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.
Acetate Production by Methanogenic Bacteria   总被引:2,自引:2,他引:0       下载免费PDF全文
Methanosarcina barkeri MS and 227 and Methanosarcina mazei S-6 produced acetate when grown on H2-CO2, methanol, or trimethylamine. Marked differences in acetate production by the two bacterial species were found, even though methane and cell yields were nearly the same. M. barkeri produced 30 to 75 μmol of acetate per mmol of CH4 formed, but M. mazei produced only 8 to 9 μmol of acetate per mmol of CH4.  相似文献   

2.
The effects of organic acid anions on the growth of Syntrophomonas wolfei was determined by varying the initial concentration of the acid anion in the medium. The addition of 15 mM acetate decreased the growth rate of a butyrate-catabolizing coculture containing Methanospirillum hungatei from 0.0085 to 0.0029 per hour. Higher initial acetate concentrations decreased the butyrate degradation rate and the yield of cells of S. wolfei per butyrate degraded. Inhibition was not due to the counter ion or the effect of acetate on the methanogen. Initial acetate concentrations above 25 mM inhibited crotonate-using pure cultures and cocultures of S. wolfei. Benzoate and lactate inhibited the growth of S. wolfei on crotonate in pure culture and coculture. Lactate was an effective inhibitor of S. wolfei cultures at concentrations greater than 10 mM. High concentrations of acetate and lactate altered the electron flow in crotonate-catabolizing cocultures, resulting in the formation of less methane and more butyrate and caproate. The inclusion of the acetate-using methanogen, Methanosarcina barkeri, in a methanogenic butyrate-catabolizing coculture increased both the yield of S. wolfei cells per butyrate degraded and the efficacy of butyrate degradation. Butyrate degradation by acetate-inhibited cocultures occurred only after the addition of Methanosarcina barkeri. These results showed that the metabolism of S. wolfei was inhibited by high levels of organic acid anions. The activity of acetate-using methanogens is important for the syntrophic degradation of fatty acids when high levels of acetate are present.  相似文献   

3.
Abstract The minimum threshold concentrations of acetate utilization and the enzymes responsible for acetate activation of several methanogenic bacteria were investigated and compared with literature data. The minimum acetate concentrations reached by hydrogenotrophic methane bacteria, which require acetate as carbon source, were between 0.4 and 0.6 mM. The acetoclastic Methanosarcina achieves acetate concentrations between 0.2 and 1.2 mM and Methanothrix between 7 and 70 μM. For the activation of acetate most of the hydrogenotrophic methane bacteria investigated use an acetyl-CoA synthetase with a relatively low K m (40–90 μM) for acetate. although the affinity for acetate was high, the hydrogenotrophic methane bacteria were not able to remove acetate to lower concentrations than the acetoclastic methane bacteria, neither in pure cultures nor in anaerobic granular sludge samples. Based on these observations, it is not likely that hydrogenotrophic methanogens compete strongly for acetate with the acetoclastic methane bacteria.  相似文献   

4.
The affinity of Methanosarcina barkeri 227 for acetate and hydrogen at different incubation temperatures was investigated. Increasing the temperature from 20 to 37°C resulted in a 4.5-fold increase in Km for acetate and a 4.8-fold increase for hydrogen. The corresponding increase in Vmax for acetate was 8.3-fold (5.4-fold for hydrogen). This response implied a decrease in the temperature coefficient (Q10) and hence a decrease in the temperature dependency as a function of decreasing substrate concentration.  相似文献   

5.
When grown in the absence of added sulfate, cocultures of Desulfovibrio desulfuricans or Desulfovibrio vulgaris with Methanobrevibacter smithii (Methanobacterium ruminantium), which uses H2 and CO2 for methanogenesis, degraded lactate, with the production of acetate and CH4. When D. desulfuricans or D. vulgaris was grown in the absence of added sulfate in coculture with Methanosarcina barkeri (type strain), which uses both H2-CO2 and acetate for methanogenesis, lactate was stoichiometrically degraded to CH4 and presumably to CO2. 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 CH4. Later, acetate was degraded to CH4 and presumably to CO2. 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 H2 was detected in cocultures which contained D. desulfuricans and M. barkeri until after all lactate was degraded. The addition of H2, but not of lactate, to the growth medium inhibited acetate degradation by pure cultures of M. barkeri. Pure cultures of M. barkeri produced CH4 from acetate at a rate equivalent to that observed for cocultures containing M. barkeri. Inocula of M. barkeri grown with H2-CO2 as the methanogenic substrate produced CH4 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 H2 produced by the Desulfovibrio species during growth with lactate inhibited acetate degradation by M. barkeri.  相似文献   

6.
The kinetics of acetate utilization were examined for washed concentrated cell suspensions of two thermophilic acetotrophic methanogens isolated from a 58°C anaerobic digestor. Progress curves for acetate utilization by cells of Methanosarcina sp. strain CALS-1 showed that the utilization rate was concentration independent (zero order) above concentrations near 3 mM and that acetate utilization ceased when a threshold concentration near 1 mM was reached. Acetate utilization by cells of Methanothrix sp. strain CALS-1 was concentration independent down to 0.1 to 0.2 mM, and threshold values of 12 to 21 μM were observed. Typical utilization rates in the concentration-independent stage were 210 and 130 nmol min−1 mg of protein−1 for the methanosarcina and the methanothrix, respectively. These results are in agreement with a general model in which high acetate concentrations favor Methanosarcina spp., while low concentrations favor Methanothrix spp. However, acetate utilization by these two strains did not follow simple Michaelis-Menton kinetics.  相似文献   

7.
Methanosarcina sp. strain TM-1 and Methanosarcina acetivorans produced and consumed H2 to maintain H2 partial pressures of 16 to 92 Pa in closed cultures during growth on acetate. Strain TM-1 produced H2 continuously when H2 was continuously removed from the culture. The potential physiological significance of H2 in acetate metabolism to methane is discussed.  相似文献   

8.
Nutritional Requirements of Methanosarcina sp. Strain TM-1   总被引:2,自引:1,他引:1       下载免费PDF全文
Methanosarcina sp. strain TM-1, an acetotrophic, thermophilic methanogen isolated from an anaerobic sludge digestor, was originally reported to require an anaerobic sludge supernatant for growth. It was found that the sludge supernatant could be replaced with yeast extract (1 g/liter), 6 mM bicarbonate-30% CO2, and trace metals, with a doubling time on methanol of 14 h. For growth on either methanol or acetate, yeast extract could be replaced with CaCl2 · 2H2O (13.6 μM minimum) and the vitamin p-aminobenzoic acid (PABA, ca. 3 nM minimum), with a doubling time on methanol of 8 to 9 h. Filter-sterilized folic acid at 0.3 μM could not replace PABA. The antimetabolite sulfanilamide (20 mM) inhibited growth of and methanogenesis by Methanosarcina sp. strain TM-1, and this inhibition was reversed by the addition of 0.3 μM PABA. When a defined medium buffered with 20 mM N,N-bis(2-hydroxyethyl)-2-aminoethanesulfonic acid was used, it was shown that Methanosarcina sp. strain TM-1 required 6 mM bicarbonate-30% CO2 for optimal growth and methanogenesis from methanol. Cells growing on acetate were less dependent on bicarbonate-CO2. When we used a defined medium in which the only organic compounds present were methanol or acetate, nitrilotriacetic acid (0.2 mM), and PABA, it was possible to limit batch cultures of Methanosarcina sp. strain TM-1 for nitrogen at NH4+ concentrations at or below 2.0 mM, in marked contrast with Methanosarcina barkeri 227, which fixes dinitrogen when grown under NH4+ limitation.  相似文献   

9.
Methanosarcina barkeri and Desulfobacter postgatei are ubiquitous anaerobic bacteria which grow on acetate or acetate plus sulfate, respectively, as sole energy sources. Their apparent K s values for acetate were determined and found to be approximately 0.2 mM for the sulfate-reducing bacterium and 3 mM for the methanogenic bacterium. In mixed cell suspensions of the two bacteria (adjusted to equal V max) the rate of acetate consumption by D. postgatei approached 15-fold the rate of M. barkeri at low acetate concentrations. The apparent inhibition of methanogenesis was of the same order as expected from the different K s value for acetate. Difference in substrate affinities can thus account for the inhibition of methanogenesis from acetate in sulfate-rich environments, where the acetate concentration is well below 1 mM.  相似文献   

10.
Methanosarcina barkeri 227 and Methanosarcina mazei S-6 grew with acetate as the substrate; we found little effect of H2 on the rate of aceticlastic growth in the presence of various H2 pressures between 2 and 810 Pa. We used physical (H2 addition or flushing the headspace to remove H2) and biological (H2-producing or -utilizing bacteria in cocultures) methods for controlling H2 pressure in Methanosarcina cultures growing on acetate. Added H2 (ca. 100 Pa) was removed rapidly (a few hours) by M. barkeri and slowly (within a day) by M. mazei. When the H2 produced by the aceticlastic methanogens was removed by coculturing with an H2-using Desulfovibrio sp., the H2 pressure was about 2.2 Pa. Under these conditions the stoichiometry of aceticlastic methanogenesis did not change. H2-grown inocula of M. barkeri grew with acetate as the sole catabolic substrate if the inoculum culture was transferred during logarithmic growth to acetate-containing medium or if the transfer was accomplished within 1 or 2 days after exhaustion of H2. H2-grown cultures incubated for 4 or more days after exhaustion of H2 were able to grow with H2 but not with acetate as the sole catabolic substrate. Addition of small quantities of H2 to acetate-containing medium permitted these cultures to initiate growth on acetate.  相似文献   

11.
Abstract Acetate conversion to methane in a steady-state, thermophilic (60°C) anaerobic sewage-sludge digestor and in a thermophilic (60°C) acetate chemostat inoculated with anaerobic thermophilic sewage sludge, was investigated by use of radiotracer methodology. When the acetate pool in the sewage-sludge digestor was 1–2 mM, 4.1% of 2-labeled acetate was converted to CO2. However, when acetate was consumed to less than 1.0 mM, prior to isotopic examinations, this increased to 14.1%. Microscopic observations showed a shift in the acetate-degrading populations during start-up of the acetate-limited chemostat inoculated from the sewage-sludge digestor. Large numbers of Methanosarcina -aggregates were seen during the first 100–150 days of operation, while Methanosaeta -like rods were not observed. The Methanosarcina -aggregates disappeared concurrently with a decrease in the acetate concentration to approx. 0.4 mM, and the culture consisted mainly of a large number of autofluorescent, short rods together with fewer and longer, non-fluorescent, rods. Non-aceticlastic oxidation of acetate to methane was the mechanism of the acetate conversion in the chemostat after 7 months of operation. Our results indicate that the concentration of acetate can influence the mechanism of acetate conversion during thermophilic anaerobic digestion of organic matter.  相似文献   

12.
Acetate kinase catalyzes the reversible phosphorylation of acetate (CH3COO + ATPCH3CO2PO32− + ADP). A mechanism which involves a covalent phosphoryl-enzyme intermediate has been proposed, and chemical modification studies of the enzyme from Escherichia coli indicate an unspecified glutamate residue is phosphorylated (J. A. Todhunter and D. L. Purich, Biochem. Biophys. Res. Commun. 60:273–280, 1974). Alignment of the amino acid sequences for the acetate kinases from E. coli (Bacteria domain), Methanosarcina thermophila (Archaea domain), and four other phylogenetically divergent microbes revealed high identity which included five glutamates. These glutamates were replaced in the M. thermophila enzyme to determine if any are essential for catalysis. The histidine-tagged altered enzymes were produced in E. coli and purified to electrophoretic homogeneity by metal affinity chromatography. Replacements of E384 resulted in either undetectable or extremely low kinase activity, suggesting E384 is essential for catalysis which supports the proposed mechanism. Replacement of E385 influenced the Km values for acetate and ATP with only moderate decreases in kcat, which suggests that this residue is involved in substrate binding but not catalysis. The unaltered acetate kinase was not inactivated by N-ethylmaleimide; however, replacement of E385 with cysteine conferred sensitivity to N-ethylmaleimide which was prevented by preincubation with acetate, acetyl phosphate, ATP, or ADP, suggesting that E385 is located near the active site. Replacement of E97 decreased the Km value for acetate but not ATP, suggesting this residue is involved in binding acetate. Replacement of either E32 or E334 had no significant effects on the kinetic constants, which indicates that neither residue is essential for catalysis or significantly influences the binding of acetate or ATP.  相似文献   

13.
Methanosarcina barkeri strain Fusaro was grown on a mixed substrate medium of methanol and acetate. When 50 mM of acetate was added to the methanol basal medium (250 mM), the rates of methane production, methanol consumption, cell growth and corrinoid production were stimulated 3.2, 2.7, 3.5, and 2.4 times, respectively compared with those in methanol alone. Addition of acetate also has significant influence on corrinoid distribution decreasing the intracellular corrinoid content from 6.8 to 3.0 mg/g dry cell and increasing the extracellular corrinoid concentration from 4.0 to 5.4 mg/l. The carbon balance analysis for methanogenesis and cellular growth with or without acetate addition revealed that about 50% of the utilized acetate carbon might be incorporated in the cellular materials and the remaining might be oxidized to generate the electrons which stimulate the methanol reduction to methane, accelerating the metabolic activities of the methanogenesis from methanol consequently enhancing the rates of methane and corrinoid production, and cell growth.  相似文献   

14.
Oil production by water injection can cause souring in which sulfate in the injection water is reduced to sulfide by resident sulfate-reducing bacteria (SRB). Sulfate (2 mM) in medium injected at a rate of 1 pore volume per day into upflow bioreactors containing residual heavy oil from the Medicine Hat Glauconitic C field was nearly completely reduced to sulfide, and this was associated with the generation of 3 to 4 mM acetate. Inclusion of 4 mM nitrate inhibited souring for 60 days, after which complete sulfate reduction and associated acetate production were once again observed. Sulfate reduction was permanently inhibited when 100 mM nitrate was injected by the nitrite formed under these conditions. Pulsed injection of 4 or 100 mM nitrate inhibited sulfate reduction temporarily. Sulfate reduction resumed once nitrate injection was stopped and was associated with the production of acetate in all cases. The stoichiometry of acetate formation (3 to 4 mM formed per 2 mM sulfate reduced) is consistent with a mechanism in which oil alkanes and water are metabolized to acetate and hydrogen by fermentative and syntrophic bacteria (K. Zengler et al., Nature 401:266–269, 1999), with the hydrogen being used by SRB to reduce sulfate to sulfide. In support of this model, microbial community analyses by pyrosequencing indicated SRB of the genus Desulfovibrio, which use hydrogen but not acetate as an electron donor for sulfate reduction, to be a major community component. The model explains the high concentrations of acetate that are sometimes found in waters produced from water-injected oil fields.  相似文献   

15.
We compared the metabolism of methanol and acetate when Methanosarcina barkeri was grown in the presence and absence of Desulfovibrio vulgaris. The sulfate reducer was not able to utilize methanol or acetate as the electron donor for energy metabolism in pure culture, but was able to grow in coculture. Pure cultures of M. barkeri produced up to 10 μmol of H2 per liter in the culture headspace during growth on acetate or methanol. In coculture with D. vulgaris, the gaseous H2 concentration was ≤2 μmol/liter. The fractions of 14CO2 produced from [14C]methanol and 2-[14C]acetate increased from 0.26 and 0.16, respectively, in pure culture to 0.59 and 0.33, respectively, in coculture. Under these conditions, approximately 42% of the available electron equivalents derived from methanol or acetate were transferred and were utilized by D. vulgaris to reduce approximately 33 μmol of sulfate per 100 μmol of substrate consumed. As a direct consequence, methane formation in cocultures was two-thirds that observed in pure cultures. The addition of 5.0 mM sodium molybdate or exogenous H2 decreased the effects of D. vulgaris on the metabolism of M. barkeri. An analysis of growth and carbon and electron flow patterns demonstrated that sulfate-dependent interspecies H2 transfer from M. barkeri to D. vulgaris resulted in less methane production, increased CO2 formation, and sulfide formation from substrates not directly utilized by the sulfate reducer as electron donors for energy metabolism and growth.  相似文献   

16.
The morphology of Methanosarcina mazei was controlled by magnesium, calcium, and substrate concentrations and by inoculum size; these factors allowed manipulation of the morphology and interconversions between pseudosarcinal aggregates and individual, coccoid cells. M. mazei grew as aggregates in medium with a low concentration of catabolic substrate (either 50 mM acetate, 50 mM methanol, or 10 mM trimethylamine) unless Ca2+ and Mg2+ concentrations were high. Growth in medium high in Ca2+, Mg2+, and substrate (i.e., 150 mM acetate, 150 mM methanol, or 40 mM trimethylamine) converted pseudosarcinal aggregates to individual cocci. In such media, aggregates separated into individual cells which continued to grow exclusively as single cells during subsequent transfers. Conversion of single cells back to aggregates was complicated, because conditions which supported the aggregated morphology (e.g., low calcium or magnesium concentration) caused lysis of coccoid inocula. We recovered aggregates from coccoid cells by inoculating serial dilutions into medium high in calcium and magnesium. Cells from very dilute inocula grew into aggregates which disaggregated on continued incubation. However, timely transfer of the aggregates to medium low in calcium, magnesium, and catabolic substrates allowed continued growth as aggregates. We demonstrated the activity of the enzyme (disaggregatase) which caused the dispersion of aggregates into individual cells; disaggregatase was produced not only during disaggregation but also in growing cultures of single cells. Uronic acids, the monomeric constituents of the Methanosarcina matrix, were also produced during disaggregation and during growth as coccoids.  相似文献   

17.
Acetate kinase (ACK) catalyzes the reversible synthesis of acetyl phosphate by transfer of the γ-phosphate of ATP to acetate. Here we report the first biochemical and kinetic characterization of a eukaryotic ACK, that from the protist Entamoeba histolytica. Our characterization revealed that this protist ACK is the only known member of the ASKHA structural superfamily, which includes acetate kinase, hexokinase, and other sugar kinases, to utilize inorganic pyrophosphate (PPi)/inorganic phosphate (Pi) as the sole phosphoryl donor/acceptor. Detection of ACK activity in E. histolytica cell extracts in the direction of acetate/PPi formation but not in the direction of acetyl phosphate/Pi formation suggests that the physiological direction of the reaction is toward acetate/PPi production. Kinetic parameters determined for each direction of the reaction are consistent with this observation. The E. histolytica PPi-forming ACK follows a sequential mechanism, supporting a direct in-line phosphoryl transfer mechanism as previously reported for the well-characterized Methanosarcina thermophila ATP-dependent ACK. Characterizations of enzyme variants altered in the putative acetate/acetyl phosphate binding pocket suggested that acetyl phosphate binding is not mediated solely through a hydrophobic interaction but also through the phosphoryl group, as for the M. thermophila ACK. However, there are key differences in the roles of certain active site residues between the two enzymes. The absence of known ACK partner enzymes raises the possibility that ACK is part of a novel pathway in Entamoeba.  相似文献   

18.
Acetate Inhibition of Methanogenic, Syntrophic Benzoate Degradation   总被引:4,自引:4,他引:0       下载免费PDF全文
Acetate inhibited benzoate degradation by a syntrophic coculture of an anaerobic benzoate degrader (strain BZ-2) and Methanospirillum strain PM-1; the apparent Ki for acetate was approximately 40 mM. The addition of acetate resulted in a decrease in the hydrogen concentration in the coculture, indicating that phenomena related to interspecies hydrogen transfer affected this value and that the effect of acetate on the benzoate-degrading partner was probably greater than the apparent Ki for the coculture suggests.  相似文献   

19.
20.
Acetate kinase, a member of the acetate and sugar kinase-Hsp70-actin (ASKHA) enzyme superfamily1-5, is responsible for the reversible phosphorylation of acetate to acetyl phosphate utilizing ATP as a substrate. Acetate kinases are ubiquitous in the Bacteria, found in one genus of Archaea, and are also present in microbes of the Eukarya6. The most well characterized acetate kinase is that from the methane-producing archaeon Methanosarcina thermophila7-14. An acetate kinase which can only utilize PPi but not ATP in the acetyl phosphate-forming direction has been isolated from Entamoeba histolytica, the causative agent of amoebic dysentery, and has thus far only been found in this genus15,16.In the direction of acetyl phosphate formation, acetate kinase activity is typically measured using the hydroxamate assay, first described by Lipmann17-20, a coupled assay in which conversion of ATP to ADP is coupled to oxidation of NADH to NAD+ by the enzymes pyruvate kinase and lactate dehydrogenase21,22, or an assay measuring release of inorganic phosphate after reaction of the acetyl phosphate product with hydroxylamine23. Activity in the opposite, acetate-forming direction is measured by coupling ATP formation from ADP to the reduction of NADP+ to NADPH by the enzymes hexokinase and glucose 6-phosphate dehydrogenase24.Here we describe a method for the detection of acetate kinase activity in the direction of acetate formation that does not require coupling enzymes, but is instead based on direct determination of acetyl phosphate consumption. After the enzymatic reaction, remaining acetyl phosphate is converted to a ferric hydroxamate complex that can be measured spectrophotometrically, as for the hydroxamate assay. Thus, unlike the standard coupled assay for this direction that is dependent on the production of ATP from ADP, this direct assay can be used for acetate kinases that produce ATP or PPi.  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