首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
Kinetics of butyrate, acetate, and hydrogen metabolism were determined with butyrate-limited, chemostat-grown tricultures of a thermophilic butyrate-utilizing bacterium together with Methanobacterium thermoautotrophicum and the TAM organism, a thermophilic acetate-utilizing methanogenic rod. Kinetic parameters were determined from progress curves fitted to the integrated form of the Michaelis-Menten equation. The apparent half-saturation constants, Km, for butyrate, acetate, and dissolved hydrogen were 76 μM, 0.4 mM, and 8.5 μM, respectively. Butyrate and hydrogen were metabolized to a concentration of less than 1 μM, whereas acetate uptake usually ceased at a concentration of 25 to 75 μM, indicating a threshold level for acetate uptake. No significant differences in Km values for butyrate degradation were found between chemostat- and batch-grown tricultures, although the maximum growth rate was somewhat higher in the batch cultures in which the medium was supplemented with yeast extract. Acetate utilization was found to be the rate-limiting reaction for complete degradation of butyrate to methane and carbon dioxide in continuous culture. Increasing the dilution rate resulted in a gradual accumulation of acetate. The results explain the low concentrations of butyrate and hydrogen normally found during anaerobic digestion and the observation that acetate is the first volatile fatty acid to accumulate upon a decrease in retention time or increase in organic loading of a digestor.  相似文献   

2.
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   

3.
The butyrate-oxidizing, proton-reducing, obligately anaerobic bacterium NSF-2 was grown in batch cocultures with either the hydrogen-oxidizing bacterium Methanospirillum hungatei PM-1 or Desulfovibrio sp. strain PS-1. Metabolism of butyrate occurred in two phases. The first phase exhibited exponential growth kinetics (phase a) and had a doubling time of 10 h. This value was independent of whether NSF-2 was cultured with a methanogen or a sulfate reducer and likely represents the maximum specific growth rate of NSF-2. This exponential growth phase was followed by a second phase with a nearly constant rate of degradation (phase b) which dominated the time course of butyrate degradation. The specific activity of H2 uptake by the hydrogen-oxidizing bacterium controlled the bioenergetic conditions of metabolism in phase b. During this phase both the Gibbs free energy (ΔG′) and the butyrate degradation rate (v) were greater for NSF-2-Desulfovibrio sp. strain PS-1 (ΔG′ = −17.0 kJ/mol; v = 0.20 mM/h) than for NSF-2-M. hungatei PM-1 (ΔG′ = −3.8 kJ/mol, v = 0.12 mM/h). The ΔG′ value remained stable and characteristic of the two hydrogen oxidizers during phase b. The stable ΔG′ resulted from the close coupling of the rates of butyrate and H2 oxidation. The addition of 2-bromoethanesulfonate to a NSF-2-methanogen coculture resulted in the total inhibition of butyrate degradation; the inhibition was relieved when Desulfovibrio sp. strain PS-1 was added as a new H2 sink. When the specific activity of H2 consumption was increased by adding higher densities of the Desulfovibrio sp. to 2-bromoethanesulfonate-inhibited NSF-2-methanogen cocultures, lower H2 pool sizes and higher rates of butyrate degradation resulted. Thus, it is the kinetic parameters of H2 consumption, not the type of H2 consumer per se, that establishes the thermodynamic conditions which in turn control the rate of fatty acid degradation. The bioenergetic homeostasis we observed in phase b was a result of the kinetics of the coculture members and the feedback inhibition by hydrogen which prevents butyrate degradation rates from reaching their theoretical Vmax.  相似文献   

4.
The effects of 2-bromoethanesulfonate, an inhibitor of methanogenesis, on metabolism in sludge from a thermophilic (58°C) anaerobic digestor were studied. It was found from short-term experiments that 1 μmol of 2-bromoethanesulfonate per ml completely inhibited methanogenesis from 14CH3COO, whereas 50 μmol/ml was required for complete inhibition of 14CO2 reduction. When 1 μmol of 2-bromoethanesulfonate per ml was added to actively metabolizing sludge which was then incubated for 24 h. it caused a 60% reduction in methanogenesis and a corresponding increase in acetate accumulation; at 50 μmol/ml it caused complete inhibition of methanogenesis and accumulation of acetate. H2, and ethanol.  相似文献   

5.
CO and H2 have been implicated in methanogenesis from acetate, but it is unclear whether they are directly involved in methanogenesis or electron transfer in acetotrophic methanogens. We compared metabolism of H2, CO, and formate by cultures of the thermophilic acetotrophic methanogens Methanosarcina thermophila TM-1 and Methanothrix sp. strain CALS-1. M. thermophila accumulated H2 to partial pressures of 40 to 70 Pa (1 Pa = 0.987 × 10-5 atm), as has been previously reported for this and other Methanosarcina cultures. In contrast, Methanothrix sp. strain CALS-1 accumulated H2 to maximum partial pressures near 1 Pa. Growing cultures of Methanothrix sp. strain CALS-1 initially accumulated CO, which reached partial pressures near 0.6 Pa (some CO came from the rubber stopper) during the middle of methanogenesis; this was followed by a decrease in CO partial pressures to less than 0.01 Pa by the end of methanogenesis. Accumulation or consumption of CO by cultures of M. thermophila growing on acetate was not detected. Late-exponential-phase cultures of Methanothrix sp. strain CALS-1, in which the CO partial pressure was decreased by flushing with N2-CO2, accumulated CO to 0.16 Pa, whereas cultures to which ca. 0.5 Pa of CO was added consumed CO until it reached this partial pressure. Cyanide (1 mM) blocked CO consumption but not production. High partial pressures of H2 (40 kPa) inhibited methanogenesis from acetate by M. thermophila but not by Methanothrix sp. strain CALS-1, and 2 kPa of CO was not inhibitory to M. thermophila but was inhibitory to Methanothrix sp. strain CALS-1. Levels of CO dehydrogenase, hydrogenase, and formate dehydrogenase in Methanothrix sp. strain CALS-1 were 9.1, 0.045, and 5.8 μmol of viologen reduced min-1 mg of protein-1. These results suggest that CO plays a role in Methanothrix sp. strain CALS-1 similar to that of H2 in M. thermophila and are consistent with the conclusion that CO is an intermediate in a catabolic or anabolic pathway in Methanothrix sp. strain CALS-1; however, they could also be explained by passive equilibration of CO with a metabolic intermediate.  相似文献   

6.
The levels of seven intermediary enzymes involved in acetate and butyrate formation from acetyl coenzyme A in the saccharolytic anaerobe Clostridium acetobutylicum were investigated as a function of time in solvent-producing batch fermentations. Phosphate acetyltransferase and acetate kinase, which are known to form acetate from acetyl coenzyme A, both showed a decrease in specific activity when the organism reached the solvent formation stage. The three consecutive enzymes thiolase, β-hydroxybutyrylcoenzyme A dehydrogenase, and crotonase exhibited a coordinate expression and a maximal activity after growth had ceased. Only low levels of butyryl coenzyme A dehydrogenase activity were found. Phosphate butyryltransferase activity rapidly decreased after 20 h from 5 to 11 U/mg of protein to below the detection limit (1 mU/mg). Butyrate no longer can be formed, and the metabolic flux may be diverted to butanol. Butyrate kinase showed a 2.5- to 10-fold increase in specific activity after phosphate butyryltransferase activity no longer could be detected. These results suggest that the uptake of acetate and butyrate during solvent formation can not proceed via a complete reversal of the phosphate transferase and kinase reactions. The activities of all enzymes investigated as a function of time in vitro are much higher than the metabolic fluxes through them in vivo. This indicates that none of the maximal activities of the enzymes assayed is rate limiting in C. acetobutylicum.  相似文献   

7.
Sludge from a thermophilic, 55 degrees C digester produced methane without a lag period when enriched with butyrate. The sludge was found by most-probable-number enumeration to have ca. 5 x 10 butyrate-utilizing bacteria per ml. A thermophilic butyrate-utilizing bacterium was isolated in coculture with Methanobacterium thermoautotrophicum. This bacterium was a gram-negative, slightly curved rod, occurred singly, was nonmotile, and did not appear to produce spores. When this coculture was incubated with Methanospirillum hungatei at 37 degrees C, the quantity of methane produced was less than 5% of the methane produced when the coculture was incubated at 55 degrees C, the routine incubation temperature. The coculture required clarified digester fluid. The addition of yeast extract to medium containing 5% clarified digester fluid stimulated methane production when a Methanosarcina sp. was present. Hydrogen in the gas phase prevented butyrate utilization. However, when the hydrogen was removed, butyrate utilization began. Penicillin G and d-cycloserine caused the complete inhibition of butyrate utilization by the coculture. The ability of various ecosystems to convert butyrate to methane was studied. Marine sediments enriched with butyrate required a 2-week incubation period before methanogenesis began. Hypersaline sediments did not produce methane after 3 months when enriched with butyrate.  相似文献   

8.
The effects of acetate and butyrate during glycerol fermentation to 1,3-propanediol at pH 7.0 by Clostridium butyricum CNCM 1211 were studied. At pH 7.0, the calculated quantities of undissociated acetic and butyric acids were insufficient to inhibit bacterial growth. The initial addition of acetate or butyrate at concentrations of 2.5 to 15 gL−1 had distinct effects on the metabolism and growth of Clostridium butyricum. Acetate increased the biomass and butyrate production, reducing the lag time and 1,3-propanediol production. In contrast, the addition of butyrate induced an increase in 1,3-propanediol production (yield: 0.75 mol/mol glycerol, versus 0.68 mol/mol in the butyrate-free culture), and reduced the biomass and butyrate production. It was calculated that reduction of butyrate production could provide sufficient NADH to increase 1,3-propanediol production. The effects of acetate and butyrate highlight the metabolic flexibility of Cl. butyricum CNCM 1211 during glycerol fermentation. Received: 2 January 2001 / Accepted: 6 February 2001  相似文献   

9.
Inhibition of the fermentation of propionate to methane and carbon dioxide by hydrogen, acetate, and propionate was analyzed with a mesophilic propionate-acclimatized sludge that consisted of numerous flocs (size, 150 to 300 μm). The acclimatized sludge could convert propionate to methane and carbon dioxide stoichiometrically without accumulating hydrogen and acetate in a propionate-minimal medium. Inhibition of propionate utilization by propionate could be analyzed by a second-order substrate inhibition model (shown below) given that the substrate saturation constant, Ks, was 15.9 μM; the substrate inhibition constant, Ki, was 0.79 mM; and the maximum specific rate of propionate utilization, qm, was 2.15 mmol/g of mixed-liquor volatile suspended solids (MLVSS) per day: qs = qmS/[Ks + S + (S2/Ki)], where qs is the specific rate of propionate utilization and S is the initial concentration of undissociated propionic acid. For inhibition by hydrogen and acetate to propionate utilization, a noncompetitive product inhibition model was used: qs = qm/[1 + (P/Kp)n], where P is the initial concentration of hydrogen or undissociated acetic acid and Kp is the inhibition constant. Kinetic analysis gave, for hydrogen inhibition, Kp(H2) = 0.11 atm (= 11.1 kPa, 71.5 μM), qm = 2.40 mmol/g of MLVSS per day, and n = 1.51 and, for acetate inhibition, Kp(HAc) = 48.6 μM, qm = 1.85 mmol/g of MLVSS per day, and n = 0.96. It could be concluded that the increase in undissociated propionic acid concentration was a key factor in inhibition of propionate utilization and that hydrogen and acetate cooperatively inhibited propionate degradation, suggesting that hydrogenotrophic and acetoclastic methanogens might play an important role in enhancing propionate degradation to methane and carbon dioxide.  相似文献   

10.
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   

11.
Recent evidence suggests that the metabolism of some organisms, such as Escherichia coli, is remarkably efficient, producing close to the maximum amount of biomass per unit of nutrient consumed. This observation raises the question of what regulatory mechanisms enable such efficiency. Here, we propose that simple product-feedback inhibition by itself is capable of leading to such optimality. We analyze several representative metabolic modules—starting from a linear pathway and advancing to a bidirectional pathway and metabolic cycle, and finally to integration of two different nutrient inputs. In each case, our mathematical analysis shows that product-feedback inhibition is not only homeostatic but also, with appropriate feedback connections, can minimize futile cycling and optimize fluxes. However, the effectiveness of simple product-feedback inhibition comes at the cost of high levels of some metabolite pools, potentially associated with toxicity and osmotic imbalance. These large metabolite pool sizes can be restricted if feedback inhibition is ultrasensitive. Indeed, the multi-layer regulation of metabolism by control of enzyme expression, enzyme covalent modification, and allostery is expected to result in such ultrasensitive feedbacks. To experimentally test whether the qualitative predictions from our analysis of feedback inhibition apply to metabolic modules beyond linear pathways, we examine the case of nitrogen assimilation in E. coli, which involves both nutrient integration and a metabolic cycle. We find that the feedback regulation scheme suggested by our mathematical analysis closely aligns with the actual regulation of the network and is sufficient to explain much of the dynamical behavior of relevant metabolite pool sizes in nutrient-switching experiments.  相似文献   

12.
A mixed methanogenic culture fed with glucose was perturbed with butyrate and formate to investigate the role of formate in the acetogenesis of butyrate. A free energy analysis suggests that formate rather than H2 was the interspecies electron carrier for butyrate conversion into acetate for the culture studied.  相似文献   

13.
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 mumol of H(2) per liter in the culture headspace during growth on acetate or methanol. In coculture with D. vulgaris, the gaseous H(2) concentration was 相似文献   

14.
Hydrogen production by nitrogen-limited cultures of a thermophilic blue-green alga (cyanobacterium), Mastigocladus laminosus, was studied to develop the concept of a high-temperature biophotolysis system. Biophotolytic production of hydrogen by solar radiation was also demonstrated. Hydrogen consumption activity in these cultures was relatively high and is the present limiting factor on both the net rate and duration of hydrogen production.  相似文献   

15.
The control over ice crystal growth, melting, and shaping is important in a variety of fields, including cell and food preservation and ice templating for the production of composite materials. Control over ice growth remains a challenge in industry, and the demand for new cryoprotectants is high. Naturally occurring cryoprotectants, such as antifreeze proteins (AFPs), present one solution for modulating ice crystal growth; however, the production of AFPs is expensive and inefficient. These obstacles can be overcome by identifying synthetic substitutes with similar AFP properties. Zirconium acetate (ZRA) was recently found to induce the formation of hexagonal cavities in materials prepared by ice templating. Here, we continue this line of study and examine the effects of ZRA and a related compound, zirconium acetate hydroxide (ZRAH), on ice growth, shaping, and recrystallization. We found that the growth rate of ice crystals was significantly reduced in the presence of ZRA and ZRAH, and that solutions containing these compounds display a small degree of thermal hysteresis, depending on the solution pH. The compounds were found to inhibit recrystallization in a manner similar to that observed in the presence of AFPs. The favorable properties of ZRA and ZRAH suggest tremendous potential utility in industrial applications.  相似文献   

16.
Methanogenesis and the anaerobic metabolism of acetate were examined in the sediment and water column of Knaack Lake, a small biogenic meromictic lake located in central Wisconsin. The lake was sharply stratified during the summer and was anaerobic below a depth of 3 m. Large concentrations (4,000 μmol/liter) of dissolved methane were detected in the bottom waters. A methane concentration maximum occurred at 4 m above the sediment. The production of 14CH4 from 14C-labeled HCOOH, HCO3, and CH3OH and [2-14C]acetate demonstrated microbial methanogenesis in the water column of the lake. The maximum rate of methanogenesis calculated from reduction of H14CO3 by endogenous electron donors in the surface sediment (depth, 22 m) was 7.6 nmol/h per 10 ml and in the water column (depth, 21 m) was 0.6 nmol/h per 10 ml. The methyl group of acetate was simultaneously metabolized to CH4 and CO2 in the anaerobic portions of the lake. Acetate oxidation was greatest in surface waters and decreased with water depth. Acetate was metabolized primarily to methane in the sediments and water immediately above the sediment. Sulfide inhibition studies and temperature activity profiles demonstrated that acetate metabolism was performed by several microbial populations. Sulfide additions (less than 5 μg/ml) to water from 21.5 m stimulated methanogenesis from acetate, but inhibited CO2 production. Sulfate addition (1 mM) had no significant effect on acetate metabolism in water from 21.5 m, whereas nitrate additions (10 to 14,000 μg/liter) completely inhibited methanogenesis and stimulated CO2 formation.  相似文献   

17.
沼泽红假单胞菌乙酸光合放氢研究   总被引:21,自引:0,他引:21  
依据光合细菌生长代谢特性和有机废水降解主要产物类型,11种有机物被用于沼泽红假单胞菌(Rhodopseudomonas palustris)Z菌株的光合产氢研究,其中,乙酸反应体系产氢活性最高。在此基础上,研究了该菌株的生长与产氢动力学行为,探求了影响该菌株光合放氢的主要限制性影响因素。结果表明,该菌株产氢与生长部分相关。种子培养基和菌龄对产氢活性有明显影响。细胞最适产氢和生长所需要的光照强度和温度基本一致。当种子来源于硫酸铵高菌龄预培养物或谷氨酸钠对数期预培养物时,该菌株产氢活性显著增加,产氢延滞期明显缩短。氧浓度和接种量对产氢活性也有显著影响。供氢体和氮源浓度直接决定细胞的生长与光放氢活性。在低于70 mmol/L乙酸钠和15 mmol/L谷氨酸钠时,产氢活性随底物浓度的增加而增强。谷氨酸钠浓度高于15mmol/L时,由于游离NH4+的出现,产氢活性受到抑制,但却明显刺激细胞的生长。在标准状况下,该菌株的最大产氢速率可达19.4 mL·L-1·h-1。  相似文献   

18.
Hydrogen Metabolism by Rhodomicrobium vannielii   总被引:1,自引:0,他引:1  
Under appropriate cultural conditions, cell suspensions of Rhodomicrobium vannielii effect two distinct photoreactions involving molecular hydrogen: (i) the photoreduction of carbon dioxide, and (ii) the photoproduction of hydrogen.  相似文献   

19.
Isolation of thermophilic hydrogen bacteria was performed at 50°C using enrichment culture method. One of the four strains isolated, strain TH-1 grew most rapidly. Culture conditions of strain TH-1 were investigated. Optimum temperature and pH for growth proved to be 52°C and 7.0, respectively. There existed a positive correlation between the specific growth rate and the partial pressure of carbon dioxide in the gas phase. Ammonium and nitrate are the good nitrogen sources in that order. Effect of concentrations of nitrogen source, magnesium, ferrous and phosphate ions on the cell growth was also investigated. The maximum specific growth rate (μmax) of strain TH-1 was determined as 0.68 hr?1 by the cultivation at 52°C in a jar fermentor containing the optimal medium at pH 7.0.  相似文献   

20.
有机酸代谢途径在丁醇发酵过程中具有重要的作用,对细胞内碳流的分配和产物的合成影响显著。在7 L厌氧发酵罐中,进行了间歇添加乙酸或丁酸的发酵实验。结果表明,乙、丁酸的添加显著提高了总溶剂的生产效率,分别提高了47.1%和39.2%;此外,丁醇/丙酮比在添加丁酸的批次中提高了21.7%,在添加乙酸的批次中降低了16.2%;厌氧瓶中的发酵实验也证实了以上结果。有机酸代谢计算的结果表明,乙、丁酸的添加基本上阻断了相应有机酸闭环的吸收途径。基于相关报道和代谢计算结果,构建了针对乙、丁酸添加批次的图论模型,并利用该模型对不同发酵条件下的溶剂浓度和丁醇/丙酮比进行了计算。结果表明,该模型很好地预测了实验结果,合理地构建了乙、丁酸添加批次的信号传递线图。  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