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.
2.
3.
4.
5.
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   

6.
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   

7.
Lactococcus lactis, a gram-positive bacterium widely used by the dairy industry to manufacture cheeses, is subject to infection by a diverse population of virulent phages. We have previously determined the structures of three receptor binding proteins (RBPs) from lactococcal phages TP901-1, p2, and bIL170, each of them having a distinct host range. Virulent phages p2 and bIL170 are classified within the 936 group, while the temperate phage TP901-1 is a member of the genetically distinct P335 polythetic group. These RBPs comprise three domains: the N-terminal domain, binding to the virion particle; a β-helical linker domain; and the C-terminal domain, bearing the receptor binding site used for host recognition. Here, we have designed, expressed, and determined the structure of an RBP chimera in which the N-terminal and linker RBP domains of phage TP901-1 (P335) are fused to the C-terminal RBP domain of phage p2 (936). This chimera exhibits a stable structure that closely resembles the parental structures, while a slight displacement of the linker made RBP domain adaptation efficient. The receptor binding site is structurally indistinguishable from that of native p2 RBP and binds glycerol with excellent affinity.A broad number of products are manufactured by large-scale bacterial fermentation, including the value-added fermented dairy products. Most bacterial fermentation industries have experienced problems with phage contamination. Phage outbreaks are costly and time-consuming because they can slow or arrest the fermentation process and adversely affect product quality (15). For decades, the dairy industry has relied on an array of strategies to control this natural phenomenon, including rotation of their bacterial cultures (11, 24, 25). However, in spite of these efforts, new virulent lactococcal phages keep emerging. A better understanding of the various mechanisms affecting the genetic diversity of the phage population is necessary for optimal phage control strategies (18).Lactococcal phages are among the most studied bacterial viruses because of the economic importance of their hosts. Hundreds of lactococcal phages have been isolated, and the vast majority of them have a long, contractile tail, thereby belonging to the Siphoviridae family (1). Lactococcus lactis phages are currently classified into 10 genetically distinct groups (10), but only members of 3 of them are highly adapted to multiply in milk, namely, the 936, c2, and P335 groups (11, 24, 25). The first step for such an effective viral infection is host recognition, which necessitates the interaction between the adsorption device located at the distal tail end of the phage and the cell surface receptor (32). Members of the 936 and P335 groups recognize their host through an interaction between their receptor binding protein (RBP) (13) and receptors, probably lipoteichoic acids, at the host cell surface (27, 29-31).We have previously determined the crystal structures of three RBPs, from the virulent lactococcal phages p2 (30, 31) and bIL170 (936 group) (27) and from the temperate phage TP901-1 (P335 group) (29). The RBPs of these phages have a similar architecture of three protomers related by a threefold axis. Each protomer comprises three domains: the N terminus (named shoulders in p2), the interlaced β-prism linker (the “neck” domain), and the jelly-roll domain (2) at the C terminus (the “head” domain). This last domain harbors a saccharide binding site likely involved in host recognition, as it binds with high affinity to phosphoglycerol, a component of teichoic acid (8, 19, 27, 29-31). We have previously shown that the shoulder and neck domains are highly conserved in the RBPs of 936-like phages (8, 19, 27, 29-31). The individuality of the RBP C-terminal domain sequence likely dictates phage specificity for the receptor, which may specifically recognize different substitutions (H, GlcNAc, or d-Ala) of the phosphoglycerol moieties of the L. lactis teichoic acid polymers. Recently, the complete genomic sequence of the reference virulent phage P335 was determined, and comparative analysis revealed that the C terminus of its RBP showed homology to the RBP of the virulent lactococcal phage P475 of the 936 group (17). Such homology between RBP head domains was surprising because the two lactococcal phage groups rarely shared common genes or domains. This observation suggested that modular shuffling of domains can occur between these otherwise genetically distinct phage groups.The overall fold of the N-terminal RBP domain is different in 936- and P335-like phages. In the P335 group, the N-terminal domain comprises a unique helix that fits into the rest of the phage baseplate (28, 29) (Fig. (Fig.1A),1A), while in the 936 group, this 140-residue domain is a large β-sandwich with an external α-helix (30) (Fig. (Fig.1B).1B). Nonetheless, the N-terminal domains of the two RBPs may still be, related because both appear to be built using a coiled coil, although the 936-like phages have an additional β-sandwich. The β-prism linkers (neck domain) of the two phage groups also differ in sequence and in radius, but they have a similar fold, the latter being also close to that of T4 phage short fiber (33). The linker domain of phage TP901-1 is wider than that of p2 and exhibits a repeated motif (G-X-Y-X-Y, where X is polar and Y nonpolar). Finally, the C-terminal domains of both species share the same fold, a jelly-roll motif (2) also found in adenovirus (5) and reovirus (3, 4, 6).Open in a separate windowFIG. 1.Structures and sequences of RBPs from lactococcal phages. (A) Three-dimensional structure of the RBP from phage TP901-1 (P335 group; blue). (B) Three-dimensional structure of the RBP from phage p2 (936 group; magenta). (C) View of a model associating domains of TP901-1 (N terminus and linker domain, below red line, blue) and p2 (head, above red line, magenta) RBPs. (D) Three-dimensional crystal structure of chimera form 1 (yellow) assembled according to the model in panel C. (E) Sequence alignment of the RBPs of p2 (part) and TP901-1. The secondary structure is described above the alignment. The binding residues are shown with blue dots. The hinge proline (Pro 162/63) is identified by a red arrow. The chimera is composed of the N-terminal domain (residues 17 to 33) and the linker domain residues (residues 34 to 63) from phage TP901-1 RBP and the C-terminal domain (residues 163 to 264) from phage p2 RBP.The question addressed here was whether exchange between the C-terminal domains of two phage groups would lead to a stable protein with conserved binding capacity. To answer this question, we have generated an RBP chimera comprising the N-terminal and linker domains of phage TP901-1 fused to the C-terminal domain of phage p2. We have produced this chimera and determined its crystal structure and its sugar binding capacity. These results indicate that straightforward domain exchange produced a stable chimera with a conserved binding capacity and a structure close to that of each of the parental parts.  相似文献   

8.
We have sequenced the double-stranded DNA genomes of six lactococcal phages (SL4, CB13, CB14, CB19, CB20, and GR7) from the 936 group that were isolated over a 9-year period from whey samples obtained from a Canadian cheese factory. These six phages infected the same two industrial Lactococcus lactis strains out of 30 tested. The CB14 and GR7 genomes were found to be 100% identical even though they were isolated 14 months apart, indicating that a phage can survive in a cheese plant for more than a year. The other four genomes were related but notably different. The length of the genomes varied from 28,144 to 32,182 bp, and they coded for 51 to 55 open reading frames. All five genomes possessed a 3′ overhang cos site that was 11 nucleotides long. Several structural proteins were also identified by nano-high-performance liquid chromatography-tandem mass spectrometry, confirming bioinformatic analyses. Comparative analyses suggested that the most recently isolated phages (CB19 and CB20) were derived, in part, from older phage isolates (CB13 and CB14/GR7). The organization of the five distinct genomes was similar to the previously sequenced lactococcal phage genomes of the 936 group, and from these sequences, a core genome was determined for lactococcal phages of the 936 group.The manufacture of cheeses requires the inoculation of carefully selected bacterial cultures, known as starter cultures, at concentrations of at least 107 live bacteria per ml of heat-treated milk. The purpose of this process is to control the fermentation and to obtain high-quality fermented products (29). Starter cultures are a combination of lactic acid bacteria (LAB), of which one of the most important species is Lactococcus lactis. L. lactis is a low-GC gram-positive bacterium used to metabolize lactose into lactic acid during the production of several cheese varieties. Because large amounts of lactococcal cells are cultivated each day in large-scale fermentation vats and because these cells are susceptible to bacteriophage infection, it is not surprising that most cheese factories have experienced problems with phage contamination (13). Even a single phage infecting a starter strain is enough to begin a chain reaction that can eventually inhibit bacterial growth and cause production delays, taste and texture variations, and even complete fermentation failures (1, 29).Phage infections are unpredictable in food fermentations. Their presence and persistence in a dairy factory can be explained in many ways. First, raw milk can introduce new phages into an industrial plant (25). Madera et al. (22) also reported that newly isolated lactococcal phages were more resistant to pasteurization. Whey, a liquid by-product of cheese manufacturing, is another reservoir that can spread phages in a factory environment (25). Airborne phage dissemination may also be important since concentrations of up to 106 PFU/m3 have been observed close to a functional whey separation tank (32).For decades, the dairy industry has been working to curtail the propagation of virulent phages using a variety of practical strategies, including, among others, sanitation, optimized factory design, air filtration units, rotation of bacterial strains, and the use of phage resistance systems (13). Yet new virulent phages emerge on a regular basis. Indeed, large-scale industrial milk fermentation processes can be slowed down by virulent phages of the Caudovirales order. Members of three lactococcal phage groups, namely, 936, c2, and P335, are mostly found in dairy plants. The 936-like phages are by far the most predominant worldwide (3, 18, 22, 27).Phages of the 936 group have a double-stranded DNA genome and possess a long noncontractile tail connected to a capsid with icosahedral symmetry characteristic of the Siphoviridae family. Currently, six complete phage genomes of the lactococcal 936 group are available in public databases, including sk1 (6), bIL170 (10), jj50, 712, P008 (23), and bIBB29 (16). Their comparative analysis revealed a conserved gene organization despite being isolated from different countries. Most of the differences have been observed in the early gene module, where insertions, deletions, and point mutations likely occurred (16, 23). Moreover, it is assumed that these phages can also exchange DNA through recombination with other bacterial viruses present in the same ecosystem.Because new members of this lactococcal phage group are regularly isolated, a better understanding of their evolution is warranted to better control them. A cheese factory is a particular man-made niche where rapidly growing bacterial strains encounter ubiquitous phages. Such active environments provide ample opportunities for phage evolution, especially to dodge phage resistance mechanisms that may be present in host cells. Nonetheless, the evolutionary dynamics that shape the diversity of lactococcal phage populations are still not well understood.In this study, we analyzed the genome and structural proteome of six 936-group phages (SL4, CB13, CB14, CB19, CB20, and GR7) that infected the same L. lactis strains and were isolated over a 9-year period from a cheese factory.  相似文献   

9.
10.
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   

11.
Shiga toxin-converting bacteriophages (Stx phages) are involved in the pathogenicity of some enteric bacteria, such as Escherichia coli O157:H7. Stx phages are released from their bacterial hosts after lytic induction and remain free in the environment. Samples were analyzed for the presence of free Stx phages by an experimental approach based on the use of real-time quantitative PCR (qPCR), which enables stx to be detected in the DNA from the viral fraction of each sample. A total of 150 samples, including urban raw sewage samples, wastewater samples with fecal contamination from cattle, pigs, and poultry, and fecal samples from humans and diverse animals, were used in this study. Stx phages were detected in 70.0% of urban sewage samples (10 to 103 gene copies [GC] per ml) and in 94.0% of animal wastewater samples of several origins (10 to 1010 GC per ml). Eighty-nine percent of cattle fecal samples were positive for Stx phages (10 to 105 GC per g of sample), as were 31.8% of other fecal samples of various origins (10 to 104 GC per g of sample). The stx2 genes and stx2 variants were detected in the viral fraction of some of the samples after sequencing of stx2 fragments amplified by conventional PCR. The occurrence and abundance of Stx phages in the extraintestinal environment confirm the role of Stx phages as a reservoir of stx in the environment.Shiga toxin-producing Escherichia coli (STEC) is associated with diarrhea, hemorrhagic enterocolitis, and hemolytic-uremic syndrome in humans (46). Escherichia coli serotype O157:H7 is the main cause of these diseases, although other serotypes of E. coli and other enterobacteria species have been described (36). These E. coli serotypes produce at least two immunologically distinct Shiga toxins, called Stx1 and Stx2. In addition to these, several variations of these toxins have been reported in recent years, showing differences in virulence and distribution in the host populations examined (48, 51). Shiga toxin genes are carried by temperate bacteriophages (19, 35). Stx-encoding bacteriophages investigated to date consist of double-stranded DNA and have lambdoid genetic structures (19, 27, 32, 37, 47). The induction and regulation of these phages are directly involved in the production of toxin and, therefore, in the pathogenicity of the strains (8, 50). Stx phages are efficient vectors for the transfer of toxin genes, being able to convert nonpathogenic bacterial hosts into Stx-producing strains by transduction of stx, as has been demonstrated under various conditions (1, 4, 27, 28, 41, 49).Most of the reported outbreaks of STEC infections are associated with cattle products (10, 17), with the consumption of contaminated foods (10, 34), and with several waterborne infections (30). Stx phages are present within fecally contaminated aquatic environments (9, 28, 30, 32, 45). Moreover, a high percentage of STEC strains present in extraintestinal environments carry inducible Stx phages (14, 30).As individuals infected with STEC strains shed large quantities of Stx phages in feces, Stx phages should be prevalent in the environment, as are other viruses transmitted by the fecal-oral route (5, 11) or bacteriophages infecting bacteria present in the intestinal tract (16, 23). Moreover, those STEC strains isolated from food and animals carry inducible Stx phages (24, 27, 42). The virulence profiles of STEC strains isolated from food also suggest the presence of inducible Stx phages (10).Stx phages in sewage have been detected by nested PCR (28, 29, 31). However, to quantify them, the most probable number (MPN) method was applied, which allows only a rough estimate of the amount of Stx phages present in the sample. To assess the number of Stx phages accurately, real-time quantitative PCR (qPCR) technology is a useful tool. This technology is both sensitive and specific, and it gives accurate quantitative results (25). Comparison with a standard enables the number of copies of stx to be quantified, which can then be translated into the number of Stx phage particles.Little is known about the prevalence of phages carrying stx in fecal samples. The data available on the numbers of these phages in fecally contaminated water samples were only roughly estimated. The first step to evaluate the role of Stx phages in the environment as lateral gene transfer vectors is to know the extent of these viruses in the environment. The aim of this study is to report quantitative data on the abundance of Stx phages in urban sewage samples, in wastewater samples from cattle, pigs, and poultry, and in diverse fecal samples, calculated by means of a methodology based on qPCR.  相似文献   

12.
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   

13.
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   

14.
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   

15.
To determine the relative importance of temperate bacteriophage in the horizontal gene transfer of fitness and virulence determinants of Enterococcus faecalis, a panel of 47 bacteremia isolates were treated with the inducing agents mitomycin C, norfloxacin, and UV radiation. Thirty-four phages were purified from culture supernatants and discriminated using pulsed-field gel electrophoresis (PFGE) and restriction mapping. From these analyses the genomes of eight representative phages were pyrosequenced, revealing four distinct groups of phages. Three groups of phages, ΦFL1 to 3, were found to be sequence related, with ΦFL1A to C and ΦFL2A and B sharing the greatest identity (87 to 88%), while ΦFL3A and B share 37 to 41% identity with ΦFL1 and 2. ΦFL4A shares 3 to 12% identity with the phages ΦFL1 to 3. The ΦFL3A and B phages possess a high DNA sequence identity with the morphogenesis and lysis modules of Lactococcus lactis subsp. cremoris prophages. Homologs of the Streptococcus mitis platelet binding phage tail proteins, PblA and PblB, are encoded on each sequenced E. faecalis phage. Few other phage genes encoding potential virulence functions were identified, and there was little evidence of carriage of lysogenic conversion genes distal to endolysin, as has been observed with genomes of many temperate phages from the opportunist pathogens Staphylococcus aureus and Streptococcus pyogenes. E. faecalis JH2-2 lysogens were generated using the eight phages, and these were examined for their relative fitness in Galleria mellonella. Several lysogens exhibited different effects upon survival of G. mellonella compared to their isogenic parent. The eight phages were tested for their ability to package host DNA, and three were shown to be very effective for generalized transduction of naive host cells of the laboratory strains OG1RF and JH2-2.Enterococcus faecalis is a member of the natural flora of humans and colonizes the gastrointestinal and vaginal tracts and the oral cavity. In recent years it has emerged as an important opportunistic nosocomial pathogen and is a causative agent of bacteremia, infective endocarditis, and surgical wound and urinary tract infections. The accumulation of acquired antibiotic resistance determinants, in addition to its intrinsic resistance and tenacity, has given rise to the evolution of clinical isolates of E. faecalis that are therapeutically problematic (19). Greater notoriety was afforded to this species following the observed transfer of the conjugative transposon Tn1546 to Staphylococcus aureus, imparting vancomycin resistance (11). Subsequent analysis has revealed that multiple independent E. faecalis-dependent vanA transfers had occurred in the United States prior to 2007 (50). This places enterococci in an important and dynamic position within the health care system, warranting their increased study.The specific determinants that are proposed to contribute to the virulence of E. faecalis are not universally present, and expression of the cognate genes is variable (21, 37). For example, in a recent study of 106 clonally diverse strains of E. faecalis the metallopeptidase gelatinase (GelE) was shown to be expressed in less than 60% of 106 genotypically positive isolates, whereas expression of cytolysin was less frequently observed (expression in ∼25% of isolates, with 30% being genotypically positive) (33). A proposed pathogenicity island identified with E. faecalis V583 (49) is composed of a variable gene set encoding the virulence determinants enterococcal surface protein, cytolysin, and aggregation substance. This highly variable 150-kb mobile element contains many components of unknown function that are hypothesized to facilitate survival and/or transmission in the health care setting (34, 40, 49).Two sequenced and annotated genomes of E. faecalis have been completed and published to date. These are the blood isolate and first-observed vancomycin-resistant strain V583 (40) and the oral isolate OG1RF, used as a common laboratory strain (8). A major difference between these genomes is the presence in V583 of seven regions containing phage-associated sequences. In contrast, OG1RF contains only one phage remnant, which was proposed by McBride et al. (33) to form part of the core genome, a theory supported by the presence in OG1RF of this phage remnant region together with two CRISPR loci. CRISPR sequences provide sequence-specific resistance to bacteriophages via the assembly of phage DNA sequences interspersed as spacers between repeats, in concert with associated cas genes, which collectively operate as an RNA-based gene silencing mechanism (5, 6, 28, 30, 36, 42). This elegant heritable mechanism is proposed to limit horizontal gene transfer of bacteriophage, transposable elements, and conjugative plasmids (9, 10, 32).Within the firmicute division of Gram-positive bacteria, temperate bacteriophages are key vectors for the horizontal transfer of virulence genes. In Staphylococcus aureus, bacteriophages encode and mobilize an impressive array of immune evasion genes (54, 55) and Panton-Valentine leukocidin (43). Several bacteriophage-encoded virulence determinants also contribute to pathogenesis in group A Streptococcus (2, 3, 4).The role of bacteriophages in the virulence of E. faecalis is not clear. Encoded within seven phage-related sequences of strain V583, there are multiple reported homologs of the Streptococcus mitis platelet-binding proteins PblA and PblB (7) and a ferrochelatase (40). In contrast, the absence of mobile genetic elements (MGEs) in strain OG1RF led Bourgogne et al. (8) to speculate that they did not engender virulence in E. faecalis.In this study we determined the morphology and complete genome sequences of eight induced bacteriophages purified from clinical isolates of E. faecalis. We sought to determine the potential carriage of genes that might contribute to the virulence or fitness of this organism and characterize the capacity of these phages to participate in transduction.  相似文献   

16.
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   

17.
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   

18.
The Burkholderia cepacia complex (BCC) is made up of at least 17 species of Gram-negative opportunistic bacterial pathogens that cause fatal infections in patients with cystic fibrosis and chronic granulomatous disease. KS9 (vB_BcenS_KS9), one of a number of temperate phages isolated from BCC species, is a prophage of Burkholderia pyrrocinia LMG 21824. Transmission electron micrographs indicate that KS9 belongs to the family Siphoviridae and exhibits the B1 morphotype. The 39,896-bp KS9 genome, comprised of 50 predicted genes, integrates into the 3′ end of the LMG 21824 GTP cyclohydrolase II open reading frame. The KS9 genome is most similar to uncharacterized prophage elements in the genome of B. cenocepacia PC184 (vB_BcenZ_ PC184), as well as Burkholderia thailandensis phage φE125 and Burkholderia pseudomallei phage φ1026b. Using molecular techniques, we have disrupted KS9 gene 41, which exhibits similarity to genes encoding phage repressors, producing a lytic mutant named KS9c. This phage is incapable of stable lysogeny in either LMG 21824 or B. cenocepacia strain K56-2 and rescues a Galleria mellonella infection model from experimental B. cenocepacia K56-2 infections at relatively low multiplicities of infection. These results readily demonstrate that temperate phages can be genetically engineered to lytic form and that these modified phages can be used to treat bacterial infections in vivo.The Burkholderia cepacia complex (BCC) is a group of at least 17 Gram-negative species, the first identified strains of which were characterized as onion pathogens by W. H. Burkholder (9). Although these bacteria have a number of beneficial activities, including the promotion of crop growth and the degradation of organic pollutants, they have gained notoriety in the last two decades as serious opportunistic pathogens (19, 21, 25). BCC species, particularly B. multivorans and B. cenocepacia, cause serious respiratory infections in patients with cystic fibrosis and chronic granulomatous disease (42, 7). These infections are especially problematic due to symptom severity, the inherent antibiotic resistance of Bcc species, and the potential for rapid spread through susceptible patient populations (25, 23). Difficulties in treating these infections have led to the unfortunate practice of segregating patients, which has high economic, social, and psychological costs (18).Because of these clinical difficulties, interest in the isolation and characterization of Burkholderia-specific bacteriophages (or phages) has increased in recent years, with the apparent potential for using phages as therapeutic agents. Phage therapy is the clinical application of phages to prevent and/or to treat infections, which offers a promising alternative to antibiotic treatment for resistant bacteria such as those of the BCC (33, 39). A second benefit of these phage studies is that they may provide insight into the possible mechanisms of BCC virulence. For example, BcepMu, a transposable phage that specifically infects strains of B. cenocepacia, was found to carry genes similar to exeA, involved in toxin secretion, and mdmB and oafA, two acyltransferases (44). Finally, as Burkholderia phages tend to be underrepresented in comparative studies with respect to Escherichia coli and lactic acid bacteria phages, BCC-specific phage studies provide novel information about a relatively uncharacterized group of viruses.Although phage therapy using temperate virions can be effective (39), there are several reasons why lytic phages are generally considered the most appropriate candidates for use in phage therapy. One of the concerns is that phage integration can lead to lysogenic conversion and enhanced virulence (8). A second concern is that integration of temperate phages results in superinfection immunity due to expression of the phage repressor from the prophage. This protein binds to the operators of infecting phage DNA and represses gene expression, preventing both the initiation of the lytic cycle and the establishment of lysogeny (14). A third concern is that lysogeny affects the kinetics of infection. When a phage infects a cell and undergoes lysogeny instead of entering the lytic cycle, the cell survives, and no new phage particles are released (27). A final problem is that prophages can lead to specialized transduction after induction. Specialized transduction occurs after inexact excision of a prophage from the bacterial chromosome. Bacterial DNA flanking the prophage is packaged into the capsid, and this sequence, which can potentially encode virulence factors, can subsequently recombine into the chromosome of a new host (14).It has been estimated that more than half of tailed phages have evolved a temperate lifestyle, although some estimates have been greater than 90% (1, 22). This situation makes the isolation of naturally lytic phages extremely difficult, particularly when they must have a specific host range that includes clinically relevant bacterial species, such as B. cenocepacia (24). The use of classical genetics to produce lytic phage variants, for example, by plating temperate phages on lysogens and screening for clear plaque vir mutants, is complicated by the fact that such mutations are undefined.This report describes the characterization of KS9 (vB_BcenS_KS9), a prophage of Burkholderia pyrrocinia LMG 21824 (41), and its conversion to a lytic phage through specific molecular modification of gene 41 encoding its putative lytic phase repressor. Preliminary characterization of short sequences by Seed and Dennis (41) indicated that the genome of KS9, whose host range includes Bcc B. cenocepacia K56-2, shows similarity to the genomes of two non-BCC Burkholderia phages: φE125, a prophage of Burkholderia thailandensis E125 (47), and φ1026b, a prophage of Burkholderia pseudomallei 1026b (17). However, no phages closely related to KS9 have been functionally tested to demonstrate that proteins similar to gp41 function as true phage repressors. In the present study, we have used the BCC infection model of Galleria mellonella (40) to assess both the contribution of the KS9 prophage to BCC host virulence and the ability of a genetically modified KS9 to treat B. cenocepacia infections without stably integrating into the host bacterial chromosome as a prophage.  相似文献   

19.
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   

20.
The temperate phage φSLT of Staphylococcus aureus carries genes for Panton-Valentine leukocidin. Here, we identify ORF636, a constituent of the phage tail tip structure, as a recognition/adhesion protein for a poly(glycerophosphate) chain of lipoteichoic acid on the cell surface of S. aureus. ORF636 bound specifically to S. aureus; it did not bind to any other staphylococcal species or to several gram-positive bacteria.Staphylococcus aureus, a ubiquitous and harmful human pathogen, produces three types of bicomponent pore-forming cytotoxins, namely, γ-hemolysin (LukF and Hlg2), leukocidin (LukF and LukS), and Panton-Valentine leukocidin (PVL) (LukF-Pv and LukS-Pv) (16). Of these, PVL has been investigated as a virulence-related factor of some S. aureus infectious diseases (7, 11, 23, 24, 31, 37). PVL shows high cytolytic specificity against human polymorphonuclear leukocytes and macrophages, and it is closely associated with most cutaneous necrotic lesions, such as furuncles or primary abscesses, and severe necrotic skin infection (24, 31), as well as with severe necrotic hemorrhagic pneumonia (11, 23). LukF-Pv and LukS-Pv are expressed by the PVL locus (pvl), which is distinct from the γ-hemolysin locus (hlg) (16, 32). In previous research, we found that pvl genes are located in the genome of the lysogenic bacteriophage φPVL (17, 18). We also found another PVL-carrying temperate elongated-head Siphoviridae phage, φSLT, which has the ability to convert S. aureus to the PVL-producing strain from a clinical isolate (29). These findings indicated that at least two types of staphylococcal temperate phages are involved in the horizontal transfer of pvl genes among S. aureus strains (16, 29). Recently, the emergence of a single clonal community-acquired methicillin-resistant S. aureus (CA-MRSA), which produces PVL, was reported (7). Most CA-MRSA strains isolated in the United States and Australia carry the staphylococcal cassette chromosome mec (SCCmec) IV, and they were divided into five clonal complexes by multilocus sequence typing (30). The analysis of the CA-MRSA clones confirmed the presence of PVL genes and SCCmec IV in CA-MRSA and suggested that various CA-MRSA strains have arisen from the diverse genetic backgrounds associated with each geographic origin, rather than from the worldwide spread of a single clone (30, 37). Although there is great debate as to whether PVL is an important virulence factor, numerous studies support the hypothesis that PVL plays an important role in the pathogenesis of CA-MRSA necrotizing pneumonia (3, 6). In regard to the acquisition of PVL gene clusters and the proliferation of PVL-carrying CA-MRSA, the horizontal transfer of PVL via PVL-carrying phages, as well as that of SCCmec, has become the focus of intense research interest. To understand the horizontal transfer of PVL, the analysis of the infection ability of a PVL-carrying phage is important. If the phage has a wide host range, the PVL-carrying phage might threaten to become a source of emerging PVL-positive bacteria. Phage infection starts from an interaction between a phage virion and its host cell surface receptor. Nevertheless, little is known about phage receptors on the surface of S. aureus, and the mechanism of host cell-specific binding of staphylococcal phages has been poorly characterized. In addition, there is no information about staphylococcal phage proteins involved in host cell recognition and/or binding. Here, we identify ORF636, with a mass of 66 kDa, as a structural protein of the φSLT tail and determine that it acts as a protein for recognition/adhesion of a poly(glycerophosphate) moiety of lipoteichoic acid (LTA) on the cell surface of the host S. aureus in the first stage of infection by φSLT.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号