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.
A plasmid in the archaebacterium Methanobacterium thermoautotrophicum   总被引:16,自引:0,他引:16  
The archaebacterium Methanobacterium thermoautotrophicum Marburg (DSM 2133) was found to contain a plasmid (pME2001) in covalently closed circular form. It was isolated by CsCl gradient centrifugation of total DNA in the presence of ethidium bromide. Multimers up to the hexamer were observed upon agarose gel electrophoresis and electron microscopy of a purified plasmid preparation. A restriction map was constructed. The length of plasmid pME2001 was determined to be approximately 4,500 bp. Southern hybridization of plasmid DNA to DNA extracted from Methanobacterium thermoautotrophicum delta H (DSM1053) revealed the presence of a plasmid with homologous sequences in the delta H strain.  相似文献   

2.
Methanobacterium thermoautotrophicum delta H and Marburg were adapted to grow in medium containing up to 0.65 M NaCl. From 0.01 to 0.5 M NaCl, there was a lag before cell growth which increased with increasing external NaCl. The effect of NaCl on methane production was not significant once the cells began to grow. Intracellular solutes were monitored by nuclear magnetic resonance (NMR) spectroscopy as a function of osmotic stress. In the delta H strain, the major intracellular small organic solutes, cyclic-2,3-diphosphoglycerate and glutamate, increased at most twofold between 0.01 and 0.4 M NaCl and decreased when the external NaCl was 0.5 M. M. thermoautotrophicum Marburg similarly showed a decrease in solute (cyclic-2,3-diphosphoglycerate, 1,3,4,6-tetracarboxyhexane, and L-alpha-glutamate) concentrations for cells grown in medium containing > 0.5 M NaCl. At 0.65 M NaCl, a new organic solute, which was visible in only trace amounts at the lower NaCl concentrations, became the dominant solute. Intracellular potassium in the delta H strain, detected by atomic absorption and 39K NMR, was roughly constant between 0.01 and 0.4 M and then decreased as the external NaCl increased further. The high intracellular K+ was balanced by the negative charges of the organic osmolytes. At the higher external salt concentrations, it is suggested that Na+ and possibly Cl- ions are internalized to provide osmotic balance. A striking difference of strain Marburg from strain delta H was that yeast extract facilitated growth in high-NaCl-containing medium. The yeast extract supplied only trace NMR-detectable solutes (e.g., betaine) but had a large effect on endogenous glutamate levels, which were significantly decreased. Exogenous choline and glycine, instead of yeast extract, also aided growth in NaCl-containing media. Both solutes were internalized with the choline converted to betaine; the contribution to osmotic balance of these species was 20 to 25% of the total small-molecule pool. These results indicate that M. thermoautotrophicum shows little changes in its internal solutes over a wide range of external NaCl. Furthermore, they illustrate the considerable differences in physiology in the delta H and Marburg strains of this organism.  相似文献   

3.
A wild-type strain of Methanobacterium thermoautotrophicum Marburg was transformed by DNA from strains resistant to 5-fluorouracil. Recipient cells were grown without selection on gellan gum (GELRITE) plates with DNA. Drug-resistant cells were recovered by replica plating the resulting colonies onto drug plates. Transformation required high-molecular-weight DNA with appropriate markers and was not observed on agar or in liquid media under a variety of conditions.  相似文献   

4.
The virulent bacteriophage psi M1 of Methanobacterium thermoautotrophicum Marburg mediated transduction of a resistance marker and of three biosynthesis markers. Transductants were observed at frequencies of 6 x 10(-4) to 5 x 10(-6)/PFU.  相似文献   

5.
D P Nagle  Jr  R Teal    A Eisenbraun 《Journal of bacteriology》1987,169(9):4119-4123
Growth of Methanobacterium thermoautotrophicum Marburg is inhibited by the pyrimidine, 5-fluorouracil (FU). It was shown previously that methanogenesis is not inhibited to the same extent as growth. A spontaneously occurring FU-resistant strain (RTAE-1) was isolated from a culture of strain Marburg. The growth of both strains was inhibited by 5-fluorodeoxyuridine but not 5-fluorocytosine, and the wild type was more susceptible to inhibition by 5-azauracil and 6-azauracil than was strain RTAE-1. The cellular targets for the pyrimidine analogs are not known. When the accumulation of 14C-labeled uracil or FU by the two strains was compared, the wild type took up 15-fold more radiolabel per cell than did the FU-resistant strain. In the wild type, radiolabel from uracil was incorporated into the soluble pool, RNA, and DNA. The metabolism of uracil appeared to involve a uracil phosphoribosyltransferase activity. Strain Marburg extracts contained this enzyme, whereas FU-resistant strain RTAE-1 extracts had less than 1/10 as much activity. Although it is possible that a change in permeability to the compounds plays a role in the stable resistance of strain RTAE-1, the fact that it lacks the ability to metabolize pyrimidines to nucleotides is sufficient to account for its phenotype.  相似文献   

6.
Abstract Growth of Methanosarcina barkeri (strain Fusaro) was found to be inhibited by 5-fluorouracil (FU) only at relatively high concentrations (>50 μg / ml ). Inhibition could not be relieved by uracil. Therefore, FU probably did not exert its effect via inhibition of DNA synthesis as is the case in other organisms. Control experiments with Methanobacterium thermoautotrophicum (strain Marburg) on the other hand revealed that the effect of FU on this archaebacterium is probably exerted at the level of nucleic acid synthesis. The M. thermoautotrophicum cultures rapidly acquired resistance towards the pyramidine analog.  相似文献   

7.
Formaldehyde conversion into methyl-coenzyme M involves (a) reaction of the substrate with 5,6,7,8-tetrahydromethanopterin (H4MPT) giving 5,10-methylene-H4MPT, followed by its reduction to 5-methyl-H4MPT and (b) transfer of the methyl group from the latter compound to coenzyme M. The reactions were studied in a resolved system from Methanobacterium thermoautotrophicum strain delta H. The first part (a) of the reactions was catalyzed by the 55% ammonium sulfate supernatant of cell-free extracts. The methyltransferase step (b) was dependent on an oxygen-sensitive enzyme, called methyltransferase a (MTa). Isolation of MTa was achieved by gel filtration on Sephacryl S-400. MTa was a high-molecular-weight complex of at least 2000 kDa and between 900 to 1500 kDa when purified in the absence and presence of the detergent CHAPS, respectively. The enzyme consisted of 100 kDa units composed of three subunits in an alpha beta gamma configuration with apparent molecular masses of 35, 33 and 31 kDa, respectively. The corrinoid, 5-hydroxybenzymidazolyl cobamide (B12HBI, Factor III) copurified with MTa and the latter contained 2 nmol B12HBI per mg protein. B12HBI present in MTa could be methylated under the appropriate conditions by 5-methyl-H4MPT. These findings suggest that the corrinoid is a prosthetic group of MTa. MTa may be homologous to the corrinoid membrane protein purified before from M. thermoautotrophicum strain Marburg (Schulz, H., Albracht, S.P.J., Coremans, J.M.C.C. and Fuchs, G. (1988) Eur. J. Biochem. 171, 589-597).  相似文献   

8.
A physical map of the Methanobacterium thermoautotrophicum Marburg chromosome was constructed by using pulsed-field gel electrophoresis of restriction fragments generated by NotI, PmeI, and NheI. The order of the fragments was deduced from Southern blot hybridization of NotI fragment probes to various restriction digests and from partial digests. The derived map is circular, and the genome size was estimated to be 1,623 kb. Several cloned genes were hybridized to restriction fragments to locate their positions on the map. Genes coding for proteins involved in the methanogenic pathway were located on the same segment of the circular chromosome. In addition, the genomes of a variety of thermophilic Methanobacterium strains were treated with restriction enzymes and analyzed by pulsed-field gel electrophoresis. The sums of the fragment sizes varied from 1,600 to 1,728 kb among the strains, and widely different macrorestriction patterns were observed.  相似文献   

9.
10.
Methyl-coenzyme M reductase (MCR) catalyzes the methane-forming step in methanogenic archaebacteria. The reductase has been characterized in detail from Methanobacterium thermoautotrophicum strain Marburg and delta H, which grow on H2 and CO2 as energy source. During purification of the enzyme we have now discovered a second methyl-coenzyme M reductase (MCR II) in the two strains, which elutes at lower salt concentration from anion-exchange columns than the enzyme (MCR I) previously characterized. MCR II is similar to MCR I in that it is also composed of three different subunits alpha, beta, and gamma but distinct from MCR I in that the gamma subunit is 5 kDa smaller, as revealed by sodium dodecyl sulfate/polyacrylamide gel electrophoresis. The N-terminal amino acid sequences of the alpha, beta, and gamma subunits of MCR II and MCR I were found to be different in several amino acid positions. The respective sequences showed, however, strong similarities indicating that MCR II was not derived from MCR I by limited proteolysis. The relative amounts of MCR I and MCR II present in the cells were affected by the growth conditions. When the cultures were supplied with sufficient H2 and and CO2 and the cells grew exponentially, essentially only MCR II was found. When growth was limited by the gas supply, MCR I predominated.  相似文献   

11.
Formate auxotroph of Methanobacterium thermoautotrophicum Marburg.   总被引:3,自引:1,他引:2       下载免费PDF全文
A formate-requiring auxotroph of Methanobacterium thermoautotrophicum Marburg was isolated after hydroxylamine mutagenesis and bacitracin selection. The requirement for formate is unique and specific; combined pools of other volatile fatty acids, amino acids, vitamins, and nitrogen bases did not substitute for formate. Compared with those of the wild type, cell extracts of the formate auxotroph were deficient in formate dehydrogenase activity, but cells of all of the strains examined catalyzed a formate-carbon dioxide exchange activity. All of the strains examined took up a small amount (200 to 260 mumol/liter) of formate (3 mM) added to medium. The results of the study of this novel auxotroph indicate a role for formate in biosynthetic reactions in this methanogen. Moreover, because methanogenesis from H2-CO2 is not impaired in the mutant, free formate is not an intermediate in the reduction of CO2 to CH4.  相似文献   

12.
The N5,N10-methylenetetrahydromethanopterin dehydrogenase from Methanobacterium thermoautotrophicum strain Marburg has been purified with reasonable yield and much higher specific activity than previously reported. For the first time it has been shown that both N5,N10-methylenetetrahydromethanopterin dehydrogenase and N5,N10-methenyltetrahydromethanopterin cyclohydrolase activities were stable under air and could be purified using aerobic operations. The dehydrogenase activity from Methanobacterium thermoautotrophicum Marburg was stable in phosphate buffer with or without glycerol or ammonium sulfate under both aerobic and anaerobic conditions. However, the presence of either 2-mercaptoethanol or dithiothreitol in the enzyme solution destroyed the enzyme activity during both aerobic and anaerobic incubations. Dehydrogenase was purified 62-fold using Phenyl-Sepharose and DEAE-Sephadex chromatography in succession under air. Both of these chromatographic methods separated dehydrogenase activity from N5,N10-methenyltetrahydromethanopterin cyclohydrolase; DEAE-Sephadex provided the best separation. Phenyl-Sepharose chromatography of the supernatant of cell extracts containing ammonium sulfate at 60% of saturation provided a 4.7-fold purification and 98% recovery of cyclohydrolase; this result established the air stability of N5,N10-methenyltetrahydromethanopterin cyclohydrolase from Methanobacterium thermoautotrophicum Marburg.  相似文献   

13.
Abstract Evidence shows the presence on the chromosome of Methanobacterium wolfei of a defective prophage which, by DNA-DNA hybridization, is closely related to the virulent archaeophage ψM1 of Methanobacterium thermoautotrophicum Marburg. Partial sequencing of a M. wolfei 16S rRNA gene and phylogenetic analysis indicated that this organism is more closely related to other representatives of the genus Methanobacterium than to M. thermoautotrophicum Marburg. The chromosomal region of M. wolfei encoding the putative prophage was found to be deleted for two non-contiguous segments of the phage ψM1 genome and thus encompassed only 80 to 90% of the ψM1 DNA. The prophage region was mapped to a 30 kb restriction fragment on the physical map of the M. wolfei chromosome. A randomly chosen DNA fragment was cloned from phage ψM1 DNA, as was its homologous counterpart from the chromosome of M. wolfei . The 126-bp region present in both clones exhibited 100% sequence identity.  相似文献   

14.
Abstract A gene ( sod ) encoding Superoxide dismutase (SOD) was isolated from the strictly anaerobic archaeon Methanobacterium thermoautotrophicum Marburg. Its identity was confirmed by functional complementation of an Escherichia coli mutant strain lacking SOD activity and by DNA sequence analysis of a cloned fragment. Upstream of sod , separated by a 5-bp intergenic region, lies the open reading frame orfk which potentially codes for a protein of 209 amino acid residues. The amino acid sequence for this presumptive product had a similarity coefficient of 55.5% to a subunit of the alkyl hydroperoxide reductase (encoded by the ahpC gene) from Salmonella typhimurium .  相似文献   

15.
Methyl-coenzyme M reductase (= component C) from Methanobacterium thermoautotrophicum (strain Marburg) was highly purified via anaerobic fast protein liquid chromatography on columns of Mono Q and Superose 6. The enzyme was found to catalyze the reduction of methylcoenzyme M (CH3-S-CoM) with N-7-mercaptoheptanoylthreonine phosphate (H-S-HTP = component B) to CH4. The mixed disulfide of H-S-CoM and H-S-HTP (CoM-S-S-HTP) was the other major product formed. The specific activity was up to 75 nmol min-1 mg protein-1. In the presence of dithiothreitol and of reduced corrinoids or titanium(III) citrate the specific rate of CH3-S-CoM reduction to CH4 with H-S-HTP increased to 0.5-2 mumol min-1 mg protein-1. Under these conditions the CoM-S-S-HTP formed from CH3-S-CoM and H-S-HTP was completely reduced to H-S-CoM and H-S-HTP. Methyl-CoM reductase was specific for H-S-HTP as electron donor. Neither N-6-mercaptohexanoylthreonine phosphate (H-S-HxoTP) nor N-8-mercaptooctanoylthreonine phosphate (H-S-OcoTP) nor any other thiol compound could substitute for H-S-HTP. On the contrary, H-S-HxoTP (apparent Ki = 0.1 microM) and H-S-OcoTP (apparent Ki = 15 microM) were found to be effective inhibitors of methyl-CoM reductase, inhibition being non-competitive with CH3-S-CoM and competitive with H-S-HTP.  相似文献   

16.
The gene for component A2 of the methylcoenzyme M reductase system from Methanobacterium thermoautotrophicum delta H was cloned, and its nucleotide sequence was determined. The gene for A2, designated atwA, encodes an acidic protein of 59,335 Da. Amino acid sequence analysis revealed partial homology of A2 to a number of eucaryotic and bacterial proteins in the ATP-binding cassette (ABC) family of transport systems. Component A2 possesses two ATP-binding domains. A 2.2-kb XmaI-BamHI fragment containing atwA and the surrounding open reading frames was cloned into pGEM-7Zf(+). A cell extract from this strain replaced purified A2 from M. thermoautotrophicum delta H in an in vitro methylreductase assay.  相似文献   

17.
18.
Riboflavin synthase was purified by a factor of about 1,500 from cell extract of Methanobacterium thermoautotrophicum. The enzyme had a specific activity of about 2,700 nmol mg(-1) h(-1) at 65 degrees C, which is relatively low compared to those of riboflavin synthases of eubacteria and yeast. Amino acid sequences obtained after proteolytic cleavage had no similarity with known riboflavin synthases. The gene coding for riboflavin synthase (designated ribC) was subsequently cloned by marker rescue with a ribC mutant of Escherichia coli. The ribC gene of M. thermoautotrophicum specifies a protein of 153 amino acid residues. The predicted amino acid sequence agrees with the information gleaned from Edman degradation of the isolated protein and shows 67% identity with the sequence predicted for the unannotated reading frame MJ1184 of Methanococcus jannaschii. The ribC gene is adjacent to a cluster of four genes with similarity to the genes cbiMNQO of Salmonella typhimurium, which form part of the cob operon (this operon contains most of the genes involved in the biosynthesis of vitamin B12). The amino acid sequence predicted by the ribC gene of M. thermoautotrophicum shows no similarity whatsoever to the sequences of riboflavin synthases of eubacteria and yeast. Most notably, the M. thermoautotrophicum protein does not show the internal sequence homology characteristic of eubacterial and yeast riboflavin synthases. The protein of M. thermoautotrophicum can be expressed efficiently in a recombinant E. coli strain. The specific activity of the purified, recombinant protein is 1,900 nmol mg(-1) h(-1) at 65 degrees C. In contrast to riboflavin synthases from eubacteria and fungi, the methanobacterial enzyme has an absolute requirement for magnesium ions. The 5' phosphate of 6,7-dimethyl-8-ribityllumazine does not act as a substrate. The findings suggest that riboflavin synthase has evolved independently in eubacteria and methanobacteria.  相似文献   

19.
2,3-Diphosphoglycerate (2,3-DPG) has been found to be the product of the enzymatic degradation of cyclic 2,3-diphosphoglycerate (cDPG) in the archaebacterium Methanobacterium thermoautotrophicum delta H. Although 2,3-DPG has not previously been detected as a major soluble component of M. thermoautotrophicum, large pools accumulated at an incubation temperature of 50 degrees C (below the optimum growth temperature of 62 degrees C). Under these conditions, cellular activity was significantly decreased; a return of the culture to the optimum growth temperature restored the 2,3-DPG pool back to original low levels and caused steady-state cDPG levels to increase again. While 13CO2-pulse/12CO2-chase experiments at 50 degrees C showed that the cDPG turned over, the appearance of 2,3-DPG at NMR-visible concentrations required at least 10 h. Production of 2,3-DPG in vivo was prevented by exposure of the cells to O2. The enzyme responsible for this hydrolysis of cDPG was purified by affinity chromatography and appears to be a 33-kDa protein. Activity was detected in the presence of oxygen and was enhanced by a solution of 1 M KCl, 25 mM MgCl2, and dithiothreitol. Both Km and Vmax have been determined at 37 degrees C; kinetics also indicate that in vitro the product, 2,3-DPG, is an inhibitor of cDPG hydrolysis. These findings are discussed in view of a proposed role for cDPG in methanogens.  相似文献   

20.
The enzymes involved in the purine interconversion pathway of wild-type and purine analog-resistant strains of Methanobacterium thermoautotrophicum Marburg were assayed by radiometric and spectrophotometric methods. Wild-type cells incorporated labeled adenine, guanine, and hypoxanthine, whereas mutant strains varied in their ability to incorporate these bases. Adenine, guanine, hypoxanthine, and xanthine were activated by phosphoribosyltransferase activities present in wild-type cell extracts. Some mutant strains simultaneously lost the ability to convert both guanine and hypoxanthine to the respective nucleotide, suggesting that the same enzyme activates both bases. Adenosine, guanosine, and inosine phosphorylase activities were detected for the conversion of base to nucleoside. Adenine deaminase activity was detected at low levels. Guanine deaminase activity was not detected. Nucleoside kinase activities for the conversion of adenosine, guanosine, and inosine to the respective nucleotides were detected by a new assay. The nucleotide-interconverting enzymes AMP deaminase, succinyl-AMP synthetase, succinyl-AMP lyase, IMP dehydrogenase, and GMP synthetase were present in extracts; GMP reductase was not detected. The results indicate that this autotrophic methanogen has a complex system for the utilization of exogenous purines.  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