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 35 毫秒
1.
Methanogenesis in thermophilic biogas reactors   总被引:2,自引:0,他引:2  
Methanogenesis in thermophilic biogas reactors fed with different wastes is examined. The specific methanogenic activity with acetate or hydrogen as substrate reflected the organic loading of the specific reactor examined. Increasing the loading of thermophilic reactors stabilized the process as indicated by a lower concentration of volatile fatty acids in the effluent from the reactors. The specific methanogenic activity in a thermophilic pilot-plant biogas reactor fed with a mixture of cow and pig manure reflected the stability of the reactor. The numbers of methanogens counted by the most probable number (MPN) technique with acetate or hydrogen as substrate were further found to vary depending on the loading rate and the stability of the reactor. The numbers of methanogens counted with antibody probes in one of the reactor samples was 10 times lower for the hydrogen-utilizing methanogens compared to the counts using the MPN technique, indicating that other non-reacting methanogens were present. Methanogens that reacted with the probe againstMethanobacterium thermoautotrophicum were the most numerous in this reactor. For the acetate-utilizing methanogens, the numbers counted with the antibody probes were more than a factor of 10 higher than the numbers found by MPN. The majority of acetate utilizing methanogens in the reactor wereMethanosarcina spp. single cells, which is a difficult form of the organism to cultivatein vitro. No reactions were observed with antibody probes raised againstMethanothrix soehngenii orMethanothrix CALS-1 in any of the thermophilic biogas reactors examined. Studies using 2-14C-labeled acetate showed that at high concentrations (more than approx. 1 mM) acetate was metabolized via the aceticlastic pathway, transforming the methyl-group of acetate into methane. When the concentration of acetate was less than approx. 1 mM, most of the acetate was oxidized via a two-step mechanism (syntrophic acetate oxidation) involving one organism oxidizing acetate into hydrogen and carbon dioxide and a hydrogen-utilizing methanogen forming the products of the first microorganism into methane. In thermophilic biogas reactors, acetate oxidizing cultures occupied the niche ofMethanothrix species, aceticlastic methanogens which dominate at low acetate concentrations in mesophilic systems. Normally, thermophilic biogas reactors are operated at temperatures from 52 to 56° C. Experiments using biogas reactors fed with cow manure showed that the same biogas yield found at 55° C could be obtained at 61° C after a long adaptation period. However, propionate degradation was inhibited by increasing the temperature.  相似文献   

2.
The effects of Cd, Cu, and Ni on pure cultures of thermophilic methanogenic bacteria were studied. The bacteria used wereMethanobacterium thermoautotrophicum and TAM, a thermophilic, acetate-decarboxylating, methanogenic bacterium. Much lower concentrations of heavy metals were needed to cause initial inhibition of TAM (1 mg/liter Cu and Cd; 5 mg/liter Ni) compared withM. thermoautotrophicum (10 mg/liter Cu and Cd; and 100 mg/liter Ni). No growth of TAM occurred at 5 mg/liter Cu and 25 mg/liter Ni, while the corresponding values forM. thermoautotrophicum were 50 mg/liter Cu and 200 mg/liter Ni. Cd (50 mg/liter) was totally inhibitory toM. thermoautotrophicum but allowed minimal growth of TAM. Ni stimulated both organisms at an optimal concentration of 5 mg/liter forM. thermoautotrophicum and 1 mg/liter for TAM. The toxicity of Cd and Cu was found to depend upon the presence of Ni in the medium.  相似文献   

3.
Summary A 1.6-kb fragment of DNA from the thermophilic, methane-producing, anaerobic archaebacteriumMethanobacterium thermoautotrophicum H has been cloned and sequenced. This DNA complements mutations in both the purE1 and purE2 loci ofEscherichia coli. The sequence of theM. thermoautotrophicum DNA predicts that complementation inE. coli results from the synthesis of a polypeptide with a molecular weight of 36,249. A polypeptide apparently of this molecular weight is synthesized inE. coli minicells containing recombinant plasmids that carry the cloned fragment of methanogen DNA. We have previously cloned and sequenced a purE-complementing gene from the mesophilic methanogenMethanobrevibacter smithii. The two methanogen-derived purE-complementing genes are 53% homologous and encode polypeptides that are 45% homologous in their amino acid sequences but would be 74% homologous if conservative amino acid substitutions were considered as maintaining sequence homology. The genome ofM. thermoautotrophicum has a molar G+C content of 49.7%, whereas the genome ofM. smithii is 30.6% G+C. Conservation of encoded amino acids while accommodating the very different G+C contents is accomplished by use of different codons that encode the same amino acid. The majority of base changes occur at the third codon position. The intergenic regions of the clonedM. thermoautotrophicum DNA contain sequences previously identified as ribosome binding sites and as putative methanogen promoters. Although the two purE-complementing genes are apparently derived from a common ancestor, only the gene fromM. smithii maintains a codon usage that conforms to the RNY rule.  相似文献   

4.
Benzene, toluene, ethylbenzene and xylene (BTEX) substrate interactions for a mesophilic (25°C) and thermophilic (50°C) toluene-acclimatized composted pine bark biofilter were investigated. Toluene, benzene, ethylbenzene, o-xylene, m-xylene and p-xylene removal efficiencies, both individually and in paired mixtures with toluene (1:1 ratio), were determined at a total loading rate of 18.1 g m–3 h–1 and retention time ranges of 0.5–3.0 min and 0.6–3.8 min for mesophilic and thermophilic biofilters, respectively. Overall, toluene degradation rates under mesophilic conditions were superior to degradation rates of individual BEX compounds. With the exception of p-xylene, higher removal efficiencies were achieved for individual BEX compounds compared to toluene under thermophilic conditions. Overall BEX compound degradation under mesophilic conditions was ranked as ethylbenzene >benzene >o-xylene >m-xylene >p-xylene. Under thermophilic conditions overall BEX compound degradation was ranked as benzene >o-xylene >ethylbenzene >m-xylene >p-xylene. With the exception of o-xylene, the presence of toluene in paired mixtures with BEX compounds resulted in enhanced removal efficiencies of BEX compounds, under both mesophilic and thermophilic conditions. A substrate interaction index was calculated to compare removal efficiencies at a retention time of 0.8 min (50 s). A reduction in toluene removal efficiencies (negative interaction) in the presence of individual BEX compounds was observed under mesophilic conditions, while enhanced toluene removal efficiency was achieved in the presence of other BEX compounds, with the exception of p-xylene under thermophilic conditions.  相似文献   

5.
30 strains of xylanolytic thermophilic actinomycetes were isolated from composted grass and cattle manure and identified as members of the generaThermomonospora, Saccharomonospora, Microbispora, Streptomyces andActinomadura. Screening of these strains for extracellular xylanase indicated that strains ofSaccharomonospora andMicrobispora generally were poor xylanase producers (0.5–1.5 U/ml) whereas relatively high activities were observed in cultures ofStreptomyces andActionomadura (4–12 U/ml).A preliminary characterization of the enzymes of strains of the latter genera suggested that xylanases of all the strains ofActinomadura exhibited higher thermostabilities than those ofStreptomyces. To evaluate the potential of thermophilicActinomadura for industrial applications, xylanases of three strains were studied in more detail. The highest activity levels for xylanases were observed in cultures grown on xylan and wheat bran. The optimal pH and temperature for xylanase activities ranged from 6.0 to 7.0 and 70 to 80°C. The enzymes exhibited considerable thermostability at their optimum temperature. The half-lives at 75°C were in the range from 6.5 to 17h. Hydrolysis of xylan by extracellular xylanases yielded xylobiose, xylose and arabinose as principal products. Estimated by the amount of reducing sugars liberated the degree of hydrolysis was 55 to 65%. Complete utilization of xylan is presumably achieved by -xylosidase activities which could be shown to be largely cell-associated in the 3Actinomadura strains.  相似文献   

6.
Summary Estimates of bacterial numbers from raw sewage sludge and sludge treated by thermophilic aerobic digestion were compared with simple indicators of sludge quality and concentrations of potential substrates. Significant differences were found between sludge types for all but one of the variables examined (frequency of dividing cells). During a stable period of digestor operation, the average number of viable obligate thermophiles present in digested sludge (1.63 × 106 ml–1) was approximately 102-fold greater than in feed sludge (1.10 × 104 ml–1). Total numbers of bacteria were slightly greater in digested sludge (3.24 × 1010 ml–1) than in feed sludge (2.39 × 10 ml–10), as were viable counts of bacteria at incubation temperatures of 37°C and 55°C. Significant correlation was found between viable counts of bacteria at 37°C and 55°C for digested sludge, and 65°C and 55°C for feed sludge. The numbers of obligate thermophiles present and the total of bacteria present were related to the temperature and pH of the digested sludge and inversely related to the numbers ofEscherichia coli and coliforms present, which were not detected at temperatures greater than 50°C.  相似文献   

7.
The pyrimidine analogue 5-fluorouracil was shown to be a potent inhibitor of the growth ofMethanobacterium thermoautotrophicum strain Marburg (50% inhibition of growth at 1 g ml–1). The nucleoside, 5-fluorodeoxyuridine, also inhibited growth, but the nucleotide 5-fluorodeoxyuridylate did not inhibit, nor did 5-fluorocytosine. Several nucleobases and nucleosides were used as potential antagonists of fluorouracil and fluorodeoxyuridine. Of these, only uracil in excess over fluorouracil relieved the inhibition of growth. These results imply that a pyrimidine salvage pathway is present inM. thermoautotrophicum. 5-Fluorouracil does not inhibit methane production. Although treated cultures produced less methane than did controls, more than twice as much methane was synthesized per cell. This result suggests that methanogenesis is uncoupled from growth by 5-fluorouracil.  相似文献   

8.
DNA ligase genes of the thermophilic archaeae Pyrococcus abyssi (Pab DNA ligase) and Methanobacterium thermoautotrophicum (Mth DNA ligase) were cloned in Escherichia coli. The resulting recombinant enzymes were tested for activity in a ligation mixture with two oligonucleotides, one containing a preformed hairpin structure. The yield of the reaction products was maximal at temperatures close to 70°C for either enzyme; their accumulation reached a plateau at 70–75% of the theoretical yield at a stoichiometric enzyme-to-substrate ratio. The enzymes differed in thermal stability. The half-life of Pab DNA ligase was approximately 60 min at 90°C. Mth DNA ligase was completely inactivated within 10 min at this temperature. The recombinant DNA ligases from P. abyssi and M. thermoautotrophicum remained stabile during long-term storage at 4°C.  相似文献   

9.
The isolation of a thermophilic biosurfactant producing Bacillus SP   总被引:3,自引:0,他引:3  
Summary A thermophilic Bacillus strain has been isolated on a hydrocarbon containing medium and grew at up to 50°C. This strain produced biosurfactant and its 20h old culture broth had low surface and interfacial tension (27–29 and 1.5 mN/m, respectively). It emulsified Kerosene and other hydrocarbons efficiently (E–24 = 95 %) and was able to recover more than 95 % of the residual oil from sandpack columns. Potential uses in oil industries are discussed.  相似文献   

10.
Recent work on biosurfactant release by thermophilic dairy streptococci is reviewed. There is a suggestion thatStreptococcus thermophilus isolates may release biosurfactants that stimulate detachment of already-adhering cells and leave an anti-adhesive coating on a substratum. A previously published rapid screening method is described for the identification of biosurfactant-releasing microorganisms, and growth medium supplements to enhance biosurfactant release by thermophilic dairy streptococci are reported. New experimental work described includes the isolation and purification of biosurfactants from dairy isolates by thin layer chromatography. Many compounds isolated were extremely surface-active and reduced the water surface tension to values around 30 mJ m–2 at a concentration of 10 mg ml–1. Most importantly, the thin layer chromatograms of various isolates resembled each other, and an adsorbed purified compound from one isolate retarded the deposition to glass of another isolate by a factor of two. Provided our findings implicate that these biosurfactants could also be adsorbed to heat exchanger plates in pasteurizers and thereby retard colonization by thermophilic streptococci, these compounds may have major economic implications. Further work is required, however.  相似文献   

11.
A complete gene, xyl10C, encoding a thermophilic endo-1,4-β-xylanase (XYL10C), was cloned from the acidophilic fungus Bispora sp. MEY-1 and expressed in Pichia pastoris. XYL10C shares highest nucleotide and amino acid sequence identities of 57.3 and 49.7%, respectively, with a putative xylanase from Aspergillus fumigatus Af293 of glycoside hydrolase family 10. A high expression level in P. pastoris (73,400 U ml−1) was achieved in a 3.7–l fermenter. The purified recombinant XYL10C was thermophilic, exhibiting maximum activity at 85°C, which is higher than that reported from any fungal xylanase. The enzyme was also highly thermostable, exhibiting ~100% of the initial activity after incubation at 80°C for 60 min and >87% of activity at 90°C for 10 min. The half lives of XYL10C at 80 and 85°C were approximately 45 and 3 h, respectively. It had two activity peaks at pH 3.0 and 4.5–5.0 (maximum), respectively, and was very acid stable, retaining more than 80% activity after incubation at pH 1.5−6.0 for 1 h. The enzyme was resistant to Co2+, Mn2+, Cr3+ and Ag+. The specific activity of XYL10C for oat spelt xylan was 18,831 U mg−1. It also had wide substrate specificity and produced simple products (65.1% xylose, 25.0% xylobiose and 9.9% xylan polymer) from oat spelt xylan.  相似文献   

12.
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   

13.
Strains of thermophilic bacilli were screened for cellulolytic activity by gel diffusion assay on selective medium at 55°C. Strain B-41361, identified as a strain of Bacillus licheniformis, displayed activity against carboxymethylcellulose. Zymogram analysis demonstrated several catalytically active polypeptides with the most prominent species having a mass of 37 kDa. The enzyme was purified 60-fold with a 17% yield and specific activity of 183 U/mg. The amino terminal sequence was homologous to members of glycoside hydrolase family 5. Optimal temperature was 65°C (measured over 30 min), but the enzyme was most stable at 60°C, retaining greater than 90% activity after one hour. The enzyme had a broad pH range, with maximal activity at pH 6.0, 75% maximal activity at pH 4.5, and 40% at pH 10. The enzyme hydrolyzed p-nitrophenylcellobioside, barley β-glucan, and lichenan, but no activity was detected against avicel or acid-swollen cellulose.Mention of a trade name or commercial products in this publication is solely for the purpose of providing specific information and does not imply recommendation or endorsement by the U.S. Department of Agriculture.  相似文献   

14.
We report the production, purification and characterization of a DNA ligase encoded by the thermophilic archaeon Methanobacterium thermoautotrophicum. The 561 amino acid Mth ligase catalyzed strand-joining on a singly nicked DNA in the presence of a divalent cation (magnesium, manganese or cobalt) and ATP (Km 1.1 µM). dATP can substitute for ATP, but CTP, GTP, UTP and NAD+ cannot. Mth ligase activity is thermophilic in vitro, with optimal nick-joining at 60°C. Mutational analysis of the conserved active site motif I (KxDG) illuminated essential roles for Lys251 and Asp253 at different steps of the ligation reaction. Mutant K251A is unable to form the covalent ligase–adenylate intermediate (step 1) and hence cannot seal a 3′-OH/5′-PO4 nick. Yet, K251A catalyzes phosphodiester bond formation at a pre-adenylated nick (step 3). Mutant D253A is active in ligase–adenylate formation, but defective in activating the nick via formation of the DNA–adenylate intermediate (step 2). D253A is also impaired in phosphodiester bond formation at a pre-adenylated nick. A profound step 3 arrest, with accumulation of high levels of DNA–adenylate, could be elicited for the wild-type Mth ligase by inclusion of calcium as the divalent cation cofactor. Mth ligase sediments as a monomer in a glycerol gradient. Structure probing by limited proteolysis suggested that Mth ligase is a tightly folded protein punctuated by a surface-accessible loop between nucleotidyl transferase motifs III and IIIa.  相似文献   

15.
Escherichia coli TG1, transformed with an expression plasmid pAQN carrying the aqualysin I (AQI) gene derived from Thermus aquaticus YT-1 under the control of the tac promoter, was cultivated under various conditions in order to find fermentation conditions for the efficient production of the thermophilic protease, AQI. The amount of AQI produced was closely related to the growth phase at the time of isopropyl--d-thiogalactopyranoside (IPTG) induction, and the highest production was obtained when it was added during the exponential growth phase. The addition of yeast extract had a greater effect on AQI production than did Polypeptone or casamino acids, and AQI productivity increased from 1.1 × 103 kU/g to 2.7 × 103 kU/g cells when 2 g/l yeast extract was supplied. Furthermore, the specific growth rate improved from 0.35 h–1 to 0.89 h–1 when 5 g/l yeast extract was supplied. The culture temperature also affected AQI gene expression. When the temperature was shifted from 37°C to 34°C at the time of IPTG induction, 19 kU/ml enzymatically active AQI was obtained, corresponding to a 28% increase over the amount produced in a batch culture without a shift. This is about a 44-fold higher yield than was obtained from the original strain, T. aquaticus YT-1.  相似文献   

16.
Inorganic nitrogen metabolism in the obligate anaerobic thermophiles Chlostridium thermosaccharolyticum and Clostridium thermoautotrophicum differs in several respects. C. thermosaccharolyticum contains a nitrogenase as inferred from NH 4 + repressible C2H2 reduction, a glutamine synthetase which is partially repressed by ammonium, very labile glutamate synthase activities with both NADH and NADPH, NADPH-dependent glutamate dehydrogenase, and NH 4 + -dependent asparagine synthetase. C. thermoautotrophicum contains no nitrogenase, but glutamine synthetase, no glutamate synthase, no glutamate dehydrogenase, but a NADH-dependent alanine dehydrogenase and a NH 4 + -dependent asparagine synthetase.Abbreviation GOGAT glutamine-oxoglutarate amidotransferase amidotransferase (glutamate synthase)  相似文献   

17.
Biosurfactant production of eight Streptococcus thermophilus strains, isolated from heat exchanger plates in the downstream side of the regenerator section of pasteurizers in the dairy industry has been measured using axisymmetric drop shape analysis by profile (ADSA-P). Strains were grown in M17 broth with either lactose, saccharose or glucose added. After harvesting, cells were suspended in water or in 10 mm potassium phosphate buffer, pH 7.0, and suspension droplets were put on a piece of FEP-Teflon. Changes in droplet profile were analysed by ADSA-P to yield the surface tension decrease due to biosurfactant production as a function of time. Surface tension decreases larger than 8 mJ·m–2 were taken as indicative of biosurfactant production. Only five strains produced biosurfactants in water, solely when saccharose was added to the growth medium. In buffer, all strains produced biosurfactants and production was generally greater than in water. Also, most strains suspended in buffer produced maximally when saccharose was added to the growth medium, whereas one strain produced maximally in buffer upon the addition of glucose. Four strains suspended in buffer produced biosurfactants when glucose was added and only two strains when lactose was added. The possible role of these biosurfactants as anti-adhesives in the dairy industry and for the survival of these strains in natural systems is discussed.Correspondence to: H. J. Busscher  相似文献   

18.
2,4-Dinitrophenol and gramicidin D completely inhibited growth and methanogenesis inMethanobacterium thermoautotrophicum. At low K+ concentrations valinomycin inhibited growth and methanogenesis relatively slightly, at high K+ concentrations (0.1m KCl) growth was inhibited completely and methanogenesis by about 50%. Monensin and nigericin inhibited growth completely; methanogenesis was inhibited like with valinomycin at high K+ concentrations. The results can be interpreted in terms of Mitchell’s chemiosmotic theory as follows. The protonmotive force inM. thermoautotrophicum is the basic source of energy for endergonic processes. Dissipation of the electrical component of protonmotive force may probably be compensated by an increased generation of the proton gradient. However, the osmotic component is essential for growth ofM. thermoautotrophicum.  相似文献   

19.
Thermostable enzymes from thermophiles have attracted extensive studies. However, little is known about thermophilic lysin of bacteriophage obtained from deep-sea hydrothermal vent. In this study, a lysin from deep-sea thermophilic bacteriophage Geobacillus virus E2 (GVE2) was characterized for the first time. It was found that the GVE2 lysin was highly homologous with N-acetylmuramoyl-L-alanine amidases. After expression in Escherichia coli, the recombinant GVE2 lysin was purified. The recombinant lysin was active over a range of temperature from 40 °C to 80 °C, with an optimum at 60 °C. Its optimal pH was 6.0, and it was stable over a wide range of pH from 4.0 to 10.0. The lysin was highly active when some enzyme inhibitors or detergents (phenylmethylsulfonyl fluoride, Tween 20, Triton X-100, and chaps) were used. However, it was strongly inhibited by sodium dodecyl sulfate and ethylene diamine tetraacetic acid. Its enzymatic activity could be slightly stimulated in the presence of Na+ and Li+. But the metal ions Mg2+, Ba2+, Zn2+, Fe3+, Ca2+, and Mn2+ at concentrations of 1 or 10 mM showed inhibitions to the lysin activity. Our study demonstrated the first characterization of lysin from deep-sea thermophilic bacteriophage.  相似文献   

20.
Methanogenesis was studied using stirred, bench-top fermentors of 3-1 working volume fed on a semi-continuous basis with waste obtained from cattle fed a high grain, finishing diet. Digestion was carried out at 40 and 60°C. CH4 production was 11.8, 18.3, 61.9 and 84.5% higher in the thermophilic than the mesophilic digestor at the 3, 6, 9 and 12 g volatile solids (VS) l–1 reactor volume loading rates, respectively. When compared on an energetic basis CH4 production was 7.4, 18.3, 72.9 and 107.3 kJ day higher in the thermophilic than the mesophilic digestor. CH4 production decreased more rapidly with each increase in VS loading rate and decrease in retention time (RT) in the mesophilic than the thermophilic digestor. When expressed as l g–1 VS fed or as kJ kJ–1 fed, the amount of CH4 was 49% less at the highest compared to the lowest loading rate in the mesophilic digestor. In the thermophilic digestor the decrease was only 16%. Propionate accumulated in the mesophilic digestor at the two highest loading rates, reaching concentrations of about 50 mM, but were only about 13 mM in the thermophilic digestor. Isobutyrate, isovalerate plus 2-methylbutyrate, and valerate also accumulated at the higher loading rates.  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