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.
Stability of plant tumor-inducing (Ti) plasmids differs among strains. A high level of stability prevents basic and applied studies including the development of useful strains. The nopaline type Ti plasmid pTiC58 significantly reduces the transconjugant efficiency for incoming incompatible plasmids relative to the other type, such as octopine-type plasmids. In this study we identified a region that increases the incompatibility and stability of the plasmid. This region was located on a 4.3-kbp segment about 38 kbp downstream of the replication locus, repABC. We named two open reading frames in the segment, ietA and ietS, both of which were essential for the high level of incompatibility and stability. Plasmid stabilization by ietAS was accomplished by a toxin-antitoxin (TA) mechanism, where IetS is the toxin and IetA is the antitoxin. A database search revealed that putative IetA and IetS proteins are highly similar to AAA-ATPases and subtilisin-like serine proteases, respectively. Amino acid substitution experiments in each of the highly conserved characteristic residues, in both putative enzymes, suggested that the protease activity is essential and that ATP binding activity is important for the operation of the TA system. The ietAS-containing repABC plasmids expelled Ti plasmids even in strains which were tolerant to conventional Ti-curing treatments.Agrobacterium tumefaciens strains bearing a tumor-inducing (Ti) plasmid are the etiological agents of crown gall disease. Most genes required for pathogenicity are located on the plasmids (17, 33). Ti plasmids are kept stable at a low copy number equivalent to that of the chromosomal DNA in the bacterial cells (32) due to the repABC locus (16, 30, 34). The stability of Ti plasmids differs among strains (11).Many genes for keeping plasmids stable have been reported in eubacteria, and these are divided into three categories based on their mechanism: multimer resolution systems, active partitioning systems, and toxin-antitoxin (TA) systems (15). Multimer resolution systems increase the number of plasmid molecules by resolving a multimer plasmid into monomers, resulting in a higher probability of plasmid distribution to daughter cells during cell division even when plasmid distribution occurs randomly (29). Active partitioning systems deliver plasmid copies to each progeny cell at cell division (21). In the repABC locus, the RepA and RepB proteins and parS site(s) ensure stable plasmid inheritance by the active partitioning system (2). TA systems contribute to plasmid maintenance in cell populations by initiating growth inhibition or death of plasmid-free cells and are widely distributed among eubacterial and archaeal plasmids as well as their chromosomes (9). Generally, the TA module consists of two genes which encode toxin and antitoxin. The antitoxin neutralizes the action of a cognate toxin by interaction with the toxin or its target molecules. When a plasmid harboring the TA module is lost from a host cell, the antitoxin molecules decrease to an ineffective level because the antitoxin is degraded quickly or diluted by cell division (15). Thereafter, the toxin exerts its toxicity and inhibits the host cell growth. RNA antitoxins can suppress toxin expression by binding to the toxin mRNA as an antisense RNA or repress toxicity effects by an unknown mechanism (6, 4). In pTi-SAKURA, the Ti plasmid in the A. tumefaciens strain MAFF301001, it was shown that the tiorf24 and tiorf25 module increased plasmid stability by the TA mechanism (40; also S. Yamamoto, unpublished data).Differences in Ti plasmid stability are critical for plasmid engineering and evolution (33). However, little is known about the stability factors of Ti plasmids other than the repABC locus. In our previous study (40), tiorf24 and tiorf25 were shown to increase the segregational stability and incompatibility of Ti plasmids and reduce the efficiency of transconjugants by the introduction of incompatible plasmids into host cells. The two genes are located 2.5-kbp downstream of repABC (8). Generally, incompatibility has been defined as a situation where two plasmids contain a related replication and/or partitioning system and are unable to exist in a cell simultaneously without external selection (1). A. tumefaciens strain C58, which contains a Ti plasmid pTiC58, allows entry of an incompatible repABC plasmid into the cell 60-fold less efficiently than a derivative of C58. The derivative of C58 harbors a small repABC vector instead of pTiC58 (40). This suggests the presence of incompatibility-enhancing genes on pTiC58.In this study, we located the responsible genes in pTiC58 and found that the novel genes ietA and ietS enhance the incompatibility and stability of the plasmid by the TA mechanism.  相似文献   

2.
3.
4.
Immunogold localization revealed that OmcS, a cytochrome that is required for Fe(III) oxide reduction by Geobacter sulfurreducens, was localized along the pili. The apparent spacing between OmcS molecules suggests that OmcS facilitates electron transfer from pili to Fe(III) oxides rather than promoting electron conduction along the length of the pili.There are multiple competing/complementary models for extracellular electron transfer in Fe(III)- and electrode-reducing microorganisms (8, 18, 20, 44). Which mechanisms prevail in different microorganisms or environmental conditions may greatly influence which microorganisms compete most successfully in sedimentary environments or on the surfaces of electrodes and can impact practical decisions on the best strategies to promote Fe(III) reduction for bioremediation applications (18, 19) or to enhance the power output of microbial fuel cells (18, 21).The three most commonly considered mechanisms for electron transfer to extracellular electron acceptors are (i) direct contact between redox-active proteins on the outer surfaces of the cells and the electron acceptor, (ii) electron transfer via soluble electron shuttling molecules, and (iii) the conduction of electrons along pili or other filamentous structures. Evidence for the first mechanism includes the necessity for direct cell-Fe(III) oxide contact in Geobacter species (34) and the finding that intensively studied Fe(III)- and electrode-reducing microorganisms, such as Geobacter sulfurreducens and Shewanella oneidensis MR-1, display redox-active proteins on their outer cell surfaces that could have access to extracellular electron acceptors (1, 2, 12, 15, 27, 28, 31-33). Deletion of the genes for these proteins often inhibits Fe(III) reduction (1, 4, 7, 15, 17, 28, 40) and electron transfer to electrodes (5, 7, 11, 33). In some instances, these proteins have been purified and shown to have the capacity to reduce Fe(III) and other potential electron acceptors in vitro (10, 13, 29, 38, 42, 43, 48, 49).Evidence for the second mechanism includes the ability of some microorganisms to reduce Fe(III) that they cannot directly contact, which can be associated with the accumulation of soluble substances that can promote electron shuttling (17, 22, 26, 35, 36, 47). In microbial fuel cell studies, an abundance of planktonic cells and/or the loss of current-producing capacity when the medium is replaced is consistent with the presence of an electron shuttle (3, 14, 26). Furthermore, a soluble electron shuttle is the most likely explanation for the electrochemical signatures of some microorganisms growing on an electrode surface (26, 46).Evidence for the third mechanism is more circumstantial (19). Filaments that have conductive properties have been identified in Shewanella (7) and Geobacter (41) species. To date, conductance has been measured only across the diameter of the filaments, not along the length. The evidence that the conductive filaments were involved in extracellular electron transfer in Shewanella was the finding that deletion of the genes for the c-type cytochromes OmcA and MtrC, which are necessary for extracellular electron transfer, resulted in nonconductive filaments, suggesting that the cytochromes were associated with the filaments (7). However, subsequent studies specifically designed to localize these cytochromes revealed that, although the cytochromes were extracellular, they were attached to the cells or in the exopolymeric matrix and not aligned along the pili (24, 25, 30, 40, 43). Subsequent reviews of electron transfer to Fe(III) in Shewanella oneidensis (44, 45) appear to have dropped the nanowire concept and focused on the first and second mechanisms.Geobacter sulfurreducens has a number of c-type cytochromes (15, 28) and multicopper proteins (12, 27) that have been demonstrated or proposed to be on the outer cell surface and are essential for extracellular electron transfer. Immunolocalization and proteolysis studies demonstrated that the cytochrome OmcB, which is essential for optimal Fe(III) reduction (15) and highly expressed during growth on electrodes (33), is embedded in the outer membrane (39), whereas the multicopper protein OmpB, which is also required for Fe(III) oxide reduction (27), is exposed on the outer cell surface (39).OmcS is one of the most abundant cytochromes that can readily be sheared from the outer surfaces of G. sulfurreducens cells (28). It is essential for the reduction of Fe(III) oxide (28) and for electron transfer to electrodes under some conditions (11). Therefore, the localization of this important protein was further investigated.  相似文献   

5.
6.
Analysis of Lyme borreliosis (LB) spirochetes, using a novel multilocus sequence analysis scheme, revealed that OspA serotype 4 strains (a rodent-associated ecotype) of Borrelia garinii were sufficiently genetically distinct from bird-associated B. garinii strains to deserve species status. We suggest that OspA serotype 4 strains be raised to species status and named Borrelia bavariensis sp. nov. The rooted phylogenetic trees provide novel insights into the evolutionary history of LB spirochetes.Multilocus sequence typing (MLST) and multilocus sequence analysis (MLSA) have been shown to be powerful and pragmatic molecular methods for typing large numbers of microbial strains for population genetics studies, delineation of species, and assignment of strains to defined bacterial species (4, 13, 27, 40, 44). To date, MLST/MLSA schemes have been applied only to a few vector-borne microbial populations (1, 6, 30, 37, 40, 41, 47).Lyme borreliosis (LB) spirochetes comprise a diverse group of zoonotic bacteria which are transmitted among vertebrate hosts by ixodid (hard) ticks. The most common agents of human LB are Borrelia burgdorferi (sensu stricto), Borrelia afzelii, Borrelia garinii, Borrelia lusitaniae, and Borrelia spielmanii (7, 8, 12, 35). To date, 15 species have been named within the group of LB spirochetes (6, 31, 32, 37, 38, 41). While several of these LB species have been delineated using whole DNA-DNA hybridization (3, 20, 33), most ecological or epidemiological studies have been using single loci (5, 9-11, 29, 34, 36, 38, 42, 51, 53). Although some of these loci have been convenient for species assignment of strains or to address particular epidemiological questions, they may be unsuitable to resolve evolutionary relationships among LB species, because it is not possible to define any outgroup. For example, both the 5S-23S intergenic spacer (5S-23S IGS) and the gene encoding the outer surface protein A (ospA) are present only in LB spirochete genomes (36, 43). The advantage of using appropriate housekeeping genes of LB group spirochetes is that phylogenetic trees can be rooted with sequences of relapsing fever spirochetes. This renders the data amenable to detailed evolutionary studies of LB spirochetes.LB group spirochetes differ remarkably in their patterns and levels of host association, which are likely to affect their population structures (22, 24, 46, 48). Of the three main Eurasian Borrelia species, B. afzelii is adapted to rodents, whereas B. valaisiana and most strains of B. garinii are maintained by birds (12, 15, 16, 23, 26, 45). However, B. garinii OspA serotype 4 strains in Europe have been shown to be transmitted by rodents (17, 18) and, therefore, constitute a distinct ecotype within B. garinii. These strains have also been associated with high pathogenicity in humans, and their finer-scale geographical distribution seems highly focal (10, 34, 52, 53).In this study, we analyzed the intra- and interspecific phylogenetic relationships of B. burgdorferi, B. afzelii, B. garinii, B. valaisiana, B. lusitaniae, B. bissettii, and B. spielmanii by means of a novel MLSA scheme based on chromosomal housekeeping genes (30, 48).  相似文献   

7.
8.
9.
10.
Many species of Rickettsia are well-known mammalian pathogens transmitted by blood-feeding arthropods. However, molecular surveys are continually uncovering novel Rickettsia species, often in unexpected hosts, including many arthropods that do not feed on blood. This study reports a systematic molecular characterization of a Rickettsia infecting the psocid Liposcelis bostrychophila (Psocoptera: Liposcelidae), a common and cosmopolitan household pest. Surprisingly, the psocid Rickettsia is shown to be Rickettsia felis, a human pathogen transmitted by fleas that causes serious morbidity and occasional mortality. The plasmid from the psocid R. felis was sequenced and was found to be virtually identical to the one in R. felis from fleas. As Liposcelis insects are often intimately associated with humans and other vertebrates, it is speculated that they acquired R. felis from fleas. Whether the R. felis in psocids causes disease in vertebrates is not known and warrants further study.Many species of Rickettsia are well-known mammalian pathogens that are transmitted by blood-feeding arthropods via bites or feces and can cause mild to fatal diseases in humans (33). Some species are also considered potential bioterrorism agents (4). Most Rickettsia research has focused on pathogens that are found in two closely related species groups, the typhus and spotted fever groups, such as Rickettsia prowazekii, Rickettsia rickettsii, and Rickettsia typhi, the causal agents of epidemic typhus, Rocky Mountain spotted fever, and murine typhus, respectively (3, 4, 33). However, recent surveys suggest that Rickettsia bacteria are much more widespread than previously suspected and that they are being detected in novel hosts, the vast majority of which are arthropods, including many that do not feed on blood (29, 45).The number of new rickettsial species that cause diseases in humans is rapidly increasing (33). One such species that has been generating much interest in recent years is Rickettsia felis, the causative agent of a murine typhus-like disease (1, 2, 13, 16, 17, 28, 44). The disease is often unrecognized, and even though it is considered clinically mild, it can cause severe illness and death in older patients and in cases of delayed diagnosis (2). R. felis was identified only in 1990 (1) and has since been found worldwide in fleas, where it is maintained transovarially and can reach high infection rates (e.g., 86% to 94% in cat fleas) (2, 3, 44), as well as in ticks and mites (34). While experimental infections have confirmed that R. felis is transmitted to vertebrate hosts via blood feeding and that R. felis occurs in an infectious extracellular state (39), it is not known whether transmission can also occur through contamination of broken skin by infected vector feces, as in R. typhi (3, 34).A number of features distinguish R. felis from species in both the typhus and spotted fever groups. Lately, it has been proposed that R. felis be in its own group, allied with Rickettsia akari and Rickettsia australis, the causal agents of rickettsial pox and Queensland tick typhus, respectively, and a number of recently discovered strains infecting insects that do not feed on blood (16, 17, 29, 45). Moreover, R. felis was the first Rickettsia species shown to have a plasmid (28). While plasmids now appear to be quite widespread in the genus, the R. felis plasmid stands out with respect to its relatively large size and distinctive gene content (5, 6, 9, 14, 17).This study reports that a common and cosmopolitan insect, the psocid Liposcelis bostrychophila (Psocoptera: Liposcelidae) harbors R. felis. Liposcelids are the closest free-living relatives of parasitic lice (19) and are well-known for their close proximity to humans, particularly as pests in houses and grain storage facilities (8, 41). Through 16S rRNA gene sequencing, L. bostrychophila was recently shown to harbor a strain of Rickettsia (29, 30, 42). A systematic molecular characterization of this Rickettsia was conducted, demonstrating that it is authentic R. felis. Furthermore, the psocid symbiont plasmid was sequenced and was shown to be virtually identical to the plasmid from R. felis that infects cat fleas.  相似文献   

11.
12.
We report here the isolation and recombinational cloning of a large plasmid, pZL12, from endophytic Streptomyces sp. 9R-2. pZL12 comprises 90,435 bp, encoding 112 genes, 30 of which are organized in a large operon resembling bacteriophage genes. A replication locus (repA) and a conjugal transfer locus (traA-traC) were identified in pZL12. Surprisingly, the supernatant of a 9R-2 liquid culture containing partially purified phage particles infected 9R-2 cured of pZL12 (9R-2X) to form plaques, and a phage particle (φZL12) was observed by transmission electron microscopy. Major structural proteins (capsid, portal, and tail) of φZL12 virions were encoded by pZL12 genes. Like bacteriophage P1, linear φZL12 DNA contained ends from a largely random pZL12 sequence. There was also a hot end sequence in linear φZL12. φZL12 virions efficiently infected only one host, 9R-2X, but failed to infect and form plaques in 18 other Streptomyces strains. Some 9R-2X spores rescued from lysis by infection of φZL12 virions contained a circular pZL12 plasmid, completing a cycle comprising autonomous plasmid pZL12 and lytic phage φZL12. These results confirm pZL12 as the first example of a plasmid-phage in Streptomyces.Streptomyces species, a major source of antibiotics and pharmacologically active metabolites, are Gram-positive, mycelial bacteria with high G+C content in their DNA (15). They usually harbor conjugative circular and/or linear plasmids, propagating in autonomous and/or chromosomally integrated forms (14). Most Streptomyces circular plasmids reported are small (8 to 14 kb), including rolling-circle-replication (RCR) plasmids (pIJ101, pJV1, pSG5, pSN22, pSVH1, pSB24.2, pSY10, pSNA1, pSLG33, pEN2701, etc.) (12, 14) and chromosomally integrating/autonomous plasmids (SLP1 and pSAM2) (4, 27, 28). Some theta replication plasmids are of intermediate size (31 to 39 kb), such as SCP2, pFP1, and pFP11 (13, 40). These theta replication loci comprise a rep gene and an adjacent noncoding or iteron sequence, to which Rep protein binds specifically in vitro (10, 40). The occurrence of an ∼163-kb large plasmid, pSV1, in Streptomyces violaceoruber SANK95570 was confirmed (1, 37), but this plasmid could not be physically isolated by standard procedures for plasmid preparation (17). In contrast to more than 30 genes for conjugal transfer on the Escherichia coli F plasmid (20), Streptomyces plasmids usually need a single tra gene (encoding a DNA translocase containing a cell division FtsK/SpoIIIE domain) (15, 29). The transfer of Streptomyces circular plasmids involves binding of the nonnicked double-stranded DNA (dsDNA) by multimers of Tra proteins at a noncoding sequence and ATP hydrolysis-dependent translocation of this DNA through the hyphal tips of the Streptomyces mycelium (15, 32).Numerous Streptomyces phages have been described, including φC31 (22), SAt1 (26), TG1 (11), FP43 (24), φSPK1 (19), φSC623 (34), DAH2/DAH4/DAH5/DAH6 (6), and mu1/6 (9). They range in size from 36 kb (19) to 121 kb (6), with 50 to 71.2% GC content (9, 23, 35). Streptomyces phages often have a wide host range; for example, 16 of 27 Streptomyces strains are susceptible to infection by φSPK1 (19), and phage FP43 transduces species of Streptoverticillium, Chainia, and Sacchropolyspora (24). φC31 is the most-studied Streptomyces phage and cloning vector (8). The sequences of the φC31 head proteins (e.g., portal, capsid, and head protease) resemble those of other bacterial dsDNA phages, suggesting evolutionary relationships to other viruses (35).We report here the isolation and recombinational cloning of a 90,435-bp plasmid, pZL12, from endophytic Streptomyces sp. 9R-2 and the characterization of its replication and transfer. Surprisingly, the supernatant of 9R-2 liquid culture infected 9R-2 cured of pZL12 to form plaques. A cycle comprising autonomous plasmid pZL12 and lytic phage φZL12 is described.  相似文献   

13.
A conjugative plasmid from the catheter-associated urinary tract infection strain Escherichia coli MS2027 was sequenced and annotated. This 42,644-bp plasmid, designated pMAS2027, contains 58 putative genes and is most closely related to plasmids belonging to incompatibility group X (IncX1). Plasmid pMAS2027 encodes two important virulence factors: type 3 fimbriae and a type IV secretion (T4S) system. Type 3 fimbriae, recently found to be functionally expressed in E. coli, played an important role in biofilm formation. Biofilm formation by E. coli MS2027 was specifically due to expression of type 3 fimbriae and not the T4S system. The T4S system, however, accounted for the conjugative ability of pMAS2027 and enabled a non-biofilm-forming strain to grow as part of a mixed biofilm following acquisition of this plasmid. Thus, the importance of conjugation as a mechanism to spread biofilm determinants was demonstrated. Conjugation may represent an important mechanism by which type 3 fimbria genes are transferred among the Enterobacteriaceae that cause device-related infections in nosocomial settings.Bacterial biofilms are complex communities of bacterial cells living in close association with a surface (17). Bacterial cells in these protected environments are often resistant to multiple factors, including antimicrobials, changes in the pH, oxygen radicals, and host immune defenses (19, 38). Biofilm formation is a property of many bacterial species, and a range of molecular mechanisms that facilitate this process have been described (2, 3, 11, 14, 16, 29, 33, 34). Often, the ability to form a biofilm is dependent on the production of adhesins on the bacterial cell surface. In Escherichia coli, biofilm formation is enhanced by the production of certain types of fimbriae (e.g., type 1 fimbriae, type 3 fimbriae, F1C, F9, curli, and conjugative pili) (14, 23, 25, 29, 33, 39, 46), cell surface adhesins (e.g., autotransporter proteins such as antigen 43, AidA, TibA, EhaA, and UpaG) (21, 34, 35, 40, 43), and flagella (22, 45).The close proximity of bacterial cells in biofilms creates an environment conducive for the exchange of genetic material. Indeed, plasmid-mediated conjugation in monospecific and mixed E. coli biofilms has been demonstrated (6, 18, 24, 31). The F plasmid represents the best-characterized conjugative system for biofilm formation by E. coli. The F pilus mediates adhesion to abiotic surfaces and stabilizes the biofilm structure through cell-cell interactions (16, 30). Many other conjugative plasmids also contribute directly to biofilm formation upon derepression of the conjugative function (16).One example of a conjugative system employed by gram-negative Enterobacteriaceae is the type 4 secretion (T4S) system. The T4S system is a multisubunit structure that spans the cell envelope and contains a secretion channel often linked to a pilus or other surface filament or protein (8). The Agrobacterium tumefaciens VirB-VirD4 system is the archetypical T4S system and is encoded by 11 genes in the virB operon and one gene (virD4) in the virD operon (7, 8). Genes with strong homology to genes in the virB operon have also been identified on other conjugative plasmids. For example, the pilX1 to pilX11 genes on the E. coli R6K IncX plasmid and the virB1 to virB11 genes are highly conserved at the nucleotide level (28).We recently described identification and characterization of the mrk genes encoding type 3 fimbriae in a uropathogenic strain of E. coli isolated from a patient with a nosocomial catheter-associated urinary tract infection (CAUTI) (29). The mrk genes were located on a conjugative plasmid (pMAS2027) and were strongly associated with biofilm formation. In this study we determined the entire sequence of plasmid pMAS2027 and revealed the presence of conjugative transfer genes homologous to the pilX1 to pilX11 genes of E. coli R6K (in addition to the mrk genes). We show here that biofilm formation is driven primarily by type 3 fimbriae and that the T4S apparatus is unable to mediate biofilm growth in the absence of the mrk genes. Finally, we demonstrate that conjugative transfer of pMAS2027 within a mixed biofilm confers biofilm formation properties on recipient cells due to acquisition of the type 3 fimbria-encoding mrk genes.  相似文献   

14.
Factors potentially contributing to the lower incidence of Lyme borreliosis (LB) in the far-western than in the northeastern United States include tick host-seeking behavior resulting in fewer human tick encounters, lower densities of Borrelia burgdorferi-infected vector ticks in peridomestic environments, and genetic variation among B. burgdorferi spirochetes to which humans are exposed. We determined the population structure of B. burgdorferi in over 200 infected nymphs of the primary bridging vector to humans, Ixodes pacificus, collected in Mendocino County, CA. This was accomplished by sequence typing the spirochete lipoprotein ospC and the 16S-23S rRNA intergenic spacer (IGS). Thirteen ospC alleles belonging to 12 genotypes were found in California, and the two most abundant, ospC genotypes H3 and E3, have not been detected in ticks in the Northeast. The most prevalent ospC and IGS biallelic profile in the population, found in about 22% of ticks, was a new B. burgdorferi strain defined by ospC genotype H3. Eight of the most common ospC genotypes in the northeastern United States, including genotypes I and K that are associated with disseminated human infections, were absent in Mendocino County nymphs. ospC H3 was associated with hardwood-dominated habitats where western gray squirrels, the reservoir host, are commonly infected with LB spirochetes. The differences in B. burgdorferi population structure in California ticks compared to the Northeast emphasize the need for a greater understanding of the genetic diversity of spirochetes infecting California LB patients.In the United States, Lyme borreliosis (LB) is the most commonly reported vector-borne illness and is caused by infection with the spirochete Borrelia burgdorferi (3, 9, 52). The signs and symptoms of LB can include a rash, erythema migrans, fever, fatigue, arthritis, carditis, and neurological manifestations (50, 51). The black-legged tick, Ixodes scapularis, and the western black-legged tick, Ixodes pacificus, are the primary vectors of B. burgdorferi to humans in the United States, with the former in the northeastern and north-central parts of the country and the latter in the Far West (9, 10). These ticks perpetuate enzootic transmission cycles together with a vertebrate reservoir host such as the white-footed mouse, Peromyscus leucopus, in the Northeast and Midwest (24, 35), or the western gray squirrel, Sciurus griseus, in California (31, 46).B. burgdorferi is a spirochete species with a largely clonal population structure (14, 16) comprising several different strains or lineages (8). The polymorphic ospC gene of B. burgdorferi encodes a surface lipoprotein that increases expression within the tick during blood feeding (47) and is required for initial infection of mammalian hosts (25, 55). To date, approximately 20 North American ospC genotypes have been described (40, 45, 49, 56). At least four, and possibly up to nine, of these genotypes are associated with B. burgdorferi invasiveness in humans (1, 15, 17, 49, 57). Restriction fragment length polymorphism (RFLP) and, subsequently, sequence analysis of the 16S-23S rRNA intergenic spacer (IGS) are used as molecular typing tools to investigate genotypic variation in B. burgdorferi (2, 36, 38, 44, 44, 57). The locus maintains a high level of variation between related species, and this variation reflects the heterogeneity found at the genomic level of the organism (37). The IGS and ospC loci appear to be linked (2, 8, 26, 45, 57), but the studies to date have not been representative of the full range of diversity of B. burgdorferi in North America.Previous studies in the northeastern and midwestern United States have utilized IGS and ospC genotyping to elucidate B. burgdorferi evolution, host strain specificity, vector-reservoir associations, and disease risk to humans. In California, only six ospC and five IGS genotypes have been described heretofore in samples from LB patients or I. pacificus ticks (40, 49, 56) compared to approximately 20 ospC and IGS genotypes identified in ticks, vertebrate hosts, or humans from the Northeast and Midwest (8, 40, 45, 49, 56). Here, we employ sequence analysis of both the ospC gene and IGS region to describe the population structure of B. burgdorferi in more than 200 infected I. pacificus nymphs from Mendocino County, CA, where the incidence of LB is among the highest in the state (11). Further, we compare the Mendocino County spirochete population to populations found in the Northeast.  相似文献   

15.
16.
17.
In their vertebrate hosts, arboviruses such as Semliki Forest virus (SFV) (Togaviridae) generally counteract innate defenses and trigger cell death. In contrast, in mosquito cells, following an early phase of efficient virus production, a persistent infection with low levels of virus production is established. Whether arboviruses counteract RNA interference (RNAi), which provides an important antiviral defense system in mosquitoes, is an important question. Here we show that in Aedes albopictus-derived mosquito cells, SFV cannot prevent the establishment of an antiviral RNAi response or prevent the spread of protective antiviral double-stranded RNA/small interfering RNA (siRNA) from cell to cell, which can inhibit the replication of incoming virus. The expression of tombusvirus siRNA-binding protein p19 by SFV strongly enhanced virus spread between cultured cells rather than virus replication in initially infected cells. Our results indicate that the spread of the RNAi signal contributes to limiting virus dissemination.In animals, RNA interference (RNAi) was first described for Caenorhabditis elegans (27). The production or introduction of double-stranded RNA (dsRNA) in cells leads to the degradation of mRNAs containing homologous sequences by sequence-specific cleavage of mRNAs. Central to RNAi is the production of 21- to 26-nucleotide small interfering RNAs (siRNAs) from dsRNA and the assembly of an RNA-induced silencing complex (RISC), followed by the degradation of the target mRNA (23, 84). RNAi is a known antiviral strategy of plants (3, 53) and insects (21, 39, 51). Study of Drosophila melanogaster in particular has given important insights into RNAi responses against pathogenic viruses and viral RNAi inhibitors (31, 54, 83, 86, 91). RNAi is well characterized for Drosophila, and orthologs of antiviral RNAi genes have been found in Aedes and Culex spp. (13, 63).Arboviruses, or arthropod-borne viruses, are RNA viruses mainly of the families Bunyaviridae, Flaviviridae, and Togaviridae. The genus Alphavirus within the family Togaviridae contains several mosquito-borne pathogens: arboviruses such as Chikungunya virus (16) and equine encephalitis viruses (88). Replication of the prototype Sindbis virus and Semliki Forest virus (SFV) is well understood (44, 71, 74, 79). Their genome consists of a positive-stranded RNA with a 5′ cap and a 3′ poly(A) tail. The 5′ two-thirds encodes the nonstructural polyprotein P1234, which is cleaved into four replicase proteins, nsP1 to nsP4 (47, 58, 60). The structural polyprotein is encoded in the 3′ one-third of the genome and cleaved into capsid and glycoproteins after translation from a subgenomic mRNA (79). Cytoplasmic replication complexes are associated with cellular membranes (71). Viruses mature by budding at the plasma membrane (35).In nature, arboviruses are spread by arthropod vectors (predominantly mosquitoes, ticks, flies, and midges) to vertebrate hosts (87). Little is known about how arthropod cells react to arbovirus infection. In mosquito cell cultures, an acute phase with efficient virus production is generally followed by the establishment of a persistent infection with low levels of virus production (9). This is fundamentally different from the cytolytic events following arbovirus interactions with mammalian cells and pathogenic insect viruses with insect cells. Alphaviruses encode host response antagonists for mammalian cells (2, 7, 34, 38).RNAi has been described for mosquitoes (56) and, when induced before infection, antagonizes arboviruses and their replicons (1, 4, 14, 15, 29, 30, 32, 42, 64, 65). RNAi is also functional in various mosquito cell lines (1, 8, 43, 49, 52). In the absence of RNAi, alphavirus and flavivirus replication and/or dissemination is enhanced in both mosquitoes and Drosophila (14, 17, 31, 45, 72). RNAi inhibitors weakly enhance SFV replicon replication in tick and mosquito cells (5, 33), posing the questions of how, when, and where RNAi interferes with alphavirus infection in mosquito cells.Here we use an A. albopictus-derived mosquito cell line to study RNAi responses to SFV. Using reporter-based assays, we demonstrate that SFV cannot avoid or efficiently inhibit the establishment of an RNAi response. We also demonstrate that the RNAi signal can spread between mosquito cells. SFV cannot inhibit cell-to-cell spread of the RNAi signal, and spread of the virus-induced RNAi signal (dsRNA/siRNA) can inhibit the replication of incoming SFV in neighboring cells. Furthermore, we show that SFV expression of a siRNA-binding protein increases levels of virus replication mainly by enhancing virus spread between cells rather than replication in initially infected cells. Taken together, these findings suggest a novel mechanism, cell-to-cell spread of antiviral dsRNA/siRNA, by which RNAi limits SFV dissemination in mosquito cells.  相似文献   

18.
19.
Adhesive pili on the surface of the serotype M1 Streptococcus pyogenes strain SF370 are composed of a major backbone subunit (Spy0128) and two minor subunits (Spy0125 and Spy0130), joined covalently by a pilin polymerase (Spy0129). Previous studies using recombinant proteins showed that both minor subunits bind to human pharyngeal (Detroit) cells (A. G. Manetti et al., Mol. Microbiol. 64:968-983, 2007), suggesting both may act as pilus-presented adhesins. While confirming these binding properties, studies described here indicate that Spy0125 is the pilus-presented adhesin and that Spy0130 has a distinct role as a wall linker. Pili were localized predominantly to cell wall fractions of the wild-type S. pyogenes parent strain and a spy0125 deletion mutant. In contrast, they were found almost exclusively in culture supernatants in both spy0130 and srtA deletion mutants, indicating that the housekeeping sortase (SrtA) attaches pili to the cell wall by using Spy0130 as a linker protein. Adhesion assays with antisera specific for individual subunits showed that only anti-rSpy0125 serum inhibited adhesion of wild-type S. pyogenes to human keratinocytes and tonsil epithelium to a significant extent. Spy0125 was localized to the tip of pili, based on a combination of mutant analysis and liquid chromatography-tandem mass spectrometry analysis of purified pili. Assays comparing parent and mutant strains confirmed its role as the adhesin. Unexpectedly, apparent spontaneous cleavage of a labile, proline-rich (8 of 14 residues) sequence separating the N-terminal ∼1/3 and C-terminal ∼2/3 of Spy0125 leads to loss of the N-terminal region, but analysis of internal spy0125 deletion mutants confirmed that this has no significant effect on adhesion.The group A Streptococcus (S. pyogenes) is an exclusively human pathogen that commonly colonizes either the pharynx or skin, where local spread can give rise to various inflammatory conditions such as pharyngitis, tonsillitis, sinusitis, or erysipelas. Although often mild and self-limiting, GAS infections are occasionally very severe and sometimes lead to life-threatening diseases, such as necrotizing fasciitis or streptococcal toxic shock syndrome. A wide variety of cell surface components and extracellular products have been shown or suggested to play important roles in S. pyogenes virulence, including cell surface pili (1, 6, 32). Pili expressed by the serotype M1 S. pyogenes strain SF370 mediate specific adhesion to intact human tonsil epithelia and to primary human keratinocytes, as well as cultured keratinocyte-derived HaCaT cells, but not to Hep-2 or A549 cells (1). They also contribute to adhesion to a human pharyngeal cell line (Detroit cells) and to biofilm formation (29).Over the past 5 years, pili have been discovered on an increasing number of important Gram-positive bacterial pathogens, including Bacillus cereus (4), Bacillus anthracis (4, 5), Corynebacterium diphtheriae (13, 14, 19, 26, 27, 44, 46, 47), Streptococcus agalactiae (7, 23, 38), and Streptococcus pneumoniae (2, 3, 24, 25, 34), as well as S. pyogenes (1, 29, 32). All these species produce pili that are composed of a single major subunit plus either one or two minor subunits. During assembly, the individual subunits are covalently linked to each other via intermolecular isopeptide bonds, catalyzed by specialized membrane-associated transpeptidases that may be described as pilin polymerases (4, 7, 25, 41, 44, 46). These are related to the classical housekeeping sortase (usually, but not always, designated SrtA) that is responsible for anchoring many proteins to Gram-positive bacterial cell walls (30, 31, 33). The C-terminal ends of sortase target proteins include a cell wall sorting (CWS) motif consisting, in most cases, of Leu-Pro-X-Thr-Gly (LPXTG, where X can be any amino acid) (11, 40). Sortases cleave this substrate between the Thr and Gly residues and produce an intermolecular isopeptide bond linking the Thr to a free amino group provided by a specific target. In attaching proteins to the cell wall, the target amino group is provided by the lipid II peptidoglycan precursor (30, 36, 40). In joining pilus subunits, the target is the ɛ-amino group in the side chain of a specific Lys residue in the second subunit (14, 18, 19). Current models of pilus biogenesis envisage repeated transpeptidation reactions adding additional subunits to the base of the growing pilus, until the terminal subunit is eventually linked covalently via an intermolecular isopeptide bond to the cell wall (28, 41, 45).The major subunit (sometimes called the backbone or shaft subunit) extends along the length of the pilus and appears to play a structural role, while minor subunits have been detected either at the tip, the base, and/or at occasional intervals along the shaft, depending on the species (4, 23, 24, 32, 47). In S. pneumoniae and S. agalactiae one of the minor subunits acts as an adhesin, while the second appears to act as a linker between the base of the assembled pilus and the cell wall (7, 15, 22, 34, 35). It was originally suggested that both minor subunits of C. diphtheriae pili could act as adhesins (27). However, recent data showed one of these has a wall linker role (26, 44) and may therefore not function as an adhesin.S. pyogenes strain SF370 pili are composed of a major (backbone) subunit, termed Spy0128, plus two minor subunits, called Spy0125 and Spy0130 (1, 32). All three are required for efficient adhesion to target cells (1). Studies employing purified recombinant proteins have shown that both of the minor subunits, but not the major subunit, bind to Detroit cells (29), suggesting both might act as pilus-presented adhesins. Here we report studies employing a combination of recombinant proteins, specific antisera, and allelic replacement mutants which show that only Spy0125 is the pilus-presented adhesin and that Spy0130 has a distinct role in linking pili to the cell wall.  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