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 93 毫秒
1.
2.
3.
4.
Vibrio parahaemolyticus O3:K6 pandemic strains recovered in Chile frequently possess a 42-kb plasmid which is the prophage of a myovirus. We studied the prototype phage VP58.5 and show that it does not integrate into the host cell chromosome but replicates as a linear plasmid (Vp58.5) with covalently closed ends (telomeres). The Vp58.5 replicon coexists with other plasmid prophages (N15, PY54, and ΦKO2) in the same cell and thus belongs to a new incompatibility group of telomere phages. We determined the complete nucleotide sequence (42,612 nucleotides) of the VP58.5 phage DNA and compared it with that of the plasmid prophage. The two molecules share the same nucleotide sequence but are 35% circularly permuted to each other. In contrast to the hairpin ends of the plasmid, VP58.5 phage DNA contains 5′-protruding ends. The VP58.5 sequence is 92% identical to the sequence of phage VHML, which was reported to integrate into the host chromosome. However, the gene order and termini of the phage DNAs are different. The VHML genome exhibits the same gene order as does the Vp58.5 plasmid. VHML phage DNA has been reported to contain terminal inverted repeats. This repetitive sequence is similar to the telomere resolution site (telRL) of VP58.5 which, after processing by the phage protelomerase, forms the hairpin ends of the Vp58.5 prophage. It is discussed why these closely related phages may be so different in terms of their genome ends and their lifestyle.Most temperate bacteriophages integrate into the host chromosome during lysogeny. However, there are some phages (telomere phages) whose prophages are linear plasmids with covalently closed ends. Members of this group of phages are the siphoviruses N15, PY54, and ΦKO2 isolated from Escherichia coli, Yersinia enterocolitica, and Klebsiella oxytoca, respectively, and the recently described myoviruses ΦHAP-1 of Halomonas aquamarina and VP882 of Vibrio parahaemolyticus (6, 20, 23, 26, 37). Despite their different origins (enterobacteria versus marine bacterium) and morphologies, all known telomere phages share similar genome organizations and some protein similarities. The linear DNA of each phage is a circular permutation of the respective linear plasmid prophage. For the generation of the terminal hairpins of the linear plasmid, the protelomerase (Tel) is essential (8). This enzyme has cleaving/joining activity; its target is a large palindromic DNA sequence called the telomere resolution site (telRL) located upstream of tel on the phage genome. After cleaving telRL by staggered cuts, the resulting self-complementary single-stranded DNA overhangs fold back and are rejoined by the protelomerase (9). Besides tel, all telomere phages possess the gene repA, encoding a multifunctional replication protein. repA of N15 and PY54 was shown to harbor the prophage replication origin and to function as a circular minimal replicon (35, 42). Compatibility studies demonstrated that the N15 and ΦKO2 plasmids belong to the same incompatibility group, whereas the PY54 plasmid is able to coexist with these two prophages in doubly lysogenic E. coli and Y. enterocolitica hosts (19).There are some reports on the presence of tel and repA in prophages (VP882, VHML, and Vp58.5) of marine Vibrio strains (28, 41). V. parahaemolyticus phage VP882 is a close relative of the Halomonas phage ΦHAP-1 (26). VHML was isolated from a toxin-producing Vibrio harveyi strain, pathogenic for some crustaceans and fish (30). Similarly to ΦHAP-1 and VP882, VHML has a myovirus-like morphology. The phage contains genes for products similar to Tel and RepA, suggesting that its prophage is a linear plasmid with terminal hairpins. However, it was surmised that VHML integrates into the Vibrio chromosome (28, 29). Phage VP58.5 was isolated from a V. parahaemolyticus strain belonging to the serovar O3:K6 pandemic clonal complex (41). During the last several years, this clone has been associated with many seafood-borne diarrhea outbreaks in Southeast Asia and South America, particularly Chile (5, 12, 13, 15). Up to 33% of the Chilean isolates harbored a 42-kb plasmid which was shown to be the prophage of a myovirus inducible by mitomycin C. VP58.5 is the prototype of these phages.In this work we demonstrate that VP58.5 is closely related to the V. harveyi phage VHML but that its prophage is a linear plasmid with covalently closed ends. The Vp58.5 prophage belongs to a new incompatibility group of telomere phages.  相似文献   

5.
Wolbachia endosymbionts are ubiquitously found in diverse insects including many medical and hygienic pests, causing a variety of reproductive phenotypes, such as cytoplasmic incompatibility, and thereby efficiently spreading in host insect populations. Recently, Wolbachia-mediated approaches to pest control and management have been proposed, but the application of these approaches has been hindered by the lack of genetic transformation techniques for symbiotic bacteria. Here, we report the genome and structure of active bacteriophages from a Wolbachia endosymbiont. From the Wolbachia strain wCauB infecting the moth Ephestia kuehniella two closely related WO prophages, WOcauB2 of 43,016 bp with 47 open reading frames (ORFs) and WOcauB3 of 45,078 bp with 46 ORFs, were characterized. In each of the prophage genomes, an integrase gene and an attachment site core sequence were identified, which are putatively involved in integration and excision of the mobile genetic elements. The 3′ region of the prophages encoded genes with sequence motifs related to bacterial virulence and protein-protein interactions, which might represent effector molecules that affect cellular processes and functions of their host bacterium and/or insect. Database searches and phylogenetic analyses revealed that the prophage genes have experienced dynamic evolutionary trajectories. Genes similar to the prophage genes were found across divergent bacterial phyla, highlighting the active and mobile nature of the genetic elements. We suggest that the active WO prophage genomes and their constituent sequence elements would provide a clue to development of a genetic transformation vector for Wolbachia endosymbionts.Members of the genus Wolbachia are endosymbiotic bacteria belonging to the Alphaproteobacteria and infecting a wide range of arthropods, including over 60% of insect species, and some filarial nematodes. They are vertically transmitted through the maternal germ line of their host and are known to distort host reproduction by causing cytoplasmic incompatibility (CI), parthenogenesis, male killing, or feminization. The ability of Wolbachia to cause these reproductive phenotypes is thought to be responsible for their efficient and rapid spread into host populations (5, 21, 35, 51).Recently, Wolbachia-mediated pest control approaches have been proposed. A number of insect pests that have important medical and hygienic consequences, such as tsetse flies and mosquitoes that vector devastating human pathogens including African sleeping disease trypanosomes, malaria plasmodia, dengue viruses, Japanese encephalitis viruses, and others, often also carry Wolbachia infections (8, 24, 25, 34). In theory, if maternally transmitted genetic elements coinherited with a CI-inducing Wolbachia, such as mitochondria, the Wolbachia itself, or other coinfecting endosymbionts, are transformed with a gene of interest (like a gene that confers resistance of the vector insect against the pathogen infection), the genetic trait is expected to be spread and fixed in the host insect population, driven by the symbiont-induced reproductive phenotype (1, 2, 10, 11, 13, 32, 43, 44). The paratransgenesis and Wolbachia-driven population replacement approaches are, although potentially promising in controlling such insect-borne diseases, still at a conceptual stage mainly because no technique has been available for Wolbachia transformation.For genetic transformation of bacteria, mobile genetic elements such as plasmids, bacteriophages, and transposons have been used successfully. For example, pUC plasmids, λ phages, and transposons have been widely utilized for transforming Escherichia coli and other model bacterial species (38). While few plasmids and transposons have been reported from Wolbachia, a family of bacteriophages, called WO phages, has been detected from a diverse array of Wolbachia strains (3, 6, 7, 12, 17, 18, 31, 39, 49). For example, in the genomes of the Wolbachia strains wMel from the fruit fly Drosophila melanogaster and wPip from the mosquito Culex quinquefasciatus, three and five WO prophages are present, respectively (26, 52). Many of the prophages are pseudogenized and inactive while some are active and capable of producing phage particles (4, 7, 15, 17, 30, 40). Such active WO phage elements may provide tools for genetic transformation of Wolbachia endosymbionts.λ phage and many other temperate bacteriophages alternate between lytic phase and lysogenic phase in their life cycles. In the lytic phase, phage particles are produced and released via host cell lysis for infection to new host cells. In the lysogenic phase, the phage genome is integrated into the host genome via a site-specific recombination process, and the integrated phage genome, called prophage, is maintained in the host genome and multiplies together with the host DNA replication (38). Upon infection and lysogenic integration of λ phage, both ends of the linear phage genomic DNA are connected by DNA ligase, and the resultant circular phage genome is inserted into the E. coli genome by site-specific recombination at a region containing a core sequence of an attachment (att) site (28). att sites on the phage genome and the bacterial genome are called attP (phage att site) and attB (bacterial att site), respectively. After integration, attP and attB are located on both ends of the prophage, called attL (left prophage att site) and attR (right prophage att site), respectively. The integration and excision processes are mediated by a site-specific recombinase, called λ integrase, encoded in the phage genome (see Fig. S1 in the supplemental material) (27, 50). Hence, the att site and the integrase are the pivotal functional elements that mediate site-specific integration and excision of λ phage. Considering the structural similarity between λ phage and WO phage (31), identification of the att site and integrase from WO phage is of interest in that these elements could be utilized for delivering foreign genes into the Wolbachia genome.In order to identify a functional att site and integrase of WO phage, the complete genome sequences of active prophage elements producing phage particles should be determined. Here, the Wolbachia strain wCauB derived from the almond moth Cadra cautella was investigated because wCauB was reported to actively produce phage particles, and a partial genome sequence of its WO phage has been determined (15). In the original host insect, C. cautella, wCauB coexists with another Wolbachia strain wCauA, and both cause CI phenotypes and produce phage particles (15, 41). Not to be confounded by the coinfecting Wolbachia strains, we used a transfected line of the Mediterranean flour moth Ephestia kuehniella infected with wCauB only, which was generated by interspecific ooplasm transfer (42). It should be noted that a mass preparation procedure for WO phage particles by centrifugation has been established for the wCauB-infected E. kuehniella (15).In this study, we determined the complete genome sequences of two active WO prophages, named WOcauB2 and WOcauB3, that are capable of producing phage particles and that are located on the genome of the Wolbachia strain wCauB. Furthermore, we identified core sequences of att sites and integrase genes of these WO phages that are putatively involved in integration of the genetic elements into the Wolbachia genome.  相似文献   

6.
7.
Listeria monocytogenes epidemic clone II (ECII) has been responsible for two multistate outbreaks in the United States in 1998-1999 and in 2002, in which contaminated ready-to-eat meat products (hot dogs and turkey deli meats, respectively) were implicated. However, ecological adaptations of ECII strains in the food-processing plant environment remain unidentified. In this study, we found that broad-host-range phages, including phages isolated from the processing plant environment, produced plaques on ECII strains grown at 37°C but not when the bacteria were grown at lower temperatures (30°C or below). ECII strains grown at lower temperatures were resistant to phage regardless of the temperature during infection and subsequent incubation. In contrast, the phage susceptibility of all other tested strains of serotype 4b (including epidemic clone I) and of strains of other serotypes and Listeria species was independent of the growth temperature of the bacteria. This temperature-dependent phage susceptibility of ECII bacteria was consistently observed with all surveyed ECII strains from outbreaks or from processing plants, regardless of the presence or absence of cadmium resistance plasmids. Phages adsorbed similarly on ECII bacteria grown at 25°C and at 37°C, suggesting that resistance of ECII strains grown at 25°C was not due to failure of the phage to adsorb. Even though the underlying mechanisms remain to be elucidated, temperature-dependent phage resistance may represent an important ecological adaptation of L. monocytogenes ECII in processed, cold-stored foods and in the processing plant environment, where relatively low temperatures prevail.Listeria monocytogenes is responsible for an estimated 2,500 cases of serious food-borne illness (listeriosis) and 500 deaths annually in the United States. It affects primarily pregnant women, newborns, the elderly, and adults with weakened immune systems. L. monocytogenes is frequently found in the environment and can grow at low temperatures, thus representing a serious hazard for cold-stored, ready-to-eat foods (18, 31).Two multistate outbreaks of listeriosis in the United States, in 1998-1999 and in 2002, respectively, were caused by contaminated ready-to-eat meats (hot dogs and turkey deli meats, respectively) contaminated by serotype 4b strains that represented a novel clonal group, designated epidemic clone II (ECII) (3, 4). ECII strains have distinct genotypes as determined by pulsed-field gel electrophoresis and various other subtyping tools, and harbor unique genetic markers (6, 8, 11, 19, 34). The genome sequencing of one of the isolates (L. monocytogenes H7858) from the 1998-1999 outbreak revealed the presence of a plasmid of ca. 80 kb (pLM80), which harbored genes mediating resistance to the heavy metal cadmium as well as genes conferring resistance to the quaternary ammonium disinfectant benzalkonium chloride (10, 29).Listeria phages (listeriaphage) have long been used for subtyping purposes (33), and extensive research has focused on the genomic characterization (2, 24, 26, 35), transducing potential (14), and biotechnological applications of selected phages (25). In addition, applications of listeriaphage as biocontrol agents in foods and the processing plant environment have been investigated (12, 15, 22). However, limited information exists on phages from processing plant environments and on the impact of environmental conditions on susceptibility of L. monocytogenes strains representing the major epidemic-associated clonal groups to such phages. We have found that strains harboring ECII-specific genetic markers can indeed be recovered from the environment of turkey-processing plants (9). Furthermore, environmental samples from such processing plants yielded phages with broad host range, which were able to infect L. monocytogenes strains of various serotypes, and different Listeria species (20). In this study, we describe the impact of growth temperature on susceptibility of L. monocytogenes ECII strains to phages, including phages isolated from turkey-processing plant environmental samples.  相似文献   

8.
9.
Virulent phage 1358 is the reference member of a rare group of phages infecting Lactococcus lactis. Electron microscopy revealed a typical icosahedral capsid connected to one of the smallest noncontractile tails found in a lactococcal phage of the Siphoviridae family. Microbiological characterization identified a burst size of 72 virions released per infected host cell and a latent period of 90 min. The host range of phage 1358 was limited to 3 out of the 60 lactococcal strains tested. Moreover, this phage was insensitive to four Abi systems (AbiK, AbiQ, AbiT, and AbiV). The genome of phage 1358 consisted of a linear, double-stranded, 36,892-bp DNA molecule containing 43 open reading frames (ORFs). At least 14 ORFs coded for structural proteins, as identified by SDS-PAGE coupled to liquid chromatography-tandem mass spectrometry (LC-MS/MS) analyses. The genomic organization was similar to those of other siphophages. All genes were on the same coding strand and in the same orientation. This lactococcal phage was unique, however, in its 51.4% GC content, much higher than those of other phages infecting this low-GC Gram-positive host. A bias for GC-rich codons was also observed. Comparative analyses showed that several phage 1358 structural proteins shared similarity with two Listeria monocytogenes phages, P35 and P40. The possible origin and evolution of lactococcal phage 1358 is discussed.The first sequenced genome of a phage infecting Lactococcus lactis (bIL67) was reported in 1994 (57). Its genomic characterization was performed with the prospect of a better understanding of lactococcal phage biology. L. lactis is a Gram-positive bacterium added to milk to produce an array of fermented dairy products. In this human-made environment, substantial amounts of lactococcal cells are cultivated on a daily basis in large fermentation vats, and these added cells randomly encounter virulent phages present in heat-treated but nonsterile milk. Moreover, it is widely acknowledged that the increased use of the same bacterial strains within existing dairy facilities inevitably leads to milk fermentation failures due to the multiplication of virulent phages. This biotechnological problem reduces yields and lowers the quality of fermented products (51).Over 700 lactococcal phage isolates have been reported in the literature (3). To date, more than 25 complete genome sequences of lactococcal phages are publicly available in the NCBI database, and the sequencing of others is under way. These numbers indicate that Lactococcus phages are among the most studied of the bacterial viruses. All lactococcal phages belong to the order Caudovirales and are included within two families according to their tail morphology: the Siphoviridae (long noncontractile tail [most lactococcal phages]) and the Podoviridae (short noncontractile tail [few lactococcal phages]) (14). Currently, phages infecting L. lactis strains have been divided into 10 genetically distinct groups (14). The complete genomic sequence is available for at least one representative of 8 of the groups.Early sequencing efforts concentrated on the genomes of lactococcal phages belonging to the 936, c2, and P335 groups (Siphoviridae), because members of these groups were regularly isolated in dairy plants (8, 36, 50). PCR-based methods were also devised to rapidly classify these phages (41). These Siphoviridae phages pose a significant risk to the dairy industry, and their characterization is important for developing adapted antiphage strategies to limit their propagation and evolution.In recent years, representatives of the less recognized lactococcal phage groups have been characterized, including phages Q54 (22), KSY1 (13), 1706 (23), asccφ28 of the P034 group (39), and P087 (63). Their molecular characterizations were aimed at understanding why some phage groups (936, c2, and P335) predominate while the others have remained marginal, at best. However, it was recently reported that P034-like phages may be emerging in certain regions (52). Genomic and microbiological analyses indicated that members of these rare phage groups were likely the result of recombination between different lactococcal phages and phages infecting other Gram-positive bacteria, and they may not be fit to multiply rapidly in milk. For example, lactococcal phage 1706 shares similarities with Ruminococcus and Clostridium prophages (23). Similarly, L. lactis phage P087 structural proteins share identity with gene products found in a prophage in the Enterococcus faecalis genome (63). It was also shown previously that lactococcal phage asccφ28 was related to Streptococcus pneumoniae phage Cp-1 and Bacillus subtilis φ29-like phages (39). It was suggested that phages 1706, asccφ28, and P087 acquired a receptor-binding protein complex from another lactococcal phage that enabled them to infect a L. lactis host.Here, we report the complete genome sequence and analysis of phage 1358, a virulent representative of the 9th lactococcal phage group.  相似文献   

10.
We report here the characterization of the nonstructural protein ORF12 of the virulent lactococcal phage p2, which belongs to the Siphoviridae family. ORF12 was produced as a soluble protein, which forms large oligomers (6- to 15-mers) in solution. Using anti-ORF12 antibodies, we have confirmed that ORF12 is not found in the virion structure but is detected in the second half of the lytic cycle, indicating that it is a late-expressed protein. The structure of ORF12, solved by single anomalous diffraction and refined at 2.9-Å resolution, revealed a previously unknown fold as well as the presence of a hydrophobic patch at its surface. Furthermore, crystal packing of ORF12 formed long spirals in which a hydrophobic, continuous crevice was identified. This crevice exhibited a repeated motif of aromatic residues, which coincided with the same repeated motif usually found in tape measure protein (TMP), predicted to form helices. A model of a complex between ORF12 and a repeated motif of the TMP of phage p2 (ORF14) was generated, in which the TMP helix fitted exquisitely in the crevice and the aromatic patches of ORF12. We suggest, therefore, that ORF12 might act as a chaperone for TMP hydrophobic repeats, maintaining TMP in solution during the tail assembly of the lactococcal siphophage p2.During industrial milk fermentation, Lactococcus lactis cells are added to transform milk into an array of fermented products such as cheese. However, this manufacturing process may be impaired by lytic phages present in the factory environment as well as in the milk itself (30). Due to the destructive effects of phage infections on bacterial fermentation, much effort has been undertaken to isolate and study the biodiversity of these bacteriophages. Lactococcal bacteriophages belong to at least 10 different genetically distinct species of double-stranded DNA viruses (9). Of them, three lactococcal phage species, all belonging to the Siphoviridae family, are the major source of problems in milk fermentation, namely, the 936, P335, and c2 species (7, 28, 29). Furthermore, members of the 936 species are by far responsible for the majority of infections (50 to 80%) (1, 24, 41). Numerous phages of the 936 species have been isolated, and several have been characterized at the genome level (25). However, little is known concerning their molecular mechanisms of infection, although we recently solved the structure of the receptor-binding protein (RBP) of our model 936-like phage, namely, the virulent phage p2 (38, 43), and of phages belonging to the P335 species (27, 34, 37, 38).As with all viruses, bacteriophage genomes are quite compact, leaving little room for noncoding sequences (4). In fact, phage genes are disposed in an operon-type organization (4), and the order of genes corresponds to the different phases of the infection cycle. Moreover, genes are often in clusters (referred to as modules), with gene products from adjacent genes generally found to interact with each other. Interestingly, phage genome organization, including individual gene order, is often conserved within a given species, particularly within the Siphoviridae family. In the case of L. lactis virulent phages belonging to the 936 or P335 species, this principle applies particularly to the morphogenesis gene module, which includes all the genes coding for the phage structural protein genes. For the tail assembly, a module comprises a set of genes between the portal protein, which is connecting the tail to the capsid, and the RBP, which is located at the tip of the tail and is involved in host recognition (39, 43).The characterization of tail assembly genes of lactococcal phages has been more extensive for temperate siphophages belonging to the P335 species (27, 34, 37, 38). Because of the similarities in genome organization, the findings in this phage species can, in some cases, be used as clues toward understanding the morphology of 936-like phages. For the temperate phage Tuc2009 (P335 species), all structural proteins required for tail and baseplate assembly have been identified (27, 34, 37, 38). Genes located between those coding for the tape measure protein (TMP) and BppL (RBP) were identified as corresponding to components of the baseplate structure, located at the tail distal end. Furthermore, a gene coding for the major tail protein (MTP) was also identified at a position upstream from tmp. Between the genes coding for the MTP and those coding for the TMP in Tuc2009 are two gene products identified as gpG and gpGT, which are not present in the phage particle. These two proteins were named based on their likely role analogous to the tail assembly proteins present in coliphage lambda, a model virus belonging to the Siphoviridae family (21, 27, 47). gpGT has an essential role in lambda tail assembly, acting prior to tail shaft assembly, while the role of gpG in tail assembly is not known (21). Both gpG and gpGT are also absent from mature lambda virions (21). It has been argued that they may act as assembly chaperones (47).A close examination of 936 genomes indicates the presence of two genes coding for gpG and gpGT-like proteins. Analysis of the phage p2 genome, closely related to that of lactococcal phage sk1 (6), revealed that the putative tail assembly proteins could correspond to gene products ORF12 and ORF13. These two genes are followed by the TMP gene corresponding to orf14, other genes coding for other structural proteins, and the RBP gene orf18. During our ongoing investigation of the structure of phage p2, we report here the cloning, expression, and crystal structure of ORF12 in order to decipher its role in the tail assembly process.  相似文献   

11.
12.
A collection of 54 clinical and agricultural isolates of Burkholderia cenocepacia was analyzed for genetic relatedness by using multilocus sequence typing (MLST), pathogenicity by using onion and nematode infection models, antifungal activity, and the distribution of three marker genes associated with virulence. The majority of clinical isolates were obtained from cystic fibrosis (CF) patients in Michigan, and the agricultural isolates were predominantly from Michigan onion fields. MLST analysis resolved 23 distinct sequence types (STs), 11 of which were novel. Twenty-six of 27 clinical isolates from Michigan were genotyped as ST-40, previously identified as the Midwest B. cenocepacia lineage. In contrast, the 12 agricultural isolates represented eight STs, including ST-122, that were identical to clinical isolates of the PHDC lineage. In general, pathogenicity to onions and the presence of the pehA endopolygalacturonase gene were detected only in one cluster of related strains consisting of agricultural isolates and the PHDC lineage. Surprisingly, these strains were highly pathogenic in the nematode Caenorhabditis elegans infection model, killing nematodes faster than the CF pathogen Pseudomonas aeruginosa PA14 on slow-kill medium. The other strains displayed a wide range of pathogenicity to C. elegans, notably the Midwest clonal lineage which displayed high, moderate, and low virulence. Most strains displayed moderate antifungal activity, although strains with high and low activities were also detected. We conclude that pathogenicity to multiple hosts may be a key factor contributing to the potential of B. cenocepacia to opportunistically infect humans both by increasing the prevalence of the organism in the environment, thereby increasing exposure to vulnerable hosts, and by the selection of virulence factors that function in multiple hosts.The betaproteobacterium Burkholderia cenocepacia, 1 of now 17 classified species belonging to the Burkholderia cepacia complex (BCC), is ubiquitous and extremely versatile in its metabolic capabilities and interactions with other organisms (38, 40, 57, 58). Strains of B. cenocepacia are pathogens of onion and banana plants, opportunistic pathogens of humans, symbionts of numerous plant rhizospheres, contaminants of pharmaceutical and industrial products, and inhabitants of soil and surface waters (14, 29, 33, 34, 37, 45). Originally described as a pathogen of onions (8), organisms of the BCC emerged in the past 3 decades as serious human pathogens, capable of causing devastating chronic lung infections in persons with cystic fibrosis (CF) or chronic granulomatous disease (21, 24, 28). Infections due to BCC are a serious concern to CF patients due to their inherent antibiotic resistance and high potential for patient-to-patient transmission (23). Although 16 of the BCC species have been recovered from respiratory secretions of CF patients in many countries (46, 58), B. cenocepacia has been the most common species isolated in North America, detected in 50% of 606, 83% of 447, and 45.6% of 1,218 patients in recent studies (35, 46, 52).The epidemiology of infectious disease caused by B. cenocepacia appears to involve patient-to-patient spread of genetically distinct lineages. B. cenocepacia lineages, such as ET12, Midwest, and PHDC, have been identified from large numbers of individuals in disease outbreaks in North America and Europe (11, 32, 54). A recently developed multilocus sequence typing (MLST) scheme has been shown to be a reliable epidemiologic tool for differentiating between the five subgroups (IIIA to IIIE) of B. cenocepacia, and strains representing three of these subgroups (IIIA, IIIB, and IIID) have been recovered from CF patients (2). Outside of the patient-to-patient transmission of clonal lineages, the mode of acquisition of strains causing sporadic cases of B. cenocepacia in CF patients remains unclear, although environmental sources are a logical reservoir for infection. Previously, an isolate of B. cenocepacia indistinguishable from the PHDC epidemic clonal lineage by using standard typing methods (e.g., repetitive-sequence-based PCR, randomly amplified polymorphic DNA, pulsed-field gel electrophoresis) was detected in an agricultural soil sample (34). Similarly, three distinct MLST sequence types containing both clinical and environmental (plant and soil) B. cenocepacia isolates were identified (1). These findings suggest that natural populations of B. cenocepacia in soil or associated with plants are a potential reservoir for the emergence of new human pathogenic lineages.Experimental models for the study of virulence potential and traits of B. cenocepacia include mouse and rat models with genetic defects allowing chronic lung infections to be established (e.g., see reference 48). Nematode (Caenorhabditis elegans), alfalfa (Medicago sativa), and onion (Allium cepa) models have also been routinely utilized for the identification of virulence factors (5, 29, 31). C. elegans has been extensively used to study the pathogenesis and virulence factors of a wide variety of bacterial and fungal pathogens (9, 15, 42, 51, 56). In several pathogens, including Pseudomonas (56) and Burkholderia (20), putative virulence factors important for the pathogenesis in mammalian systems (15, 51) have been identified using the C. elegans model. The C. elegans model might be limited in the detection of host-specific virulence factors; however, several attributes, such as small size and rapid development, make it an excellent whole animal model for pathogenesis research (16, 51).The evidence that individual strains of B. cenocepacia can be pathogenic to both plants and humans and are prevalent in various environmental niches has provoked particular interest in elucidating the clinical pathogenic potential of environmental isolates. The basis of this study was to examine whether genetically related B. cenocepacia strains exhibit shared characteristics that contribute to their pathogenicity in multiple hosts and to examine the potential for circulating environmental isolates to emerge as new clinical pathogens. Here, we tested the degree of virulence in animal (nematode) and plant (onion) infection models, the production of antifungal activity, and the genetic relatedness of clinical and environmental B. cenocepacia subgroup IIIB strains predominantly isolated from Michigan.  相似文献   

13.
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   

14.
A number of bacteriophages have been identified that target the Vi capsular antigen of Salmonella enterica serovar Typhi. Here we show that these Vi phages represent a remarkably diverse set of phages belonging to three phage families, including Podoviridae and Myoviridae. Genome analysis facilitated the further classification of these phages and highlighted aspects of their independent evolution. Significantly, a conserved protein domain carrying an acetyl esterase was found to be associated with at least one tail fiber gene for all Vi phages, and the presence of this domain was confirmed in representative phage particles by mass spectrometric analysis. Thus, we provide a simple explanation and paradigm of how a diverse group of phages target a single key virulence antigen associated with this important human-restricted pathogen.Bacteriophages are dependent for their survival on the presence of susceptible host bacteria in their environment. The first stage of recognition of the bacterial host normally involves binding of a specific phage attachment protein to a receptor molecule on the bacterial surface. Bacteria can evade phage infection by various mechanisms, including accumulating escape mutations in the receptor, acquiring phage inhibitory proteins, or directly modifying the receptor, for example, lipopolysaccharide (LPS) (43). In addition, phage can adapt to recognize different receptors through a number of genetic mechanisms involving evolution of their attachment proteins (20) or by tropism switching (21, 22).Phage can exploit capsular exopolysaccharides as receptors, some of which are associated with virulence in pathogens (5, 23, 35). A notable example is the Vi capsule found in Salmonella enterica serovar Typhi (S. Typhi) and some isolates of S. Dublin and Citrobacter freundii (29). The Vi capsule of S. Typhi is an important virulence factor, facilitating the bacteria to escape opsonization and other forms of immune surveillance (14, 30) as well as potentially helping the bacteria to evade phage that would otherwise target the O:9 LPS, which the Vi capsule can, at least in part, mask (27). In the middle of the last century, a set of lytic phages were isolated that utilized the Vi capsule as a receptor (6). These Vi phages were exploited in diagnostic laboratories as a “typing set” to distinguish between different strains of S. Typhi isolated from typhoid patients (8). A secondary typing set was generated from Vi typing phage II by adapting this phage to grow on different S. Typhi hosts (6). At this time, typhoid was still common in many parts of Europe and North America, and clinicians tested some of these Vi phages for their potential in phage therapy experiments with human typhoid patients (11). Although this work showed significant promise, phage therapy gradually disappeared from clinical practice in many countries as antibiotics became readily available.S. Typhi is a monophyletic serovar of the broad enteric species S. enterica (16, 31). Interestingly, S. Typhi is host restricted to humans and has no known zoonotic source. Unlike many other S. enterica serovars, S. Typhi normally causes a systemic infection and does not persist in the intestine efficiently, where high levels of bacteriophage are present. Although it is rare in developed countries, S. Typhi is still a significant cause of mortality in many developing countries (26). Most current clinical isolates are Vi positive when first isolated (2), but it is noteworthy that the Vi capsule biosynthesis and export genes are carried by an operon within a potentially unstable island called Salmonella pathogenicity island 7 (SPI-7) (29).Although some phenotypic characterization of the Vi phage has been undertaken (1), very little has been performed at the molecular level. We previously showed that Vi typing phage II-E1 is related to the S. Typhimurium phage ES18 (4, 28), with synteny in many capsid and tail proteins. We have now further characterized the other members of this S. Typhi Vi phage collection, designated types I, III, IV, V, VI, and VII (abbreviated from here on as Vi phages I, III, IV, etc.) (6, 11), by utilizing electron microscopy and genomic analysis. This analysis shows that this collection of Vi phages represents a diverse group of bacteriophages that have adapted to growth on S. Typhi through convergent evolution within their tail spike protein genes and the acquisition of conserved acetyl esterase domains.  相似文献   

15.
Fourier transform infrared (FT-IR) spectroscopy and chemometric techniques were used to discriminate five closely related Salmonella enterica serotype Enteritidis phage types, phage type 1 (PT1), PT1b, PT4b, PT6, and PT6a. Intact cells and outer membrane protein (OMP) extracts from bacterial cell membranes were subjected to FT-IR analysis in transmittance mode. Spectra were collected over a wavenumber range from 4,000 to 600 cm−1. Partial least-squares discriminant analysis (PLS-DA) was used to develop calibration models based on preprocessed FT-IR spectra. The analysis based on OMP extracts provided greater separation between the Salmonella Enteritidis PT1-PT1b, PT4b, and PT6-PT6a groups than the intact cell analysis. When these three phage type groups were considered, the method based on OMP extract FT-IR spectra was 100% accurate. Moreover, complementary local models that considered only the PT1-PT1b and PT6-PT6a groups were developed, and the level of discrimination increased. PT1 and PT1b isolates were differentiated successfully with the local model using the entire OMP extract spectrum (98.3% correct predictions), whereas the accuracy of discrimination between PT6 and PT6a isolates was 86.0%. Isolates belonging to different phage types (PT19, PT20, and PT21) were used with the model to test its robustness. For the first time it was demonstrated that FT-IR analysis of OMP extracts can be used for construction of robust models that allow fast and accurate discrimination of different Salmonella Enteritidis phage types.Over the past 10 years there has been an increase in the incidence of gastrointestinal infections caused by Salmonella enterica serovar Enteritidis, which is now one of the leading S. enterica serotypes worldwide (21, 27). Poultry, poultry products, cattle, and dairy products are the predominant sources of Salmonella-contaminated food products that cause human salmonellosis (28). Large-scale infections continue to occur in developed countries (8). Unrestricted international movement of commercially prepared food and food ingredients and dissimilarities in government and industry food safety controls during the processing, distribution, and marketing of products have surely contributed to the increase in food-borne outbreaks. Salmonella is a tremendous challenge for the agricultural and food processing industries because of its ability to survive under adverse conditions, such as low levels of nutrients and suboptimal temperatures (4, 13).Salmonella Enteritidis isolates can be categorized for epidemiological purposes by using a variety of typing tools (13). These tools include typing techniques such as serological and phage typing (29) and antibiotic resistance patterns (25). These methods are now supplemented by molecular genetics techniques, such as DNA fingerprinting (23), plasmid profiling (16), and pulsed-field gel electrophoresis (26). Phage typing has been used to diagnose Salmonella outbreaks, including S. enterica serovar Typhi and S. enterica serovar Typhimurium outbreaks (29). It is useful to evaluate whether isolates obtained from different sources at different times are similar or distinct in terms of their reactions with a specific collection of bacteriophages used for typing. The correlation between phage type and the source of an epidemic is high (22). Although very effective, existing classification methods are time-consuming, laborious, and expensive, and they often require special training of personnel and expertise, which can prevent a rapid response to the presence of pathogenic bacterial species.Fourier transform infrared (FT-IR) spectroscopy has been successfully used for differentiation and classification of microorganisms at the species and subspecies levels (7, 9, 12, 15, 18, 19, 20). This technique has been shown to have high discriminatory power and allows identification of bacteria at distinct taxonomic levels based on differences in the infrared absorption patterns of microbial cells. FT-IR spectroscopy has been used to differentiate and characterize intact microbial cells based on outer membrane cell components, including lipopolysaccharides (LPS), lipoproteins, and phospholipids (24). Several studies in which S. enterica serotypes have been discriminated using multivariate data analysis and FT-IR spectroscopy have been performed (1, 2, 10, 11). Kim et al. (11) compared the FT-IR spectra of intact cells and the FT-IR spectra of outer membrane protein (OMP) extracts from S. enterica serotypes to discriminate serotypes. Analysis of spectra of OMP extracts in the 1,800- to 1,500-cm−1 region resulted in 100% correct classification of the serotypes investigated.Previously, there have been no reports of differentiation of Salmonella Enteritidis phage types by FT-IR spectroscopy and chemometric methods. To discriminate closely related phage types of Salmonella Enteritidis in this study, intact cells and OMP extracts of bacterial cell membranes were subjected to FT-IR analysis. The isolates analyzed included isolates belonging to five of the phage types of Salmonella Enteritidis found most frequently in Portuguese hospitals in the period from 2004 to 2006, phage type 1 (PT1), PT1b, PT4b, PT6, and PT6a (5, 14). Chemometric models were used to discriminate between phage types based on infrared spectra.  相似文献   

16.
Toxigenic Vibrio cholerae, the causative agent of the epidemic diarrheal disease cholera, interacts with diverse environmental bacteriophages. These interactions promote genetic diversity or cause selective enrichment of phage-resistant bacterial clones. To identify bacterial genes involved in mediating the phage-resistant phenotype, we screened a transposon insertion library of V. cholerae O1 El Tor biotype strain C6706 to identify mutants showing altered susceptibility to a panel of phages isolated from surface waters in Bangladesh. Mutants with insertion in cyaA or crp genes encoding adenylate cyclase or cyclic AMP (cAMP) receptor protein (CRP), respectively, were susceptible to a phage designated JSF9 to which the parent strain was completely resistant. Application of the cyaA mutant as an indicator strain in environmental phage monitoring enhanced phage detection, and we identified 3 additional phages to which the parent strain was resistant. Incorporation of the cyaA or crp mutations into other V. cholerae O1 strains caused similar alterations in their phage susceptibility patterns, and the susceptibility correlated with the ability of the bacteria to adsorb these phages. Our results suggest that cAMP-CRP-mediated downregulation of phage adsorption may contribute to a mechanism for the V. cholerae O1 strains to survive predation by multiple environmental phages. Furthermore, the cyaA or crp mutant strains may be used as suitable indicators in monitoring cholera phages in the water.Bacteriophages contribute to the evolution of bacteria by mediating horizontal gene transfer and genomic rearrangements, as well as by bactericidal selection, in which bacterial strains that are able to resist phage predation thrive over competing phage-susceptible strains (5, 10, 11). Toxigenic Vibrio cholerae, the causative agent of the epidemic diarrheal disease cholera, interacts with diverse phages, both in the aquatic environment and in the host milieu, and these interactions may promote genetic diversity and/or cause selective enrichment of particular bacterial clones (10, 11, 26, 27).Historically, cholera is an ancient disease with the occurrence of seven distinct pandemics since the first pandemic of cholera began in 1817, but the disease still affects millions of people (9, 16). The current seventh pandemic of cholera, which originated in Indonesia in 1961, is the most extensive in geographic spread and duration, and the causative agent is V. cholerae O1 of the El Tor biotype. The sixth pandemic and presumably the earlier pandemics were caused by the classical biotype, which now seems to be extinct.Molecular epidemiological surveillance has revealed continually changing relative prevalences of different clones of pathogenic V. cholerae (9), and the emergence of new clones has been attributed to possible horizontal transfer of clusters of genes associated with virulence or environmental fitness as well as resistance to different antibiotics (9, 20). The recent recognition that phage predation may play a role in the natural control of cholera epidemics (10, 11, 14) reinforces predictions that changes in this pathogen and the prevalences of different clones may also be driven by environmental phages. The emergence of certain strains is likely to be enhanced by phages through the bactericidal mechanism in which phage-sensitive strains are killed while providing a selective advantage to phage-resistant strains. Therefore, the ability to evade phage predation constitutes an important factor in attaining increased evolutionary fitness.In the present study we screened a transposon insertion library of V. cholerae O1 El Tor biotype strain C6706, to identify genes whose inactivation would enhance the susceptibility of the bacteria to environmental phages. Presumably, these genes contribute in mediating resistance to the relevant phages and thus allow the bacteria to survive phage predation. Bacteria with increased phage susceptibility due to mutations in the appropriate genes may also have application as improved indicator strains to monitor the prevalence of relevant phages in the environment.  相似文献   

17.
Bacteria are constantly challenged by bacteriophage (phage) infection and have developed multiple adaptive resistance mechanisms. These mechanisms include the abortive infection systems, which promote “altruistic suicide” of an infected cell, protecting the clonal population. A cryptic plasmid of Erwinia carotovora subsp. atroseptica, pECA1039, has been shown to encode an abortive infection system. This highly effective system is active across multiple genera of gram-negative bacteria and against a spectrum of phages. Designated ToxIN, this two-component abortive infection system acts as a toxin-antitoxin module. ToxIN is the first member of a new type III class of protein-RNA toxin-antitoxin modules, of which there are multiple homologues cross-genera. We characterized in more detail the abortive infection phenotype of ToxIN using a suite of Erwinia phages and performed mutagenesis of the ToxI and ToxN components. We determined the minimal ToxI RNA sequence in the native operon that is both necessary and sufficient for abortive infection and to counteract the toxicity of ToxN. Furthermore, site-directed mutagenesis of ToxN revealed key conserved amino acids in this defining member of the new group of toxic proteins. The mechanism of phage activation of the ToxIN system was investigated and was shown to have no effect on the levels of the ToxN protein. Finally, evidence of negative autoregulation of the toxIN operon, a common feature of toxin-antitoxin systems, is presented. This work on the components of the ToxIN system suggests that there is very tight toxin regulation prior to suicide activation by incoming phage.Interactions between bacteria and their natural parasites, bacteriophages (phage), have global-scale effects (42). Although the vast majority of the phage infections, which occur at a rate of 1025 infections per s (26), are overlooked by humans, en masse they affect environmental nutrient cycling (18) and have long been known to be vital to the spread and continued diversity of microbial genes (11). A tiny proportion of this activity can directly affect our everyday activities; the lysis of bacteria following phage infection has potential medical benefits, such as use in phage therapy (30), or can be economically damaging, as it is in cases of bacterial fermentation failure (for instance, in the dairy industry [31]).Gram-positive lactococcal strains used in dairy fermentation have been shown to naturally harbor multiple phage resistance mechanisms (16). These mechanisms can be broadly classed as systems which (i) prevent phage adsorption, (ii) interfere with phage DNA injection, (iii) restrict unmodified DNA, and (iv) induce abortive infection. There is also an increasing amount of research that focuses on new systems that use clustered regularly interspaced short palindromic repeats to mediate phage resistance (3). Clustered regularly interspaced short palindromic repeats and associated proteins, although widespread in archaea and bacteria (39), have not been identified yet in lactococcal strains (23).The abortive infection (Abi) systems induce cell death upon phage infection and often rely on a toxic protein to cause “altruistic cell suicide” in the infected host (16). Although Abi systems have been studied predominantly using lactococcal systems, because of their potential economic importance (8) they have been identified in some gram-negative species, such as Escherichia coli, Vibrio cholerae, Shigella dysenteriae, and Erwinia carotovora (9, 14, 36, 38). The prr and lit systems of E. coli have been studied at the molecular level, and their mode of action and mode of activation by incoming phage have been identified (2, 37, 38). In contrast, lactococcal Abi systems have been characterized mainly by the range of phages actively aborted and the scale of these effects, and the Abi systems have been grouped based on general modes of action (8, 12). More recently, research has begun to identify more specific lactococcal Abi activities at the molecular level (12, 17) and has revealed phage activation of two such Abi systems (6, 21).An Abi system was identified on plasmid pECA1039, which was isolated from a strain of the phytopathogen E. carotovora subsp. atroseptica (14). Designated ToxIN, this two-component Abi system operates as a novel protein-RNA toxin-antitoxin (TA) system to abort phage infection in multiple gram-negative bacteria. The toxic activity of the ToxN protein was inhibited by ToxI RNA, which consists of 5.5 direct repeats of 36 nucleotides. It is now recognized that TA loci, which were originally characterized as “plasmid addiction” modules (43), are widely distributed in the chromosomes of archaea and bacteria (19) and in phage genomes, such as that of the extrachromosomal prophage P1 (27). As a result, the precise biological role of TA systems is under debate (29). It is clear, however, that they can be effective phage resistance systems, as is the case for toxIN in E. carotovora subsp. atroseptica (14) and hok/sok and mazEF in E. coli (22, 33). Previously characterized TA systems operate with both components interacting as either RNAs (e.g., hok/sok) (type I) or proteins (e.g., MazE and MazF) (type II). In this study, a mutagenesis approach was used to further characterize the ToxI and ToxN components of the new (type III) protein-RNA TA Abi system. The regulation of the operon and the mode of phage activation were also examined.  相似文献   

18.
19.
Most bacteriophages (phages) release their progeny through the action of holins that form lesions in the cytoplasmic membrane and lysins that degrade the bacterial peptidoglycan. Although the function of each protein is well established in phages infecting Streptococcus pneumoniae, the role—if any—of the powerful bacterial autolysin LytA in virion release is currently unknown. In this study, deletions of the bacterial and phage lysins were done in lysogenic S. pneumoniae strains, allowing the evaluation of the contribution of each lytic enzyme to phage release through the monitoring of bacterial-culture lysis and phage plaque assays. In addition, we assessed membrane integrity during phage-mediated lysis using flow cytometry to evaluate the regulatory role of holins over the lytic activities. Our data show that LytA is activated at the end of the lytic cycle and that its triggering results from holin-induced membrane permeabilization. In the absence of phage lysin, LytA is able to mediate bacterial lysis and phage release, although exclusive dependence on the autolysin results in reduced virion egress and altered kinetics that may impair phage fitness. Under normal conditions, activation of bacterial LytA, together with the phage lysin, leads to greater phage progeny release. Our findings demonstrate that S. pneumoniae phages use the ubiquitous host autolysin to accomplish an optimal phage exiting strategy.Streptococcus pneumoniae (pneumococcus), a common and important human pathogen, is characterized by the high incidence of lysogeny in isolates associated with infection (34, 44). Pneumococcal bacteriophages (phages) share with the majority of bacteriophages infecting other bacterial species the “holin-lysin” system to lyse the host cell and release their progeny at the end of the lytic cycle. Genes encoding both holins and lysins (historically termed “endolysins”) are indeed found in the genomes of all known pneumococcal phages (8, 28, 31, 37). Supporting this mechanism, a lytic phenotype in the heterologous Escherichia coli system was achieved only by the simultaneous expression of the Ejh holin and the Ejl endolysin of pneumococcal phage EJ-1 (8). When these proteins were independently expressed, cellular lysis was not perceived. Similar results were shown for pneumococcal phage Cp-1, not only in E. coli, but also in the pneumococcus itself (28).Phage lysins destroy the pneumococcal peptidoglycan network due to their muralytic activity, whereas holins have been shown in S. pneumoniae to form nonspecific lesions (8), most likely upon a process of oligomerization in the cytoplasmic membrane, as observed for the E. coli phage λ (13, 14, 43). It was generally proposed that holin lesions allow access of phage lysins to the cell wall (52, 54), as the majority of phage lysins, including the pneumococcal endolysins, lack a typical N-terminal secretory signal sequence and transmembrane domains (8). However, recent evidence also highlights the possibility for a holin-independent targeting of phage lysins to the cell wall, where holin lesions seem to be crucial for the activation of the already externalized phage lysins (42, 50, 51). Regardless of the mechanism operating in S. pneumoniae to activate phage lysins, holin activity compromises membrane integrity.Pneumococcal cells present their own autolytic activity, mainly due to the presence of a powerful bacterial cell wall hydrolase, LytA (an N-acetylmuramoyl-l-alanine-amidase), responsible for bacterial lysis under certain physiological conditions (47). Although other bacterial species also encode peptidoglycan hydrolases, the extensive lysis shortly after entering stationary phase caused by LytA is a unique feature of S. pneumoniae. Interestingly, LytA is translocated across the cytoplasmic membrane to the cell wall—where it remains inactive—in spite of the absence of a canonical N-terminal sequence signal (7). In the cell wall, autolysin activities are tightly regulated by mechanisms that seem to be related to the energized state of the cell membrane. In fact, depolarizing agents are able to trigger autolysis in Bacillus subtilis (16, 17), and bacteriocin-induced depletion of membrane potential triggers autolysis of some species of the genera Lactococcus and Lactobacillus, closely related to streptococci (29). It is therefore possible that the holin-inflicted perturbations of the S. pneumoniae cytoplasmic membrane upon the induction of the lytic cycle may trigger not only the lytic activity of the phage lysin, but also that of inactive LytA located in the cell wall. Accordingly, LytA could also participate in the release of phage particles at the end of the infectious cycle, especially considering its powerful autolytic activity. Previous studies have suggested a role for the host autolytic enzyme in the release of phage progeny (11, 38), but in fact, the evidence is unclear and dubious, considering that the existence of phage-encoded lysins was unknown or very poorly understood and some of the experimental conditions used to show a role of LytA could have also affected the activity of the phage lysin (38).To clarify the possible role of the bacterial autolysin in host lysis, we used the S. pneumoniae strain SVMC28, lysogenic for the SV1 prophage (34), which contains a typical “holin-lysin” cassette, and a different host strain lysogenized with the same SV1 phage. Our results show that LytA is activated by the holin-induced membrane disruption, just like the phage endolysin. In the absence of the endolysin, LytA is capable of mediating host lysis, releasing functional phage particles able to complete their life cycle. Still, sole dependence on LytA results in an altered pattern of phage release that may reduce phage fitness. Importantly, we also show that, together with the endolysin, the concurrent LytA activation is critical for optimal phage progeny release.  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