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 novel sulfate-reducing bacterium (strain DePue) closely related to Desulfovibrio vulgaris ssp. vulgaris strain Hildenborough was isolated from the sediment of a heavy-metal impacted lake using established techniques. Although few physiological differences between strains DePue and Hildenborough were observed, pulse-field gel electrophoresis (PFGE) revealed a significant genome reduction in strain DePue. Comparative whole-genome microarray and polymerase chain reaction analyses demonstrated that the absence of genes annotated in the Hildenborough genome as phage or phage-related contributed to the significant genome reduction in strain DePue. Two morphotypically distinct temperate bacteriophage from strain Hildenborough were recovered using strain DePue as a host for plaque isolation.  相似文献   

2.
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   

3.
Desulfovibrio vulgaris Hildenborough is a Gram-negative sulfate-reducing bacterium (SRB), and the physiology of SRBs can impact many anaerobic environments including radionuclide waste sites, oil reservoirs and metal pipelines. In an attempt to understand D. vulgaris as a population that can adhere to surfaces, D. vulgaris cultures were grown in a defined medium and analysed for carbohydrate production, motility and biofilm formation. Desulfovibrio vulgaris wild-type cells had increasing amounts of carbohydrate into stationary phase and approximately half of the carbohydrate remained internal. In comparison, a mutant that lacked the 200 kb megaplasmid, strain DeltaMP, produced less carbohydrate and the majority of carbohydrate remained internal of the cell proper. To assess the possibility of carbohydrate re-allocation, biofilm formation was investigated. Wild-type cells produced approximately threefold more biofilm on glass slides compared with DeltaMP; however, wild-type biofilm did not contain significant levels of exopolysaccharide. In addition, stains specific for extracellular carbohydrate did not reveal polysaccharide material within the biofilm. Desulfovibrio vulgaris wild-type biofilms contained long filaments as observed with scanning electron microscopy (SEM), and the biofilm-deficient DeltaMP strain was also deficient in motility. Biofilms grown directly on silica oxide transmission electron microscopy (TEM) grids did not contain significant levels of an exopolysaccharide matrix when viewed with TEM and SEM, and samples stained with ammonium molybdate also showed long filaments that resembled flagella. Biofilms subjected to protease treatments were degraded, and different proteases that were added at the time of inoculation inhibited biofilm formation. The data indicated that D. vulgaris did not produce an extensive exopolysaccharide matrix, used protein filaments to form biofilm between cells and silica oxide surfaces, and the filaments appeared to be flagella. It is likely that D. vulgaris used flagella for more than a means of locomotion to a surface, but also used flagella, or modified flagella, to establish and/or maintain biofilm structure.  相似文献   

4.
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   

5.
Peculiar attributes revealed by sequencing the genome of Desulfovibrio vulgaris Hildenborough are analyzed, particularly in relation to the presence of a phosphotransferase system (PTS). The PTS is a typical bacterial carbohydrate transport system functioning via group translocation. Novel avenues for investigations are proposed emphasizing the metabolic diversity of D. vulgaris Hildenborough, especially the likely utilization of mannose-type sugars. Comparative analysis with PTS from other Gram-negative and Gram-positive bacteria indicates regulatory functions for the PTS of D. vulgaris Hildenborough, including catabolite repression and inducer exclusion. Chemotaxis towards PTS substrates is considered. Evidence suggests that this organism may not be a strict anaerobic sulfate reducer typical of the ocean, but a versatile organism capable of bidirectional transmigration and adaptation to both water and terrestrial environments.  相似文献   

6.
A library of 879 recombinant lambda phages, constructed for the genome of Desulfovibrio vulgaris Hildenborough, has been ordered by restriction fingerprinting. Restriction endonuclease HinfI digestion patterns were entered into a data base and sorted into 87 overlapping groups (contigs), with 19 clones remaining unattached. Eight of ten cloned genes of D. vulgaris, including dcrA, which encodes a transmembrane methyl-accepting protein, were assigned to contigs. Probing of a filter containing the lambda DNAs of the library with the labeled, conserved 3' end of the dcrA gene indicated hybridization to 54 clones distributed over multiple contigs. The presence of 11 additional dcr genes (dcrB to dcrL) was confirmed by direct cycled dideoxy sequencing of positive lambda clones. Since the ordered library provides only partial coverage of the D. vulgaris Hildenborough genome, we estimate that the dcr gene family has 16 members spread throughout the genome, making it the second largest gene family found in prokaryotes.  相似文献   

7.
The physiological properties of a hyd mutant of Desulfovibrio vulgaris Hildenborough, lacking periplasmic Fe-only hydrogenase, have been compared with those of the wild-type strain. Fe-only hydrogenase is the main hydrogenase of D. vulgaris Hildenborough, which also has periplasmic NiFe- and NiFeSe-hydrogenases. The hyd mutant grew less well than the wild-type strain in media with sulfate as the electron acceptor and H(2) as the sole electron donor, especially at a high sulfate concentration. Although the hyd mutation had little effect on growth with lactate as the electron donor for sulfate reduction when H(2) was also present, growth in lactate- and sulfate-containing media lacking H(2) was less efficient. The hyd mutant produced, transiently, significant amounts of H(2) under these conditions, which were eventually all used for sulfate reduction. The results do not confirm the essential role proposed elsewhere for Fe-only hydrogenase as a hydrogen-producing enzyme in lactate metabolism (W. A. M. van den Berg, W. M. A. M. van Dongen, and C. Veeger, J. Bacteriol. 173:3688-3694, 1991). This role is more likely played by a membrane-bound, cytoplasmic Ech-hydrogenase homolog, which is indicated by the D. vulgaris genome sequence. The physiological role of periplasmic Fe-only hydrogenase is hydrogen uptake, both when hydrogen is and when lactate is the electron donor for sulfate reduction.  相似文献   

8.
The genes encoding the periplasmic [Fe] hydrogenase from Desulfovibrio vulgaris subsp. oxamicus Monticello were cloned by exploiting their homology with the hydAB genes from D. vulgaris subsp. vulgaris Hildenborough, in which this enzyme is present as a heterologous dimer of alpha and beta subunits. Nucleotide sequencing showed that the enzyme is encoded by an operon in which the gene for the 46-kilodalton (kDa) alpha subunit precedes that of the 13.5-kDa beta subunit, exactly as in the Hildenborough strain. The pairs of hydA and hydB genes are highly homologous; both alpha subunits (420 amino acid residues) share 79% sequence identity, while the unprocessed beta subunits (124 and 123 amino acid residues, respectively) share 71% sequence identity. In contrast, there appears to be no sequence homology outside these coding regions, with the exception of a possible promoter element, which was found approximately 90 base pairs upstream from the translational start of the hydA gene. The recently discovered hydC gene, which may code for a 65.8-kDa fusion protein (gamma) of the alpha and beta subunits and is present immediately downstream from the hydAB genes in the Hildenborough strain, was found to be absent from the Monticello strain. The implication of this result for the possible function of the hydC gene product in Desulfovibrio species is discussed.  相似文献   

9.
C Li  H D Peck  Jr    A E Przybyla 《Journal of bacteriology》1986,165(2):644-646
A PyrF- mutant of Escherichia coli (SK1108, pyrF::Tn5 Kanr) was complemented with the Desulfovibrio vulgaris (Hildenborough) structural gene for orotidine-5'-phosphate decarboxylase (EC 4.1.1.23). Either orientation of a 1.6-kilobase-pair D. vulgaris DNA fragment (pLP3B or pLP3A) complemented the PyrF- strain suggesting that the D. vulgaris pyrF promoter was functional. The apparent product of the D. vulgaris pyrF gene was a single 26-kilodalton polypeptide. These results demonstrate the utility of E. coli cloning systems in studying metabolic and energetic pathways in sulfate-reducing bacteria.  相似文献   

10.
The sulfate-reducing bacterium Desulfovibrio vulgaris Hildenborough possesses a large number of porphyrin-containing proteins whose biosynthesis is poorly characterized. In this work, we have studied two putative CbiK cobaltochelatases present in the genome of D. vulgaris. The assays revealed that both enzymes insert cobalt and iron into sirohydrochlorin, with specific activities with iron lower than that measured with cobalt. Nevertheless, the two D. vulgaris chelatases complement an E. coli cysG mutant strain showing that, in vivo, they are able to load iron into sirohydrochlorin. The results showed that the functional cobaltochelatases have distinct roles with one, CbiK(C), likely to be the enzyme associated with cytoplasmic cobalamin biosynthesis, while the other, CbiK(P), is periplasmic located and possibly associated with an iron transport system. Finally, the ability of D. vulgaris to produce vitamin B 12 was also demonstrated in this work.  相似文献   

11.
Evidence is presented for an alternative to the superoxide dismutase (SOD)-catalase oxidative stress defense system in Desulfovibrio vulgaris (strain Hildenborough). This alternative system consists of the nonheme iron proteins, rubrerythrin (Rbr) and rubredoxin oxidoreductase (Rbo), the product of the rbo gene (also called desulfoferrodoxin). A Deltarbo strain of D. vulgaris was found to be more sensitive to internal superoxide exposure than was the wild type. Unlike Rbo, expression of plasmid-borne Rbr failed to restore the aerobic growth of a SOD-deficient strain of Escherichia coli. Conversely, plasmid-borne expression of two different Rbrs from D. vulgaris increased the viability of a catalase-deficient strain of E. coli that had been exposed to hydrogen peroxide whereas Rbo actually decreased the viability. A previously undescribed D. vulgaris gene was found to encode a protein having 50% sequence identity to that of E. coli Fe-SOD. This gene also encoded an extended N-terminal sequence with high homologies to export signal peptides of periplasmic redox proteins. The SOD activity of D. vulgaris is not affected by the absence of Rbo and is concentrated in the periplasmic fraction of cell extracts. These results are consistent with a superoxide reductase rather than SOD activity of Rbo and with a peroxidase activity of Rbr. A joint role for Rbo and Rbr as a novel cytoplasmic oxidative stress protection system in D. vulgaris and other anaerobic microorganisms is proposed.  相似文献   

12.
By using a synthetic deoxyoligonucleotide probe designed to recognize the structural gene for cytochrome cc3 from Desulfovibrio vulgaris Hildenborough, a 3.7-kb XhoI genomic DNA fragment containing the cc3 gene was isolated. The gene encodes a precursor polypeptide of 58.9 kDa, with an NH2-terminal signal sequence of 31 residues. The mature polypeptide (55.7 kDa) has 16 heme binding sites of the form C-X-X-C-H. Covalent binding of heme to these 16 sites gives a holoprotein of 65.5 kDa with properties similar to those of the high-molecular-weight cytochrome c (Hmc) isolated from the same strain by Higuchi et al. (Y. Higuchi, K. Inaka, N. Yasuoka, and T. Yagi, Biochim. Biophys. Acta 911:341-348, 1987). Since the data indicate that cytochrome cc3 and Hmc are the same protein, the gene has been named hmc. The Hmc polypeptide contains 31 histidinyl residues, 16 of which are integral to heme binding sites. Thus, only 15 of the 16 hemes can have bis-histidinyl coordination. A comparison of the arrangement of heme binding sites and coordinated histidines in the amino acid sequences of cytochrome c3 and Hmc from D. vulgaris Hildenborough suggests that the latter contains three cytochrome c3-like domains. Cloning of the D. vulgaris Hildenborough hmc gene into the broad-host-range vector pJRD215 and subsequent conjugational transfer of the recombinant plasmid into D. desulfuricans G200 led to expression of a periplasmic Hmc gene product with covalently bound hemes.  相似文献   

13.
14.
Two mutant strains of Desulfovibrio vulgaris Hildenborough lacking either the sod gene for periplasmic superoxide dismutase or the rbr gene for rubrerythrin, a cytoplasmic hydrogen peroxide (H(2)O(2)) reductase, were constructed. Their resistance to oxidative stress was compared to that of the wild-type and of a sor mutant lacking the gene for the cytoplasmic superoxide reductase. The sor mutant was more sensitive to exposure to air or to internally or externally generated superoxide than was the sod mutant, which was in turn more sensitive than the wild-type strain. No obvious oxidative stress phenotype was found for the rbr mutant, indicating that H(2)O(2) resistance may also be conferred by two other rbr genes in the D. vulgaris genome. Inhibition of Sod activity by azide and H(2)O(2), but not by cyanide, indicated it to be an iron-containing Sod. The positions of Fe-Sod and Sor were mapped by two-dimensional gel electrophoresis (2DE). A strong decrease of Sor in continuously aerated cells, indicated by 2DE, may be a critical factor in causing cell death of D. vulgaris. Thus, Sor plays a key role in oxygen defense of D. vulgaris under fully aerobic conditions, when superoxide is generated mostly in the cytoplasm. Fe-Sod may be more important under microaerophilic conditions, when the periplasm contains oxygen-sensitive, superoxide-producing targets.  相似文献   

15.
The genome of the sulphate reducing bacterium Desulfovibrio vulgaris Hildenborough, still considered a strict anaerobe, encodes two oxygen reductases of the bd and haem-copper types. The haem-copper oxygen reductase deduced amino acid sequence reveals that it is a Type A2 enzyme, which in its subunit II contains two c-type haem binding motifs. We have characterized the cytochrome c domain of subunit II and confirmed the binding of two haem groups, both with Met-His iron coordination. Hence, this enzyme constitutes the first example of a ccaa3 haem-copper oxygen reductase. The expression of D. vulgaris haem-copper oxygen reductase was found to be independent of the electron donor and acceptor source and is not altered by stress factors such as oxygen exposure, nitrite, nitrate, and iron; therefore the haem-copper oxygen reductase seems to be constitutive. The KCN sensitive oxygen reduction by D. vulgaris membranes demonstrated in this work indicates the presence of an active haem-copper oxygen reductase. D. vulgaris membranes perform oxygen reduction when accepting electrons from the monohaem cytochrome c553, thus revealing the first possible electron donor to the terminal oxygen reductase of D. vulgaris. The physiological implication of the presence of the oxygen reductase in this organism is discussed.  相似文献   

16.
A 47 kb genomic island (GEI) bracketed by 50 bp direct repeats, containing 52 annotated genes, was found to delete spontaneously from the genome of Desulfovibrio vulgaris Hildenborough. The island contains genes for site-specific recombinases and transposases, rubredoxin:oxygen oxidoreductase-1 (Roo1) and hybrid cluster protein-1 (Hcp1), which promote survival in air and nitrite stress. The numbering distinguishes these from the Roo2 and Hcp2 homologues for which the genes are located elsewhere in the genome. Cells with and without the island (GEI+ and GEI- cells respectively) were obtained by colony purification. GEI- cells arise in anaerobic cultures of colony-purified GEI+ cells, indicating that the site-specific recombinases encoded by the island actively delete this region. GEI+ cells survive better in microaerophilic conditions due to the presence of Roo1, whereas the Hcps appear to prevent inhibition by sulfur and polysulfide, which are formed by chemical reaction of sulfide and nitrite. Hence, the island confers resistance to oxygen and nitrite stress. However, GEI- cells have a higher growth rate in anaerobic media. Microarrays and enzyme activity stains indicated that the GEI- cells have increased expression of genes, which promote anaerobic energy conservation, explaining the higher growth rate. Hence, while lowering the efficiency of anaerobic metabolism, the GEI increases the fitness of D. vulgaris under stress conditions, a feature reminiscent of pathogenicity islands which allow more effective colonization of environments provided by the targeted hosts.  相似文献   

17.
Desulfovibrio vulgaris Hildenborough is a good model organism to study hydrogen metabolism in sulfate-reducing bacteria. Hydrogen is a key compound for these organisms, since it is one of their major energy sources in natural habitats and also an intermediate in the energy metabolism. The D. vulgaris Hildenborough genome codes for six different hydrogenases, but only three of them, the periplasmic-facing [FeFe], [FeNi]1, and [FeNiSe] hydrogenases, are usually detected. In this work, we studied the synthesis of each of these enzymes in response to different electron donors and acceptors for growth as well as in response to the availability of Ni and Se. The formation of the three hydrogenases was not very strongly affected by the electron donors or acceptors used, but the highest levels were observed after growth with hydrogen as electron donor and lowest with thiosulfate as electron acceptor. The major effect observed was with inclusion of Se in the growth medium, which led to a strong repression of the [FeFe] and [NiFe]1 hydrogenases and a strong increase in the [NiFeSe] hydrogenase that is not detected in the absence of Se. Ni also led to increased formation of the [NiFe]1 hydrogenase, except for growth with H2, where its synthesis is very high even without Ni added to the medium. Growth with H2 results in a strong increase in the soluble forms of the [NiFe]1 and [NiFeSe] hydrogenases. This study is an important contribution to understanding why D. vulgaris Hildenborough has three periplasmic hydrogenases. It supports their similar physiological role in H2 oxidation and reveals that element availability has a strong influence in their relative expression.  相似文献   

18.
Sulfate-reducing bacteria, like Desulfovibrio vulgaris Hildenborough, use the reduction of sulfate as a sink for electrons liberated in oxidation reactions of organic substrates. The rate of the latter exceeds that of sulfate reduction at the onset of growth, causing a temporary accumulation of hydrogen and other fermentation products (the hydrogen or fermentation burst). In addition to hydrogen, D. vulgaris was found to produce significant amounts of carbon monoxide during the fermentation burst. With excess sulfate, the hyd mutant (lacking periplasmic Fe-only hydrogenase) and hmc mutant (lacking the membrane-bound, electron-transporting Hmc complex) strains produced increased amounts of hydrogen from lactate and formate compared to wild-type D. vulgaris during the fermentation burst. Both hydrogen and CO were produced from pyruvate, with the hyd mutant producing the largest transient amounts of CO. When grown with lactate and excess sulfate, the hyd mutant also exhibited a temporary pause in sulfate reduction at the start of stationary phase, resulting in production of 600 ppm of headspace hydrogen and 6,000 ppm of CO, which disappeared when sulfate reduction resumed. Cultures with an excess of the organic electron donor showed production of large amounts of hydrogen, but no CO, from lactate. Pyruvate fermentation was diverse, with the hmc mutant producing 75,000 ppm of hydrogen, the hyd mutant producing 4,000 ppm of CO, and the wild-type strain producing no significant amount of either as a fermentation end product. The wild type was most active in transient production of an organic acid intermediate, tentatively identified as fumarate, indicating increased formation of organic fermentation end products in the wild-type strain. These results suggest that alternative routes for pyruvate fermentation resulting in production of hydrogen or CO exist in D. vulgaris. The CO produced can be reoxidized through a CO dehydrogenase, the presence of which is indicated in the genome sequence.  相似文献   

19.
20.
Sulfate-reducing bacteria, like Desulfovibrio vulgaris Hildenborough, have developed a set of reactions allowing them to survive in oxic environments and even to reduce molecular oxygen to water. D. vulgaris contains a cytoplasmic superoxide reductase (SOR) and a periplasmic superoxide dismutase (SOD) involved in the elimination of superoxide anions. To assign the function of SOD, the periplasmic [Fe] hydrogenase activity was followed in both wild-type and sod deletant strains. This activity was lower in the strain lacking the SOD than in the wild-type when the cells were exposed to oxygen for a short time. The periplasmic SOD is thus involved in the protection of sensitive iron-sulfur-containing enzyme against superoxide-induced damages. Surprisingly, production of the periplasmic [Fe] hydrogenase was higher in the cells exposed to oxygen than in those kept in anaerobic conditions. A similar increase in the amount of [Fe] hydrogenase was observed when an increase in the redox potential was induced by addition of chromate. Viability of the strain lacking the gene encoding [Fe] hydrogenase after exposure to oxygen for 1 h was lower than that of the wild-type. These data reveal for the first time that production of the periplasmic [Fe] hydrogenase is up-regulated in response to an oxidative stress. A new function of the periplasmic [Fe] hydrogenase in the protective mechanisms of D. vulgaris Hildenborough toward an oxidative stress is proposed.  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