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 46 毫秒
1.
Upflow anaerobic sludge blanket bioreactor granules were used as an experimental model microbial consortium to study the dynamics and distribution of methanogens. Immunologic methods revealed a considerable diversity of methanogens that was greater in mesophilic granules than in the same granules 4 months after a temperature shift from 38 to 55°C. During this period, the sizes of the methanogenic subpopulations changed with distinctive profiles after the initial reduction caused by the shift. Methanogens antigenically related to Methanobrevibacter smithii PS and ALI, Methanobacterium hungatei JF1, and Methanosarcina thermophila TM1 increased rapidly, reached a short plateau, and then fell to lower concentrations that persisted for the duration of the experiment. A methanogen related to Methanogenium cariaci JR1 followed a similar profile at the beginning, but it soon diminished below detection levels. Methanothrix rods weakly related to the strain Opfikon increased rapidly, reaching a high-level, long-lasting plateau. Two methanogens related to Methanobrevibacter arboriphilus AZ and Methanobacterium thermoautotrophicum ΔH emerged from very low levels before the temperature shift and multiplied to attain their highest numbers 4 months after the shift. Histochemistry and immunohistochemistry revealed thick layers, globular clusters, and lawns of variable density which were distinctive of the methanogens related to M. thermoautotrophicum ΔH, M. thermophila TM1, and M. arboriphilus AZ and M. soehngenii Opfikon, respectively, in thin sections of granules grown at 55°C for 4 months. Mesophilic granules showed a different pattern of methanogenic subpopulations.  相似文献   

2.
The thermophilic, anaerobic, propionate-oxidizing bacterial populations present in the methanogenic granular sludge in a thermophilic (55°C) upflow anaerobic sludge blanket reactor were studied by cultivation and in situ hybridization analysis. For isolation of propionate-degrading microbes, primary enrichment was made with propionate as the sole energy source at 55°C. After several attempts to purify the microbes, a thermophilic, syntrophic, propionate-oxidizing bacterium, designated strain SI, was isolated in both pure culture and coculture with Methanobacterium thermoautotrophicum. Under thermophilic (55°C) conditions, strain SI oxidized propionate, ethanol, and lactate in coculture with M. thermoautotrophicum. In pure culture, the isolate was found to ferment pyruvate. 16S ribosomal DNA sequence analysis revealed that the strain was relatively close to members of the genus Desulfotomaculum, but it was only distantly related to any known species. To elucidate the abundance and spatial distribution of organisms of the strain SI type within the sludge granules, a 16S rRNA-targeted oligonucleotide probe specific for strain SI was developed and applied to thin sections of the granules. Fluorescence in situ hybridization combined with confocal laser scanning microscopy revealed that a number of rod-shaped cells were present in the middle and inner layers of the thermophilic granule sections and that they formed close associations with hydrogenotrophic methanogens. They accounted for approximately 1.1% of the total cells in the sludge. These results demonstrated that strain SI was one of the significant populations in the granular sludge and that it was responsible for propionate oxidation in the methanogenic granular sludge in the reactor.  相似文献   

3.
Oxidation of succinate to fumarate is an energetically difficult step in the biochemical pathway of propionate oxidation by syntrophic methanogenic cultures. Therefore, the effect of fumarate on propionate oxidation by two different propionate-oxidizing cultures was investigated. When the methanogens in a newly enriched propionate-oxidizing methanogenic culture were inhibited by bromoethanesulfonate, fumarate could act as an apparent terminal electron acceptor in propionate oxidation. 13C-nuclear magnetic resonance experiments showed that propionate was carboxylated to succinate while fumarate was partly oxidized to acetate and partly reduced to succinate. Fumarate alone was fermented to succinate and CO2. Bacteria growing on fumarate were enriched and obtained free of methanogens. Propionate was metabolized by these bacteria when either fumarate or Methanospirillum hungatii was added. In cocultures with Syntrophobacter wolinii, such effects were not observed upon addition of fumarate. Possible slow growth of S. wolinii on fumarate could not be demonstrated because of the presence of a Desulfovibrio strain which grew rapidly on fumarate in both the absence and presence of sulfate.  相似文献   

4.
Experiments document the ability of two species of autotrophic methanogens to assimilate and utilize organic substrates as the nutrient sulfur or nitrogen source and as a carbon source during growth on H2-CO2. Methanobacterium thermoautotrophicum strain ΔH and the mesophilic species Methanobacterium sp. strain Ivanov grew with glutamine as the nitrogen source or cysteine as the sulfur source. M. thermoautotrophicum also utilized urea as the nitrogen source and as a carbon precursor for methane and cell synthesis. Methanobacterium sp. strain Ivanov grew with methionine as the sulfur source. The growth rate of two different Methanobacterium species was lower on an organic N or S source than on ammonium or sulfide. 35S and 14C tracer studies demonstrated that amino acid or urea assimilation correlated with time and amount of growth. The rate of [35S]cysteine incorporation was similar in strain ΔH (34 nmol h−1 mg of cells−1) and strain Ivanov (23 nmol h−1 mg of cells−1). However, the rate of [14C]acetate incorporation was dramatically different (17 versus 208 nmol h−1 mg of cells−1 in strains ΔH and Ivanov, respectively). [14C]acetate accounted for 1.3 and 21.2% of the total cell carbon synthesized by strains ΔH and Ivanov, respectively. Amino acids and urea were mainly assimilated into the cell protein fraction, but accounted for less than 2.0% of the total cell carbon synthesized. The data suggest that a biochemical-genetic approach to understanding cell carbon synthesis in methanogens is feasible; mutants that are auxotrophic for either acetate, glutamine, cysteine, or methionine are suggested as future targets for genetic studies.  相似文献   

5.
Hydrogen partial pressures were measured in a thermophilic coculture comprised of a eubacterial rod which oxidized acetate to H2 and CO2 and a hydrogenotrophic methanogen, Methanobacterium sp. strain THF. Zinder and Koch (S. H. Zinder and M. Koch, Arch. Microbiol. 138:263-272, 1984) originally predicted, on the basis of calculations of Gibbs free energies of reactions, that the H2 partial pressure near the midpoint of growth of the coculture should be near 4 Pa (ca. 4 × 10−5 atm; ca. 0.024 μM dissolved H2) for both organisms to be able to conserve energy for growth. H2 partial pressures in the coculture were measured to be between 20 and 50 Pa (0.12 to 0.30 μM) during acetate utilization, approximately one order of magnitude higher than originally predicted. However, when ΔGf (free energy of formation) values were corrected for 60°C by using the relationship ΔGf = ΔHfTΔSHf is the enthalpy or heat of formation, ΔS is the entropy value, and T is the temperature in kelvins), the predicted value was near 15 Pa, in closer agreement with the experimentally determined values. The coculture also oxidized ethanol to acetate, a more thermodynamically favorable reaction than oxidation of acetate to CO2. During ethanol oxidation, the H2 partial pressure reached values as high as 200 Pa. Acetate was not used until after the ethanol was consumed and the H2 partial pressure decreased to 40 to 50 Pa. After acetate utilization, H2 partial pressures fell to approximately 10 Pa and remained there, indicating a threshold for H2 utilization by the methanogen. Axenic cultures of the acetate-oxidizing organism were combined with pure cultures of either Methanobacterium sp. strain THF or Methanobacterium thermoautotrophicum ΔH to form reconstituted acetate-oxidizing cocultures. The H2 partial pressures measured in both of these reconstituted cocultures were similar to those measured in the original acetate-oxidizing rod coculture. Since M. thermoautotrophicum ΔH did not use formate as a substrate, formate is not necessarily involved in interspecies electron transfer in this coculture.  相似文献   

6.
Microbiology of Methanogenesis in Thermal, Volcanic Environments   总被引:33,自引:5,他引:28       下载免费PDF全文
Microbial methanogenesis was examined in thermal waters, muds, and decomposing algal-bacterial mats associated with volcanic activity in Yellowstone National Park. Radioactive tracer studies with [14C]glucose, acetate, or carbonate and enrichment culture techniques demonstrated that methanogenesis occurred at temperatures near 70°C but below 80°C and correlated with hydrogen production from either geothermal processes or microbial fermentation. Three Methanobacterium thermoautotrophicum strains (YT1, YTA, and YTC) isolated from diverse volcanic habitats differed from the neotype sewage strain ΔH in deoxyribonucleic acid guanosine-plus-cytosine content and immunological properties. Microbial methanogenesis was characterized in more detail at a 65°C site in the Octopus Spring algal-bacterial mat ecosystem. Here methanogenesis was active, was associated with anaerobic microbial decomposition of biomass, occurred concomitantly with detectable microbial hydrogen formation, and displayed a temperature activity optimum near 65°C. Enumeration studies estimated more than 109 chemoorganotrophic hydrolytic bacteria and 106 chemolithotrophic methanogenic bacteria per g (dry weight) of algal-bacterial mat. Enumeration, enrichment, and isolation studies revealed that the microbial population was predominantly rod shaped and asporogenous. A prevalent chemoorganotrophic organism in the mat that was isolated from an end dilution tube was a taxonomically undescribed gram-negative obligate anaerobe (strain HTB2), whereas a prevalent chemolithotrophic methanogen isolated from an end dilution tube was identified as M. thermoautotrophicum (strain YTB). Taxonomically recognizable obligate anaerobes that were isolated from glucose and xylose enrichment cultures included Thermoanaerobium brockii strain HTB and Clostridium thermohydrosulfuricum strain 39E. The nutritional properties, growth temperature optima, growth rates, and fermentation products of thermophilic bacterial strains 39E, HTB2, and YTB were determined.  相似文献   

7.
Three microtitration plate enzyme-linked immunosorbent assays (ELISAs) have been developed: a competitive ELISA and a two-site (or indirect sandwich) ELISA for Methanosarcina mazei S6 and a two-site ELISA for Methanobacterium bryantii FR-2. The assays were sensitive, with limits of cell protein detection of 3 ng ml−1, 5 ng ml−1, and 50 ng ml−1, respectively, and showed good precision. The M. mazei assays used monoclonal antibodies and were entirely species specific, showing no cross-reaction with methanogens of other genera or with other species of the same genus. The Methanobacterium bryantii assay, which used two polyclonal antisera, showed only a slight cross-reaction with one other Methanobacterium species but no cross-reaction with methanogens of other genera. The use of the ELISAs for quantitative analysis of mixed cultures and of sewage sludge samples was investigated. Sludge diluted at 1:103 or more caused no significant interference in any of the three ELISAs. Various cultures of bacteria, methanogens, and nonmethanogens at a protein concentration of 50 μg ml−1 showed no significant interference in the M. mazei competitive assay and the Methanobacterium bryantii two-site assay, although they did cause falsely low results in the M. mazei two-site assay.  相似文献   

8.
Several methanogenic bacteria were shown to produce ethane, ethylene, and acetylene when exposed to the halogenated hydrocarbons bromoethane, dibromo- or dichloroethane, and 1,2-dibromoethylene, respectively. They also produced ethylene when exposed to the coenzyme M analog and specific methanogenic inhibitor bromoethanesulfonic acid. The production of these gases from halogenated hydrocarbons has a variety of implications concerning microbial ecology, agriculture, and toxic waste treatment. All halogenated aliphatic compounds tested were inhibitory to methanogens. Methanococcus thermolithotrophicus, Methanococcus deltae, and Methanobacterium thermoautotrophicum ΔH and Marburg were completely inhibited by 7 μM 1,2-dibromoethane and, to various degrees, by 51 to 1,084 μM 1,2-dichloroethane, 1,2-dibromoethylene, 1,2-dichloroethylene, and trichloroethylene. In general, the brominated compounds were more inhibitory. The two Methanococcus species were fully inhibited by 1 μM bromoethanesulfonic acid, whereas both Methanobacterium strains were only partly inhibited by 2,124 μM. Coenzyme M protected cells from bromoethanesulfonic acid but not from any of the other inhibitors.  相似文献   

9.
A thermophilic syntrophic bacterium, Pelotomaculum thermopropionicum strain SI, was grown in a monoculture or coculture with a hydrogenotrophic methanogen, Methanothermobacter thermautotrophicus strain ΔH. Microscopic observation revealed that cells of each organism were dispersed in a monoculture independent of the growth substrate. In a coculture, however, these organisms coaggregated to different degrees depending on the substrate; namely, a large fraction of the cells coaggregated when they were grown on propionate, but relatively few cells coaggregated when they were grown on ethanol or 1-propanol. Field emission-scanning electron microscopy revealed that flagellum-like filaments of SI cells played a role in making contact with ΔH cells. Microscopic observation of aggregates also showed that extracellular polymeric substance-like structures were present in intercellular spaces. In order to evaluate the importance of coaggregation for syntrophic propionate oxidation, allowable average distances between SI and ΔH cells for accomplishing efficient interspecies hydrogen transfer were calculated by using Fick's diffusion law. The allowable distance for syntrophic propionate oxidation was estimated to be approximately 2 μm, while the allowable distances for ethanol and propanol oxidation were 16 μm and 32 μm, respectively. Considering that the mean cell-to-cell distance in the randomly dispersed culture was approximately 30 μm (at a concentration in the mid-exponential growth phase of the coculture of 5 × 107 cells ml−1), it is obvious that close physical contact of these organisms by coaggregation is indispensable for efficient syntrophic propionate oxidation.  相似文献   

10.
Studies on product inhibition of a thermophilic butyrate-degrading bacterium in syntrophic association with Methanobacterium thermoautotrophicum showed that a gas phase containing more than 2 × 10−2 atm (2.03 kPa) of hydrogen prevented growth and butyrate consumption, while a lower hydrogen partial pressure of 1 × 10−3 to 2 × 10−2 atm (0.1 to 2.03 kPa) gradually inhibited the butyrate consumption of the coculture. No inhibition of butyrate consumption was found on the addition of 0.75 × 10−3 atm (76 Pa) of hydrogen to the gas phase. A slight inhibition of butyrate consumption by the coculture occurred at an acetate concentration of 16.4 mM. Inhibition gradually increased with increasing acetate concentration up to 81.4 mM, when complete inhibition of butyrate consumption occurred. When the culture contained an acetate-utilizing methanogen in addition to M. thermoautotrophicum, the inhibition of the triculture by acetate was gradually reversed as the acetate concentration was lowered by the aceticlastic methanogen. The results show that optimal growth conditions for the thermophilic butyrate-degrading bacterium depend on both hydrogen and acetate removal.  相似文献   

11.
Cell-free extracts of Methanobacterium thermoautotrophicum were found to contain high activities of the following oxidoreductases (at 60°C): pyruvate dehydrogenase (coenzyme A acetylating), 275 nmol/min per mg of protein; α-ketoglutarate dehydrogenase (coenzyme A acylating), 100 nmol/min per mg; fumarate reductase, 360 nmol/min per mg; malate dehydrogenase, 240 nmol/min per mg; and glyceraldehyde-3-phosphate dehydrogenase, 100 nmol/min per mg. The kinetic properties (apparent Vmax and KM values), pH optimum, temperature dependence of the rate, and specificity for electron acceptors/donors of the different oxidoreductases were examined. Pyruvate dehydrogenase and α-ketoglutarate dehydrogenase were shown to be two separate enzymes specific for factor 420 rather than for nicotinamide adenine dinucleotide (NAD), NADP, or ferredoxin as the electron acceptor. Both activities catalyzed the reduction of methyl viologen with the respective α-ketoacid and a coenzyme A-dependent exchange between the carboxyl group of the α-ketoacid and CO2. The data indicate that the two enzymes are similar to pyruvate synthase and α-ketoglutarate synthase, respectively. Fumarate reductase was found in the soluble cell fraction. This enzyme activity coupled with reduced benzyl viologen as the electron donor, but reduced factor 420, NADH, or NADPH was not effective. The cells did not contain menaquinone, thus excluding this compound as the physiological electron donor for fumarate reduction. NAD was the preferred coenzyme for malate dehydrogenase, whereas NADP was preferred for glyceraldehyde-3-phosphate dehydrogenase. The organism also possessed a factor 420-dependent hydrogenase and a factor 420-linked NADP reductase. The involvement of the described oxidoreductases in cell carbon synthesis is discussed.  相似文献   

12.
Despite the fact that rice paddy fields (RPFs) are contributing 10 to 25% of global methane emissions, the organisms responsible for methane production in RPFs have remained uncultivated and thus uncharacterized. Here we report the isolation of a methanogen (strain SANAE) belonging to an abundant and ubiquitous group of methanogens called rice cluster I (RC-I) previously identified as an ecologically important microbial component via culture-independent analyses. To enrich the RC-I methanogens from rice paddy samples, we attempted to mimic the in situ conditions of RC-I on the basis of the idea that methanogens in such ecosystems should thrive by receiving low concentrations of substrate (H2) continuously provided by heterotrophic H2-producing bacteria. For this purpose, we developed a coculture method using an indirect substrate (propionate) in defined medium and a propionate-oxidizing, H2-producing syntroph, Syntrophobacter fumaroxidans, as the H2 supplier. By doing so, we significantly enriched the RC-I methanogens and eventually obtained a methanogen within the RC-I group in pure culture. This is the first report on the isolation of a methanogen within RC-I.  相似文献   

13.
Methanogenic bacteria contain high activities of fumarate reductase. An interesting hypothesis has recently been advanced that this enzyme, in cooperation with a succinate dehydrogenase, functions in a fumarate-succinate cycle for ATP synthesis. This hypothesis was tested by determining whether [2, 3-3H] succinate loses3H when taken up by growing cells.Methanobacterium thermoautotrophicum was grown on H2 plus CO2 in the presence of [U-14C, 2,3-3H] succinate. The double labelled dicarboxylic acid was found to be incorporated into cell material with the loss of only 30% of tritium. Neither was3H released into H2O in significant amounts. This finding excludes a catabolic oxidation of succinate to fumarate in the growing cells and thus the operation of a fumaratesuccinate cycle. It is shown that the function of fumarate reductase inM. thermoautotrophicum is to provide the cells with succinate for the synthesis of -ketoglutarate, an intermediate in glutamate, arginine and proline synthesis.  相似文献   

14.
The kinetics of formate metabolism in Methanobacterium formicicum and Methanospirillum hungatei were studied with log-phase formate-grown cultures. The progress of formate degradation was followed by the formyltetrahydrofolate synthetase assay for formate and fitted to the integrated form of the Michaelis-Menten equation. The Km and Vmax values for Methanobacterium formicicum were 0.58 mM formate and 0.037 mol of formate h−1 g−1 (dry weight), respectively. The lowest concentration of formate metabolized by Methanobacterium formicicum was 26 μM. The Km and Vmax values for Methanospirillum hungatei were 0.22 mM and 0.044 mol of formate h−1 g−1 (dry weight), respectively. The lowest concentration of formate metabolized by Methanospirillum hungatei was 15 μM. The apparent Km for formate by formate dehydrogenase in cell-free extracts of Methanospirillum hungatei was 0.11 mM. The Km for H2 uptake by cultures of Methanobacterium formicicum was 6 μM dissolved H2. Formate and H2 were equivalent electron donors for methanogenesis when both substrates were above saturation; however, H2 uptake was severely depressed when formate was above saturation and the dissolved H2 was below 6 μM. Formate-grown cultures of Methanobacterium formicicum that were substrate limited for 57 h showed an immediate increase in growth and methanogenesis when formate was added to above saturation.  相似文献   

15.
A mesophilic acetogenic bacterium (MPOB) oxidized propionate to acetate and CO2 in cocultures with the formate- and hydrogen-utilizing methanogens Methanospirillum hungatei and Methanobacterium formicicum. Propionate oxidation did not occur in cocultures with two Methanobrevibacter strains, which grew only with hydrogen. Tricultures consisting of MPOB, one of the Methanobrevibacter strains, and organisms which are able to convert formate into H2 plus CO2 (Desulfovibrio strain G11 or the homoacetogenic bacterium EE121) also degraded propionate. The MPOB, in the absence of methanogens, was able to couple propionate conversion to fumarate reduction. This propionate conversion was inhibited by hydrogen and by formate. Formate and hydrogen blocked the energetically unfavorable succinate oxidation to fumarate involved in propionate catabolism. Low formate and hydrogen concentrations are required for the syntrophic degradation of propionate by MPOB. In triculture with Methanospirillum hungatei and the aceticlastic Methanothrix soehngenii, propionate was degraded faster than in biculture with Methanospirillum hungatei, indicating that low acetate concentrations are favorable for propionate oxidation as well.  相似文献   

16.
Lipids can be anaerobically digested to methane, but methanogens are often considered to be highly sensitive to the long-chain fatty acids (LCFA) deriving from lipids hydrolysis. In this study, the effect of unsaturated (oleate [C18:1]) and saturated (stearate [C18:0] and palmitate [C16:0]) LCFA toward methanogenic archaea was studied in batch enrichments and in pure cultures. Overall, oleate had a more stringent effect on methanogens than saturated LCFA, and the degree of tolerance to LCFA was different among distinct species of methanogens. Methanobacterium formicicum was able to grow in both oleate- and palmitate-degrading enrichments (OM and PM cultures, respectively), whereas Methanospirillum hungatei only survived in a PM culture. The two acetoclastic methanogens tested, Methanosarcina mazei and Methanosaeta concilii, could be detected in both enrichment cultures, with better survival in PM cultures than in OM cultures. Viability tests using live/dead staining further confirmed that exponential growth-phase cultures of M. hungatei are more sensitive to oleate than are M. formicicum cultures; exposure to 0.5 mM oleate damaged 99% ± 1% of the cell membranes of M. hungatei and 53% ± 10% of the cell membranes of M. formicicum. In terms of methanogenic activity, M. hungatei was inhibited for 50% by 0.3, 0.4, and 1 mM oleate, stearate, and palmitate, respectively. M. formicicum was more resilient, since 1 mM oleate and >4 mM stearate or palmitate was needed to cause 50% inhibition on methanogenic activity.  相似文献   

17.
Methanobacterium espanolae, an acidiphilic methanogen, required acetate for maximal growth on H2-CO2. In the presence of 5 to 15 mM acetate, at a growth pH of 5.5, the μmax was 0.05 h-1. M. espanolae consumed 12.3 mM acetate during 96 h of incubation at 35°C with shaking at 100 rpm. At initial acetate levels of 2.5 to 10.0 mM, the amount of biomass produced was dependent on the amount of acetate in the medium. 13C nuclear magnetic resonance spectra of protein hydrolysates obtained from cultures grown on [1-13C]- or [2-13C]acetate indicated that an incomplete tricarboxylic acid pathway, operating in the reductive direction, was functional in this methanogen. The amino acids were labeled with a very high degree of specificity and at greater than 90% enrichment levels. Less than 2% label randomization occurred between positions primarily labeled from either the carboxyl or methyl group of acetate, and very little label was transferred to positions primarily labeled from CO2. The labeling pattern of carbohydrates was typical for glucogenesis from pyruvate. This methanogen, by virtue of the properties described above and its ability to incorporate all of the available acetate (10 mM or lower) from the growth medium, has advantages over other microorganisms for use in the production of specifically labeled compounds.  相似文献   

18.
19.
Metabolism of Formate in Methanobacterium formicicum   总被引:24,自引:10,他引:14       下载免费PDF全文
Methanobacterium formicicum strain JF-1 was cultured with formate as the sole energy source in a pH-stat fermentor. Growth was exponential, and both methane production and formate consumption were linear functions of the growth rate. Hydrogen was produced in only trace amounts, and the dissolved H2 concentration of the culture medium was below 1 μM. The effect of temperature or pH on the rate of methane formation was studied with a single fermentor culture in mid-log phase that was grown with formate under standard conditions at 37°C and pH 7.6. Methane formation from formate occurred over the pH range from 6.5 to 8.6, with a maximum at pH 8.0. The maximum temperature of methanogenesis was 56°C. H2 production increased at higher temperatures. Hydrogen and formate were consumed throughout growth when both were present in saturating concentrations. The molar growth yields were 1.2 ± 0.06 g (dry weight) per mol of formate and 4.8 ± 0.24 g (dry weight) per mol of methane. Characteristics were compared for cultures grown with either formate or H2-CO2 as the sole energy source at 37°C and pH 7.6; the molar growth yield for methane of formate cultures was 4.8 g (dry weight) per mol, and that of H2-CO2 cultures was 3.5 g (dry weight) per mol. Both formate and H2-CO2 cultures had low efficiencies of electron transport phosphorylation; formate-cultured cells had greater specific activities of coenzyme F420 than did H2-CO2-grown cultures. Hydrogenase, formate dehydrogenase, chromophoric factor F342, and low levels of formyltetrahydrofolate synthetase were present in cells cultured with either substrate. Methyl viologen-dependent formate dehydrogenase was found in the soluble fraction from broken cells.  相似文献   

20.
The sensitivity of three methanogenic bacteria towards ultraviolet irradiation was similar to the UV-sensitivity of Escherichia coli. The lethal effects of UV-irradiation in Methanobacterium thermoautotrophicum Marburg and in Methanobacterium thermoautotrophicum H but not in Methanococcus vannielii were reversed by exposure to visible light. In cell suspensions of Methanobacterium thermoautotrophicum that had been irradiated to 0.1% survival, 90% of the UV-caused damage was photorepairable. The in vivo action spectrum for photoreactivation suggests that in this organism a deazaflavin, probably F420, functions as the chromophore of the photoreactivating enzyme.  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