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.
In contrast to the situation in enteric bacteria, chemotaxis in Rhodobacter sphaeroides requires transport and partial metabolism of chemoattractants. A chemotaxis operon has been identified containing homologues of the enteric cheA , cheW , cheR genes and two homologues of the cheY gene. However, mutations in these genes have only minor effects on chemotaxis. In enteric species, CheW transmits sensory information from the chemoreceptors to the histidine protein kinase, CheA. Expression of R. sphaeroides cheW in Escherichia coli showed concentration-dependent inhibition of wild-type behaviour, increasing counter-clockwise rotation and thus smooth swimming — a phenotype also seen when E. coli cheW is overexpressed in E. coli . In contrast, overexpression of R. sphaeroides cheW in wild-type R. sphaeroides inhibited motility completely, the equivalent of inducing tumbly motility in E. coli . Expression of R. sphaeroides cheW in an E. coli Δ cheW chemotaxis mutant complemented this mutation, confirming that CheW is involved in chemosensory signal transduction. However, unlike E. coli Δ cheW mutants, in-frame deletion of R. sphaeroides cheW did not affect either swimming behaviour or chemotaxis to weak organic acids, although the responses to sugars were enhanced. Therefore, although CheW may act as a signal-transduction protein in R. sphaeroides , it may have an unusual role in controlling the rotation of the flagellar motor. Furthermore, the ability of a Δ cheW mutant to swim normally and show wild-type responses to weak acids supports the existence of additional chemosensory signal-transduction pathways.  相似文献   

2.
The Escherichia coli chemotaxis signal transduction pathway has: CheA, a histidine protein kinase; CheW, a linker between CheA and sensory proteins; CheY, the effector; and CheZ, a signal terminator. Rhodobacter sphaeroides has multiple copies of these proteins (2 x CheA, 3 x CheW and 3 x CheY, but no CheZ). In this study, we found a fourth cheY and expressed these R. sphaeroides proteins in E. coli. CheA2 (but not CheA1) restored swarming to an E. coli cheA mutant (RP9535). CheW3 (but not CheW2) restored swarming to a cheW mutant of E. coli (RP4606). R. sphaeroides CheYs did not affect E. coli lacking CheY, but restored swarming to a cheZ strain (RP1616), indicating that they can act as signal terminators in E. coli. An E. coli CheY, which is phosphorylated but cannot bind the motor (CheY109KR), was expressed in RP1616 but had no effect. Overexpression of CheA2, CheW2, CheW3, CheY1, CheY3 and CheY4 inhibited chemotaxis of wild-type E. coli (RP437) by increasing its smooth-swimming bias. While some R. sphaeroides proteins restore tumbling to smooth-swimming E. coli mutants, their activity is not controlled by the chemosensory receptors. R. sphaeroides possesses a phosphorelay cascade compatible with that of E. coli, but has additional incompatible homologues.  相似文献   

3.
Rhodobacter sphaeroides has multiple homologues of most of the Escherichia coli chemotaxis genes, organized in two major operons and other, unlinked, loci. These include cheA1 and cheW1 (che Op1) and cheA2, cheW2 and cheW3 (che Op2). We have deleted each of these cheA and cheW homologues in-frame and examined the chemosensory behaviour of these strains on swarm plates and in tethered cell assays. In addition, we have examined the effect of these deletions on the polar localization of the chemoreceptor McpG. In E. coli, deletion of either cheA or cheW results in a non-chemotactic phenotype, and these strains also show no receptor clustering. Here, we demonstrate that CheW2 and CheA2 are required for the normal localization of McpG and for normal chemotactic responses under both aerobic and photoheterotrophic conditions. Under aerobic conditions, deletion of cheW3 has no significant effect on McpG localization and only has an effect on chemotaxis to shallow gradients in swarm plates. Under photoheterotrophic conditions, however, CheW3 is required for McpG localization and also for chemotaxis both on swarm plates and in the tethered cell assay. These phenotypes are not a direct result of delocalization of McpG, as this chemoreceptor does not mediate chemotaxis to any of the compounds tested and can therefore be considered a marker for general methyl-accepting chemotaxis protein (MCP) clustering. Thus, there is a correlation between the normal localization of McpG (and presumably other chemoreceptors) and chemotaxis. We propose a model in which the multiple different MCPs in R. sphaeroides are contained within a polar chemoreceptor cluster. Deletion of cheW2 and cheA2 under both aerobic and photoheterotrophic conditions, and cheW3 under photoheterotrophic conditions, disrupts the cluster and hence reduces chemotaxis to any compound sensed by these MCPs.  相似文献   

4.
5.
6.
Two chemotaxis-defective mutants of Pseudomonas aeruginosa, designated PC3 and PC4, were selected by the swarm plate method after N-methyl-N'-nitro-N-nitrosoguanidine mutagenesis. These mutants were not complemented by the P. aeruginosa cheY and cheZ genes, which had been previously cloned (Masduki et al., J. Bacteriol., 177, 948-952, 1995). DNA sequences downstream of the cheY and cheZ genes were able to complement PC3 but not PC4. Sequence analysis of a 9.7-kb region directly downstream of the cheZ gene found three chemotaxis genes, cheA, cheB, and cheW, and seven unknown open reading frames (ORFs). The predicted translation products of the cheA, cheB, and cheW genes showed 33, 36, and 31% amino acid identity with Escherichia coli CheA, CheB, and CheW, respectively. Two of the unknown ORFs, ORF1 and ORF2, encoded putative polypeptides that resembled Bacillus subtilis MotA (40% amino acid identity) and MotB (34% amino acid identity) proteins, respectively. Although P. aeruginosa was found to have proteins similar to the enteric chemotaxis proteins CheA, CheB, CheW, CheY, and CheZ, the gene encoding a CheR homologue did not reside in the chemotaxis gene cluster. The P. aeruginosa cheR gene could be cloned by phenotypic complementation of the PC4 mutant. This gene was located at least 1,800 kb away from the chemotaxis gene cluster and encoded a putative polypeptide that had 32% amino acid identity with E. coli CheR.  相似文献   

7.
Rhodobacter sphaeroides chemotaxis is significantly more complex than that of enteric bacteria. Rhodobacter sphaeroides has multiple copies of chemotaxis genes (two cheA, one cheB, two cheR, three cheW, five cheY but no cheZ), controlling a single 'stop-start' flagellum. The growth environment controls the level of expression of different groups of genes. Tethered cell analysis of mutants suggests that CheY(4) and CheY(5) are the motor-binding response regulators. The histidine protein kinase CheA(2) mediates an attractant ('normal') response via CheY(4), while CheA(1) and CheY(5) appear to mediate a repellent ('inverted') response. CheY(3) facilitates signal termination, possibly acting as a phosphate sink, although CheY(1) and CheY(2) can substitute. The normal and inverted responses may be initiated by separate sets of chemoreceptors with their relative strength dependent on growth conditions. Rhodobacter sphaeroides may use antagonistic responses through two chemosensory pathways, expressed at different levels in different environments, to maintain their position in a currently optimum environment. Complex chemotaxis systems are increasingly being identified and the strategy adopted by R.sphaeroides may be common in the bacterial kingdom.  相似文献   

8.
Pseudomonas aeruginosa, a gamma-proteobacterium, is motile by means of a single polar flagellum and is chemotactic to a variety of organic compounds and phosphate. P. aeruginosa has multiple homologues of Escherichia coli chemotaxis genes that are organized into five gene clusters. Previously, it was demonstrated that genes in cluster I and cluster V are essential for chemotaxis. A third cluster (cluster II) contains a complete set of che genes, as well as two genes, mcpA and mcpB, encoding methyl-accepting chemotaxis proteins. Mutations were constructed in several of the cluster II che genes and in the mcp genes to examine their possible contributions to P. aeruginosa chemotaxis. A cheB2 mutant was partially impaired in chemotaxis in soft-agar swarm plate assays. Providing cheB2 in trans complemented this defect. Further, overexpression of CheB2 restored chemotaxis to a completely nonchemotactic, cluster I, cheB-deficient strain to near wild-type levels. An mcpA mutant was defective in chemotaxis in media that were low in magnesium. The defect could be relieved by the addition of magnesium to the swarm plate medium. An mcpB mutant was defective in chemotaxis when assayed in dilute rich soft-agar swarm medium or in minimal-medium swarm plates containing any 1 of 60 chemoattractants. The mutant phenotype could be complemented by the addition of mcpB in trans. Overexpression of either McpA or McpB in P. aeruginosa or Escherichia coli resulted in impairment of chemotaxis, and these cells had smooth-swimming phenotypes when observed under the microscope. Expression of P. aeruginosa cheA2, cheB2, or cheW2 in E. coli K-12 completely disrupted wild-type chemotaxis, while expression of cheY2 had no effect. These results indicate that che cluster II genes are expressed in P. aeruginosa and are required for an optimal chemotactic response.  相似文献   

9.
Bacterial chemotaxis is a colonization factor for the ulcer-causing pathogen Helicobacter pylori. H. pylori contains genes encoding the chemotaxis signalling proteins CheW, CheA and CheY; CheW couples chemoreceptors to the CheA kinase and is essential for chemotaxis. While characterizing a cheW mutant, we isolated a spontaneous, chemotactic variant (Che+). We determined that this phenotype was caused by a genetic change unlinked to the original cheW mutation. To locate the underlying Che+ mutation, we compared total protein profiles of the non-chemotactic mutant (cheW) with those from the cheW Che+ variant by two-dimensional differential in-gel electrophoresis. One protein was found only in the cheW Che+ variant. This protein was identified by MS/MS as HP0170, a hypothetical protein with no known function. DNA sequencing verified that hp0170 was mutated in the cheW Che+ suppressor, and deletion of this open reading frame in the cheW background nearly recapitulated the Che+ suppressor phenotype. Using hidden Markov models, we found that HP0170 is a remote homologue of E. coli CheZ. CheZ interacts with phosphorylated CheY and stimulates its autodephosphorylation. CheZ was not predicted to be present in epsilon-proteobacteria. We found that chemotaxis in the cheW Che+ suppressor depended on both cheY and cheA. We hypothesize that a small amount of phosphorylated CheY is generated via CheA in the cheW mutant, and this amount is sufficient to affect flagellar rotation when HP0170 is removed. Our results suggest that HP0170 is a remote homologue of CheZ, and that CheZ homologues are found in a broader range of bacteria than previously supposed.  相似文献   

10.
Z Y Jiang  H Gest    C E Bauer 《Journal of bacteriology》1997,179(18):5720-5727
The chemotaxis gene cluster from the photosynthetic bacterium Rhodospirillum centenum contains five open reading frames (ORFs) that have significant sequence homology to chemotaxis genes from other bacteria. To elucidate the functions of each ORF, we have made various mutations in the gene cluster and analyzed their phenotypic defects. Deletion of the entire che operon (delta che), as well as nonpolar disruptions of cheAY, cheW, and cheR, resulted in a smooth-swimming phenotype, whereas disruption of cheB resulted in a locked tumbly phenotype. Each of these mutants was defective in chemotactic response. Interestingly, disruption of cheY resulted in a slight increase in the frequency of tumbling/reversal with no obvious defects in chemotactic response. In contrast to observations with Escherichia coli and several other bacteria, we found that all of the che mutant cells were capable of differentiating into hyperflagellated swarmer cells when plated on a solid agar surface. When viewed microscopically, the smooth-swimming che mutants exhibited active surface motility but were unable to respond to a step-down in light intensity. Both positive and negative phototactic responses were abolished in all che mutants, including the cheY mutant. These results indicate that eubacterial photosensory perception is mediated by light-generated signals that are transmitted through the chemotaxis signal transduction cascade.  相似文献   

11.
Hybrid Escherichia coli ColE1 plasmids carrying the genes for motility (mot) and chemotaxis (che) were transferred to a minicell-producing strain. The mot and che genes on the hybrid plasmid directed protein synthesis in minicells. Polypeptides synthesized in minicells were identical to the products of the motA, motB, cheA, cheW, cheM, cheX, cheB, cheY, and cheZ genes previously identified by using hybrid lambda and ultraviolet-irradiated host cells (Silverman and Simon, J. Bacteriol. 130:1317-1325, 1977), thus confirming these gene product assignments. The products of some che genes (cheA and cheM) appeared as more than one band on polyacrylamide gel electrophoresis, but analysis of partial peptide digests of these polypeptides suggested that the multiple forms were coded for by a single gene. Measurement of the physical length of the hybrid plasmids allowed an estimate of the amount of coding capacity of the cloned deoxyribonucleic acid, which was devoted to the synthesis of the mot and che gene products. These estimates were also consistent with the hypothesis that the multiple polypeptides corresponding to cheA and cheM were the products of single genes.  相似文献   

12.
Rhodobacter sphaeroides has multiple homologues of most of the Escherichia coli chemotaxis genes, organized in three major operons and other, unlinked, loci. These include cheA(1) and cheR(1) (che Op(1)) and cheA(2), cheR(2), and cheB(1) (che Op(2)). In-frame deletions of these cheR and cheB homologues were constructed and the chemosensory behaviour of the resultant mutants examined on swarm plates and in tethered cell assays. Under the conditions tested, CheR(2) and CheB(1) were essential for normal chemotaxis, whereas CheR(1) was not. cheR(2) and cheB(1), but not cheR(1), were also able to complement the equivalent E. coli mutants. However, none of the proteins were required for the correct polar localization of the chemoreceptor McpG in R. sphaeroides. In E. coli, CheR binds to the NWETF motif on the high-abundance receptors, allowing methylation of both high- and low-abundance receptors. This motif is not contained on any R. sphaeroides chemoreceptors thus far identified, although 2 of the 13 putative chemoreceptors, McpA and TlpT, do have similar sequences. This suggests that CheR(2) either interacts with the NWETF motif of E. coli methyl-accepting chemotaxis proteins (MCPs), even though its native motif may be slightly different, or with another conserved region of the MCPs. Methanol release measurements show that R. sphaeroides has an adaptation system that is different from that of Bacillus subtilis and E. coli, with methanol release measurable on the addition of attractant but not on its removal. Intriguingly, CheA(2), but not CheA(1), is able to phosphorylate CheB(1), suggesting that signaling through CheA(1) cannot initiate feedback receptor adaptation via CheB(1)-P.  相似文献   

13.
The nucleotide sequence of a region downstream of the Listeria monocytogenes flagellin gene, flaA, revealed two putative chemotaxis genes, cheY and cheA. These genes have been shown to be transcribed as a bicistronic unit. In this study Tn916 delta E mutagenesis was used to generate two mutants, PF10 and PF16, which contain transposon inserts in the promoter region of this operon. These mutants were motile in liquid, but had reduced flagellin expression and were unable to burrow or swarm on soft agar plates. Complementation of the single transposon-copy mutant PF16 with cloned cheY and cheA in trans partially restored microaerotaxis and swarming on soft agar. The complemented strain did not exhibit any increase in flagellin production. Both PF10 and PF16 appear deficient in their ability to attach to the mouse fibroblast cell line 3T3.  相似文献   

14.
A chemotaxis signal transduction pathway (hereafter called Che1) has been previously identified in the alphaproteobacterium Azospirillum brasilense. Previous experiments have demonstrated that although mutants lacking CheB and/or CheR homologs from this pathway are defective in chemotaxis, a mutant in which the entire chemotaxis pathway has been mutated displayed a chemotaxis phenotype mostly similar to that of the parent strain, suggesting that the primary function of this Che1 pathway is not the control of motility behavior. Here, we report that mutants carrying defined mutations in the cheA1 (strain AB101) and the cheY1 (strain AB102) genes and a newly constructed mutant lacking the entire operon [Δ(cheA1-cheR1)::Cm] (strain AB103) were defective, but not null, for chemotaxis and aerotaxis and had a minor defect in swimming pattern. We found that mutations in genes of the Che1 pathway affected the cell length of actively growing cells but not their growth rate. Cells of a mutant lacking functional cheB1 and cheR1 genes (strain BS104) were significantly longer than wild-type cells, whereas cells of mutants impaired in the cheA1 or cheY1 genes, as well as a mutant lacking a functional Che1 pathway, were significantly shorter than wild-type cells. Both the modest chemotaxis defects and the observed differences in cell length could be complemented by expressing the wild-type genes from a plasmid. In addition, under conditions of high aeration, cells of mutants lacking functional cheA1 or cheY1 genes or the Che1 operon formed clumps due to cell-to-cell aggregation, whereas the mutant lacking functional CheB1 and CheR1 (BS104) clumped poorly, if at all. Further analysis suggested that the nature of the exopolysaccharide produced, rather than the amount, may be involved in this behavior. Interestingly, mutants that displayed clumping behavior (lacking cheA1 or cheY1 genes or the Che1 operon) also flocculated earlier and quantitatively more than the wild-type cells, whereas the mutant lacking both CheB1 and CheR1 was delayed in flocculation. We propose that the Che1 chemotaxis-like pathway modulates the cell length as well as clumping behavior, suggesting a link between these two processes. Our data are consistent with a model in which the function of the Che1 pathway in regulating these cellular functions directly affects flocculation, a cellular differentiation process initiated under conditions of nutritional imbalance.  相似文献   

15.
Rhodospirillum centenum is a photosynthetic bacterium capable of undergoing swim cell to swarm cell differentiation that allows this species to be motile on both liquid and solid media. Previous experiments have demonstrated that the che1 operon is required for the control of chemotactic and phototactic behaviour of both swim and swarm cells. In this report, we analyse the function of a second che-like gene cluster in R. centenum, the che2 gene cluster. In-frame deletion mutants of cheW2, cheB2, cheR2, cheY2, and of the entire che2 operon, exhibit defects in swim and swarm cell motility. Analysis of these strains demonstrates that they are non-motile, and that the non-motile phenotype is resulting from reduced polar and lateral flagella synthesis. Additionally, mutations in mcp2, ORF204, cheA2 and ORF74 remain chemotacticly and phototacticly competent at both high and low growth temperatures. Mutations in these che2 genes result in elevated levels of flagellin proteins giving rise to a hyperflagellate phenotype. We propose a model in which R. centenum utilizes a che-like signal transduction pathway (che2) for regulating flagellum synthesis in order to optimize swim cell-swarm cell differentiation in response to changing environmental conditions.  相似文献   

16.
Flagellar proteins controlling motility and chemotaxis in Escherichia coli were selectively labeled in vivo with [35S]methionine. This distribution of these proteins in subcellular fractions was examined by sodium dodecyl sulfatepolyacrylamide gel electrophoresis and autoradiography. The motA, motB, cheM, and cheD gene products were found to be confined exclusively to the inner cytoplasmic membrane fraction, whereas the cheY, cheW, and cheA (66,000 daltons) polypeptides appeared only in the soluble cytoplasmic fraction. The cheB, cheX, cheZ, and cheA (76,000 daltons) proteins, however, were distributed in both the cytoplasm and the inner membrane fractions. The hag gene product (flagellin) was the only flagellar protein examined that copurified with the outer lipopolysaccharide membrane. Differences in the intracellular locations of the che and mot gene prodcuts presumably reflect the functional attributes of these components.  相似文献   

17.
Two chemotaxis-defective mutants of Pseudomonas aeruginosa, designated PC1 and PC2, were selected by the swarm plate method after N-methyl-N'-nitro-N-nitrosoguanidine mutagenesis. These mutants were fully motile but incapable of swarming, suggesting that they had a defect in the intracellular signalling pathway. Computer-assisted capillary assays confirmed that they failed to show behavioral responses to chemical stimuli, including peptone, methyl thiocyanate, and phosphate. Two chemotaxis genes were cloned by phenotypic complementation of PC1 and PC2. From nucleotide sequence analysis, one gene was found to encode a putative polypeptide that was homologous to the enteric CheZ protein, while the other gene was cheY, which had been previously reported (M. N. Starnbach and S. Lory, Mol. Microbiol. 6:459-469, 1992). Deletion and complementation analysis showed that PC1 was a cheY mutant, whereas PC2 had a double mutation in the cheY and cheZ genes. A chromosomal cheZ mutant, constructed by inserting a kanamycin resistance gene cassette into the wild-type gene, changed its swimming direction much more frequently than did wild-type strain PAO1. In contrast, cheY mutants were found to rarely reverse their swimming directions.  相似文献   

18.
19.
An expressed sequence tag homologous to cheA was previously isolated by random sequencing of Thermotoga maritima cDNA clones (C. W. Kim, P. Markiewicz, J. J. Lee, C. F. Schierle, and J. H. Miller, J. Mol. Biol. 231: 960-981, 1993). Oligonucleotides complementary to this sequence tag were synthesized and used to identify a clone from a T. maritima lambda library by using PCR. Two partially overlapping restriction fragments were subcloned from the lambda clone and sequenced. The resulting 5,251-bp sequence contained five open reading frames, including cheA, cheW, and cheY. In addition to the chemotaxis genes, the fragment also encodes a putative protein isoaspartyl methyltransferase and an open reading frame of unknown function. Both the cheW and cheY genes were individually cloned into inducible Escherichia coli expression vectors. Upon induction, both proteins were synthesized at high levels. T. maritima CheW and CheY were both soluble and were easily purified from the bulk of the endogenous E. coli protein by heat treatment at 80 degrees C for 10 min. CheY prepared in this way was shown to be active by the demonstration of Mg(2+)-dependent autophosphorylation with [32P]acetyl phosphate. In E. coli, CheW mediates the physical coupling of the receptors to the kinase CheA. The availability of a thermostable homolog of CheW opens the possibility of structural characterization of this small coupling protein, which is among the least well characterized proteins in the bacterial chemotaxis signal transduction pathway.  相似文献   

20.
Myxococcus xanthus co-ordinates cell movement during its complex life cycle using multiple chemotaxis-like signal transduction pathways. These pathways regulate both type IV pilus-mediated social (S) motility and adventurous (A) motility. During a search for new chemoreceptors, we identified the che4 operon, which encodes homologues to a MCP (methyl-accepting chemotaxis protein), two CheWs, a hybrid CheA-CheY, a response regulator and a CheR. Deletion of the che4 operon did not cause swarming or developmental defects in either the wild-type (A(+)S(+)) strain or in a strain sustaining only A motility (A(+)S(-)). However, in a strain displaying only S motility (A(-)S(+)), deletion of the che4 operon or the gene encoding the response regulator, cheY4, caused enhanced vegetative swarming and prevented aggregation and sporulation. In contrast, deletion of mcp4 caused reduced vegetative swarming and enhanced development compared with the parent strain. Single-cell analysis of the motility of the A(-)S(+) parent strain revealed a previously unknown inverse correlation between velocity and reversal frequency. Thus, cells that moved at higher velocities showed a reduced reversal frequency. This co-ordination of reversal frequency and velocity was lost in the mcp4 and cheY4 mutants. The structural components of the S motility apparatus were unaffected in the che4 mutants, suggesting that the Che4 system affects reversal frequency of cells by modulating the function of the type IV pilus.  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