formyltransferases from methanogens and an open reading frame from Methylobacterium extorquens AM1 are discussed. Received: 7 September 1995 / Accepted: 7 November 1995  相似文献   

19.
Targeting methanopterin biosynthesis to inhibit methanogenesis   总被引:5,自引:0,他引:5  
This paper describes the design, synthesis, and successful employment of inhibitors of 4-(beta-D-ribofuranosyl)aminobenzene-5'-phosphate (RFA-P) synthase, which catalyzes the first committed step in the biosynthesis of methanopterin, to specifically halt the growth of methane-producing microbes. RFA-P synthase catalyzes the first step in the synthesis of tetrahydromethanopterin, a key cofactor required for methane formation and for one-carbon transformations in methanogens. A number of inhibitors, which are N-substituted derivatives of p-aminobenzoic acid (pABA), have been synthesized and their inhibition constants with RFA-P synthase have been determined. Based on comparisons of the inhibition constants among various inhibitors, we propose that the pABA binding site in RFA-P synthase has a relatively large hydrophobic pocket near the amino group. These enzyme-targeted inhibitors arrest the methanogenesis and growth of pure cultures of methanogens. Supplying pABA to the culture relieves the inhibition, indicating a competitive interaction between pABA and the inhibitor at the cellular target, which is most likely RFAP synthase. The inhibitors do not adversely affect the growth of pure cultures of the bacteria (acetogens) that play a beneficial role in the rumen. Inhibitors added to dense ruminal fluid cultures (artificial rumena) halt methanogenesis; however, they do not inhibit volatile fatty acid (VFA) production and, in some cases, VFA levels are slightly elevated in the methanogenesis-inhibited cultures. We suggest that inhibiting methanopterin biosynthesis could be considered in strategies to decrease anthropogenic methane emissions, which could have an environmental benefit since methane is a potent greenhouse gas.  相似文献   

20.
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   

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