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.
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   

2.
The rate of methane formation from H2 and CO2, the intracellular ATP content and the electrochemical proton potential (delta mu H+) were determined in cell suspensions of Methanobacterium thermoautotrophicum, which were permeabilized for K+ with valinomycin (1.2 mumol/mg protein). In the absence of extracellular K+ the cells formed methane at a rate of 4 mumol min-1 (mg protein)-1, the intracellular ATP content was 20 nmol/mg protein and the delta mu H+ was 200 mV (inside negative). When K+ was added to the suspensions the measured delta mu H+ decreased to the value calculated from the [K+]in/[K+]out ratio. Using this method of delta mu H+ adjustment, it was found that lowering delta mu H+ from 200 mV ([K+]in/[K+]out = 1000) to 100 mV ([K+]in/[K+]out = 40) had no effect on the rate of methane formation and on the intracellular ATP content. At delta mu H+ values below 100 mV ([K+]in/[K+]out less than 40) both the rate of methanogenesis and the ATP content decreased. Methanogenesis completely ceased and the ATP content was 2 nmol/mg when delta mu H+ was adjusted to values lower 50 mV ([K+]in/[K+]out less than 7). The data show that methanogenesis from H2 and CO2 and ATP synthesis in M. thermoautotrophicum are possible at relatively low electrochemical proton potentials. Similar results were obtained with Methanosarcina barkeri. Protonophoric uncouplers like 3,5,3',4'-tetrachlorosalicylanilide (TCS) or 3,5-di-tert-butyl-4-hydroxy-benzylidenemalononitrile (SF 6847) were found not to dissipate delta mu H+ below 100 mV in M. thermoautotrophicum even when used at high concentrations (400 nmol/mg protein). This finding explains the observed uncoupler insensitivity of methanogenesis and ATP synthesis in this organism.  相似文献   

3.
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   

4.
CH4 formation from CO2 and H2 rather than from formaldehyde and H2 in methanogenic bacteria is inhibited by uncouplers, indicating that CO2 reduction to the formaldehyde level is energy-driven. We report here that in Methanosarcina barkeri the driving force is a primary electrochemical sodium potential (delta mu Na+) generated by formaldehyde reduction to CH4. This is concluded from the following findings. 1. CO2 reduction to CH4 was insensitive towards protonophores, when the Na+/H+ antiporter was inhibited; under these conditions delta mu Na+ was 120 mV (inside negative), whereas both delta mu H+ and the cellular ATP content were low. 2. CO2 reduction to CH4, rather than formaldehyde reduction, was sensitive towards Na+ ionophores, which dissipated delta mu Na+. 3. CO2 reduction to CH4, in the presence of protonophores and Na+/H+ antiport inhibitors, was coupled with the extrusion of 1-2 mol Na+/mol CH4, and formaldehyde reduction to CH4 was coupled with the extrusion of 3-4 mol Na+/mol CH4. Thus during CO2 reduction to the formaldehyde level 2-3 mol Na+ were consumed.  相似文献   

5.
Methanogenesis from methanol by cell suspensions of Methanosarcina barkeri was inhibited by the uncoupler tetrachlorosalicylanilide. This inhibition was reversed by the addition of formaldehyde. 14C labeling experiments revealed that methanol served exclusively as the electron acceptor, whereas formaldehyde was mainly oxidized to CO2 under these conditions. These data support the hypothesis (M. Blaut and G. Gottschalk, Eur. J. Biochem. 141: 217-222, 1984) that the first step in methanol oxidation depends on the proton motive force or a product thereof. Cell extracts of M. barkeri converted methanol and formaldehyde to methane under an H2 atmosphere. Under an N2 atmosphere, however, formaldehyde was disproportionated to CH4 and CO2, whereas methanol was metabolized to a very small extent only, irrespective of the presence of ATP. It was concluded that cell extracts of M. barkeri are not able to oxidize methanol. In further experiments, the sodium dependence of methanogenesis and ATP formation by whole cells was investigated. Methane formation from methanol alone and the corresponding increase in the intracellular ATP content were strictly dependent on Na+. If, in contrast, methanol was utilized together with H2, methane and ATP were synthesized in the absence of Na+. The same is true for the disproportionation of formaldehyde to methane and carbon dioxide. From these experiments, it is concluded that in M. barkeri, Na+ is involved not in the process of ATP synthesis but in the first step of methanol oxidation.  相似文献   

6.
Cell suspensions of Methanosarcina barkeri were found to oxidize formaldehyde to CO2 and 2H2 (delta G0' = -27 kJ/mol CO2), when methanogenesis was inhibited by 2-bromoethanesulfonate. We report here that this reaction is coupled with (a) primary electrogenic Na+ translocation at a stoichiometry of 2-3 Na+/CO2, (b) with secondary H+ translocation via a Na+/H+ antiporter and (c) with ATP synthesis driven by an electrochemical proton potential. This is concluded from the following findings. Formaldehyde oxidation to CO2 and 2H2 was dependent on Na+ ions, 2-3 mol Na+/mol formaldehyde oxidized were extruded. Na+ translocation was inhibited by Na+ ionophores, but not affected by protonophores of Na+/H+ antiport inhibitors. Formaldehyde oxidation was associated with the build up of a membrane potential in the order of 100 mV (inside negative), which could be dissipated by sodium ionophores rather than by protonophores. Formaldehyde oxidation was coupled with ATP synthesis, which could be inhibited by Na+ ionophores, Na+/H+ antiport inhibitors, by protonophores and by the H+-translocating-ATP-synthase inhibitor, dicyclohexylcarbodiimide. With cell suspensions of Methanobacterium thermoautotrophicum similar results were obtained.  相似文献   

7.
The membrane potential (delta psi) of whole cells of Methanobacterium thermoautotrophicum strain delta H was estimated under different external conditions using a TPP(+)-sensitive electrode. The results show that the delta psi values of M. thermoautotrophicum at alkaline pHout (8.5) are comparable with delta psi values under slightly acidic conditions (pH 6.8; 230 and 205 mV, respectively). On the other hand, the size of colonies on Petri dishes was remarkably smaller at pH 8.5 than at 6.8. The delta psi was insensitive to relevant ATPase inhibitors. At pH 6.8, the protonophore 3,3',4',5-tetrachlorosalicylanilide (TCS) strongly inhibited delta psi formation and ATP synthesis driven by methanogenic electron transport. On the other hand, at pH 8.5 the CH4 formation and ATP synthesis were insensitive to TCS and a protonophore-resistant delta psi of approximately 150 mV was determined. The finding of a protonophore-resistant delta psi at pH 8.5 indicates that at alkaline pHout these cells can switch from H(+)-energetics to Na(+)-energetics, when the delta [symbol: see text] H+ becomes limited. The results strongly support the hypothesis that at alkaline pHout Na+ ions might fully substitute for H+ in these cells as the coupling ions.  相似文献   

8.
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   

9.
Cell suspensions of acetate-grown Methanosarcina barkeri mediate the conversion of CO and H2O to CO2 and H2. The reaction is coupled with the phosphorylation of ADP. Evidence is presented that CO oxidation by the cells is associated with the transient acidification of the suspension medium. Up to 2 mol vectorial protons were measured/mol CO oxidized when the transmembrane electrical gradient was kept low by the addition of valinomycin (20 nmol/mg protein) and KCl (200 mM) or of KSCN (50 mM). No transient acidification was observed in the presence of the protonophore tetrachlorosalicylanilide which stimulated rather than inhibited CO oxidation. Proton extrusion remained unaltered when the proton-translocating ATPase was specifically inhibited by dicyclohexylcarbodiimide. The latter finding indicates that proton translocation is associated with CO conversion to CO2 and H2 rather than with ATP hydrolysis in the cells. The data substantiate that the coupling of CO oxidation with ADP phosphorylation in M. barkeri occurs via a chemiosmotic mechanism.  相似文献   

10.
Methanobacterium thermoautotrophicum, M. ruminantium, and Methanosarcina barkeri were labeled with 14CO2 (14CO2 + H14CO3- + 14CO32-) for from 2 to 45 s. Radioactivity was recovered in coenzyme M derivatives, alanine, aspartate, glutamate, and several unidentified compounds. The properties of one important structurally unidentified intermediate (yellow fluorescent compound) displayed UV absorbance maxima at pH 1 of 290 and 335 nm, no absorbance in the visible region, and a fluorescence maximum at 460 nm. Label did not appear in organic phosphates until after 1 min. 14CH3OH was converted by M. barkeri primarily into coenzyme M derivatives at 25 s. [2-14C]acetate was assimilated by M. thermoautotrophicum mainly into alanine and succinate during 2 to 240 s, but not into coenzyme M derivatives or yellow fluorescent compound. Cell-free extracts of M. thermoautotrophicum lacked ribulose 1,5-bisphosphate carboxylase activity. The data indicated the absence of the Calvin, serine, and hexulose phosphate paths of C1 assimilation in the methanogens examined and indicated that pyruvate was an early intermediate product of net CO2 fixation. The in vivo importance of coenzyme M derivatives in methanogenesis was demonstrated.  相似文献   

11.
The results presented show that in Mycoplasma mycoides var. Capri, regulation of glucose uptake by its non-metabolizable analogue methyl alpha-D-glucoside, can be used to control intracellular ATP content. This in turn leads to a control of the rate of proton extrusion catalysed by the Mg2+-dependent ATPase (phi (cHxN)2C H+) and the respective amplitudes of the components of delta mu H+. When Mycoplasma cells are incubated with 10 mM methyl alpha-D-glucoside, the amplitude of phi (cHxN)2C H+, of the electrical potential delta psi and of the chemical gradient delta pH become continuous functions of external glucose concentration within the limits of the non-energized and fully energized states. Analysis of the relationships between graduated amplitudes of delta psi, delta pH and phi (cHxN) 2C H+ show that the primary form of energy stored by a delta mu H+ generator is the electrical potential.  相似文献   

12.
Methanogenesis from formaldehyde or formaldehyde + H2, as carried out by Methanosarcina barkeri, was strictly dependent on sodium ions whereas methane formation from methanol + H2 or methanol + formaldehyde was Na+-independent. This indicates that the reduction of formaldehyde to the formal redox level of methanol exhibits a Na+ requirement. During methanogenesis from formaldehyde, a delta pNa in the range of -62 mV to -80 mV was generated by means of a primary, electron-transport-driven sodium pump. This could be concluded from the following results obtained on cell suspensions of M. barkeri. 1. The addition of proton conductors or inhibitors of the Na+/H+ antiporter had no effect on sodium extrusion. 2. During methanogenesis from formaldehyde + H2 a delta psi of -60 mV to -70 mV was generated even in the presence of proton conductors. 3. ATPase inhibitors, applied in the presence of proton conductors, had no effect on primary sodium extrusion or generation of a delta psi. Evidence for a Na+-translocating ATPase could not be obtained.  相似文献   

13.
We describe a K+ transport system in Methanospirillum hungatei cells depleted of cytoplasmic K+ via an ammonia/K+ exchange reaction (Sprott, G. D., Shaw, K. M., and Jarrell, K. F. (1984) J. Biol. Chem. 259, 12602-12608). Ammonia-treated cells contained low concentrations of ATP and were unable to make CH4 or to transport 86Rb+. All of these properties were restored by CaCl2, MgCl2, or MnCl2, and not by CoCl2 or NiCl2. The Rb+ transport system had a Km of 0.42 and Vmax of 29 nmol/min X mg; K+ inhibited competitively. Both H2 and CO2 were required for appreciable transport, whereas air, valinomycin, or nigericin were potent inhibitors. The influx of Rb+ was electrogenic and associated with proton efflux, producing a delta pH (alkaline inside) in acidic media. In the absence of K+ (or Rb+), the activation of CH4 synthesis by Mg2+ produced little change in the cytoplasmic pH, showing that methanogenesis did not elicit a net efflux of protons. The pH optimum for transport was in the range 6.0-7.3 where the transmembrane pH gradient would contribute minimally to the proton motive force. Protonophores at pH 6.3 caused a partial decline in CH4 synthesis and the ATP content and dramatically collapsed Rb+ transport. These and other inhibitor experiments, coupled with the fact that the Rb+ gradient was too large to be in equilibrium with the proton motive force alone, suggest a role for both ATP and the proton motive force in Rb+ transport. Also, a role for K+ in osmoregulation is indicated.  相似文献   

14.
Electron transport phosphorylation has been demonstrated to drive ATP synthesis for the methanogenic archaebacterium Methanolobus tindarius: Protonophores evoked uncoupler effects and lowered the membrane potential delta psi. Under the influence of N,N'-dicyclohexylcarbodiimide [(cHxN)2C] the membrane potential increased while methanol turnover was inhibited. 2-Bromoethanesulfonate, an inhibitor of methanogenesis, had no effect on the membrane potential but, like (cHxN)2C and protonophores, decreased the intracellular ATP concentration. Labeling experiments with (cHxN)2(14)C showed membranes to contain a proteolipid, with a molecular mass of 5.5 kDa, that resembles known (cHxN)2C-binding proteins of F0-F1 ATPases. The (cHxN)2-sensitive membrane ATPase hydrolysed Mg.ATP at a pH optimum of 5.0 with a Km (ATP) of 2.5 mM (V = 77 mU/mg). It was inhibited competitively by ADP; Ki (ADP) = 0.65 mM. Azide or vanadate caused no significant loss in ATPase activity, but millimolar concentrations of nitrate showed an inhibitory effect, suggesting a relationship to ATPases from vacuolar membranes. In contrast, no inhibition occurred in the presence of bafilomycin A1. The ATPase was extractable with EDTA at low salt concentrations. The purified enzyme consists of four different subunits, alpha (67 kDa), beta (52 kDa), gamma (20 kDa) and beta (less than 10 kDa), as determined from SDS gel electrophoresis.  相似文献   

15.
Methanogenic bacteria are considered to couple methane formation with the synthesis of ATP by a chemiosmotic mechanism. This hypothesis was tested with Methanobacterium thermoautotrophicum. Methane formation from H2 and CO2 (2.5 - 3 mumol X min-1 X mg cells-1) by cell suspensions of this organism resulted in the formation of an electrochemical proton potential (delta mu H +) across the cytoplasmic membrane of 230 mV (inside negative) and in the synthesis of ATP up to an intracellular concentration of 5 - 7 nmol/mg. The addition of ionophores at concentrations which completely dissipated delta mu H + without inhibiting methane formation did not result in an inhibition of ATP synthesis. It thus appears that delta mu H + across the cytoplasmic membrane is not the driving force for the synthesis of ATP in M. thermoautotrophicum.  相似文献   

16.
N-carboxymethanofuran (carbamate) formation from unprotonated methanofuran (MFR) and CO2 is the first reaction in the reduction of CO2 to methane in methanogenic archaea. The reaction proceeds spontaneously. We address here the question whether the rate of spontaneous carbamate formation is high enough to account for the observed rate of methanogenesis from CO2. The rates of carbamate formation (v1) and cleavage (v2) were determined under equilibrium conditions via 2D proton exchange NMR spectroscopy (EXSY). At pH 7.0 and 300 K the second order rate constant k1* of carbamate formation from 'MFR'(MFR + MFRH+) and 'CO2' (CO2 + H2CO3 + HCO3-+ CO32-) was found to be 7 M-1.s-1 (v1 = k1* ['MFR'] ['CO2']) while the pseudo first order rate constant k2* of carbamate cleavage was 12 s-1 (v2 = k2* [carbamate]). The equilibrium constant K* = k1*/k2* = [carbamate]/['MFR']['CO2'] was 0.6 M-1 at pH 7.0 corresponding to a free energy change DeltaG degrees ' of + 1.3 kJ.mol-1. The pH and temperature dependence of k1*, of k2* and of K* were determined. From the second order rate constant k1* it was calculated that under physiological conditions the rate of spontaneous carbamate formation is of the same order as the maximal rate of methane formation and as the rate of spontaneous CO2 formation from HCO3- in methanogenic archaea, the latter being important as CO2 is mainly present as HCO3- which has to be converted to CO2 before it can react with MFR. An enzyme catalyzed carbamate formation thus appears not to be required for methanogenesis from CO2. Consistent with this conclusion is our finding that the rate of carbamate formation was not enhanced by cell extracts of Methanosarcina barkeri and Methanobacterium thermoautotrophicum or by purified formylmethanofuran dehydrogenase which catalyzes the reduction of N-carboxymethanofuran to N-formylmethanofuran. From the concentrations of 'CO2' and of 'MFR' determined by 1D-NMR spectroscopy and the pKa of H2CO3 and of MFRH+ the concentrations of CO2 and of MFR were obtained, allowing to calculate k1 (v1 = k1 [MFR] [CO2]). The second order rate constant k1 was found to be approximately 1000 M-1 x s-1 at 300 K and pH values between 7.0 and 8. 0 which is in the order of k1 values determined for other carbamate forming reactions by stopped flow.  相似文献   

17.
A mutant of Methanosarcina barkeri (Fusaro) is able to grow on pyruvate as the sole carbon and energy source. During growth, pyruvate is converted to CH4 and CO2, and about 1.5 mol of ATP per mol of CH4 is formed (A.-K. Bock, A. Prieger-Kraft, and P. Schönheit, Arch. Microbiol. 161:33-46, 1994). The pyruvate-utilizing mutant of M. barkeri could also grow on pyruvate when methanogenesis was completely inhibited by bromoethanesulfonate (BES). The mutant grew on pyruvate (80 mM) in the presence of 2 mM BES with a doubling time of about 30 h up to cell densities of about 400 mg (dry weight) of cells per liter. During growth on pyruvate, the major fermentation products were acetate and CO2 (about 0.9 mol each per mol of pyruvate). Small amounts of acetoin, acetolactate, alanine, leucine, isoleucine, and valine were also detected. CH4 was not formed. The molar growth yield (Yacetate) was about 9 g of cells (dry weight) per mol of acetate, indicating an ATP yield of about 1 mol/mol of acetate formed. Growth on pyruvate in the presence of BES was limited; after six to eight generations, the doubling times increased and the final cell densities decreased. After 9 to 11 generations, growth stopped completely. In the presence of BES, suspensions of pyruvate-grown cells fermented pyruvate to acetate, CO2, and H2. CH4 was not formed. Conversion of pyruvate to acetate, in the complete absence of methanogenesis, was coupled to ATP synthesis. Dicyclohexylcarbodiimide, an inhibitor of H(+)-translocating ATP synthase, did not inhibit ATP formation. In the presence of dicyclohexylcarbodiimide, stoichiometries of up to 0.9 mol of ATP per mol of acetate were observed. The uncoupler arsenate completely inhibited ATP synthesis, while the rates of acetate, CO2, and H2 formation were stimulated up to fourfold. Cell extracts of M. barkeri grown on pyruvate under nonmethenogenic conditions contained pyruvate: ferredoxin oxidoreductase (0.5 U/mg), phosphate acetyltransferase (12 U/mg), and acetate kinase (12 U/mg). From these data it is concluded that ATP was synthesized by substrate level phosphorylation during growth of the M. barkeri mutant on pyruvate in the absence of methanogenesis. This is the first report of growth of a methanogen under nonmethanogenic conditions at the expense of a fermentative energy metabolism.  相似文献   

18.
Crude inside-out vesicles from the methanogenic strain G?1 were prepared via protoplasts. These vesicles catalyzed methane formation from methyl-CoM and H2 at a maximal rate of 35 nmol/min.mg protein. Methane formation by the vesicles did not depend on the addition of ATP. This was in contrast to conventionally prepared crude extracts from the same organism or from Methanosarcina barkeri which exhibited strict ATP dependence of methanogenesis. ATP analogues inhibited methanogenesis by extracts to a much higher extent than that by vesicles. Both, particulate and soluble components prepared from the crude vesicles by ultracentrifugation were necessary for ATP-independent methane formation from methyl-CoM and H2. Hydrogenase activity was mainly associated with the particulate fraction whereas methyl-CoM methylreductase could be assigned to the soluble fraction. The detergent sulfobetaine inhibited methane formation from methyl-CoM without affecting hydrogenase or titanium citrate-dependent methylreductase activities, indicating that an additional membraneous component is involved in methanogenesis for methyl-CoM and H2.  相似文献   

19.
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   

20.
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   

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