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.
The McpS chemoreceptor of Pseudomonas putida KT2440 recognizes six different tricarboxylic acid (TCA) cycle intermediates. However, the magnitude of the chemotactic response towards these compounds differs largely, which has led to distinguish between strong attractants (malate, succinate, fumarate, oxaloacetate) and weak attractants (citrate, isocitrate). Citrate is abundantly present in plant tissues and root exudates and can serve as the only carbon source for growth. Citrate is known to form complexes with divalent cations which are also abundantly present in natural habitats of this bacterium. We have used isothermal titration calorimetry to study the formation of citrate-metal ion complexes. In all cases binding was entropy driven but significant differences in affinity were observed ranging from K(D)=157 μM (for Mg(2+)) to 3 μM (for Ni(2+)). Complex formation occurred over a range of pH and ionic strength. The ligand binding domain of McpS (McpS-LBD) was found to bind free citrate, but not complexes with physiologically relevant Mg(2+) and Ca(2+). In contrast, complexes with divalent cations which are present as trace elements (Co(2+), Cd(2+) and Ni(2+)) were recognized by McpS-LBD. This discrimination differs from other citrate sensing proteins. These results are discussed in the context of the three dimensional structure of free citrate and its complex with Mg(2+). Chemotaxis assays using P. putida revealed that taxis towards the strong attractant malate is strongly reduced in the presence of free citrate. However, this reduction is much less important in the presence of citrate-Mg(2+) complexes. The physiological relevance of these findings is discussed.  相似文献   

2.
Bacterial chemotaxis may have a significant impact on the structure and function of bacterial communities. Quantification of chemotactic motion is necessary to identify chemoeffectors and to determine the bacterial transport parameters used in predictive models of chemotaxis. When the chemotactic bacteria consume the chemoeffector, the chemoeffector gradient to which the bacteria respond may be significantly perturbed by the consumption. Therefore, consumption of the chemoeffector can confound chemotaxis measurements if it is not accounted for. Current methods of quantifying chemotaxis use bacterial concentrations that are too high to preclude chemoeffector consumption or involve ill-defined conditions that make quantifying chemotaxis difficult. We developed a method of quantifying bacterial chemotaxis at low cell concentrations (~105 CFU/ml), so metabolism of the chemoeffector is minimized. The method facilitates quantification of bacterial-transport parameters by providing well-defined boundary conditions and can be used with volatile and semivolatile chemoeffectors.  相似文献   

3.
Bacterial chemotaxis may have a significant impact on the structure and function of bacterial communities. Quantification of chemotactic motion is necessary to identify chemoeffectors and to determine the bacterial transport parameters used in predictive models of chemotaxis. When the chemotactic bacteria consume the chemoeffector, the chemoeffector gradient to which the bacteria respond may be significantly perturbed by the consumption. Therefore, consumption of the chemoeffector can confound chemotaxis measurements if it is not accounted for. Current methods of quantifying chemotaxis use bacterial concentrations that are too high to preclude chemoeffector consumption or involve ill-defined conditions that make quantifying chemotaxis difficult. We developed a method of quantifying bacterial chemotaxis at low cell concentrations ( approximately 10(5) CFU/ml), so metabolism of the chemoeffector is minimized. The method facilitates quantification of bacterial-transport parameters by providing well-defined boundary conditions and can be used with volatile and semivolatile chemoeffectors.  相似文献   

4.
Salmonella, but not Escherichia coli, was attracted to citrate, a distinction that is understandable in view of the inability of E. coli to transport tricarboxylic acids. The Salmonella response to citrate and to two previously described attractants, aspartate and malate, was mutually noncompetitive. Citrate taxis different from citrate uptake in that it did not require Na+, was constitutive, and was not repressible by glucose.  相似文献   

5.
6.
Citrate uptake in Bacillus subtilis is stimulated by a wide range of divalent metal ions. The metal ions were separated into two groups based on the expression pattern of the uptake system. The two groups correlated with the metal ion specificity of two homologous B. subtilis secondary citrate transporters, CitM and CitH, upon expression in Escherichia coli. CitM transported citrate in complex with Mg(2+), Ni(2+), Mn(2+), Co(2+), and Zn(2+) but not in complex with Ca(2+), Ba(2+), and Sr(2+). CitH transported citrate in complex with Ca(2+), Ba(2+), and Sr(2+) but not in complex with Mg(2+), Ni(2+), Mn(2+), Co(2+), and Zn(2+). Both transporters did not transport free citrate. Nevertheless, free citrate uptake could be demonstrated in B. subtilis, indicating the expression of at least a third citrate transporter, whose identity is not known. For both the CitM and CitH transporters it was demonstrated that the metal ion promoted citrate uptake and, vice versa, that citrate promoted uptake of the metal ion, indicating that the complex is the transported species. The results indicate that CitM and CitH are secondary transporters that transport complexes of divalent metal ions and citrate but with a complementary metal ion specificity. The potential physiological function of the two transporters is discussed.  相似文献   

7.
The chemoreceptor Tcp of Salmonella enterica serovar Typhimurium can sense citrate and a metal-citrate complex as distinct attractants. In this study, we tried to investigate the molecular mechanism of this discrimination. That citrate binds directly to Tcp was verified by the site-specific thiol modification assays using membrane fractions prepared from Escherichia coli cells expressing the mutant Tcp receptors in which single Cys residues were introduced at positions in the putative ligand-binding pocket. To determine the region responsible for the ligand discrimination, we screened for mutations defective in taxis to magnesium in the presence of citrate. All of the isolated mutants from random mutagenesis with hydroxylamine were defective in both citrate and metal-citrate sensing, and the mutated residues are located in or near the alpha1-alpha2 and alpha3-alpha4 loops within the periplasmic domain. Further analyses with site-directed replacements around these regions demonstrated that the residue Asn(67), which is presumed to lie at the subunit interface of the Tcp homodimer, plays a critical role in the recognition of the metal-citrate complex but not that of citrate. Various amino acids at this position differentially affect the citrate and metal-citrate sensing abilities. Thus, for the first time, the abilities to sense the two attractants were genetically dissected. Based on the results obtained in this study, we propose models in which the discrimination of the metal-citrate complex from citrate involves cooperative interaction at Asn(67) and allosteric switching.  相似文献   

8.
The bacterial pathogen Salmonella penetrates the intestinal epithelium by inducing its own phagocytosis into epithelial cells. The dramatic reorganization of the actin cytoskeleton required for internalization is driven by bacterial manipulation of host signaling pathways, including activation of the Rho family GTPase Rac1 and subsequent activation of the Arp2/3 complex. However, the mechanisms linking these two events remain poorly understood. Rac1 is thought to promote activation of the Arp2/3 complex through its interaction with suppressor of cAMP receptor/WASP family verprolin-homologous (SCAR/WAVE) family proteins, but this interaction is apparently indirect. Two different Rac1 effectors have been shown to bind WAVE2: IRSp53, the SH3 domain of which binds the WAVE2 proline-rich domain, and PIR121/Sra-1, which forms a pentameric complex containing WAVE, Abi1, Nap1, and HSPC300. However, the extent to which each of these complexes contributes to Arp2/3 complex activation in the context of Salmonella infection is unclear. Here, we show that WAVE2 is necessary for efficient invasion of epithelial cells by Salmonella typhimurium. We found that although Salmonella infection strongly promotes the formation of an IRSp53/WAVE2 complex, IRSp53 is not necessary for bacterial internalization. In contrast, disruption of the PIR121/Nap1/Abi1/WAVE2/HSPC300 complex potently inhibits bacterial uptake. These results indicate that WAVE2 is an important component in signaling pathways leading to Salmonella invasion. Although infection leads to the formation of an IRSp53/WAVE2 complex, it is the association of WAVE2 with the Abi1/Nap1/PIR121/HSPC300 complex that regulates bacterial internalization.  相似文献   

9.
The lethal action of streptonigrin on strains of Escherichiacoli is greatly enhanced by citrate (10?2 M). Desferrioxamine (2×10?4 M), when added with streptonigrin and citrate, eliminates the citrate enhancement. These observations point to a role for iron in the bactericidal mechanism of streptonigrin. Extracellular citrate is known to promote the acquisition of iron by E.coli by delivering it as a ferric citrate complex to a specific transport apparatus on the cell envelope. Therefore, it may promote action of streptonigrin by increasing the intracellular concentration of available iron. Desferrioxamine, which forms a much stronger complex with ferric ion than does citrate, would be expected to suppress the ferric citrate effect, and this was observed.  相似文献   

10.
The growth of 68 strains of Salmonella typhi , 697 other Salmonella strains and 213 strains of other Gram negative intestinal bacteria on 8 plating media was assessed semi-quantitatively. These media were found to be differentially inhibitory to different Salmonella serotypes. The combined use of two plating media, brilliant green MacConkey agar and deoxycholate citrate agar, allowed potentially the recovery of the maximum number of Salmonella strains. If only one medium was used, brilliant green MacConkey agar would appear to be the best plating medium for the isolation of non-typhoid salmonellas in general and S. choleraesuis in particular. Xylose lysine deoxycholate agar, on which a certain proportion of salmonellas failed to yield typical, recognizable colonies, was found not to be a good plating medium for their isolation.  相似文献   

11.
Secondary transporters of the bacterial CitMHS family transport citrate in complex with a metal ion. Different members of the family are specific for the metal ion in the complex and have been shown to transport Mg(2+)-citrate, Ca(2+)-citrate or Fe(3+)-citrate. The Fe(3+)-citrate transporter of Streptococcus mutans clusters on the phylogenetic tree on a separate branch with a group of transporters found in the phylum Firmicutes which are believed to be involved in anaerobic citrate degradation. We have cloned and characterized the transporter from Enterococcus faecalis EfCitH in this cluster. The gene was functionally expressed in Escherichia coli and studied using right-side-out membrane vesicles. The transporter catalyzes proton-motive-force-driven uptake of the Ca(2+)-citrate complex with an affinity constant of 3.5 microm. Homologous exchange is catalyzed with a higher efficiency than efflux down a concentration gradient. Analysis of the metal ion specificity of EfCitH activity in right-side-out membrane vesicles revealed a specificity that was highly similar to that of the Bacillus subtilis Ca(2+)-citrate transporter in the same family. In spite of the high sequence identity with the S. mutans Fe(3+)-citrate transporter, no transport activity with Fe(3+) (or Fe(2+)) could be detected. The transporter of E. faecalis catalyzes translocation of citrate in complex with Ca(2+), Sr(2+), Mn(2+), Cd(2+) and Pb(2+) and not with Mg(2+), Zn(2+), Ni(2+) and Co(2+). The specificity appears to correlate with the size of the metal ion in the complex.  相似文献   

12.
Thermosensitive H1 plasmids determining citrate utilization.   总被引:6,自引:0,他引:6  
Twelve thermosensitive H1 plasmids from strains of Salmonella typhi that had caused outbreaks of chloramphenicol-resistant typhoid fever in Vietnam, Thailand and India mediated citrate utilization (Cit+) in a prototrophic Escherichia coli K12 strain but not in the S. typhi strains from which they were derived. Four H1 plasmids from a similar outbreak in Mexico differed from the Far Eastern plasmids in not mediating citrate utlization but in mediating mercury resistance. H1 plasmids resembling the Far Eastern and the Mexican plasmids in regard to citrate utilization and mercury resistance were found in sewage in Britain. Citrate utilization was transferred to eight pathogenic strains of E. coli and to one strain each of Shigella flexneri and Shigella sonnei. Cultures of Cit+ bacteria grew more rapidly in citrate media at 28 degrees C than at 37 degrees C. Plasmid mutants that were more efficient at utilizing citrate were present in all such cultures--they grew equally well or better at 37 degrees C than at 28 degrees C. None of 222 strains of E. coli or Shigella that contained a variety of different plasmids were able to utilize citrate. This property was not transferred to the prototrophic E. coli K12 strain from Citrobacter (3 strains), Salmonella (39 strains), Proteus (44 strains), Klebsiella pneumoniae (33 strains) or Pseudomonas aeruginosa (44 strains).  相似文献   

13.
Citrate transport in Salmonella typhimurium involves inducible periplasmic components. Two forms of a tricarboxylate-binding protein, C1 and C2, were isolated, in high yield, from the periplasm of a cyclic AMP phosphodiesterase mutant. These immunologically cross-reactive Mr = 29,000 proteins were crystallized using ammonium sulfate. CD measurements indicated considerable secondary structure: 24% a helix, and 12% beta structure. The amino acid compositions of C1 and C2 were identical. The NH2-terminal sequence of C1 was determined; C2 was found to have a blocked NH2 terminus (pyroglutamate). C1 and C2 are products of the same gene (Somers, J. M., and Kay, W. W. (1983) Mol. Gen. Genet. 190, 20-26). C1 and C2 bound a variety of citrate analogues and organic acids, with a predominant specificity for tricarboxylates (citrate KD 1.4 X 10(-7) M), and both required a deprotonated central carboxyl group for binding. Citrate was not bound to C protein as either a salt or metal ion complex.  相似文献   

14.
D C Fry  K G Strothkamp 《Biochemistry》1983,22(21):4949-4953
Irradiation of Agaricus bisporus tyrosinase in the presence of citrate at pH 5.6 with 300-420-nm light results in a loss of both catecholase activity and cresolase activity. The light-sensitive species appears to be an enzyme-citrate complex, most likely involving coordination of citrate to the active site copper. One copper ion from each binuclear active site can be removed from the inactivated enzyme, resulting in the formation of a met apo derivative. The electron spin resonance spectrum of met apo tyrosinase resembles that of met apo hemocyanin and half-met Neurospora tyrosinase. It is consistent with a distorted square-planar geometry around the copper and with either nitrogen or nitrogen and oxygen ligands. Amino acid analysis indicates that four histidines on the heavy subunit are destroyed during the inactivation process. Some or all of these histidines may serve as ligands to the copper ion which becomes labile after inactivation. Photoinactivation results in decarboxylation of citrate and does not require the presence of oxygen. The reaction may involve generation of a free radical from the citrate which then attacks nearby histidine residues.  相似文献   

15.
Specific inhibition of flagellar rotation reversal was observed after exposure of chemotactic Salmonella typhimurium to citrate autoclaved at neutral pH. The presence of a rotation reversal inactivator was established in autoclaved citrate-containing media and nutrient broth. Since modulation of flagellar rotation by attractants and repellents is the basis of chemotactic behavior, a specific inhibitor of rotation reversal, which is essential for tumble generation, provides a useful probe into the molecular mechanism of bacterial chemotaxis. The inactivator inhibits clockwise rotation without affecting counterclockwise rotation, speed of rotation, or the capacity of the cells to grow and divide. Inactivation of clockwise rotation is gradual and irreversible, differing from the transient inhibition of clockwise rotation by attractants, which is characterized by an immediate suppression followed by a return to normal rotation patterns. The rotation reversal inactivator is stable to acidification, rotary evaporation, lyophilization, and rehydration.  相似文献   

16.
Involvement of transport in Rhodobacter sphaeroides chemotaxis.   总被引:11,自引:9,他引:2       下载免费PDF全文
The chemotactic response to a range of chemicals was investigated in the photosynthetic bacterium Rhodobacter sphaeroides, an organism known to lack conventional methyl-accepting sensory transduction proteins. Strong attractants included monocarboxylic acids and monovalent cations. Results suggest that the chemotactic response required the uptake of the chemoeffector, but not its metabolism. If a chemoeffector could block the uptake of another attractant, it also inhibited chemotaxis to that attractant. Sodium benzoate was not an attractant but was a competitive inhibitor of the propionate uptake system. Binding in an active uptake system was therefore insufficient to cause a chemotactic response. At different concentrations, benzoate either blocked propionate chemotaxis or reduced the sensitivity of propionate chemotaxis, an effect consistent with its role as a competitive inhibitor of uptake. Bacteria only showed chemotaxis to ammonium when grown under ammonia-limited conditions, which derepressed the ammonium transport system. Both chemotaxis and uptake were sensitive to the proton ionophore carbonyl cyanide m-chlorophenylhydrazone, suggesting an involvement of the proton motive force in chemotaxis, at least at the level of transport. There was no evidence for internal pH as a sensory signal. These results suggest a requirement for the uptake of attractants in chemotactic sensing in R. sphaeroides.  相似文献   

17.
The synthesis of aconitase in Bacillus subtilis wild-type and different citric acid cycle mutants has been studied and the influence of various growth conditions examined. Aconitase is induced by citrate and precursors of citrate and repressed by glutamate. Induction and repression counteract each other, and at equimolar concentrations of citrate and glutamate, aconitase synthesis is unaffected. Induction by citrate can partly overcome catabolite repression of aconitase. Isocitrate dehydrogenase show endogenous induction of aconitase due to citrate accumulation. Leaky mutants defective in citrate synthase and aconitase cannot be induced by citrate, which indicates that they carry a regulatory mutation. The complex regulation of aconitase is discussed with reference to the participation of this enzyme in glutamate biosynthesis and energy metabolism.  相似文献   

18.
Citrate binding of Al3+ and Fe3+   总被引:4,自引:0,他引:4  
Citrate occurs at about 0.1 mM in blood plasma and is the most likely small molecule plasma binder of both Al3+ and Fe3+. This paper assesses published stability constants for citrate binding to each metal ion. From pH 2 to 5 Al3+ forms a neutral complex with citrate that may be absorbed into the body in the upper regions of the gastrointestinal tract. It is especially dangerous to ingest aluminum-containing antacids with citrus fruit or juices. Ignoring the likely occurrence of a competing 2:1 citrate-Fe3+ complex necessitates adjustments in reported stability constants for Fe3+ binding to transferrin. In the blood, plasma transferrin steals both Fe3+ and Al3+ from citrate.  相似文献   

19.
The role of fatty acid and citrate on the interaction of the AMP deaminase (EC 3.5.4.6) reaction with glycolysis was investigated using permeabilized yeast cells. (a) Linolenate and citrate inhibited glycolytic flux and the recovery of the adenylate energy charge; however, linolenate remarkably retarded the depletion of the total adenylate pool, which was not at all affected by the addition of citrate. (b) Linolenate inhibited AMP deaminase activity in situ, resulting in the subsequent decrease in ammonium production, which reduced the activity of 6-phosphofructokinase (EC 2.7.1.11), whereas linolenate itself had no ability to inhibit the phosphofructokinase activity in the presence of excess ammonium concentration. (c) Citrate inhibited the activity of phosphofructokinase in situ in the presence and absence of ammonium ion, followed by an inhibition of glycolysis; however, AMP deaminase activity was not inhibited by citrate. The inhibition of glycolysis by fatty acids can be accounted for by the lowered activity of phosphofructokinase as a result of the decreased level of ammonium ion through the inhibition of the AMP deaminase reaction by these ligands, whereas the effect of citrate on glycolysis is a direct inhibition of phosphofructokinase without affecting the activity of AMP deaminase. Fatty acid and citrate, a principal metabolic product of fatty acid oxidation, can be responsible for the control of glycolysis in two different manners.  相似文献   

20.
A cryptic citrate transport gene (citA) from Salmonella typhimurium chromosome was cloned and its nucleotide sequence was determined. The cloned plasmid conferred citrate-utilizing ability on wild-type Escherichia coli, which cannot grow on citrate as the sole source of carbon. The resultant E. coli transformant was able to transport citrate. A 1,302-base-pair open reading frame with a preceding ribosomal binding site was found in the cloned DNA fragment. The 434-amino-acid protein that could be translated from this open reading frame is highly hydrophobic (69% nonpolar amino acid residues), consistent with the fact that the transport protein is an intrinsic membrane protein. The molecular weight of this protein was calculated to be 47,188. The gene sequence determined is highly homologous to those of Cit+ plasmid-mediated citrate transport gene, citA, from E. coli, the chromosomal citA gene from Citrobacter amalonaticus and the chromosomal cit+ gene from Klebsiella pneumoniae. The hydropathy profile of the deduced amino acid sequence suggests that this carrier has 12 hydrophobic segments, which may span the membrane lipid bilayer.  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