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.
A preliminary endonuclease restriction map of a bacteriophage isolated from Desulfovibrio vulgaris has been established. BamHI cleaved whole phage DNA into four fragments while HindIII cut the same DNA into seven fragments. Mapping studies succeeded in linking the four BamHI fragments into two DNA segments; however, no linkage between the two segments was detected. These data imply that two phages were induced from cultures of D. vulgaris and that the two segments represented the DNA from these phages. Support for this hypothesis came from size approximation of restriction enzyme fragments, electron micrographs, and density gradients.  相似文献   

2.
Periplasmic hydrogenase [hydrogen:ferricytochrome c3 oxidoreductase, EC 1.12.2.1] from Desulfovibrio vulgaris Miyazaki K (MK) was purified to homogeneity. Its chemical and immunological properties were examined and compared with those of other Desulfovibrio hydrogenases. The pure enzyme showed a specific activity of 1,000 mumol H2 evolution min-1 (mg protein)-1. The enzyme had a molecular weight of 50,000 as estimated by gel filtration and consisted of a single polypeptide chain. The absorption spectrum of the enzyme was characteristic of an iron-sulfur protein and the extinction coefficients at 400 and 280 nm were 34 and 104 mM-1. cm-1, respectively. It contained 9.4 mol iron and 6.9 mol of acid-labile sulfide per mol. The amino acid composition of the preparation was very similar to the value reported for D. desulfuricans NRC 49001 hydrogenase. Rabbit antisera were prepared against the enzyme of D. vulgaris MK. Ouchterlony double diffusion and immunotitration tests of crude extracts from several strains of Desulfovibrio revealed that the enzyme from MK cells was immunologically identical with those from D. vulgaris Hildenborough and D. desulfuricans NRC 49001, but different from those from D. vulgaris Miyazaki F (MF) and Miyazaki Y, and D. desulfuricans Essex 6 strains. It is concluded that among Desulfovibrio hydrogenases, those from D. vulgaris MK, D. vulgaris Hildenborough and D. desulfuricans NRC 49001 form one group in terms of both subunit structure and antigenicity.  相似文献   

3.
Cytochrome c3 (Mr 26,000) has been characterized in Desulfovibrio vulgaris (Hildenborough) and its properties compared with polyhemic cytochromes c isolated from the same organism and from D. desulfuricans (Norway). It can be described as an octaheme cytochrome c3 constituted of two identical subunits. Absorption spectrum is similar to cytochrome c3 (Mr 13,000) and individual redox potentials have an average value of -180 mV.3 The N terminal sequence is compared with an homologous cytochrome isolated from D. desulfuricans Norway.  相似文献   

4.
The genomes of Desulfovibrio gigas and D. vulgaris   总被引:8,自引:0,他引:8  
Two-dimensional electrophoresis of sequential double-restriction digests showed that the genome of Desulfovibrio gigas compromised 1.63 x 10(6) bp (1.09 x 10(9) Dal) of DNA; an ammonia-limited chemostat population possessed an average of nine genomes per cell and a multiplying batch culture possessed approximately 17 genomes per cell. The genome size of D. vulgaris (Hildenborough) was 1.72 x 10(6) bp (1.14 x 10(9) Dal); a population from an ammonia-limited batch culture contained four genomes per cell. Control digestions and analyses with Escherichia coli GM4 agreed reasonably with published values: a genome size of 3.95 x 10(6) bp and approximately two genomes per cell from a stationary batch culture in glucose minimal medium. Desulfovibrio gigas carried two plasmids of approximately 70 MDal (1.05 x 10(5) bp) and approximately 40 MDal (6 x 10(4) bp); D. vulgaris (Hildenborough) contained one of approximately 130 MDal (1.95 x 10(5) bp). Single plasmids were also detected in a second strain of D. vulgaris and in strain Berre sol of D. desulfuricans but not in 10 other desulfovibrios including representatives of D. desulfuricans, D. vulgaris, D. salexigens and D. africanus.  相似文献   

5.
The cytochrome c(3) of Desulfovibrio desulfuricans and that of D. vulgaris were purified to homogeneity as judged by disc gel electrophoresis and by ultracentrifugation. Both cytochromes had an oxidation-reduction potential of -205 +/- 5 mv at pH 7.0 and showed characteristic absorption bands at 525 and 553 nm in the reduced state. The molecular weights of the two cytochromes (calculated from sedimentation and diffusion data) were similar, with values of 13,500 to 14,300 for D. desulfuricans and 13,800 to 14,700 for D. vulgaris. The two cytochromes differed in their electrophoretic properties on Geon and polyacrylamide gel electrophoresis and did not share a common precipitating antigenic determinant as judged by immunodiffusion data.  相似文献   

6.
Two ferredoxins, Fd I and Fd II, were isolated and purified from Desulfovibrio vulgaris Miyazaki. The major component, Fd I, is an iron-sulfur protein of Mr 12,000, composed of two identical subunits. The absorption spectra of Fd I and Fd II have a broad absorption shoulder near 400 nm characteristic of iron-sulfur proteins. The purity index, A400/A280, of Fd I is 0.69, and its millimolar absorption coefficient at 400 nm is 3.73 per Fe. It contains two redox centers with discrete redox behaviors. The amino acid composition and the N-terminal sequence of Fd I are similar to those of Fd III of Desulfovibrio africanus Benghazi and Fd II of Desulfovibrio desulfuricans Norway. Fd I does not serve as an electron carrier for the hydrogenase of D. vulgaris Miyazaki, but it serves as a carrier for pyruvate dehydrogenase of this bacterium. The evolution of H2 from pyruvate was observed by a reconstructed system containing purified hydrogenase, cytochrome C3, Fd I, partially purified pyruvate dehydrogenase, and CoA. The H2-sulfite reducing system can be reconstructed from the purified hydrogenase, cytochrome C3, Fd I and desulfoviridin (sulfite reductase), but the reaction rate is very slow compared to that of the crude extract at the same molar ratio of the components.  相似文献   

7.
The different nutritional properties of several Desulfovibrio desulfuricans strains suggest that either the strains are misclassified or there is a high degree of phenotypic diversity within the genus Desulfovibrio. The results of partial 16S rRNA and 23S rRNA sequence determinations demonstrated that Desulfovibrio desulfuricans ATCC 27774 and "Desulfovibrio multispirans" are closely related to the type strain (strain Essex 6) and that strains ATCC 7757, Norway 4, and El Agheila Z are not. Therefore, these latter three strains of Desulfovibrio desulfuricans are apparently misclassified. A comparative analysis of nearly complete 16S rRNA sequences in which we used a least-squares analysis method for evolutionary distances, an unweighted pair group method, a signature analysis method, and maximum parsimony was undertaken to further investigate the phylogeny of Desulfovibrio species. The species analyzed were resolved into two branches with origins deep within the delta subdivision of the purple photosynthetic bacteria. One branch contained five deep lineages, which were represented by (i) Desulfovibrio salexigens and Desulfovibrio desulfuricans El Agheila Z; (ii) Desulfovibrio africanus; (iii) Desulfovibrio desulfuricans ATCC 27774, Desulfomonas pigra, and Desulfovibrio vulgaris; (iv) Desulfovibrio gigas; and (v) Desulfomicrobium baculatus (Desulfovibrio baculatus) and Desulfovibrio desulfuricans Norway 4. A correlation between 16S rRNA sequence similarity and percentage of DNA relatedness showed that these five deep lineages are related at levels below the minimum genus level suggested by Johnson (in Bergey's Manual of Systematic Bacteriology, vol. 1, 1984). We propose that this branch should be grouped into a single family, the Desulfovibrionaceae. The other branch includes other genera of sulfate-reducing bacteria (e.g., Desulfobacter and Desulfococcus) and contains Desulfovibrio sapovorans and Desulfovibrio baarsii as separate, distantly related lineages.  相似文献   

8.
The properties of the periplasmic hydrogenase from Desulfovibrio desulfuricans ATCC 7757, previously reported to be a single-subunit protein [Glick, B. R., Martin, W. G., and Martin, S. M. (1980) Can. J. Microbiol. 26, 1214-1223] were reinvestigated. The pure enzyme exhibited a molecular mass of 53.5 kDa as measured by analytical ultracentrifugation and was found to comprise two different subunits of 42.5 kDa and 11 kDa, with serine and alanine as N-terminal residues, respectively. The N-terminal amino acid sequences of its large and small subunits, determined up to 25 residues, were identical to those of the Desulfovibrio vulgaris Hildenborough [Fe]-hydrogenase. D. desulfuricans ATCC 7757 hydrogenase was free of nickel and contained 14.0 atoms of iron and 14.4 atoms of acid-labile sulfur/molecule and had E400, 52.5 mM-1.cm-1. The purified hydrogenase showed a specific activity of 62 kU/mg of protein in the H2-uptake assay, and the H2-uptake activity was higher than H2-evolution activity. The enzyme isolated under aerobic conditions required incubation under reducing conditions to express its maximum activity both in the H2-uptake and 2H2/1H2 exchange reaction. The ratio of the activity of activated to as-isolated hydrogenase was approximately 3. EPR studies allowed the identification of two ferredoxin-type [4Fe-4S]1+ clusters in hydrogenase samples reduced by hydrogen. In addition, an atypical cluster exhibiting a rhombic signal (g values 2.10, 2.038, 1.994) assigned to the H2-activating site in other [Fe]-hydrogenases was detected in partially reduced samples. Molecular properties, EPR spectroscopy, catalytic activities with different substrates and sensitivity to hydrogenase inhibitors indicated that D. desulfuricans ATCC 7757 periplasmic hydrogenase is a [Fe]-hydrogenase, similar in most respects to the well characterized [Fe]-hydrogenase from D. vulgaris Hildenborough.  相似文献   

9.
Strains of two species of Desulfovibrio were examined for enzymes of the tricarboxylic acid cycle and related pathways. Pyruvate carboxylase (EC6.4.1.1) is present, and alpha-ketoglutarate is formed via the tricarboxylic acids. Glutamate, but not succinyl-CoA, arises from alpha-ketoglutarate. A pathway exists from pyruvate by malic enzyme (EC 1.1.1.39) activity to malate, then fumarate and succinate, again with no evidence of succinyl-CoA formation. The enzymes concerned with metabolism of these dicarboxylic acids show greater activity in the strains that can grow by fumarate dismutation. Glutamate (or glutamine), alpha-ketoglutarate, and yeast extract repress the enzymes that metabolize the tricarboxylic acids. There appears to be no glyoxylate cycle in Desulfovibrio vulgaris or D. desulfuricans.  相似文献   

10.
The interaction between hydrogenases from either Desulfovibrio desulfuricans or Clostridium pasteurianum and electron donors methyl viologen or polymeric viologens was examined. Extracts from each organism contained a single gel electophoretic band of active hydrogenase. The hydrogenase of D. desulfuricans was much more stable than that of Cl. pasteurianum. With methyl viologen apparent Km and Vm values were 0.5 mM and 0.62 mumole H2/min per milligram protein for the Cl. pasteurianum and 0.7 and 6.2 mumole H2/min per milligram protein, respectively, for the D. desulfuricans enzyme. The hydrogenases bound the polymeric viologens more tightly than methyl viologen, more so for the enzyme of D. desulfuricans than for Cl. pasteurianum. Maximal rate of hydrogen production was less with the polymeric than with methyl viologen. The results suggest that the D. desulfuricans enzyme in conjunction wiion than that from Cl. pasteurianum.  相似文献   

11.
Plasmid pJRDC800-1, containing the cyc gene encoding cytochrome c3 from Desulfovibrio vulgaris subsp. vulgaris Hildenborough, was transferred by conjugation from Escherichia coli DH5 alpha to Desulfovibrio desulfuricans G200. The G200 strain produced an acidic cytochrome c3 (pI = 5.8), which could be readily separated from the Hildenborough cytochrome c3 (pI = 10.5). The latter was indistinguishable from cytochrome c3 produced by D. vulgaris subsp. vulgaris Hildenborough with respect to a number of chemical and physical criteria.  相似文献   

12.
The gene coding for the flavodoxin protein from Desulfovibrio desulfuricans [Essex 6] (ATCC 29577) has been cloned and sequenced. The gene was identified on Southern blots of HindIII-digested genomic DNA by hybridization to the coding region for the flavodoxin from Desulfovibrio vulgaris [Hildenborough] (Krey, G.D., Vanin, E.F. and Swenson, R.P. (1988) J. Biol. Chem. 263, 15436-15443). Ultimately, a 1.8 kb TaqI fragment was cloned which contains an open reading frame of 447 nucleotides coding for an acidic protein of 148 amino acids and calculated molecular weight of 15,726. The derived amino acid sequence of this protein is 47% identical to the flavodoxin from D. vulgaris. Regions of the polypeptide which form the flavin mononucleotide binding site are largely homologous; however, some perhaps significant differences are noted. The aromatic amino acid residues that flank the flavin isoalloxazine ring in the D. vulgaris structure, i.e., tryptophan-60 and tyrosine-98, are conserved in this flavodoxin.  相似文献   

13.
This article aims to study hydrogen production/consumption in Desulfovibrio (D.) desulfuricans strain New Jersey, a sulfate reducer isolated from a medium undergoing active biocorrosion and to compare its hydrogen metabolism with two other Desulfovibrio species, D. gigas and D. vulgaris Hildenborough. Hydrogen production was followed during the growth of these three bacterial species under different growth conditions: no limitation of sulfate and lactate, sulfate limitation, lactate limitation, pyruvate/sulfate medium and in the presence of molybdate. Hydrogen production/consumption by D. desulfuricans shows a behavior similar to that of D. gigas but a different one from that of D. vulgaris, which produces higher quantities of hydrogen on lactate/sulfate medium. The three species are able to increase the hydrogen production when the sulfate became limiting. Moreover, in a pyruvate/sulfate medium hydrogen production was lower than on lactate/sulfate medium. Hydrogen production by D. desulfuricans in presence of molybdate is extremely high. Hydrogenases are key enzymes on production/consumption of hydrogen in sulfate reducing organisms. The specific activity, number and cellular localization of hydrogenases vary within the three Desulfovibrio species used in this work, which could explain the differences observed on hydrogen utilization.  相似文献   

14.
Aldehyde oxidoreductase (AOR) activity has been found in a number of sulfate-reducing bacteria. The enzyme that is responsible for the conversion of aldehydes to carboxylic acids is a mononuclear molybdenum enzyme belonging to the xanthine oxidase family. We report here the purification and characterization of AOR isolated from the sulfate-reducing bacterium Desulfovibrio (D.) aminophilus DSM 12254, an aminolytic strain performing thiosulfate dismutation. The enzyme is a homodimer (ca. 200 kDa), containing a molybdenum centre and two [2Fe-2S] clusters per monomer. UV/Visible and electron paramagnetic resonance (EPR) spectra of D. aminophilus AOR recorded in as-prepared and reduced states are similar to those obtained in AORs from Desulfovibrio gigas, Desulfovibrio desulfuricans and Desulfovibrio alaskensis. Despite AOR from D. aminophilus is closely related to other AORs, it presents lower activity towards aldehydes and no activity towards N-heterocyclic compounds, which suggests another possible role for this enzyme in vivo. A comparison of the molecular and EPR properties of AORs from different Desulfovibrio species is also included.  相似文献   

15.
Arbitrarily primed polymerase chain reaction (AP-PCR) method was applied to the differentiation of 15 (soil and intestinal) Desulfovibrio desulfuricans strains. The primer M 13, which is a core sequence of phage M 13, was found to be appropriate for the differentiation of isolates of this species. The analysis revealed characteristic band patterns for all of the examined strains of which two soil strains (DV-7 and DV-8) showed identical DNA fingerprints. According to Jaccard's coefficient, the soil bacterial group as well as intestinal bacterial group formed two different clusters. Furthermore, the soil strains showed greater variability than the intestinal isolates. Based on the AP-PCR fingerprints D. desulfuricans strains were differentiated depending on their origin. This study demonstrates that the typing method AP-PCR can be useful in epidemiologic investigations as a rapid and valuable tool for differentiation of the strains of D. desulfuricans species.  相似文献   

16.
A comparative study of 30 phages of Streptococcus thermophilus was performed based on DNA restriction profiles, DNA homology, structural proteins, packaging mechanisms, and host range data. All phages exhibited distinct DNA restriction profiles, with some phages displaying similarly sized restriction fragments. DNA homology was shown to be present among all 30 phages. The phages could be divided into two groups on the basis of their packaging mechanism as was derived from the appearance of submolar DNA fragments in restriction enzyme digests and the presence (cos-containing phages) or absence (pac-containing phages) of cohesive genomic extremities. Interestingly, the 19 identified cos-containing phages possessed two major structural proteins (32 and 26 kDa) in contrast to the remaining 11 pac-containing phages, which possessed three major structural proteins (41, 25, and 13 kDa). Southern hybridization demonstrated that all pac-containing phages tested contain homologs of the genes encoding the three major structural proteins of the pac-containing phage O1205, whereas all cos-containing phages tested exhibit homology to the gene specifying one of the structural components of the cos-containing phage phi 7201. Fifty-seven percent of the phages (both cos and pac containing) possessed the previously identified 2.2-kb EcoRI fragment of the temperate S. thermophilus phage Sfi18 (H. Brüssow, A. Probst, M. Frémont, and J. Sidoti, Virology 200:854-857, 1994). No obvious correlation was detected between grouping based on packaging mechanism and host range data obtained with 39 industrial S. thermophilus strains.  相似文献   

17.
The phosphorus metabolism of sulfate-reducing bacteria was, for the first time, probed by in vivo 31P NMR. A novel phosphoric anhydride diester compound was detected in Desulfovibrio desulfuricans ATCC 27774 at intracellular concentrations up to 5 mM. The compound has been extracted and partially purified by anion-exchange chromatography and analysed by 31P, 13C and 1H NMR. These studies show that the novel phosphorus-containing compound is formed by five carbon atoms and is probably cyclic, with a Mr of approximately 300. Various Desulfovibrio strains were examined in vivo for the presence of this phosphorus-containing compound. Detectable amounts of the novel metabolite were found in D. desulfuricans ATCC 27774 when grown on lactate/sulfate, lactate/thiosulfate or pyruvate/sulfate. The phosphorus-containing compound was not detected when this strain of D. desulfuricans was grown on lactate/nitrate or pyruvate; neither was it detected in two other strains which, like D. desulfuricans ATCC 27774, have the capability of utilizing nitrate as a terminal electron acceptor.  相似文献   

18.
When grown in the absence of added sulfate, cocultures of Desulfovibrio desulfuricans or Desulfovibrio vulgaris with Methanobrevibacter smithii (Methanobacterium ruminantium), which uses H(2) and CO(2) for methanogenesis, degraded lactate, with the production of acetate and CH(4). When D. desulfuricans or D. vulgaris was grown in the absence of added sulfate in coculture with Methanosarcina barkeri (type strain), which uses both H(2)-CO(2) and acetate for methanogenesis, lactate was stoichiometrically degraded to CH(4) and presumably to CO(2). 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 CH(4). Later, acetate was degraded to CH(4) and presumably to CO(2). 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 H(2) was detected in cocultures which contained D. desulfuricans and M. barkeri until after all lactate was degraded. The addition of H(2), but not of lactate, to the growth medium inhibited acetate degradation by pure cultures of M. barkeri. Pure cultures of M. barkeri produced CH(4) from acetate at a rate equivalent to that observed for cocultures containing M. barkeri. Inocula of M. barkeri grown with H(2)-CO(2) as the methanogenic substrate produced CH(4) 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 H(2) produced by the Desulfovibrio species during growth with lactate inhibited acetate degradation by M. barkeri.  相似文献   

19.
Washed cells of Desulfovibrio vulgaris strain Marburg (DSM 2119) reduced oxygen to water with H(2) as electron donor at a mean rate of 253 nmol O(2) min(-1) (mg protein)(-1). After separating the periplasm from the cells, more than 60% of the cytochrome c activity and 90% of the oxygen-reducing activity were found in the periplasmic fraction. Oxygen reduction and the reduction of cytochrome c with H(2) were inhibited by CuCl(2). After further separation of the periplasm by ultrafiltration (exclusion sizes 30, 50, and 100 kDa), oxygen reduction with H(2) occurred with the retentates only. Ascorbate plus tetramethyl-p-phenylenediamine (TMPD), however, were also oxidized by the filtrates. The stoichiometry of 1 mol O(2) reduced per 2 mol ascorbate oxidized indicated the formation of water. Our experiments present evidence that in D. vulgaris periplasmic hydrogenase and cytochrome c play a major role in oxygen reduction. Preliminary studies with other Desulfovibrio species indicated a similar function of periplasmic c-type cytochromes in D. desulfuricans CSN and D. termitidis KH1.  相似文献   

20.
The gene encoding the prismane protein from Desulfovibrio vulgaris (Hildenborough) was inserted into broad-host-range vector pSUP104. The recombinant plasmid, pJSP104, was transferred to D. vulgaris by conjugal plasmid transfer. In the transconjugant D. vulgaris cells the prismane protein was 25-fold overproduced. The overproduced prismane protein was characterized by molecular mass, isoelectric point, iron content and spectroscopical properties. Both the iron content and the ultraviolet/visible spectrum are identical to the wild-type protein indicating that iron incorporation in the overproduced protein is complete. EPR spectra of the dithionite-reduced form of the overproduced protein indicated that the Fe-S cluster might occur in a similar structure as found in inorganic model compounds containing a [6Fe-6S] prismane core. The as-isolated overproduced protein showed the presence of a second S = 1/2 spin system that was also detected in the corresponding prismane protein from D. desulfuricans (ATCC 27774), but not in the protein from wild-type D. vulgaris. This additional signal was irreversibly transformed to the 'wild-type' high-spin and low-spin systems upon two reduction/re-oxidation cycles. It is shown that the EPR spectroscopy of the overproduced prismane protein is very similar to that of the D. desulfuricans enzyme and, with the exception of the second S = 1/2 spin system, to that of the prismane protein from wild-type D. vulgaris. Contrary to claims for the D. desulfuricans protein, it is shown here that all data can be fully explained assuming a single [6Fe-6S] cluster, that might be titrated into four different redox states and occurs in up to three different spin systems in the one-electron reduced state.  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