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.
2.
Yersinia pestis KIM5 was found to be cytotoxic for the IC21 and P388D1 mouse macrophage cell lines, as well as for resident peritoneal macrophages from C57BL/6 mice. Affected cells phagocytosed KIM5 inefficiently, became spherical, detached readily from culture dishes, and retained 51Cr poorly. The cytotoxic effect was dependent on the presence of the 75-kilobase plasmid pCD1. Because this plasmid also encodes the low calcium response (LCR), three Mu d1 insertion mutants previously shown to be LCR- and of reduced virulence in mice were examined for cytotoxicity; all were found to be atoxic. The insertions in these mutants lie within three distinct LCR loci (lcrB, C, and D). Like LCR, cytotoxicity was expressed only at 37 degrees C. Unlike LCR, it was not influenced by Ca2+ concentration, indicating that the V and W antigens are probably not involved. Yersinia pseudotuberculosis was found to have a similar plasmid-dependent cytotoxicity. Thus, biological activity observed as cytotoxicity in vitro may well be a common feature contributing to virulence of the yersiniae.  相似文献   

3.
Brucella suis is a facultative intracellular pathogen of mammals, residing in macrophage vacuoles. In this work, we studied the phagosomal environment of these bacteria in order to better understand the mechanisms allowing survival and multiplication of B. suis. Intraphagosomal pH in murine J774 cells was determined by measuring the fluorescence intensity of opsonized, carboxyfluorescein-rhodamine- and Oregon Green 488-rhodamine-labeled bacteria. Compartments containing live B. suis acidified to a pH of about 4.0 to 4.5 within 60 min. Acidification of B. suis-containing phagosomes in the early phase of infection was abolished by treatment of host cells with 100 nM bafilomycin A(1), a specific inhibitor of vacuolar proton-ATPases. This neutralization at 1 h postinfection resulted in a 2- to 34-fold reduction of opsonized and nonopsonized viable intracellular bacteria at 4 and 6 h postinfection, respectively. Ammonium chloride and monensin, other pH-neutralizing reagents, led to comparable loss of intracellular viability. Addition of ammonium chloride at 7 h after the beginning of infection, however, did not affect intracellular multiplication of B. suis, in contrast to treatment at 1 h postinfection, where bacteria were completely eradicated within 48 h. Thus, we conclude that phagosomes with B. suis acidify rapidly after infection, and that this early acidification is essential for replication of the bacteria within the macrophage.  相似文献   

4.
The transformation of the enteropathogenic bacterium Yersinia pseudotuberculosis into the plague bacillus, Yersinia pestis, has been accompanied by extensive genetic loss. This study focused on chromosomal regions conserved in Y. pseudotuberculosis and lost during its transformation into Y. pestis. An extensive PCR screening of 78 strains of the two species identified five regions (R1 to R5) and four open reading frames (ORFs; orf1 to orf4) that were conserved in Y. pseudotuberculosis and absent from Y. pestis. Their conservation in Y. pseudotuberculosis suggests a positive selective pressure and a role during the life cycle of this species. Attempts to delete two ORFs (orf3 and orf4) from the chromosome of strain IP32953 were unsuccessful, indicating that they are essential for its viability. The seven remaining loci were individually deleted from the IP32953 chromosome, and the ability of each mutant to grow in vitro and to kill mice upon intragastric infection was evaluated. Four loci (orf1, R2, R4, and R5) were not required for optimal growth or virulence of Y. pseudotuberculosis. In contrast, orf2, encoding a putative pseudouridylate synthase involved in RNA stability, was necessary for the optimal growth of IP32953 at 37°C in a chemically defined medium (M63S). Deletion of R1, a region predicted to encode the methionine salvage pathway, altered the mutant pathogenicity, suggesting that the availability of free methionine is severely restricted in vivo. R3, a region composed mostly of genes of unknown functions, was necessary for both optimal growth of Y. pseudotuberculosis at 37°C in M63S and for virulence. Therefore, despite their loss in Y. pestis, five of the nine Y. pseudotuberculosis-specific chromosomal loci studied play a role in the survival, growth, or virulence of this species.  相似文献   

5.
The method of Garra and Baygorria, for localized antibody production, has been adapted for obtaining high-titered and monospecific antibodies against the murine toxin of Yersinia pestis. A hollow, perforated plastic golf ball surgically implanted under the skin of rabbits induced the formation of a granuloma. When the murine toxin of Y. pestis was injected directly into the granulomatous cavity, an increased amount of antibody was found in the granuloma fluid as compared with serum or with serum antibody obtained by conventional immunization. The granuloma antibody consisted mainly of immunoglobulin G, probably produced locally by lymphoid cells of the granuloma. The immune granuloma fluid and the granuloma tissue were rich in plasma cells and lymphocytes. The chemical composition of the granuloma fluid indicated that it was a transudate.  相似文献   

6.
A 5' nuclease PCR assay for detection of the Yersinia pestis plasminogen activator (pla) gene in human respiratory specimens with simulated Y. pestis infection was developed. An internal positive control was added to the reaction mixture in order to detect the presence of PCR inhibitors that are often found in biological samples. The assay was 100% specific for Y. pestis. In the absence of inhibitors, a sensitivity of 10(2) CFU/ml of respiratory fluid was obtained. When inhibitors were present, detection of Y. pestis DNA required a longer sample treatment time and an initial concentration of bacteria of at least 10(4) CFU/ml. The test's total turnaround time was less than 5 h. The assay described here is well suited to the rapid diagnosis of pneumonic plague, the form of plague most likely to result from a bioterrorist attack.  相似文献   

7.
Purified murine toxin (Ymt) of Yersinia pestis is highly toxic for mice and rats but less active in other animals such as guinea pigs, rabbits, dogs and monkeys. This suggested that Ymt contributes to the very low infectious dose of Y. pestis in mice. The gene encoding Ymt (ymt) is localised on the 100-kb plasmid pFra, which is unique for Y. pestis. Sequence analysis revealed that Ymt showed homology to proteins of the phospholipase D (PLD) superfamily of proteins. Y. pestis strains expressing Ymt possessed PLD activity whereas strains carrying deletions in the ymt gene showed no detectable PLD activity. Western blot analysis showed that Ymt was associated with bacteria under normal growth conditions, and immunogold EM revealed that Ymt was mainly localised in the bacterial cytoplasm. Ymt was purified to homogeneity, and the purified toxin showed a dose-dependent PLD activity. Substitution of amino acids in the PLD consensus motif of Ymt essentially abolished the enzymatic activity and these variants of the toxin were no longer toxic to mice. Interestingly, an in-frame deletion mutant of ymt in the Y pestis strain KIM was not significantly attenuated for mouse virulence. Together with the observation that expression of Ymt was higher at room temperature compared to 37 degrees C this prompted us to investigate the role of Ymt in the flea vector. Fleas were infected with isogenic ymt+ or ymt- mutant strains of Y. pestis. Preliminary results suggest that Ymt is important for survival of Y. pestis in the flea and thereby also for the flea-borne route of infection.  相似文献   

8.
Yersinia pestis, the causative agent of plague, harbors at least three plasmids necessary for full virulence of the organism, two of which are species specific. One of the Y. pestis-specific plasmids, pMT1, is thought to promote deep tissue invasion, resulting in more acute onset of symptoms and death. We determined the entire nucleotide sequence of Y. pestis KIM5 pMT1 and identified potential open reading frames (ORFs) encoded by the 100,990-bp molecule. Based on codon usage for known yersinial genes, homology with known proteins in the databases, and potential ribosome binding sites, we determined that 115 of the potential ORFs which we considered could encode polypeptides in Y. pestis. Five of these ORFs were genes previously identified as being necessary for production of the classic virulence factors, murine toxin (MT), and the fraction 1 (F1) capsule antigen. The regions of pMT1 encoding MT and F1 were surrounded by remnants of multiple transposition events and bacteriophage, respectively, suggesting horizontal gene transfer of these virulence factors. We identified seven new potential virulence factors that might interact with the mammalian host or flea vector. Forty-three of the remaining 115 putative ORFs did not display any significant homology with proteins in the current databases. Furthermore, DNA sequence analysis allowed the determination of the putative replication and partitioning regions of pMT1. We identified a single 2,450-bp region within pMT1 that could function as the origin of replication, including a RepA-like protein similar to RepFIB, RepHI1B, and P1 and P7 replicons. Plasmid partitioning function was located ca. 36 kb from the putative origin of replication and was most similar to the parABS bacteriophage P1 and P7 system. Y. pestis pMT1 encoded potential genes with a high degree of similarity to a wide variety of organisms, plasmids, and bacteriophage. Accordingly, our analysis of the pMT1 DNA sequence emphasized the mosaic nature of this large bacterial virulence plasmid and provided implications as to its evolution.

The facultative intracellular parasite Yersinia pestis harbors at least three plasmids, one of which is common to the enteropathogenic species Yersinia pseudotuberculosis and Yersinia enterocolitica (30, 72). The other two plasmids, designated pMT1 and pPCP1, are unique to Y. pestis (10) and are thought to promote the ability of this organism to penetrate deep tissues and to contribute to the acute infection associated with this species. In fact, the Y. pestis genome shares much homology with that of Y. pseudotuberculosis (2, 63), yet the infection caused by the latter organism is usually mild and self-limiting (15). Accordingly, a logical starting point to understanding the difference in the pathogenesis of Y. pestis and Y. pseudotuberculosis is to study the genes encoded on the plasmids unique to plague, pMT1 and pPCP1.The 9.5-kb plasmid pPCP1 encodes a bacteriocin termed pesticin, a pesticin immunity protein, and a plasminogen activator (89). Loss of this plasmid increases the 50% lethal dose of the organism by a factor of 105 when the organism is injected subcutaneously in the mouse model of infection (90). The only characterized virulence determinant encoded by pPCP1, the plasminogen activator, has been implicated in deep tissue invasion by Y. pestis (11) and functions in the flea vector (58). These facts demonstrate that a plasmid, specifically harbored by Y. pestis, encodes a virulence factor necessary for the acute infection caused by the organism and that a single protein can influence the life cycle of the organism at multiple stages.The largest extrachromosomal element present in Y. pestis was commonly called the cryptic plasmid until 1983. Protsenko et al. (73) demonstrated that the capsular protein fraction 1 (F1) and the murine toxin (MT) were both encoded by the ∼100-kb element now called pMT1. The genes for each of the proteins have been cloned from Y. pestis EV76 and sequenced previously (36, 37, 49). Data addressing the involvement of these proteins in plague pathogenesis are open to interpretation since the effect that mutational loss has on the 50% lethal dose depends on the animal model used in the study as well as the route of infection (8, 9). However, pMT1 does appear to contribute to the acute phase of plague infection, as evidenced by the fact that strains lacking the 100-kb plasmid demonstrate reduced morbidity (27, 80, 96).Information pertaining to the genetic characterization of the pMT1 molecule is limited. The size of the plasmid has been found to vary from approximately 90 to 288 kb in size (31). Furthermore, pMT1 has been found to integrate at multiple sites into the chromosome of Y. pestis at high frequency (74), with speculation that the observed integration of pMT1 into the chromosome may have been due to IS100 homology between the two molecules. Both F1 and MT gene activation have been characterized in relation to environmental cues such as temperature and calcium (28). F1 capsule synthesis is maximal at 37°C in the absence of extracellular calcium while murine toxin expression is induced at 26°C. F1 expression is therefore maximum under conditions similar to those that induce the expression of one of the major virulence determinants of Y. pestis (9193). In contrast, MT production is induced in an environment similar to that which Y. pestis would be expected to encounter in the flea vector. The presence of genes induced under these widely different conditions indicates the presence of at least two networks regulating expression of virulence determinants operating in plague.DNA-sequencing technology has progressed to the point that large amounts of genetic material can be sequenced in a relatively short time. Several facts make pMT1 a good candidate for large-scale DNA sequencing. First, the plasmid is unique to Y. pestis. Second, some derivative of the ∼100-kb plasmid is always present in clinical isolates (31). Third, we already know that genes regulated by two different environmental stimuli that mimic different environments encountered during the life cycle of plague are present on this molecule. Here, we report and annotate the entire DNA sequence of the pMT1 plasmid derived from the Y. pestis laboratory strain KIM.  相似文献   

9.
A Wake  M Misawa    A Matsui 《Infection and immunity》1975,12(5):1211-1213
P+ plague strains contained more siderochrome-producing organisms than P-. Siderochrome enhanced the mouse virulence of an F1+Vw+P1+P-Pu+ strains, inhibited P1 activity, and could be assayed by a paper disk titration method.  相似文献   

10.
When whole cells (stationary phase) of Yersinia pestis strain EV76 were radiolabelled with Iodogen and 131I, 16 major and 10 minor surface-exposed outer membrane proteins (OMPs) were identified. Labelling with N-hydroxysuccinimidyl 6-biotinylamino-hexanoate (biotin X-NHS) resulted in a complex protein profile detectable after blotting and developing with peroxidase-conjugated avidin. Y. pestis cell fractionation revealed that biotin X-NHS labelled not only OMPs but also proteins of inner cell compartments. Therefore, radiolabelling was the more reliable technique for identifying the OMPs of Y. pestis.  相似文献   

11.
Yersinia pestis and many other Gram-negative pathogenic bacteria use the chaperone/usher (CU) pathway to assemble virulence-associated surface fibers termed pili or fimbriae. Y. pestis has two well-characterized CU pathways: the caf genes coding for the F1 capsule and the psa genes coding for the pH 6 antigen. The Y. pestis genome contains additional CU pathways that are capable of assembling pilus fibers, but the roles of these pathways in the pathogenesis of plague are not understood. We constructed deletion mutations in the usher genes for six of the additional Y. pestis CU pathways. The wild-type (WT) and usher deletion strains were compared in the murine bubonic (subcutaneous) and pneumonic (intranasal) plague infection models. Y. pestis strains containing deletions in CU pathways y0348-0352, y1858-1862, and y1869-1873 were attenuated for virulence compared to the WT strain by the intranasal, but not subcutaneous, routes of infection, suggesting specific roles for these pathways during pneumonic plague. We examined binding of the Y. pestis WT and usher deletion strains to A549 human lung epithelial cells, HEp-2 human cervical epithelial cells, and primary human and murine macrophages. Y. pestis CU pathways y0348-0352 and y1858-1862 were found to contribute to adhesion to all host cells tested, whereas pathway y1869-1873 was specific for binding to macrophages. The correlation between the virulence attenuation and host cell binding phenotypes of the usher deletion mutants identifies three of the additional CU pathways of Y. pestis as mediating interactions with host cells that are important for the pathogenesis of plague.  相似文献   

12.
Iron uptake and iron-repressible polypeptides in Yersinia pestis.   总被引:3,自引:0,他引:3       下载免费PDF全文
Pigmented (Pgm+) cells of Yersinia pestis are virulent, are sensitive to pesticin, adsorb exogenous hemin at 26 degrees C (Hms+), produce iron-repressible outer membrane proteins, and grow at 37 degrees C in iron-deficient media. These traits are lost upon spontaneous deletion of a chromosomal 102-kb pgm locus (Pgm-). Here we demonstrate that an Hms+ but pesticin-resistant (Pst(r)) mutant acquired a 5-bp deletion in the pesticin receptor gene (psn) encoding IrpB to IrpD. Growth and assimilation of iron by Pgm- and Hms+ Pst(r) mutants were markedly inhibited by ferrous chelators at 37 degrees C; inhibition by ferric and ferrous chelators was less effective at 26 degrees C. Iron-deficient growth at 26 degrees C induced iron-regulated outer membrane proteins of 34, 28.5, and 22.5 kDa and periplasmic polypeptides of 33.5 and 30 kDa. These findings provide a basis for understanding the psn-driven system of iron uptake, indicate the existence of at least one additional 26 degrees C-dependent iron assimilation system, and define over 30 iron-repressible proteins in Y. pestis.  相似文献   

13.
Macrophages were identified to be a major source of interferon produced in murine spleen cell cultures after intravenous injection of Corynebacterium parvum (C. parvum), strain CN 6134 or Bacille Calmette Guérin (BCG). Another strain of C. parvum, CN 5888, which lacks RES stimulating activity and adjuvant activity in vivo, was not effective when injected intravenously. Protein synthesis was required for interferon activity to be produced and protein synthesis was also required for the antiviral state to be expressed. The antiviral activity was relatively stable to pH 2 and neutralized by an antiserum against virus-induced fibroblast interferon, thus exhibiting some properties of type I interferon. In vitro only CN 6134, the biologically active strain, could induce small amounts of interferon in spleen macrophage cultures. Macrophages from CN 6134 or BCG-infected athymic nu/nu mice produced similar interferon titers as their controls. It is concluded that infection with certain immunomodulators can activate splenic macrophages via a predominantly T-cell independent mechanism. Interferon in turn may operate locally as a mediator of immunoregulation.  相似文献   

14.
To evaluate the role of Yersinia outer proteins (Yops) in conferring protective immunity against plague, six yop loci from Yersinia pestis were individually amplified by PCR, cloned, and expressed in Escherichia coli. The recombinant proteins were purified and injected into mice. Most Yop-vaccinated animals succumbed to infection with either wild-type encapsulated Y. pestis or a virulent, nonencapsulated isogenic variant. Vaccination with YpkA significantly prolonged mean survival time but did not increase overall survival of mice infected with the nonencapsulated strain. The only significant protection against death was observed in YopD-vaccinated mice challenged with the nonencapsulated strain.  相似文献   

15.
Yersinia pestis, the agent of plague, has arisen from a less virulent pathogen, Yersinia pseudotuberculosis, by a rapid evolutionary process. Although Y. pestis displays a large number of virulence phenotypes, it is not yet clear which of these phenotypes descended from Y. pseudotuberculosis and which were acquired independently. Y. pestis is known to replicate in macrophages, but there is no consensus in the literature on whether Y. pseudotuberculosis shares this property. We investigated whether the ability to replicate in macrophages is common to Y. pestis and Y. pseudotuberculosis or is a unique phenotype of Y. pestis. We also examined whether a chromosomal type III secretion system (TTSS) found in Y. pestis is present in Y. pseudotuberculosis and whether this system is important for replication of Yersinia in macrophages. A number of Y. pestis and Y. pseudotuberculosis strains of different biovars and serogroups, respectively, were tested for the ability to replicate in primary murine macrophages. Two Y. pestis strains (EV766 and KIM10(+)) and three Y. pseudotuberculosis strains (IP2790c, IP2515c, and IP2666c) were able to replicate in macrophages with similar efficiencies. Only one of six strains tested, the Y. pseudotuberculosis YPIII(p(-)) strain, was defective for intracellular replication. Thus, the ability to replicate in macrophages is conserved in Y. pestis and Y. pseudotuberculosis. Our results also indicate that a homologous TTSS is present on the chromosomes of Y. pestis and Y. pseudotuberculosis and that this secretion system is not required for replication of these bacteria in macrophages.  相似文献   

16.
17.
Denaturing high-performance liquid chromatography (DHPLC) has been used extensively to detect genetic variation. We used this method to detect and identify Yersinia pestis KIM5 ciprofloxacin-resistant isolates by analyzing the quinolone resistance-determining region (QRDR) of the gyrase A gene. Sequencing of the Y. pestis KIM5 strain gyrA QRDR from 55 ciprofloxacin-resistant isolates revealed five mutation types. We analyzed the gyrA QRDR by DHPLC to assess its ability to detect point mutations and to determine whether DHPLC peak profile analysis could be used as a molecular fingerprint. In addition to the five mutation types found in our ciprofloxacin-resistant isolates, several mutations in the QRDR were generated by site-directed mutagenesis and analyzed to further evaluate this method for the ability to detect QRDR mutations. Furthermore, a blind panel of 42 samples was analyzed by screening for two mutant types to evaluate the potential diagnostic value of this method. Our results showed that DHPLC is an efficient method for detecting mutations in genes that confer antibiotic resistance.  相似文献   

18.
19.
Storage reservoirs of hemin and inorganic iron in Yersinia pestis.   总被引:8,自引:2,他引:8       下载免费PDF全文
It is established that a high-frequency chromosomal deletion of ca. 100 kb accounts for the loss of properties making up the pigmented phenotype (Pgm+) of wild-type Yersinia pestis. These determinants are known to include virulence by peripheral routes of injection, sensitivity to the bacteriocin pesticin, adsorption of exogenous hemin or Congo red at 26 degrees C, and growth in iron-sequestered medium at 37 degrees C. We have now identified the outer membrane as the primary site of exogenous hemin storage in Pgm+ cells grown at 26 degrees C. Significant outer membrane storage of hemin did not occur in Pgm- mutants or in Pgm+ cells cultivated at 37 degrees C. However, both Pgm+ and Pgm- organisms grown at 37 degrees C contained a periplasmic reservoir of hemin, which may be associated with a temperature-dependent ca. 70-kDa peptide recently equated with antigen 5. At 37 degrees C, Pgm+ and Pgm- yersiniae also utilized a cytoplasmic ca. 19-kDa bacterioferritin-like peptide for deposition of inorganic iron. Incorporation of [55Fe]hemin into pools at 37 degrees C was not significantly inhibited by competition with excess unlabeled Fe3+. However, excess unlabeled hemin modestly competed with incorporation of label from 55FeCl3. This relative independence of storage pools observed at 37 degrees C is consistent with physiological linkage to in vivo acquisition and transport of Fe3+ from ferritin and of hemin from hemoglobin, myoglobin, or hemopexin.  相似文献   

20.
The virulence of the pathogenic Yersinia species depends on a plasmid-encoded type III secretion system (T3SS) that injects six Yersinia outer protein (Yop) effector proteins into the cytosol of macrophages, leading to disruption of host defence mechanisms. Here, we report that a T3SS structural protein YscW of Yersinia pestis contributed to the induction of apoptosis of murine macrophages. The apoptotic percentage of macrophages, from both mouse peritoneal cavity and spleen, and of RAW264.7 cell line, caused by the yscW mutant strain was significantly lower than that by wild type (WT) Y. pestis and yscW complemented strain. Meanwhile, detection of caspase-3 activity in macrophages, a key apoptosis-inducing protein, showed coincident results with the changes of macrophage apoptosis induced by WT, yscW mutant and complemented strains, indicating that macrophage apoptosis was related to caspase-3 signal pathways. However, ectopic expression of YscW in RAW264.7 cells cannot increase the macrophage apoptosis and death, suggesting that YscW itself could not induce macrophage apoptosis directly. To get insight into the mechanism of this phenomenon, we investigated the secretion of YopJ, which has been thought to be the only Yop effector related to apoptosis, in WT, mutant and complemented strains, respectively. Results showed that in yscW mutant strain, secretion of YopJ was decreased significantly in the supernatant than that in WT or complemented strain. This means although YscW does not induce apoptosis directly, it can indirectly affect apoptosis through reducing the secretion of YopJ.  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