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.
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   

2.
Abstract Everted vesicles of the methanogenic strain Gö1 synthesized ATP in response to methanogenesis from methyl-coenzyme M and H2. Simultaneously, a transmembrane pH gradient (ΔpH) was generated as evident from fluorescence quenching of acridine orange. Protonophorous uncouplers prevented ΔpH generation and ATP synthesis, but did not affect methanogenesis. The ATP synthase inhibitor diethylstilbestrol (DES) inhibited ATP synthesis but had no effect on methanogenesis and on ΔpH formation, indicating the essential role of the transmembrane proton potential in ATP synthesis. Progress has also been made in assigning specific functions to membrane components in methanogenesis from methyl-CoM and H2. Separation of cell extracts into cytoplasmic and membrane fraction revealed an essential role of membrane-bound components in electron transfer: methanogenesis catalyzed by the cytoplasmic fraction from strain Gö1 was stimulated several fold by membranes from various methanogens. This stimulation was prevented if the membranes had been treated with oxidants (O2, K3[Fe(CN)6]) or SH reagents (Ag+, p -chloromercuribenzoate, iodoacetamide) pointing to the involvement of functional SH groups in methanogenesis from methyl-CoM and H2.  相似文献   

3.
Formaldehyde oxidation and methanogenesis   总被引:5,自引:4,他引:1       下载免费PDF全文
Formaldehyde oxidation by cell-free extracts of Methanobacterium thermoautotrophicum was shown to drive methanogenesis from CH3-S-coenzyme M or HCHO under a nonreductive atmosphere of N2. Under N2 when HCHO was the sole source of carbon and reducing equivalents in the reaction, it underwent oxidation and reduction events (disproportionation), the sum of the reactions being 3 HCHO + H2O----CH4 + 2 HCOO - + 2H+. This reaction predicts a CH4/HCHO ratio of 1/3, which is in agreement with the experimental finding of 1/2.9. In extracts of the mesophilic methanogen Methanococcus voltae and the extreme thermophile Methanococcus jannaschii , which exhibited formate dehydrogenase activity, the CH4/HCHO ratio was 1/2. NADPH stimulated methane formation from HCHO under N2. An unidentified, oxygen-labile cofactor, the formaldehyde activation factor, present in boiled-cell extract was discovered. Methanopterin , an oxygen-stable molecule, also substituted for boiled-cell extract.  相似文献   

4.
Reductive dechlorination of 1,2-dichloroethane (1,2-DCA) to ethylene and chloroethane (CA) by crude cell extracts of Methanobacterium thermoautotrophicum delta H with H2 as the electron donor was stimulated by Mg-ATP. The heterodisulfide of coenzyme M (CoM) and 7-mercaptoheptanoylthreonine phosphate together with Mg-ATP partially inhibited ethylene production but stimulated CA production compared Mg-ATP alone. The pH optimum for the dechlorination was 6.8 (at 60 degrees C). Michaelis-Menten kinetics for initial product formation rates with different 1,2-DCA concentrations indicated the enzymatic character of the dechlorination. Apparent Kms for 1,2-DCA of 89 and 119 microM and Vmaxs of 34 and 20 pmol/min/mg of protein were estimated for ethylene and CA production, respectively. 3-Bromopropanesulfonate, a specific inhibitor for methyl-CoM reductase, completely inhibited dechlorination of 1,2-DCA. Purified methyl-CoM reductase, together with flavin adenine dinucleotide and a crude component A fraction which reduced the nickel of factor F430 in methyl-CoM reductase, converted 1,2-DCA to ethylene and CA with H2 as the electron donor. In this system, methyl-CoM reductase was also able to transform its own inhibitor 2-bromoethanesulfonate to ethylene.  相似文献   

5.
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   

6.
Methanol (CH(3)OH), a common industrial solvent, is metabolized to toxic compounds by several enzymatic as well as free radical pathways. Identifying which process best enhances or prevents CH(3)OH-induced cytotoxicity could provide insight into the molecular basis for acute CH(3)OH-induced hepatoxicity. Metabolic pathways studied include those found in 1) an isolated hepatocyte system and 2) cell-free systems. Accelerated Cytotoxicity Mechanism Screening (ACMS) techniques demonstrated that CH(3)OH had little toxicity towards rat hepatocytes in 95% O(2), even at 2M concentration, whereas 50 mM was the estimated LC(50) (2h) in 1% O(2), estimated to be the physiological concentration in the centrilobular region of the liver and also the target region for ethanol toxicity. Cytotoxicity was attributed to increased NADH levels caused by CH(3)OH metabolism, catalyzed by ADH1, resulting in reductive stress, which reduced and released ferrous iron from Ferritin causing oxygen activation. A similar cytotoxic mechanism at 1% O(2) was previous found for ethanol. With 95% O(2), the addition of Fe(II)/H(2)O(2), at non-toxic concentrations were the most effective agents for increasing hepatocyte toxicity induced by 1M CH(3)OH, with a 3-fold increase in cytotoxicity and ROS formation. Iron chelators, desferoxamine, and NADH oxidizers and ATP generators, e.g. fructose, also protected hepatocytes and decreased ROS formation and cytotoxicity. Hepatocyte protein carbonylation induced by formaldehyde (HCHO) formation was also increased about 4-fold, when CH(3)OH was oxidized by the Fenton-like system, Fe(II)/H(2)O(2), and correlated with increased cytotoxicity. In a cell-free bovine serum albumin system, Fe(II)/H(2)O(2) also increased CH(3)OH oxidation as well as HCHO protein carbonylation. Nontoxic ferrous iron and a H(2)O(2) generating system increased HCHO-induced cytotoxicity and hepatocyte protein carbonylation. In addition, HCHO cytotoxicity was markedly increased by ADH1 and ALDH2 inhibitors or GSH-depleted hepatocytes. Increased HCHO concentration levels correlated with increased HCHO-induced protein carbonylation in hepatocytes. These results suggest that CH(3)OH at 1% O(2) involves activation of the Fenton system to form HCHO. However, at higher O(2) levels, radicals generated through Fe(II)/H(2)O(2) can oxidize CH(3)OH/HCHO to form pro-oxidant radicals and lead to increased oxidative stress through protein carbonylation and ROS formation which ultimately causes cell death.  相似文献   

7.
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   

8.
Hydrogen-limited chemostat cultures of Methanobrevibacter arboriphilus A2 were carried out. The available electron balance and carbon balance in M. arboriphilus A2 and other methanogenic strains grown on various substrates were well satisfied. This indicates that no extracellular organic products were formed during methanogenic growth. The molar growth yields for methane (Y(X/CH(4) )) were calculated as 1.06-1.42 g cell/mol CH(4) at dilution rate (0.21-0.43 day(-1)). The smaller Y(X/CH(4) ) of M. arboriphilus A2 compared with that of the other methanogenic strains was probably owing to the low growth rate of M. arboriphilus A2. The low value of Y(X/CH(4) ) may be favorable for methane fermentation because less sludge accumulation is expected. The efficiency of free energy transduction to ATP during methane formation from H(2) + CO(2) was 12-17% at the dilution rate (0.21-0.43 day(-1)) assuming that Y(ATP) was 6.5 g/mol and the free energy change of CO(2) reduction to methane with H(2) was -62.8 kJ/mol under physiological conditions.  相似文献   

9.
Temperature is an important factor controlling CH(4) production in anoxic rice soils. Soil slurries, prepared from Italian rice field soil, were incubated anaerobically in the dark at six temperatures of between 10 to 37 degrees C or in a temperature gradient block covering the same temperature range at intervals of 1 degrees C. Methane production reached quasi-steady state after 60 to 90 days. Steady-state CH(4) production rates increased with temperature, with an apparent activation energy of 61 kJ mol(-1). Steady-state partial pressures of the methanogenic precursor H(2) also increased with increasing temperature from <0.5 to 3.5 Pa, so that the Gibbs free energy change of H(2) plus CO(2)-dependent methanogenesis was kept at -20 to -25 kJ mol of CH(4)(-1) over the whole temperature range. Steady-state concentrations of the methanogenic precursor acetate, on the other hand, increased with decreasing temperature from <5 to 50 microM. Simultaneously, the relative contribution of H(2) as methanogenic precursor decreased, as determined by the conversion of radioactive bicarbonate to (14)CH(4), so that the carbon and electron flow to CH(4) was increasingly dominated by acetate, indicating that psychrotolerant homoacetogenesis was important. The relative composition of the archaeal community was determined by terminal restriction fragment length polymorphism (T-RFLP) analysis of the 16S rRNA genes (16S rDNA). T-RFLP analysis differentiated the archaeal Methanobacteriaceae, Methanomicrobiaceae, Methanosaetaceae, Methanosarcinaceae, and Rice clusters I, III, IV, V, and VI, which were all present in the rice field soil incubated at different temperatures. The 16S rRNA genes of Rice cluster I and Methanosaetaceae were the most frequent methanogenic groups. The relative abundance of Rice cluster I decreased with temperature. The substrates used by this microbial cluster, and thus its function in the microbial community, are unknown. The relative abundance of acetoclastic methanogens, on the other hand, was consistent with their physiology and the acetate concentrations observed at the different temperatures, i.e., the high-acetate-requiring Methanosarcinaceae decreased and the more modest Methanosaetaceae increased with increasing temperature. Our results demonstrate that temperature not only affected the activity but also changed the structure and the function (carbon and electron flow) of a complex methanogenic system.  相似文献   

10.
Cell suspensions of methanogenic bacteria (Methanosarcina barkeri, Methanospirillum hungatei, Methano-brevibacter arboriphilus, and Methanobacterium thermoautotrophicum) were found to form CO from CO2 and H2 according to the reaction: CO2 + H2----CO + H2O; delta G0 = +20 kJ/mol. Up to 15,000 ppm CO in the gas phase were reached which is significantly higher than the equilibrium concentration calculated from delta G0 (95 ppm under the experimental conditions). This indicated that CO2 reduction with H2 to CO is energy-driven and indeed the cells only generated CO when forming CH4. The coupling of the two reactions was studied in more detail with acetate-grown cells of M. barkeri using methanogenic substrates. The effects of the protonophore tetrachlorosalicylanilide (TCS) and of the proton-translocating ATPase inhibitor N,N'-dicyclohexylcarbodiimide (cHxN)2C were determined. TCS completely inhibited CO formation from CO2 and H2 without affecting methanogenesis from CH3OH and H2. In the presence of the protonophore the proton motive force delta p and the intracellular ATP concentration were very low. (cHxN)2C, which partially inhibited methanogenesis from CH3OH and H2, had no effect on CO2 reduction to CO. In the presence of (cHxN)2C delta p was high and the intracellular ATP content was low. These findings suggest that the endergonic formation of CO from CO2 and H2 is coupled to the exergonic formation of CH4 from CH3OH and H2 via the proton motive force and not via ATP. CO formation was not stimulated by the addition of sodium ions.  相似文献   

11.
Using molecular techniques and microsensors for H(2)S and CH(4), we studied the population structure of and the activity distribution in anaerobic aggregates. The aggregates originated from three different types of reactors: a methanogenic reactor, a methanogenic-sulfidogenic reactor, and a sulfidogenic reactor. Microsensor measurements in methanogenic-sulfidogenic aggregates revealed that the activity of sulfate-reducing bacteria (2 to 3 mmol of S(2-) m(-3) s(-1) or 2 x 10(-9) mmol s(-1) per aggregate) was located in a surface layer of 50 to 100 microm thick. The sulfidogenic aggregates contained a wider sulfate-reducing zone (the first 200 to 300 microm from the aggregate surface) with a higher activity (1 to 6 mmol of S(2-) m(-3) s(-1) or 7 x 10(-9) mol s(-1) per aggregate). The methanogenic aggregates did not show significant sulfate-reducing activity. Methanogenic activity in the methanogenic-sulfidogenic aggregates (1 to 2 mmol of CH(4) m(-3) s(-1) or 10(-9) mmol s(-1) per aggregate) and the methanogenic aggregates (2 to 4 mmol of CH(4) m(-3) s(-1) or 5 x 10(-9) mmol s(-1) per aggregate) was located more inward, starting at ca. 100 microm from the aggregate surface. The methanogenic activity was not affected by 10 mM sulfate during a 1-day incubation. The sulfidogenic and methanogenic activities were independent of the type of electron donor (acetate, propionate, ethanol, or H(2)), but the substrates were metabolized in different zones. The localization of the populations corresponded to the microsensor data. A distinct layered structure was found in the methanogenic-sulfidogenic aggregates, with sulfate-reducing bacteria in the outer 50 to 100 microm, methanogens in the inner part, and Eubacteria spp. (partly syntrophic bacteria) filling the gap between sulfate-reducing and methanogenic bacteria. In methanogenic aggregates, few sulfate-reducing bacteria were detected, while methanogens were found in the core. In the sulfidogenic aggregates, sulfate-reducing bacteria were present in the outer 300 microm, and methanogens were distributed over the inner part in clusters with syntrophic bacteria.  相似文献   

12.
1. The oxyhydrogen reaction of Anacystis nidulans was studied manometrically and polarographically in whole cells and in cell-free preparations; the activity was found to be associated with the particulate fraction. 2. Besides O2, the isolated membranes reduced artificial electron acceptors of positive redox potential; the reactions were unaffected by O2 levels less than 10--15%; aerobically the artificial acceptors were reduced simultaneously with O2. 3. H2-supported O2 uptake was inhibited by CO, KCN and 2-n-heptyl-8-hydroxyquinoline-N-oxide. Inhibition by CO was partly reversed by strong light. Uncouplers stimulated the oxyhydrogen reaction. 4. The kinetic properties of O2 uptake by isolated membranes were the same in presence of H2 and of other respiratory substrates. 5. Low rates of H2 evolution by the membrane preparations were found in presence of dithionite; methyl viologen stimulated the reaction. 6. The results indicate that under certain growth conditions Anacystis synthesizes a membrane-bound hydrogenase which appears to be involved in phosphorylative electron flow from H2 to O2 through the respiratory chain.  相似文献   

13.
The anaerobic cellulolytic rumen bacterium Ruminococcus flavefaciens normally produces succinic acid as a major fermentation product together with acetic and formic acids, H2, and CO2. When grown on cellulose and in the presence of the methanogenic rumen bacterium Methanobacterium ruminantium, acetate was the major fermentation product; succinate was formed in small amounts; little formate was detected; H2 did not accumulate; and large amounts of CH4 were formed. M. ruminantium depends for growth on the reduction of CO2 to CH4 by H2, which it can obtain directly or by producing H2 and CO2 from formate. In mixed culture, the methanobacterium utilized the H2 and possibly the formate produced by the ruminococcus and in so doing stimulated the flow of electrons generated during glycolysis by the ruminococcus toward H2 formation and away from formation of succinate. This type of interaction may be of significance in determining the flow of cellulose carbon to the normal rumen fermentation products.  相似文献   

14.
The emission of methane (1.3 mmol of CH(4) m(-2) day(-1)), precursors of methanogenesis, and the methanogenic microorganisms of acidic bog peat (pH 4.4) from a moderately reduced forest site were investigated by in situ measurements, microcosm incubations, and cultivation methods, respectively. Bog peat produced CH(4) (0.4 to 1.7 micro mol g [dry wt] of soil(-1) day(-1)) under anoxic conditions. At in situ pH, supplemental H(2)-CO(2), ethanol, and 1-propanol all increased CH(4) production rates while formate, acetate, propionate, and butyrate inhibited the production of CH(4); methanol had no effect. H(2)-dependent acetogenesis occurred in H(2)-CO(2)-supplemented bog peat only after extended incubation periods. Nonsupplemented bog peat initially produced small amounts of H(2) that were subsequently consumed. The accumulation of H(2) was stimulated by ethanol and 1-propanol or by inhibiting methanogenesis with bromoethanesulfonate, and the consumption of ethanol was inhibited by large amounts of H(2); these results collectively indicated that ethanol- or 1-propanol-utilizing bacteria were trophically associated with H(2)-utilizing methanogens. A total of 10(9) anaerobes and 10(7) hydrogenotrophic methanogens per g (dry weight) of bog peat were enumerated by cultivation techniques. A stable methanogenic enrichment was obtained with an acidic, H(2)-CO(2)-supplemented, fatty acid-enriched defined medium. CH(4) production rates by the enrichment were similar at pH 4.5 and 6.5, and acetate inhibited methanogenesis at pH 4.5 but not at pH 6.5. A total of 27 different archaeal 16S rRNA gene sequences indicative of Methanobacteriaceae, Methanomicrobiales, and Methanosarcinaceae were retrieved from the highest CH(4)-positive serial dilutions of bog peat and methanogenic enrichments. A total of 10 bacterial 16S rRNA gene sequences were also retrieved from the same dilutions and enrichments and were indicative of bacteria that might be responsible for the production of H(2) that could be used by hydrogenotrophic methanogens. These results indicated that in this acidic bog peat, (i) H(2) is an important substrate for acid-tolerant methanogens, (ii) interspecies hydrogen transfer is involved in the degradation of organic carbon, (iii) the accumulation of protonated volatile fatty acids inhibits methanogenesis, and (iv) methanogenesis might be due to the activities of methanogens that are phylogenetic members of the Methanobacteriaceae, Methanomicrobiales, and Methanosarcinaceae.  相似文献   

15.
We examined the mechanisms underlying the activation of group IVA cytosolic phospholipase A(2) (cPLA(2)alpha) contributing to the supply of fatty acids required for the formation of cholesteryl ester in oxidized low-density lipoprotein (oxLDL)-stimulated macrophages. The possible involvement of oxidized lipids was also examined. In [(3)H]arachidonic acid-labeled mouse peritoneal macrophages, oxLDL stimulated the release of arachidonic acid, which was suppressed by methyl arachidonyl fluorophosphonate (MAFP), a cPLA(2)alpha inhibitor. oxLDL induced an increase in PLA(2)alpha levels in the membrane fraction without affecting those in whole cells or the activity in the lysate. Among 13-hydroxyoctadecadienoic acid (13-HODE), 7-ketocholesterol, and 25-hydroxycholesterol, oxidized lipids present in oxLDL particles, only 13-HODE induced the release of arachidonic acid, which was also sensitive to MAFP. Under conditions where addition of Ca(2+) to the cell lysate induced an increase in cPLA(2)alpha protein in the membrane fraction, preincubation with 13-HODE facilitated the Ca(2+)-dependent translocation of cPLA(2)alpha. Furthermore, 13-HODE increased cholesteryl ester formation in the presence of [(3)H]cholesterol. These results suggest that 13-HODE mediates the oxLDL-induced activation of cPLA(2)alpha through an increase in cPLA(2)alpha protein in the membranes, thus contributing, in part, to the supply of fatty acids required for the esterification of cholesterol in macrophages.  相似文献   

16.
1. Anaerobic hydrogenase activity in whole cells and cell-free preparations of H2-induced Anacystis was studied both manometrically and spectrophotometrically in presence of physiological and artificial electron acceptors. 2. Up to 90% of the activity measured in crude extracts were recovered in the chlorophyll-containing membrane fraction after centrifugation (144 000 X g, 3 h). 3. Reduction of methyl viologen, diquat, ferredoxin, nitrite and NADP by the membranes was light dependent while oxidants of more positive redox potential were reduced also in the dark. 4. Evolution of H2 by the membranes was obtained with dithionite and with reduced methyl viologen; the reaction was stimulated by detergents. 5. Both uptake and evolution of H2 were sensitive to O2, CO, and thiolblocking agents. The H2-dependent reductions were inhibited also by the plastoquinone antagonist dibromothymoquinone, while the ferredoxin inhibitor disalicylidenepropanediamine affected the photoreduction of nitrite and NADP only. 3-(3,4-Dichlorophenyl)-1,1-dimethylurea did not inhibit any one of the H2-dependent reactions. 6. The results present evidence for a membrane-bound 'photoreduction' hydrogenase in H2-induced Anacystis. The enzyme apparently initiates a light-driven electron flow from H2 to various low-potential acceptors including endogenous ferredoxin.  相似文献   

17.
18.
The bioenergetics of methanogenesis   总被引:29,自引:0,他引:29  
The reduction of CO2 or any other methanogenic substrate to methane serves the same function as the reduction of oxygen, nitrate or sulfate to more reduced products. These exergonic reactions are coupled to the production of usable energy generated through a charge separation and a protonmotive-force-driven ATPase. For the understanding of how methanogens derive energy from C-1 unit reduction one must study the biochemistry of the chemical reactions involved and how these are coupled to the production of a charge separation and subsequent electron transport phosphorylation. Data on methanogenesis by a variety of organisms indicates ubiquitous use of CH3-S-CoM as the final electron acceptor in the production of methane through the methyl CoM reductase and of 5-deazaflavin as a primary source of reducing equivalents. Three known enzymes serve as catalysts in the production of reduced 5-deazaflavin: hydrogenase, formate dehydrogenase and CO dehydrogenase. All three are potential candidates for proton pumps. In the organisms that must oxidize some of their substrate to obtain electrons for the reduction of another portion of the substrate to methane (e.g., those using formate, methanol or acetate), the latter two enzymes may operate in the oxidizing direction. CO2 is the most frequent substrate for methanogenesis but is the only substrate that obligately requires the presence of H2 and hydrogenase. Growth on methanol requires a B12-containing methanol-CoM methyl transferase and does not necessarily need any other methanogenic enzymes besides the methyl-CoM reductase system when hydrogenase is present. When bacteria grow on methanol alone it is not yet clear if they get their reducing equivalents from a reversal of methanogenic enzymes, thus oxidizing methyl groups to CO2. An alternative (since these and acetate-catabolizing methanogens possess cytochrome b) is electron transport and possible proton pumping via a cytochrome-containing electron transport chain. Several of the actual components of the methanogenic pathway from CO2 have been characterized. Methanofuran is apparently the first carbon-carrying cofactor in the pathway, forming carboxy-methanofuran. Formyl-FAF or formyl-methanopterin (YFC, a very rapidly labelled compound during 14C pulse labeling) has been implicated as an obligate intermediate in methanogenesis, since methanopterin or FAF is an essential component of the carbon dioxide reducing factor in dialyzed extract methanogenesis. FAF also carries the carbon at the methylene and methyl oxidation levels.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

19.
The activities of pure and mixed cultures of Desulfovibrio vulgaris and Methanosarcina barkeri in the exponential growth phase were monitored by measuring changes in dissolved-gas concentration by membrane-inlet mass spectrometry. M. barkeri grown under H2-CO2 or methanol produced limited amounts of methane and practically no hydrogen from either substrate. The addition of CO resulted in a transient H2 production concomitant with CO consumption. Hydrogen was then taken up, and CH4 production increased. All these events were suppressed by KCN, which inhibited carbon monoxide dehydrogenase activity. Therefore, with both substrates, H2 appeared to be an intermediate in CO reduction to CH4. The cells grown on H2-CO2 consumed 4 mol of CO and produced 1 mol of CH4. Methanol-grown cells reduced CH3OH with H2 resulting from carbon monoxide dehydrogenase activity, and the ratio was then 1 mol of CH4 to 1 mol of CO. Only 12CH4 and no 13CH4 was obtained from 13CO, indicating that CO could not be the direct precursor of CH4. In mixed cultures of D. vulgaris and M. barkeri on lactate, an initial burst of H2 was observed, followed by a lower level of production, whereas methane synthesis was linear with time. Addition of CO to the mixed culture also resulted in transient extra H2 production but had no inhibitory effect upon CH4 formation, even when the sulfate reducer was D. vulgaris Hildenborough, whose periplasmic iron hydrogenase is very sensitive to CO. The hydrogen transfer is therefore probably mediated by a less CO-sensitive nickel-iron hydrogenase from either of both species.  相似文献   

20.
To further consider the thermochemical method as a useful approach for active transport research and to investigate the characteristic of a proton electrochemical potential (delta mu H+) across the membrane, the energetics of lactose active transport across Escherichia coli membrane vesicles coupled with an artificial electron donor (phenazine methosulfate-ascorbate) has been investigated. The results were compared with those obtained with an enzyme-associated electron donor (lactate dehydrogenase-D-lactate). The oxidation of an electron donor provided the energy necessary for the transport process. The observed higher heat of ascorbate oxidation reaction in the presence of a proton ionophore (carbonyl cyanide m-chlorophenylhydrazone) further confirmed the formation of delta mu H+ across the membrane. Part of the oxidation energy was utilized to form delta mu H+. Comparison of the energetics revealed that the magnitudes of delta Hox (the enthalpy of the oxidation reaction) and delta Hm (the enthalpy of the formation of delta mu H+) in the two energy sources were comparable (-46 kcal/mol of ascorbate to -40 kcal/mol of D-lactate for delta Hox and 9.6 kcal/mol of ascorbate to 14 kcal/mol of D-lactate for delta Hm). Comparable and low value (about 1%) was also found in the free energy transfer (defined by delta Gm/delta Gox) from the oxidation reaction to the formation of delta mu H+. These results, in combination with the close values of delta mu H+ observed in the two systems, suggested that the characteristic of the created delta mu H+ was independent of the energy source. Examination of delta Hm might provide the information on the ratio of the number of protons produced, as 1 mol of two different electron donors was oxidized. The oxidation reaction in the presence of membrane vesicles was discussed.  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