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 250 毫秒
1.
The regulation of the Rhodobacter sphaeroides lexA gene has been analyzed using both gel-mobility experiments and lacZ gene fusions. PCR-mediated mutagenesis demonstrated that the second GAAC motif in the sequence GAACN7GAACN7GAAC located upstream of the R. sphaeroides lexA gene is absolutely necessary for its DNA damage-mediated induction. Moreover, mutagenesis of either the first or the third GAAC motif in this sequence reduced, but did not abolish, the inducibility of the R. sphaeroides lexA gene. A R. sphaeroides lexA-defective (Def) mutant has also been constructed by replacing the active lexA gene with an inactivated gene copy constructed in vitro. Crude extracts of the R. sphaeroides lexA(Def) strain are unable to form any protein-DNA complex when added to the wild-type lexA promoter of R. sphaeroides. Likewise, the R. sphaeroides lexA(Def) cells constitutively express the recA and lexA genes. All these data clearly indicate that the lexA gene product is the negative regulator of the R. sphaeroides SOS response. Furthermore, the morphology, growth and viability of R. sphaeroides lexA(Def) cultures do not show any significant change relative to those of the wild-type strain. Hence, R. sphaeroides is so far the only bacterial species whose viability is known not to be affected by the presence of a lexA(Def) mutation. Received: 31 January 2000 / Accepted: 3 April 2000  相似文献   

2.
Two kinds of mutants of Rhodopseudomonas sphaeroides that should be useful in extending genetic analysis of this organism have been isolated. One is deficient in recombination and has been used to isolate derivatives of the plasmid R 68.45 which incorporate chromosomal genes of R. sphaeroides. The other is apparently defective in a DNA restriction enzyme; transfer of plasmid borne chromosomal genes of R. sphaeroides from Escherichia coli back to R. sphaeroides is greatly enhanced in these mutants.In memory of R. Y. Stanier  相似文献   

3.
Plasmid RP4 (among others) has been transferred from Escherichia coli to Rhodopseudomonas sphaeroides. Data bearing on the physical presence of the plasmid and its expression of drug resistance determinants in R. sphaeroides are presented. Conditions of transfer between R. sphaeroides strains, between R. sphaeroides and R. capsulata, and between R. sphaeroides and E. coli have been carefully defined.  相似文献   

4.
Mannitol dehydrogenase (MDH) from Rhodobacter sphaeroides Si4 was overproduced by constructing a strain that overexpresses the MDH gene and by producing high cell concentrations via fed-batch cultivation in a bioreactor. With the gene of mannitol dehydrogenase (mtlK) cloned into the expression vector pKK223-3 expression of MDH in Escherichia coli was obtained, but the specific enzyme activity was lower than in R. sphaeroides Si4. In order to overexpress mtlK in R. sphaeroides, plasmid pAK82 was constructed by cloning a DNA fragment carrying mtlK into the broad-host-range expression vector pRK415. When pAK82 was introduced into R. sphaeroides Si4 the specific mannitol dehydrogenase activity in the strain obtained was 0.48 unit (U)mg–1, 3.4-fold higher thain in the wild type. In this way the enzyme yield from cultivation in a bioreactor could be improved from 110 Ul–1 to 350 Ul–1. A further increase in productivity was obtained by fed-batch cultivation of R. sphaeroides Si4 [pAK82]. Using this cultivation method can optical density of 27.6 was reached in the bioreactor, corresponding to a dry mass of 16.6 g l–1. Since MDH formation correlated with biomass production, the MDH yield could be raised to 918 Ul–1, an 8.3-fold increase in comparison to batch cultivation of the wild-type strain.Dedicated to Prof. Fritz Wagner on the occasion of his 65th birthday.  相似文献   

5.
Rhodobacter sphaeroides is a purple non-sulfur photosynthetic bacteria that participates in the anoxic cycling of carbon both as the primary producer and as the light-stimulated consumers of the reduced organic compounds. In this study, six different organic acids, i.e. acetate, lactate, oxaloacetate, malate, succinate, and citrate, were selected and used to analyze the relationships between the organic acid source and the cell growth. The C4 compound exhibited an enhanced cell growth compared to the other organic acids, and the growth rate of R. sphaeroides that was grown with 0.03 M succinic acid was significantly 3.2-fold faster than the C6 compound of 0.03 M citrate. Additionally, the cell growth of R. sphaeroides was enhanced with increasing light intensity, and the growth rate and the dry cell weight of R. sphaeroides that were grown under the light conditions of 15 W/m2 were 2.0- and 1.2-fold higher than R. sphaeroides at 3 W/m2. Therefore, the high light intensity probably affected the growth of R. sphaeroides. Moreover, the blue-colored light emitting diode (LED) exhibited a highest growth rate and cell concentration of R. sphaeroides among the various types of LEDs, and the enhanced cell growth phenomenon under the blue LED conditions was dramatically stimulated at low concentrations of succinic acid, which was compensatory for succinic acid. Therefore, a high light intensity and a blue LED as the light source were necessary for the enhanced cell growth for the C4 organic acid, i.e. succinic acid.  相似文献   

6.
1) A number of overlapping clones have been isolated from a Rhodobacter sphaeroides gene bank. Following conjugative gene transfer from Escherichia coli these clones restore a wild type phenotype to several mutants unable to synthesise bacteriochlorophyll. 2) The insert DNA was analysed by restriction mapping and together the clones form the basis of the first restriction map of the 45 kb photosynthetic gene cluster of Rb. sphaeroides. 3) This cluster is bounded on one side by puh A encoding the reaction centre H polypeptide and on the other by the puf operon encoding reaction centre L and M apoproteins and light harvesting LH1 and polypeptides. 4) DNA fragments from the 45 kb cluster were used to probe genomic DNA from other photosynthetic bacteria. Using heterologous hybridisation conditions, a significant degree of homology is shown between Rb. sphaeroides and these other bacteria, suggesting close evolutionary links with Rb. capsulatus in particular.  相似文献   

7.
The photosynthetic bacteria are at the forefront of the study of many aspects of photosynthesis, including photopigment biosynthesis, photosynthetic-membrane assembly, light-harvesting, and reaction center photochemistry. The facultative growth of some photosynthetic bacteria, their simple photosystems, and their ease of genetic manipulation have all contributed to advances in these areas. Amongst these bacteria, the purple non-sulfur bacterium Rhodobacter sphaeroides has emerged as, arguably, the leading contender for a model system in which to integrate the studies of all the different aspects of the assembly and function of the photosynthetic apparatus. Many of the genes encoding photosynthesis-related proteins are known to be clustered within a small region of the genome in this organism. As a further aid to studying the assembly and function of the photosystem of Rb. sphaeroides, the DNA sequence for a genomic segment containing this photosynthesis gene cluster (PGC) has been assembled from previous EMBL submissions and formerly unpublished data. The Rb. sphaeroides PGC is 40.7 kb in length and consists of 38 open reading frames encoding the reaction center H, L and M subunits, the and polypeptides of the light-harvesting I (B875) complex, and the enzymes of bacteriochlorophyll and carotenoid biosynthesis. PGCs are a feature of gene organization in several photosynthetic bacteria, and the similarities between the clusters of Rb. sphaeroides and Rb. capsulatus have been apparent for some time. Here we present the first comprehensive analysis of the PGC of Rb. sphaeroides, as well as a comparison with that of Rb. capsulatus.  相似文献   

8.
A new virulent bacteriophage, designated RZ1, was isolated from a local pond on the facultative phototrophic bacteriumRhodobacter sphaeroides ZZ101. Electron microscopic studies revealed that, in general morphology, phage RZ1 resembles the bacteriophage ofEscherichia coli. The host range of phage RZ1 is limited to some strains ofR. sphaeroides. The phage genome consists of double-stranded DNA of about 44 kb lacking cohesive ends and seems to present terminal redundancy and cyclic permutation. RZ1 phage may carry out a lytic cycle only in recombination-defective mutants ofR. sphaeroides. Nevertheless, a derivative of the RZ1 phage, termed RW1, able to grow in recombination-proficient strains ofR. sphaeroides, has also been obtained. In vitro restriction analysis of both RZ1 and RW1 phages shows the presence of a rearrangement in their DNA. Generalized transduction of Strr and Rifr chromosomal markers has not been detected with either RZ1 or RW1 phages.  相似文献   

9.
The recA gene of Rhodobacter sphaeroides 2.4.1 has been isolated by complementation of a UV-sensitive RecA? mutant of Pseudomonas aeruginosa. Its complete nucleotide sequence consists of 1032 bp, encoding a polypeptide of 343 amino acids. The deduced amino acid sequence displayed highest identity to the RecA proteins from Rhizobium mehloti, Rhizobium phaseoli, and Agrobacterium tumefaciens. An Escherichia coli-like SOS consensus region, which functions as a binding site for the LexA repressor molecule was not present in the 215 by upstream region of the R. sphaeroides recA gene. Nevertheless, by using a recA-lacZ fusion, we have shown that expression of the recA gene of R. sphaeroides is inducible by DNA damage. A recA-defective strain of R. sphaeroides was obtained by replacement of the active recA gene by a gene copy inactived in vitro. The resulting recA mutant exhibited increased sensitivity to UV irradiation, and was impaired in its ability to perform homologous recombination as well as to trigger DNA damage-mediated expression. This is the first recA gene from a Gram-negative bacterium that lacks an E. coli-like SOS box but whose expression has been shown to be DNA damage-inducible and auto-regulated.  相似文献   

10.
The recA genes of Agrobacterium tumefaciens, Rhizobium meliloti, Rhizobium phaseoli and Rhodobacter sphaeroides, species belonging to the alpha-group bacteria of the Proteobacteria class, have been fused in vitro to the lacZ gene of Escherichia coli. By using a mini-Tn5 transposon derivative, each of these recA-lacZ fusions was introduced into the chromosome of each of the four species, and into that of E. coli. The recA genes of three of the alpha bacteria are induced by DNA damage when inserted in A. tumefaciens, R. phaseoli or R. meliloti chromosomes. The expression of the recA gene of R. sphaeroides is DNA damage-mediated only when present in its own chromosome; none of the genes is induced in E. coli. Likewise, the recA gene of E. coli is not induced in any of the four alpha species. These data indicate that A. tumefaciens, R. meliloti and R. phaseoli possess a LexA-like repressor, which is able to block the expression of their recA genes, as well as that of R. sphaeroides, but not the recA gene of E. coli. The LexA repressor of R. sphaeroides does not repress the recA gene of A. tumefaciens, R. meliloti, R. phaseoli or E. coli.  相似文献   

11.
Specificity of protein–protein interactions plays a vital role in signal transduction. The chemosensory pathway of Rhodobacter sphaeroides comprises multiple homologues of chemotaxis proteins characterized in organisms such as Escherichia coli. Three CheA homologues are essential for chemotaxis in R. sphaeroides under laboratory conditions. These CheAs are differentially localized to two chemosensory clusters, one at the cell pole and one in the cytoplasm. The polar CheA, CheA2, has the same domain structure as E. coli CheA and can phosphorylate all R. sphaeroides chemotaxis response regulators. CheA3 and CheA4 independently localize to the cytoplasmic cluster; each protein has a subset of the CheA domains, with CheA3 phosphorylating CheA4 together making a functional CheA protein. Interestingly, CheA3‐P can only phosphorylate two response regulators, CheY6 and CheB2. R. sphaeroides CheAs exhibit two interesting differences in specificity: (i) the response regulators that they phosphorylate and (ii) the chemosensory cluster to which they localize. Using a domain‐swapping approach we investigated the role of the P1 and P5 CheA domains in determining these specificities. We show that the P1 domain is sufficient to determine which response regulators will be phosphorylated in vitro while the P5 domain is sufficient to localize the CheAs to a specific chemosensory cluster.  相似文献   

12.
The regulation of the Rhodobacter sphaeroides lexA gene has been analyzed using both gel-mobility experiments and lacZ gene fusions. PCR-mediated mutagenesis demonstrated that the second GAAC motif in the sequence GAACN7GAACN7GAAC located upstream of the R. sphaeroides lexA gene is absolutely necessary for its DNA damage-mediated induction. Moreover, mutagenesis of either the first or the third GAAC motif in this sequence reduced, but did not abolish, the inducibility of the R. sphaeroides lexA gene. A R. sphaeroides lexA-defective (Def) mutant has also been constructed by replacing the active lexA gene with an inactivated gene copy constructed in vitro. Crude extracts of the R. sphaeroides lexA(Def) strain are unable to form any protein-DNA complex when added to the wild-type lexA promoter of R. sphaeroides. Likewise, the R. sphaeroides lexA(Def) cells constitutively express the recA and lexA genes. All these data clearly indicate that the lexA gene product is the negative regulator of the R. sphaeroides SOS response. Furthermore, the morphology, growth and viability of R. sphaeroides lexA(Def) cultures do not show any significant change relative to those of the wild-type strain. Hence, R. sphaeroides is so far the only bacterial species whose viability is known not to be affected by the presence of a lexA(Def) mutation.  相似文献   

13.
This paper describes the DNA sequence of the photosynthesis region of Rhodobacter sphaeroides 2.4.1T. The photosynthesis gene cluster is located within a ~73 kb AseI genomic DNA fragment containing the puf, puhA, cycA and puc operons. A total of 65 open reading frames (ORFs) have been identified, of which 61 showed significant similarity to genes/proteins of other organisms while only four did not reveal any significant sequence similarity to any gene/protein sequences in the database. The data were compared with the corresponding genes/ORFs from a different strain of R.sphaeroides and Rhodobacter capsulatus, a close relative of R.sphaeroides. A detailed analysis of the gene organization in the photosynthesis region revealed a similar gene order in both species with some notable differences located to the pucBAC=cycA region. In addition, photosynthesis gene regulatory protein (PpsR, FNR, IHF) binding motifs in upstream sequences of a number of photosynthesis genes have been identified and shown to differ between these two species. The difference in gene organization relative to pucBAC and cycA suggests that this region originated independently of the photosynthesis gene cluster of R.sphaeroides.  相似文献   

14.
From polluted water of a lagoon pond a new type of denitrifying photosynthetic purple bacteria was isolated. With respect to morphology, fine structure, photopigments, requirement for growth factors, the range of utilization of organic substrates for phototrophic growth and DNA base ratio, the denitrifying strains show the closest resemblance to Rhodopseudomonas sphaeroides and were therefore described as a subspecies named R. sphaeroides forma sp. denitrificans. The new isolates grow well with nitrate anaerobically in the dark accompanying the evolution of nitrogen gas. They cannot assimilate nitrate as the nitrogen source for growth.  相似文献   

15.
Several optical imaging techniques have been used to monitor bacterial tropisms for cancer. Most such techniques require genetic engineering of the bacteria to express optical reporter genes. This study investigated a novel tumor‐targeting strain of bacteria, Rhodobacter sphaeroides 2.4.1 (R. sphaeroides), which naturally emits near‐infrared fluorescence, thereby facilitating the visualization of bacterial tropisms for cancer. To determine the penetration depth of bacterial fluorescence, various numbers of cells (from 108 to 1010 CFU) of R. sphaeroides and two types of Escherichia coli, which stably express green fluorescent protein (GFP) or red fluorescent protein (RFP), were injected s.c. or i.m. into mice. Bacterial tropism for cancer was determined after i.v. injection of R. sphaeroides (108 CFU) into mice implanted s.c. with eight types of tumors. The intensity of the fluorescence signal in deep tissue (muscle) from R. sphaeroides was much stronger than from E. coli‐expressing GFP or RFP. The near‐infrared fluorescence signal from R. sphaeroides was visualized clearly in all types of human or murine tumors via accumulation of bacteria. Analyses of C‐reactive protein and procalcitonin concentrations and body weights indicated that i.v. injection of R. sphaeroides does not induce serious systemic immune reactions. This study suggests that R. sphaeroides could be used as a tumor‐targeting microorganism for the selective delivery of drugs to tumor tissues without eliciting a systemic immune reaction and for visualizing tumors.  相似文献   

16.
17.

Background  

Rhodobacter sphaeroides 2.4.1 is a metabolically versatile organism that belongs to α-3 subdivision of Proteobacteria. The present study was to identify the extent, history, and role of gene duplications in R. sphaeroides 2.4.1, an organism that possesses two chromosomes.  相似文献   

18.
The adenylate kinases (EC 2.7.4.3) from photosynthetically grown Rhodopseudomonas palustris, Rhodopseudomonas sphaeroides and Rhodospirillum rubrum were purified to homogeneity by the same procedure. The purified enzymes showed optimal rates of activity with MgCl2 at 25° C and pH 8.0. They were found to be heat labile and were characterized by pI-values of 4.5. Apparent molecular weights of 33 500 for R. palustris, 34 400 for R. sphaeroides and 32 100 for R. rubrum were determined by high performance liquid chromatography. No separation into subunits was observed by use of sodium dodecylsulfate polyacrylamide gel electrophoresis. The apparent K m -values for ADP corresponded to 0.26 mM for R. palustris, 0.27 mM for R. sphaeroides and 0.24 mM for R. rubrum. ADP in excess had a strong inhibitory effect. Competitive product inhibition was found for AMP, with K i-values of 0.017 mM for R. palustris, 0.018 mM for R. sphaeroides and 0.014 mM for R. rubrum. A competitive inhibitor likewise was P1,P5-di(adenosine-5)pentaphosphate with K i-values of 0.020 M for R. palustris and R. sphaeroides, and 0.017 M for R. rubrum. Sulfhydryl-reacting reagents like p-chloromercuribenzoate and iodoacetic acid were found to be non-inhibitory. All measurements of adenylate kinase activity were carried out with the stabilized and most sensitive luciferin-luciferase system.  相似文献   

19.
For the commercial production of CoQ10, batch-type fermentations were attempted in a 150-l fermenter using a mutant strain of R. sphaeroides. Optimum temperature and initial aeration rate were found to be 30°C and 2 vvm, respectively. Under optimum fermentation conditions, the maximum value of specific CoQ10 content was achieved reproducibly as 6.34 mg/g DCW after 24 h, with 3.02 g/l of DCW. During the fermentation, aeration shift (from the adequate aeration at the early growth phase to the limited aeration in active cellular metabolism) was a key factor in CoQ10 production for scale-up. A higher value of the specific CoQ10 content (8.12 mg/g DCW) was achieved in fed-batch fermentation and comparable to those produced by the pilot-scale fed-batch fermentations of A. tumefaciens, which indicated that the mutant strain of R. sphaeroides used in this study was a potential high CoQ10 producer. This is the first detailed study to demonstrate a pilot-scale production of CoQ10 using a mutant strain of R. sphaeroides.  相似文献   

20.
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