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.
Methanosarcina barkeri strain Fusaro was found to grow on pyruvate as sole carbon and energy source after an incubation period of 10–12 weeks in the presence of high pyruvate concentrations (100 mM). Growth studies, cell suspension experiments and enzymatic investigations were performed with pyruvate-utilizing M. barkeri. For comparison acetate-adapted cells of M. barkeri were analyzed.
  1. Pyruvate-utilizing M. barkeri grew on pyruvate (100 mM) with an initial doubling time of about 25 h (37 °C, pH 6.5) up to cell densities of about 0.8 g cell dry weight/l. The specific growth rate was linearily dependent on the pyruvate concentration up to 100 mM indicating that pyruvate was taken up by passive diffusion. Only CO2 and CH4 were detected as fermentation products. As calculated from fermentation balances pyruvate was converted to CH4 and CO2 according to following equation: Pyruvate-+H++0.5 H2O » 1.25 CH4+1.75 CO2. The molar growth yield (Ych 4) was about 14 g dry weight cells/mol CH4. In contrast the growth yield (Ych 4) of M. barkeri during growth on acctate (Acetate-+H+ » CH4+CO2) was about 3 g/mol CH4.
  2. Cell suspensions of pyruvate-grown M. barkeri catalyzed the conversion of pyruvate to CH4, CO2 and H2 (5–15 nmol pyruvate consumed/min x mg protein). At low cell concentrations (0.5 mg protein/ml) 1 mol pyruvate was converted to 1 mol CH4, 2 mol CO2 and 1 mol H2. At higher cell concentration less H2 and CO2 and more CH4 were formed due to CH4 formation from H2/CO2. The rate of pyruvate conversion was linearily dependent on the pyruvate concentration up to about 30 mM. Cell suspensions of acetate-grown M. barkeri also catalyzed the conversion of 1 mol pyruvate to 1 mol CH4, 2 mol CO2 and 1 mol H2 at similar rates and with similar affinity for pyruvate as pyruvate-grown cells.
  3. Cell extracts of both pyruvate-grown and acetate-grown M. barkeri contained pyruvate: ferredoxin oxidoreductase. The specific activity in pyruvate-grown cells (0.8 U/mg) was 8-fold higher than in acetate-grown cells (0.1 U/mg). Coenzyme F420 was excluded as primary electron acceptor of pyruvate oxidoreductase. Cell extracts of pyruvate-grown M. barkeri contained carbon monoxide dehydrogenase activity and hydrogenase activity catalyzing the reduction by carbon monoxide and hydrogen of both methylviologen and ferredoxin (from Clostridium).
This is the first report on growth of a methanogen on pyruvate as sole carbon and energy source, i.e. on a substrate more complex than acetate.  相似文献   

2.
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   

3.
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   

4.
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   

5.
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   

6.
Methanosarcina barkeri grows in defined media with acetate, methanol or carbon dioxide as carbon sources. Methanol is used for methanogenesis at a 5 times higher rate as compared with the other substrates. M. barkeri can use the substrates simultaneously, but due to acidification or alkalification of the medium during growth on methanol or acetate, respectively, growth and methanogenesis may stop before the substrates are exhausted. Growth and methanogenesis on methanol or acetate are inhibited by the presence of an excess of H2; the inhibition is abolished by the addition of carbon dioxide, which probably serves as an essential source of cell carbon, in the absence of which methano-genesis ceases.  相似文献   

7.
We examined the unitrophic metabolism of acetate and methanol individually and the mixotrophic utilization of these compounds by using detailed 14C-labeled tracer studies in a strain of Methanosarcina barkeri adapted to grow on acetate as the sole carbon and energy source. The substrate consumption rate and methane production rate were significantly lower on acetate alone than during the unitrophic or mixotrophic metabolism of methanol. Cell yields (in grams per mole of substrate) were identical during exponential growth on acetate and exponential growth on methanol. During unitrophic metabolism of acetate, the methyl moiety accounted for the majority of the CH4 produced, but 14% of the CO2 generated originated from the methyl moiety. This correlated with the concurrent reduction of equivalent amounts of the C-1 of acetate to CH4. 14CH4 was also produced from added 14CO2, although to a lesser extent than from reduction of the C-1 of acetate. During mixotrophic metabolism, methanol and acetate were catabolized simultaneously. The rates of 14CH4 and 14CO2 generation from [2-14C]acetate were logarithmic and higher in mixotrophic than in unitrophic cultures at substrate concentrations of 50 mM. A comparison of the oxidoreductase activities in cell extracts of the acetate-adapted strain grown on acetate and of strain MS grown on methanol or on H2 plus CO2 indicated that the pyruvate, α-ketoglutarate, and isocitrate dehydrogenase activities remained constant, whereas the CO dehydrogenase activity was significantly higher (5,000 nmol/min per mg of protein) in the acetate-adapted strain. These results suggested that a significant intramolecular redox pathway is possible for the generation of CH4 from acetate, that energy metabolism from acetate by M. barkeri is not catabolite repressed by methanol, and that the acetate-adapted strain is a metabolic mutant with derepressed CO dehydrogenase activity.  相似文献   

8.
Nitrogen fixation (diazotrophy) has recently been demonstrated in several methanogenic archaebacteria. To compare the process in an archaebacterium with that in eubacteria, we examined the properties of diazotrophic growth and nitrogenase activity in Methanosarcina barkeri 227. Growth yields with methanol or acetate as a growth substrate were significantly lower in N2-grown cultures than in NH4+-grown cultures, and the culture doubling times were increased, indicating that diazotrophy was energetically costly, as it is in eubacteria. Growth of nitrogen-fixing cells was inhibited when molybdenum was omitted from the medium; addition of 10 nM molybdate stimulated growth, while 1 μM molybdate restored maximum diazotrophic growth. Omission of molybdenum did not inhibit growth of ammonia-grown cells. Tungstate (100 μM) strongly inhibited growth of molybdenum-deficient diazotrophic cells, while ammonia-grown cells were unaffected. The addition of 100 nM vanadate or chromate did not stimulate diazotrophic growth of molybdenum-starved cells. These results are consistent with the presence of a molybdenum-containing nitrogenase in M. barkeri. Acetylene, the usual substrate for assaying nitrogenase activity, inhibited methanogenesis by M. barkeri and consequently needed to be used at a low partial pressure (0.3% of the headspace) when acetylene reduction by whole cells was assayed. Whole cells reduced 0.3% acetylene to ethylene at a very low rate (1 to 2 nmol h−1 mg of protein−1), and they “switched off” acetylene reduction in response to added ammonia or glutamine. Crude extracts from diazotrophic cells reduced 10% acetylene at a rate of 4 to 5 nmol of C2H4 formed h−1 mg of protein−1 when supplied with ATP and reducing power, while extracts of Klebsiella pneumoniae prepared by the same procedures had rates 100-fold higher. Acetylene reduction by extracts required ATP and was completely inhibited by 1 mM ADP in the presence of 5 mM ATP. The low rates of C2H2 reduction could be due to improper assay conditions, to switched-off enzyme, or to the nitrogenase's having lower activity towards acetylene than towards dinitrogen.  相似文献   

9.
Growth of Methanosarcina sp. strain 227 and Methanosarcina mazei on H2-CO2 and mixtures of H2-CO2 and acetate or methanol was examined. The growth yield of strain 227 on H2-CO2 in complex medium was 8.4 mg/mmol of methane produced. Growth in defined medium was characteristically slower, and cell yields were proportionately lower. Labeling studies confirmed that CO2 was rapidly reduced to CH4 in the presence of H2, and little acetate was used for methanogenesis until H2 was exhausted. This resulted in a biphasic pattern of growth similar to that reported for strain 227 grown on methanol-acetate mixtures. Biphasic growth was not observed in cultures on mixtures of H2-CO2 and methanol, and less methanol oxidation occurred in the presence of H2. In M. mazei the aceticlastic reaction was also inhibited by the added H2, but since the cultures did not immediately metabolize H2, the duration of the inhibition was much longer.  相似文献   

10.
A strictly anaerobic, homoacetogenic, Gram-positive, non spore-forming bacterium, designated strain SR12T(T=type strain), was isolated from an anaerobic methanogenic digestor fed with olive mill wastewater. Yeast extract was required for growth but could also be used as sole carbon and energy source. Strain SR12Tutilized a few carbohydrates (glucose, fructose and sucrose), organic compounds (lactate, crotonate, formate and betaine), alcohols (methanol), the methoxyl group of some methoxylated aromatic compounds, and H2+CO2. The end-products of carbohydrate fermentation were acetate, formate, butyrate, H2and CO2. End-products from lactate and methoxylated aromatic compounds were acetate and butyrate. Strain SR12Twas non-motile, formed aggregates, had a G+C content of 55 mol % and grew optimally at 35°C and pH 7.2 on a medium containing glucose. Phylogenetically, strain SR12Twas related toEubacterium barkeri, E. callanderi, andE. limosumwithE. barkerias the closest relative (similarity of 98%) with which it bears little phenotypic similarity or DNA homology (60%). On the basis of its phenotypic, genotypic, and phylogenetic characteristics, we propose to designate strain SR12TasEubacterium aggreganssp. nov. The type strain is SR12T(=DSM 12183).  相似文献   

11.
Two strains of Methanosarcina (M. Barkeri strain MS, isolated from sewage sludge, and strain UBS, isolated from lake sediments) were found to have similar cellular properties and to have DNA base compositions of 44 mol percent guanosine plus cytosine. Strain MS was selected for further studies of its one-carbon metabolism. M. barkeri grew autotrophically via H2 oxidation/CO2 reduction. The optimum temperature for growth and methanogenesis was 37°C. H2 oxidation proceeded via an F420-dependent NADP+-linked hydrogenase. A maximum specific activity of hydrogenase in cell-free extracts, using methyl viologen as electron acceptor, was 6.0 mol min · mg protein at 37°C and the optimum pH (9.0). M. barkeri also fermented methanol andmethylamine as sole energy sources for growth. Cell yields during growth on H2/CO2 and on methanol were 6.4 and 7.2 mg cell dry weight per mmol CH4 formed, respectively. During mixotrophic growth on H2/CO2 plus methanol, most methane was derived from methanol rather than from CO2. Similar activities of hydrogenase were observed in cell-free extracts from H2/CO2-grown and methanol-grown cells. Methanol oxidation apparently proceeded via carrierbound intermediates, as no methylotrophy-type of methanol dehydrogenase activity was observed in cell-free extracts. During growth on methanol/CO2, up to 48% of the cell carbon was derived from methanol indicating that equivalent amounts of cell carbon were derived from CO2 and from an organic intermediate more reduced than CO2. Cell-free extracts lacked activity for key cell carbon synthesis enzymes of the Calvin cycle, serine path, or hexulose path.Abbreviations CAPS cycloaminopropane sulfonic acid - CH3-SCoM methyl coenzyme M - DCPIP 2,6-dichlorophenolindophenol - DEAE diethylaminoethyl - dimethyl POPOP 1,4-bis-2-(4-mothyl-5-phenyloxazolyl)-benzene - DNA deoxyribonucleic acid - dpm dismtegrations per min - DTT dithiothreitol - EDTA ethylenediamine tetraacetic acid - F420 factor 420 - G+C guanosine plus cytosine - NAD+ nicotinamide adenine dinucleotide - NADP+ nicotinamide adenine dinucleotide phosphate - PBBW phosphate buffered basal Weimer - PMS phenazine methosulfate - PPO 2,5-diphenyloxazole - rRNA ribosomal ribonucleic acid - RuBP ribulose-1,5-bisphosphate - Tris tris-hydroxymethyl-aminomethane - max maximum specific growth rate  相似文献   

12.
It has been known for some time that pokeweed antiviral protein acts by enzymatically inhibiting protein synthesis on eucaryotic ribosome systems. The site of this action is known to be the ribosome itself. In this paper we show that the pokeweed antiviral protein reaction against ribosomes is a strong function of salt concentrations, where 160 mM K+ and 3 mM Mg2+ retards the reaction, while 20 mM K+ and 2 mM Mg2+ allows maximum reaction rate. It is also shown, however, that an unidentified protein in the postribosomal supernatant solution, together with ATP, allows the ribosome to be attacked even in the presence of high salt. Kinetic analysis of the antiviral protein reaction has been carried out under both sets of conditions, and reveals that the turnover number for the enzyme is about 300–400 mol/mol per min. in each case. The Km for ribosomes is 1 μM in the presence of low salt and 0.2 μM at higher salt in the presence of postribosomal supernatant factors plus ATP. The antiviral protein reaction is also shown to be pH dependent and is controlled by a residue with pKa value of approx. 7.0, apparently a histidine. Stoichiometric reaction of the enzyme with iodoacetamide results in a significant loss of antiribosomal activity.  相似文献   

13.
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   

14.
Methanosarcina barkeri was cultured on methanol, H2-CO2, and acetate, and the 13C/12C ratios of the substrates and the methane produced from them were determined. The discrimination against 13C in methane relative to substrate decreased in the order methanol > CO2 > acetate. The isotopic fractionation for methane derived from acetate was only one-third of that observed with methanol as the substrate. The data presented indicate that the last enzyme of methanogenesis, methylreductase, is not the primary site of isotopic discrimination during methanogenesis from methanol or CO2. These results also support biogeochemical interpretations that gas produced in environments in which acetate is the primary methane precursor will have higher 13C/12C ratios than those from environments where other substrates predominate.  相似文献   

15.
The strictly anaerobic homoacetogenic bacteria Acetobacterium woodii and Sporomusa sphaeroides differ with respect to their energy metabolism. Since growth as well as acetate and ATP formation of A. woodii is strictly dependent on Na+, but that of S. sphaeroides is not, the question arose whether these organisms also use different coupling ions for mechanical work, i.e. flagellar rotation. During growth on fructose in the presence of Na+ (50 mM), cells of A. woodii were vigorously motile, as judged by light microscopy. At low Na+ concentrations (0.3 mM), the growth rate decreased by only 15%, but the cells were completely non-motile. Addition of Na+ to such cultures restored motility instantaneously. Motility, as determined in swarm agar tubes, was strictly dependent on Na+; Li+, but not K+ partly substituted for Na+. Of the amilorides tested, phenamil proved to be a specific inhibitor of the flagellar motor of A. woodii. Growth and motility of S. sphaeroides was neither dependent on Na+ nor inhibited by amiloride derivatives. These results indicate that flagellar rotation is driven by ΔμNa + in A. woodii, but by ΔμH + in S. sphaeroides. Received: 30 May 1995 / Accepted: 31 August 1995  相似文献   

16.
(H+ + K+)-ATPase-enriched membranes were prepared from hog gastric mucosa by sucrose gradient centrifugation. These membranes contained Mg2+-ATPase and p-nitrophenylphosphatase activities (68 ± 9 μmol Pi and 2.9 ± 0.6 μmol p-nitrophenol/mg protein per h) which were insensitive to ouabain and markedly stimulated by 20 mM KCl (respectively, 2.2- and 14.8-fold). Furthermore, the membranes autophosphorylated in the absence of K+ (up to 0.69 ± 0.09 nmol Pi incorporated/mg protein) and dephosphorylated by 85% in the presence of this ion. Membrane proteins were extracted by 1–2% (w/v) n-octylglucoside into a soluble form, i.e., which did not sediment in a 100 000 × g × 1 h centrifugation. This soluble form precipitated upon further dilution in detergent-free buffer. Extracted ATPase represented 32% (soluble form) and 68% (precipitated) of native enzyme and it displayed the same characteristic properties in terms of K+-stimulated ATPase and p-nitrophenylphosphatase activities and K+-sensitive phosphorylation: Mg2+-ATPase (μmol Pi/mg protein per h) 32 ± 9 (basal) and 86 ± 20 (K+-stimulated); Mg2+-p-nitrophenylphosphatase (μmol p-nitrophenol/mg protein per h) 2.6 ± 0.5 (basal) and 22.2 ± 3.2 (K+-stimulated); Mg2+-phosphorylation (nmol Pi/mg protein) 0.214 ± 0.041 (basal) and 0.057 ± 0.004 (in the presence of K+). In glycerol gradient centrifugation, extracted enzyme equilibrated as a single peak corresponding to an apparent 390 000 molecular weight. These findings provide the first evidence for the solubilization of (H+ + K+)-ATPase in a still active structure.  相似文献   

17.
The absorbance change of the weak base dye probe, Acridine orange, was used to monitor alterations of pH gradients across renal brush border membrane vesicles. The presence of Na+/H+ or Li+/H+ exchange was demonstrated by diluting Na2SO4 or Li2SO4 loaded vesicles into Na+- or Li+-free solutions, which caused dye uptake. About 20% of the uptake was abolished by lipid permeable cations such as valinomycin-K+ or tetraphenylphosphonium, indicating perhaps the presence of a finite Na+ conductance smaller than electroneutral Na+/H+ exchange. The protonophore tetrachlorosalicylanilide raised the rate of dye uptake under these conditions, hence the presence of an Na+ conductance greater than the H+ conductance was suggested. K+ gradients also induced changes of pH, at about 10% of the Na+ or Li+ rate. Partial inhibition (21%) was seen with 0.1 mM amiloride indicating that K+ was a low affinity substrate for the Na+/H+ exchange. Acceleration both by tetrachlorosalicylanilide (2-fold) and valinomycin (4-fold) suggested the presence of 2 classes of vesicles, those with high and those with low K+ conductance. The larger magnitude of the valinomycin dependent signal suggested that 75% of the vesicles had a low K+ conductance. Inward Cl? gradients also induced acidification, partially inhibited by the presence of tetraphenylphosphonium, and accelerated by tetrachlorosalicylanilide. Thus both a Cl? conductance greater than the H+ conductance and a Cl?/OH? exchange were present. The rate of Na+/H+ exchange was amiloride sensitive with a pH optimum of 6.5 and an apparent Km for Na+ or Li+ of about 10 mM and an EA of 14.3 kcal per mol. A 61-fold Na2SO4 gradient resulted in a pH gradient of 1.64 units which increased to 1.8 with gramicidin. An equivalent NaCl gradient gave a much lower ΔpH even in the presence of gramicidin showing that the H+ and Cl? pathways could alter the effects of the Na+/H+ exchange.  相似文献   

18.
Cell extracts (27000xg supernatant) of acetate grown Methanosarcina barkeri were found to have carbonic anhydrase activity (0.41 U/mg protein), which was lost upon heating or incubation with proteinase K. The activity was inhibited by Diamox (apparent K i=0.5 mM), by azide (apparent K i=1 mM), and by cyanide (apparent K i=0.02 mM). These and other properties indicate that the archaebacterium contains the enzyme carbonic anhydrase (EC 4.2.1.1). Evidence is presented that the protein is probably located in the cytoplasm. Methanol or H2/CO2 grown cells of M. barkeri showed no or only very little carbonic anhydrase activity. After transfer of these cells to acetate medium the activity was induced suggesting a function of this enzyme in acetate fermentation to CO2 and CH4. Interestingly, Desulfobacter postgatei and Desulfotomaculum acetoxidans, which oxidize acetate to 2 CO2 with sulfate as electron acceptor, were also found to exhibit carbonic anhydrase activity (0.2 U/mg protein).  相似文献   

19.
《Plant Science Letters》1984,33(1):103-114
The effects of monovalent cations, inhibitors of metabolism dinitrophenol (DNP), carbonyl cyanide-p-trifluoromethoxyphenylhydrazone (FCCP), and KCN and temperature variations upon Ca2+ fluxes in intact roots of barley (Hordeum vulgare L. cv. Fergus and Herta) seedlings were investigated. 45Ca2+ influx was depressed in CaSO4-grown (low-salt) plants by the presence of NH4+, K+, or Na+ in the uptake medium. In contrast Ca2+ influx was slightly increased by Li+. In low-salt roots pretreated with KCN and in roots preloaded with K+ (high-K+ plants), the presence of K+ in the medium had no significant effect on Ca2+ influx, while in roots preloaded with Na+, the presence of K+ in the medium depressed Ca2+ influx. In absolute terms, Ca2+ influx was significantly greater in high-salt (both K+ or Na+ preloaded) than in low-salt roots.Patterns of 45Ca2+ efflux in the absence and in the presence of K+, NH4+, or Li+ in the external medium showed that these monovalent cations caused stimulation of 45Ca2+ efflux both from the cytoplasmic and vacuolar phases.It was noted that these modifications of Ca2+ fluxes by monovalent cations are transient and characteristic of a transitional stage of cation uptake by low-salt roots. We conclude that, together with stimulated active H+ efflux (another characteristic of this transitional stage), modifications of Ca2+ fluxes during monovalent cation uptake by low-salt roots is a response directed towards the maintenance of electrical neutrality.Determination of net fluxes revealed that the plants were close to Ca2+ flux equilibrium in the growth medium (0.5 mM CaSO4). Transfer of these plants to 0.5 mM CaSO4 + 0.25 mM K2SO4 caused a net release of CA2+ into the external medium.  相似文献   

20.
The effects of K+, Na+ and ATP on the gastric (H+ + K+)-ATPase were investigated at various pH. The enzyme was phosphorylated by ATP with a pseudo-first-order rate constant of 3650 min?1 at pH 7.4. This rate constant increased to a maximal value of about 7900 min?1 when pH was decreased to 6.0. Alkalinization decreased the rate constant. At pH 8.0 it was 1290 min?1. Additions of 5 mM K+ or Na+, did not change the rate constant at acidic pH, while at neutral or alkaline pH a decrease was observed. Dephosphorylation of phosphoenzyme in lyophilized vesicles was dependent on K+, but not on Na+. Alkaline pH increased the rate of dephosphorylation. K+ stimulated the ATPase and p-nitrophenylphosphatase activities. At high concentrations K+ was inhibitory. Below pH 7.0 Na+ had little or no effect on the ATPase and p-nitrophenylphosphatase, while at alkaline pH, Na+ inhibited both activities. The effect of extravesicular pH on transport of H+ was investigated. At pH 6.5 the apparent Km for ATP was 2.7 μM and increased little when K+ was added extravesicularly. At pH 7.5, millimolar concentrations of K+ increased the apparent Km for ATP. Extravesicular K+ and Na+ inhibited the transport of H+. The inhibition was strongest at alkaline pH and only slight at neutral or acidic pH, suggesting a competition between the alkali metal ions and hydrogen ions at a common binding site on the cytoplasmic side of the membrane. Two H+-producing reactions as possible candidates as physiological regulators of (H+ + K+)-ATPase were investigated. Firstly, the hydrolysis of ATP per se, and secondly, the hydration of CO2 and the subsequent formation of H+ and HCO3?. The amount of hydrogen ions formed in the ATPase reaction was highest at alkaline pH. The H+/ATP ratio was about 1 at pH 8.0. When CO2 was added to the reaction medium there was no change in the rate of hydrogen ion transport at pH 7.0, but at pH 8.0 the rate increased 4-times upon the addition of 0.4 mM CO2. The results indicate a possible co-operation in the production of acid between the H+ + K+-ATPase and a carbonic anhydrase associated with the vesicular membrane.  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