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 218 毫秒
1.
2.
We note the existence of a "partially cis-acting" regulatory protein of bacteriophage λ: the product of the phage Q gene. We suggest that there may be a complete spectrum from "all cis" to "all trans" for such regulatory proteins. This behavior might arise because a DNA-binding protein either acts at a nearby (cis) site soon after synthesis or becomes "lost" for its trans activity on another genome through nonspecific interactions with DNA. Our proposed explanation provides one evolutionary basis for the linkage of genes for regulatory proteins and the sites at which such proteins act; it also suggests a possible rationale for the "metabolic instability" of certain regulatory proteins.  相似文献   

3.

Background

The burst size of a phage is important prior to phage therapy and probiotic usage. The efficiency for a phage to burst its host bacterium can result from molecular domino effects of the phage gene expressions which dominate to control host machinery after infection. We found two Podoviridae phages, ϕA318 and ϕAs51, burst a common host V. alginolyticus with different efficiencies of 72 and 10 PFU/bacterium, respectively. Presumably, the genome sequences can be compared to explain their differences in burst sizes.

Results

Among genes in 42.5 kb genomes with a GC content of 43.5%, 16 out of 47 open-reading frames (ORFs) were annotated to known functions, including RNA polymerase (RNAP) and phage structure proteins. 11 strong phage promoters and three terminators were found. The consensus sequence for the new vibriophage promoters is AATAAAGTTGCCCTATA, where the AGTTG bases of −8 through −12 are important for the vibriophage specificity, especially a consensus T at −9 position eliminating RNAP of K1E, T7 and SP6 phages to transcribe the genes. ϕA318 and ϕAs51 RNAP shared their own specific promoters. In comparing ϕAs51 with ϕA318 genomes, only two nucleotides were deleted in the RNAP gene and three mutating nucleotides were found in the major capsid genes.

Conclusion

Subtle analyses on the residue alterations uncovered the effects of five nucleotide mutations on the functions of the RNAP and capsid proteins, which account for the host-bursting efficiency. The deletion of two nucleotides in RNAP gene truncates the primary translation due to early stop codon, while a second translational peptide starting from GTG just at deletion point can remediate the polymerase activity. Out of three nucleotide mutations in major capsid gene, H53N mutation weakens the subunit assembly between capsomeres for the phage head; E313K reduces the fold binding between β-sheet and Spine Helix inside the peptide.

Electronic supplementary material

The online version of this article (doi:10.1186/1471-2164-15-505) contains supplementary material, which is available to authorized users.  相似文献   

4.
The native product of open reading frame 112 (orf112) and a recombinant variant of the RstB protein, encoded by Vibrio cholerae pathogen-specific bacteriophages VGJφ and CTXφ, respectively, were purified to more than 90% homogeneity. Orf112 protein was shown to specifically bind single-stranded genomic DNA of VGJφ; however, RstB protein unexpectedly bound double-stranded DNA in addition to the single-stranded genomic DNA. The DNA binding properties of these proteins may explain their requirement for the rolling circle replication of the respective phages and RstB''s requirement for single-stranded-DNA chromosomal integration of CTXφ phage dependent on XerCD recombinases.Vibrio cholerae, the etiologic agent of cholera, is a gram-negative bacterium which hosts several specific filamentous phages (1, 7, 8, 9, 10, 11, 13). CTXφ phage has been the most studied due to its role in pathogenicity and horizontal gene transfer (6). This phage is usually integrated into the genomes of toxigenic strains of V. cholerae, but it is also able to replicate directly from the bacterial chromosome (6) and to produce infective phage particles with potential for transducing the cholera toxin genes into nonpathogenic environmental strains (6, 13). Another filamentous phage important for its role in horizontal gene transfer is VGJφ, which is able to recombine with the CTXφ genome to originate a hybrid phage endowed with the full potential for virulence conversion. The hybrid phage shows an increased infectivity due to its specificity for the receptor mannose-sensitive hemagglutinin (receptor mannose-sensitive hemagglutinin pilus), which is ubiquitous among environmental strains (1, 2). Therefore, elucidating the biology of these phages is crucial for understanding the evolution of bacterial pathogenesis.The genomes of CTXφ and VGJφ carry the putative homologous rstB and open reading frame 112 (orf112) genes, respectively. The requirement of rstB for the integration of CTXφ into the bacterial chromosome has been described (14). However, the biochemical function of the gene product has not been elucidated. Genes rstB and orf112 are positional and size homologues of genes encoding single-stranded DNA (ssDNA)-binding proteins (SSB) in other filamentous phages (1). It is expected that the proteins encoded by rstB and orf112 exert similar functions in the biology of their respective phages (1). Thus, we wanted to evaluate the ssDNA-binding activity of these ORF products.To asses whether the Orf112 product and RstB have SSB activity, sufficient amounts of pure proteins are required. This paper describes quick purification protocols used to obtain both protein species and the evaluation of their DNA binding activities. The Orf112 protein was obtained from V. cholerae strain 569B (serogroup O1, Inaba classical biotype) infected with VGJφ, which expresses high levels of the protein. The infected bacteria were inoculated into 300 ml of LB broth and were cultured with shaking overnight at 200 rpm and 37°C. Parallel uninfected batches of 569B were also processed. Cells were collected by centrifugation for 15 min at 9,000 × g and at 4°C and stored at −20°C until processed.A recombinant rstB gene with a hexahistidine tag coding region fused to the C terminus of the respective protein product (RstB-His) was constructed by cloning the gene into the expression vector pBAD/Myc-HisC (Invitrogen). The rstB gene was PCR amplified from the CTXφ genome using the oligonucleotides CNC06-171 (5′-AGTTCCATGGGGAAATTATGGGTGATAAT-3′) and CNC06-173 (5′-CATCAAGCTTTAATGGGT-3′), which introduce restriction sites for NcoI and HindIII at the amplicon ends. The amplified fragment was digested with both enzymes and cloned into the same sites of pBAD/Myc-HisC. In the resultant construction, named pBAD/Myc-HisC-rstB 9, expression of the recombinant protein is inducible by arabinose.Plasmid pBAD/Myc-HisC-rstB 9 was electroporated into Escherichia coli Top 10. A 1-ml sample of an overnight, 5-ml, ampicillin-supplemented LB broth culture of transformed Top 10 was inoculated into 300 ml of fresh broth. The culture was incubated with orbital shaking at 200 rpm and 37°C until it reached an optical density at 600 nm of 0.5. To induce expression of the RstB-His protein, 0.002% (wt/vol) arabinose was added and the culture was reincubated for three additional hours. Parallel batches of pBAD/Myc-HisC-transformed E. coli Top 10 were processed as a negative control. Cells were sedimented by centrifugation for 15 min at 9,000 × g and 4°C and stored at −20°C until processed.The expression of the Orf112 and RstB-His proteins was monitored by sodium dodecyl sulfate-polyacrylamide gel electrophoresis (SDS-PAGE). Cell extracts of VGJφ-infected 569B and E. coli Top 10 transformed with pBAD/Myc-HisC-rstB 9 contained proteins with apparent molecular sizes of 12.7 kDa (Fig. (Fig.1)1) and 16 kDa (Fig. (Fig.2B),2B), respectively, which are not observed in cells from control cultures. The sizes match those predicted from orf112 (12.72 kDa) (see reference 1) and recombinant rstB-his (16.8 kDa).Open in a separate windowFIG. 1.SDS-PAGE monitoring of the purification process of Orf112 protein. Lane 1, broad-range protein molecular mass markers (Promega); lane 2, cell extract of non-VGJφ-infected V. cholerae 569B; lane 3, cell extract of VGJφ-infected V. cholerae 569B; lane 4, soluble fraction of the sonicate; lane 5, insoluble fraction of the sonicate; lane 6, precipitate at 30% (NH3)2SO4; lane 7, supernatant at 30% (NH3)2SO4; lane 8, precipitate at 50% (NH3)2SO4; lane 9, supernatant at 50% (NH3)2SO4; lane 10, Orf112 protein electro-eluted after preparative SDS-PAGE.Open in a separate windowFIG. 2.Isolation and purification of RstB-His. (A) Chromatogram on a Ni-CAM HC His tag affinity column. (B) SDS-PAGE monitoring of the purification process of RstB-His. Lane 1, broad-range protein molecular mass markers (Promega); lane 2, cell extract of uninduced cultures; lane 3, cell extract of expression-induced cultures; lane 4, soluble fraction of the sonicate from expression-induced cultures; lane 5, insoluble fraction of the sonicate from expression-induced cultures; lane 6, soluble fraction of the 8 M urea extract; lane 7, RstB-His eluted from the column.These proteins are not secreted into the growth medium (not shown); thus, they were released from the cells by ultrasonic disruption as previously described (4). V. cholerae was suspended in 15 ml of 20 mM Tris-HCl buffer, pH 7.5, while E. coli cells were suspended in 15 ml of 20 mM sodium phosphate, 0.5 M NaCl, 10 mM imidazole, and 1 mM phenylmethylsulfonyl fluoride, pH 8.0. Cell lysates were cleared by centrifugation for 40 min at 9,000 × g and 4°C. SDS-PAGE detected Orf112 protein in the soluble fraction, while RstB-His remained in the insoluble fractions of cell extracts (Fig. (Fig.1).1). Subsequently, 569B cell lysate supernatants containing Orf112 protein were fractionated with ammonium sulfate. At 30% ammonium sulfate, several contaminants precipitated but Orf112 protein remained in solution, while at 50% ammonium sulfate, Orf112 precipitated and was recovered by centrifugation. The pellet was washed twice with 50% ammonium sulfate and finally resuspended into 3 ml of 20 mM Tris-HCl buffer, pH 7.5. Removal of excess salt was achieved by gel filtration before the extract was applied to a preparative SDS-PAGE gel. Briefly, a 15% polyacrylamide gel (17 by 19 by 0.5 cm) was run at a constant current intensity of 100 mA and with free voltage at 4°C, until the bromophenol blue dye migrated off the gel. The gels were negatively stained with imidazole-zinc (3), and the Orf112 protein band was identified by comparing bands with the bands of a negative control applied in a neighboring lane of the same gel, where this protein was not visible. The ORF112 protein band was cut from the gel, and the slice was fragmented and introduced into a dialysis bag with a 6- to 8-kDa molecular mass cutoff in 10 ml of 24 mM Tris-HCl-250 mM glycine-0.5% (wt/vol) SDS buffer. The protein was electro-eluted for 5 h at a current intensity of 70 mA and with free voltage at 4°C. Reverse current was applied for 5 min to release membrane-bound proteins, and gel fragments were discarded. The same sample was dialyzed against 1 liter of 0.5 M Tris-HCl, 0.25 M glycine buffer, pH 7.5, for 24 h at 4°C with constant stirring. The dialysis was repeated with 20 mM Tris-HCl, 0.5 M NaCl buffer, pH 7.5, for 24 h. No contaminants were seen when 25 μg of this Orf112 protein-dialyzed extract was checked by SDS-PAGE and Coomassie brilliant blue staining (Fig. (Fig.1,1, lane 10).RstB-His protein was recovered from the insoluble fraction of the E. coli lysate by dissolving the lysate in 15 ml of a buffer containing 8 M urea, 20 mM sodium phosphate, 0.5 M NaCl, and 10 mM imidazole, pH 8.0. The mixture was stirred overnight at 4°C and cleared by centrifugation at 9,000 × g for 40 min. The supernatant was applied to a Ni-CAM HC matrix (Sigma), and urea was removed using a linear gradient from 8 to 0 M urea as previously described (5). The presence of 10 mM imidazole in the sample and binding buffer was intended to reduce the level of contaminants bound to the column. Protein was eluted using a gradient of imidazole (10 to 250 mM) in 20 mM sodium phosphate, 0.5 M NaCl buffer, pH 8.0. Fractions were assayed by SDS-PAGE, and those containing the RstB-His protein were pooled according to purity rather than yield. RstB-His protein was obtained with 90% purity (Fig. (Fig.2B,2B, lane 7), according to a densitometry scan of Coomassie brilliant blue-stained gels, using a Gene Genius gel documentation system (Syngene Synoptics Ltd., Cambridge, United Kingdom). The gradient-based removal of urea allowed effective solubilization of RstB-His without significant precipitation of protein in the column, as described before (5).Biological activity was assayed by retardation assays of VGJφ genomic ssDNA by 0.5% agarose gel electrophoresis conducted with 20 mM EDTA, 40 mM Tris-acetate buffer. Various amounts of each protein and ssDNA from VGJφ (0.5 μg) mixed in 20% glycerol, 0.25 mM EDTA, 0.3 μM bovine serum albumin, 20 mM Tris-HCl up to a total volume of 40 μl were incubated at room temperature for 30 min and loaded into the gel for analysis. The electrophoresis was run at 100 V and 4°C until colorant exit. Ethidium bromide (1 μg/ml) was used for 30 min to stain DNA bands, which were documented in a Gene Genius gel system (Syngene Synoptics Ltd., Cambridge, United Kingdom).Orf112 protein exhibited DNA retardation activity, showing a high specificity of binding for the circular ssDNA of VGJφ, but was unable to bind double-stranded DNA (dsDNA) (Fig. (Fig.3A).3A). However, RstB was able to bind ssDNA as well as dsDNA substrates (Fig. (Fig.3B).3B). No retardation was observed with protein preparations from negative controls or when the DNA-protein mixture was inactivated with 1:1 (vol/vol) phenol-chloroform, indicating that the binding activity is intrinsic to the purified proteins.Open in a separate windowFIG. 3.Gel retardation assays of VGJφ-ssDNA by Orf112 protein or RstB-His binding, as measured by 0.5% agarose gel electrophoresis. (A) Binding of Orf112 protein. Lane 1, control of 500 ng of genomic ssDNA of VGJφ; lanes 2 to 6; same as for lane 1 plus 1.25, 2.50, 5.00, 10.0, and 20.0 μg of Orf112 protein, respectively; lane 7, same as for lane 6 but treated with phenol-chloroform; lane 8, linearized replicative-form dsDNA of VGJφ; lane 9, same as for lane 8 plus 20.0 μg of Orf112; lane 10, linearized pUC19; lane 11, same as for lane 10 plus 20.0 μg of Orf112. (B) Binding of RstB-His. Lane 1, control of 500 ng of genomic ssDNA of VGJφ; lane 2, same as for lane 1 plus 20 μg of RstB treated with phenol-chloroform; lanes 3 to 8, same as for lane 1 plus 0.62, 1.25, 2.50, 5.00, 10.0, and 20.0 μg of RstB-His, respectively; lane 9, linearized replicative-form dsDNA of VGJφ; lane 10, same as for lane 9 plus 20.0 μg of RstB-His; lane 11, linearized pUC19; lane 12, same as for lane 11 plus 20.0 μg of RstB-His.In the case of RstB, which binds to ssDNA and dsDNA substrates, we wanted to rule out the possibility that this effect was caused by the His tail fused to the recombinant protein. For this, we recloned RstB in the same vector as RstB-His but without the His tail. We used the same procedure described above for RstB-His but used oligonucleotides CNC06-171 (see above) and CNC06-172 (5′-TACTGCAGTCAAGATTTAATGGGTTG-3′) for RstB amplification. In this case, CNC06-172 introduced a restriction site for PstI, which was used for cloning into pBAD/Myc-HisC.For purification of this RstB variant, E. coli growth and protein expression induction was done as described for RstB-His. Again, RstB was recovered in the insoluble fraction after cell disruption by sonication. Inclusion bodies were resuspended in 15 ml of 50 mM phosphate buffer, pH 7.7, containing 8 M urea, and after overnight stirring at 4°C, the suspension was cleared by centrifugation (9,000 × g, 40 min). The supernatant was applied to an SP Sepharose fast-flow column (Amersham, United Kingdom), and urea was removed as described above for RstB-His. The RstB that bound to the matrix was eluted using a gradient of 0 to 2 M NaCl and was obtained with about 92% purity (data not shown). RstB without the His tail also showed binding activity toward ssDNA and dsDNA substrates (data not shown), ruling out the possibility that the His hexamer is responsible for the nonspecific binding of RstB-His.Since RstB has affinity for both ssDNA and dsDNA, the possibility exists that this protein simply binds any DNA nonspecifically due to an effect of a charge interaction with the phosphate backbone of DNA. In order to study the effect of the charge in the DNA-binding activity of RstB, NaCl was included in the reaction mixture at a concentration from 0 to 500 mM (Fig. (Fig.4).4). As can be seen in Fig. Fig.4,4, the retardation activity of ssDNA was only partially inhibited from a starting concentration of 400 mM (Fig. (Fig.4A),4A), while the retardation of dsDNA started to be partially inhibited at 300 mM NaCl (Fig. (Fig.4B).4B). Since RstB continues to bind at high salt concentrations, which should equilibrate the charge effect, these results indicate that the DNA binding activity is not due to the presence of positively charged amino acids in the protein backbone but rather to the presence of domains that specifically recognize the ssDNA or dsDNA.Open in a separate windowFIG. 4.Effect of salt concentration and nonrelated dsDNA competition on the binding of RstB. (A) RstB binding to phage ssDNA in the presence of 0 to 500 mM NaCl as detected by 0.5% agarose gel electrophoresis. Lane 1, 400 ng of genomic ssDNA of VGJφ (control); lane 2, same as for lane 1 plus 15.6 μg of RstB-His; lanes 3 to 7, same as for lane 2 plus 100, 200, 300, 400, and 500 mM NaCl, respectively. (B) RstB binding to replicative-form phage dsDNA in the presence of 0 to 500 mM NaCl. Lane 1, 400 ng of dsDNA of VGJφ (control); lane 2, same as for lane 1 plus 15.6 μg of RstB-His; lanes 3 to 7, same as for lane 2 plus 100, 200, 300, 400, and 500 mM NaCl. (C) DNA-binding competition by RstB. Lane 1, 400 ng of genomic ssDNA of VGJφ (control); lane 2, same as for lane 1 plus 15.6 μg of RstB-His; lanes 3 and 4, same as for lane 2 plus 500 and 1,000 ng of sheared calf thymus dsDNA, respectively; lane 5, 500 ng of calf thymus DNA (control).We also investigated whether RstB-His has more affinity for the phage ssDNA than for a nonrelated dsDNA; a DNA-binding competition experiment in which the protein was incubated with a constant amount of ssDNA of VGJφ phage and increasing amounts of calf thymus DNA was performed (Fig. (Fig.4C).4C). The ssDNA of VGJφ was retarded by RstB-His even in the presence of 500 and 1,000 ng of calf thymus dsDNA (Fig. (Fig.4C,4C, lanes 3 and 4). These results indicate that RstB protein has more affinity for the phage ssDNA than for a non-phage-related dsDNA.Until we know more, RstB is the first protein of a filamentous phage which shows affinity for both ss- and dsDNA, at least in vitro. It is possible that RstB needs another protein from the host or the phage itself to recognize the ssDNA in a specific manner, or perhaps the affinity of RstB for both ss- and dsDNA is an intrinsic property of the protein, which is needed on one hand for binding to genomic ssDNA during the rolling circle replication of the phage and on the other hand for holding the hairpin dsDNA secondary structure formed by the phage genome that functions as the site for integration into the bacterial chromosome (12). This hairpin structure is used by XerCD recombinases as a substrate for recombining the phage genome with the bacterial chromosomal dif site (12), and RstB may act jointly with XerCD to achieve integration. This could explain the requirement of RstB for the integration of CTXφ.It is concluded that Orf112 and RstB proteins purified by the protocols described in this paper were biologically active and obtained at a high degree of purity, which paves the way for further characterization of these proteins. The SSB activities of these two proteins are shown for the first time. Consequently, we propose to rename their respective genes gVVGJφ and gVCTXφ and their proteins pVVGJφ and pVCTXφ, to follow the denomination of genes of canonical phages of the Inovirus genus. A biochemical and chemical-physical characterization of both proteins is in progress and will be published elsewhere. Should the in vitro role demonstrated for these proteins operate in vivo as well, it might explain their role for rolling circle replication and why rstB is required for CTXφ integration.  相似文献   

5.
Summary Erythromycin (2–4 g/ml) was found to inhibit specifically multiplication of SPO1 in sporulating cells of an erythromycin-resistant, conditional asporogenous mutant of Bacillus subtilis 168 thy - trp -, Ery1040. In contrast, streptomycin (150–200 g/ml) which inhibits protein synthesis to a similar extent as erythromycin did not inhibit SPO1 multiplication severely, suggesting that the inhibition of SPO1 multiplication by erythromycin is not caused by an overall inhibition of protein synthesis. Neither phage DNA synthesis nor phage messenger RNA synthesis was affected appreciably under these conditions. However, the synthesis of three phage proteins that are synthesized 15 min after infection was preferentially inhibited by erythromycin. In addition, the inhibition of SPO1 multiplication has been correlated with the stimulation of host stable RNA synthesis exhibited by erythromycin. Possible mechanisms for the inhibition of SPO1 multiplication in Ery1040 cells are discussed.  相似文献   

6.
7.
The complete genome of EcoM-GJ1, a lytic phage that attacks porcine enterotoxigenic Escherichia coli of serotype O149:H10:F4, was sequenced and analyzed. The morphology of the phage and the identity of the structural proteins were also determined. The genome consisted of 52,975 bp with a G+C content of 44% and was terminally redundant and circularly permuted. Seventy-five potential open reading frames (ORFs) were identified and annotated, but only 29 possessed homologs. The proteins of five ORFs showed homology with proteins of phages of the family Myoviridae, nine with proteins of phages of the family Podoviridae, and six with proteins of phages of the family Siphoviridae. ORF 1 encoded a T7-like single-subunit RNA polymerase and was preceded by a putative E. coli σ70-like promoter. Nine putative phage promoters were detected throughout the genome. The genome included a tRNA gene of 95 bp that had a putative 18-bp intron. The phage morphology was typical of phages of the family Myoviridae, with an icosahedral head, a neck, and a long contractile tail with tail fibers. The analysis shows that EcoM-GJ1 is unique, having the morphology of the Myoviridae, a gene for RNA polymerase, which is characteristic of phages of the T7 group of the Podoviridae, and several genes that encode proteins with homology to proteins of phages of the family Siphoviridae.  相似文献   

8.
Intrinsically disordered proteins and regions (IDPs and IDRs) lack stable 3D structure under physiological conditions in-vitro, are common in eukaryotes, and facilitate interactions with RNA, DNA and proteins. Current methods for prediction of IDPs and IDRs do not provide insights into their functions, except for a handful of methods that address predictions of protein-binding regions. We report first-of-its-kind computational method DisoRDPbind for high-throughput prediction of RNA, DNA and protein binding residues located in IDRs from protein sequences. DisoRDPbind is implemented using a runtime-efficient multi-layered design that utilizes information extracted from physiochemical properties of amino acids, sequence complexity, putative secondary structure and disorder and sequence alignment. Empirical tests demonstrate that it provides accurate predictions that are competitive with other predictors of disorder-mediated protein binding regions and complementary to the methods that predict RNA- and DNA-binding residues annotated based on crystal structures. Application in Homo sapiens, Mus musculus, Caenorhabditis elegans and Drosophila melanogaster proteomes reveals that RNA- and DNA-binding proteins predicted by DisoRDPbind complement and overlap with the corresponding known binding proteins collected from several sources. Also, the number of the putative protein-binding regions predicted with DisoRDPbind correlates with the promiscuity of proteins in the corresponding protein–protein interaction networks. Webserver: http://biomine.ece.ualberta.ca/DisoRDPbind/  相似文献   

9.
10.
Streptococcus lactis ME2 is a dairy starter strain that is insensitive to a variety of phage, including 18. The efficiency of plating of 18 on ME2 and N1 could be increased from <1 × 10−9 to 5.0 × 10−2 and from 7.6 × 10−7 to 2.1 × 10−2, respectively, when the host strains were subcultured at 40°C before plating the phage and the phage assay plates were incubated at 40°C. Host-dependent replication was demonstrated in N1 at 30°C and in N1 and ME2 at 40°C, suggesting the operation of a temperature-sensitive restriction and modification system in ME2 and N1. The increased sensitivity of ME2 and N1 to 18 at 40°C was also demonstrated by lysis of broth cultures and increased plaque size. ME2 grown at 40°C showed an increased ability to adsorb 18, indicating a second target for temperature-dependent phage sensitivity in ME2. Challenge of N1 with a 18 preparation that had been previously modified for growth on N1 indicated that at 40°C phage development was characterized by a shorter latent period and larger burst size than at 30°C. The evidence presented suggests that the high degree of phage insensitivity expressed by ME2 consists of a variety of temperature-sensitive mechanisms, including (i) the prevention of phage adsorption, (ii) host-controlled restriction of phage, and (iii) suppression of phage development. At 30°C these factors appear to act cooperatively to prevent the successful emergence of lytic phage active against S. lactis ME2.  相似文献   

11.
The genome for the marine pseudotemperate member of the Siphoviridae HSIC has been sequenced using a combination of linker amplification library construction, restriction digest library construction, and primer walking. HSIC enters into a pseudolysogenic relationship with its host, Listonella pelagia, characterized by sigmoidal growth curves producing >109 cells/ml and >1011 phage/ml. The genome (37,966 bp; G+C content, 44%) contained 47 putative open reading frames (ORFs), 17 of which had significant BLASTP hits in GenBank, including a β subunit of DNA polymerase III, a helicase, a helicase-like subunit of a resolvasome complex, a terminase, a tail tape measure protein, several phage-like structural proteins, and 1 ORF that may assist in host pathogenicity (an ADP ribosyltransferase). The genome was circularly permuted, with no physical ends detected by sequencing or restriction enzyme digestion analysis, and lacked a cos site. This evidence is consistent with a headful packaging mechanism similar to that of Salmonella phage P22 and Shigella phage Sf6. Because none of the phage-like ORFs were closely related to any existing phage sequences in GenBank (i.e., none more than 62% identical and most <25% identical at the amino acid level), HSIC is unique among phages that have been sequenced to date. These results further emphasize the need to sequence phages from the marine environment, perhaps the largest reservoir of untapped genetic information.  相似文献   

12.
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   

13.
Vegetable fermentations rely on the proper succession of a variety of lactic acid bacteria (LAB). Leuconostoc mesenteroides initiates fermentation. As fermentation proceeds, L. mesenteroides dies off and other LAB complete the fermentation. Phages infecting L. mesenteroides may significantly influence the die-off of L. mesenteroides. However, no L. mesenteroides phages have been previously genetically characterized. Knowledge of more phage genome sequences may provide new insights into phage genomics, phage evolution, and phage-host interactions. We have determined the complete genome sequence of L. mesenteroides phage Φ1-A4, isolated from an industrial sauerkraut fermentation. The phage possesses a linear, double-stranded DNA genome consisting of 29,508 bp with a G+C content of 36%. Fifty open reading frames (ORFs) were predicted. Putative functions were assigned to 26 ORFs (52%), including 5 ORFs of structural proteins. The phage genome was modularly organized, containing DNA replication, DNA-packaging, head and tail morphogenesis, cell lysis, and DNA regulation/modification modules. In silico analyses showed that Φ1-A4 is a unique lytic phage with a large-scale genome inversion (∼30% of the genome). The genome inversion encompassed the lysis module, part of the structural protein module, and a cos site. The endolysin gene was flanked by two holin genes. The tail morphogenesis module was interspersed with cell lysis genes and other genes with unknown functions. The predicted amino acid sequences of the phage proteins showed little similarity to other phages, but functional analyses showed that Φ1-A4 clusters with several Lactococcus phages. To our knowledge, Φ1-A4 is the first genetically characterized L. mesenteroides phage.Bacteriophages are the most abundant biological entities (estimated to be on the order of ≥1031) on the planet (9, 18). Phages are ubiquitous in nature and can influence the microbial ecology and genetics of bacteria. Because of their small (usually <60 kb) genomes, phages can provide an excellent model system for studying many biological processes, including DNA replication and genetic evolution. Despite this, many phages remain uncharacterized. Very little is known about phage diversity and phage-host interactions owing to the small number of sequenced phages. Furthermore, the existing phage sequence database is highly biased toward a limited spectrum of phage hosts, namely, Enterobacteriaceae, Bacillus, Staphylococcus, Pseudomonas, Vibrio cholerae, Lactococcus, Streptococcus thermophilus, and S. pyogenes. The majority of host species for sequenced phages are either pathogenic or dairy-related bacteria. Most of the newly sequenced phage genes have no assigned functions or matches in the GenBank database (7).Vegetable fermentations rely on a variety of lactic acid bacteria (LAB). The proper succession of LAB directly determines the quality and safety of the final fermentation products. Leuconostoc mesenteroides initiates most vegetable fermentations. It converts the sugars in vegetables (primarily glucose and fructose) to lactic acid, acetic acid, ethanol, CO2, and other flavor compounds (22, 58, 59, 60, 61). Acid production lowers the pH of fermenting vegetables and inhibits the growth of many microorganisms, including pathogens. CO2 production promotes the establishment of an anaerobic environment which favors the growth of other LAB. The metabolites produced by L. mesenteroides largely determine the flavor characteristics of the final products. As fermentation proceeds, L. mesenteroides rapidly dies off. Other LAB, including Lactobacillus plantarum, take over and complete the fermentation.It has been a widely held view that the disappearance of L. mesenteroides and the subsequent bacterial succession in sauerkraut fermentations are due to the inhibitory effect of acids that accumulate during fermentation (54, 61). Little is known about other factors that may play a role in bacterial succession. Recent studies have shown that phages are present in the vegetable fermentations (4, 47, 48, 74, 75). Because of the rapid lytic cycle of these phages, they may significantly impact starter cultures and bacterial succession in vegetable fermentations (56). Phages active against L. mesenteroides have been isolated and characterized (48); however, genome sequences have not been reported.L. mesenteroides phage 1-A4 (designated Φ1-A4) is of particular interest. Φ1-A4 is a lytic phage that was repeatedly isolated during the initial stages of a commercial sauerkraut fermentation. As a result, Φ1-A4 may significantly influence the survival of L. mesenteroides and flavor development during sauerkraut fermentation. It was found that Φ1-A4 infects at least three different strains of L. mesenteroides (48), and therefore it may also promote genetic exchange and genetic diversity in microbial communities (34).The objectives of this study were to determine and analyze the complete genome sequence of Φ1-A4, to experimentally identify the structural protein genes, and to compare the genome organization with that of related phages. To our knowledge, this study represents the first complete genomic and molecular characterization of Leuconostoc phage. The results from this study may provide new insights into our understanding of phage genetics. This study may aid the development of phage control technologies in vegetable and other fermentations that are susceptible to phage attack.  相似文献   

14.
  1. Download : Download high-res image (89KB)
  2. Download : Download full-size image
Highlights
  • •The proteomes of L. lactis MG1363 and phage p2 at different stages of infection were characterized.
  • •16% (226/1412) of the bacterial proteins detected were unique to infected cultures.
  • •A targeted approach using synthetic peptides improved the coverage of phage p2 proteome.
  • •By means of proteogenomics, we uncovered a conserved phage protein coded by a previously unannotated gene.
  • •Deletion of the bacterial gene llmg_0219 (unknown function) impedes phage p2 infection.
  相似文献   

15.
16.
The purification and properties of a lipid-containing bacteriophage, 6, are described. The phage contains a lipid envelope which is probably essential for infection. Infectivity of 6 was lost in the presence of organic solvents, sodium deoxycholate, and phospholipase A. The fatty acid composition of the phage lipid was similar to that of the Pseudomonas phaseolicola host cells. The phage was composed of about 25% lipid, 13% RNA, and 62% protein. The buoyant density of 6 was 1.27 g/ml in cesium chloride. The morphology of 6 was unusual; it had a polyhedral head of about 60 nm surrounded by a membranous, compressible envelope which appeared to assume an elongated configuration upon attachment to pili. The adsorption rate constant was 3.3 × 10−10 ml/min in a semi-synthetic medium and 3.8 × 10−10 ml/min in a nutrient broth-yeast extract medium. The latent period was shorter in the former medium (80-115 min compared with 120-160 min), and the average burst size was larger (250-400 compared with 125-150). The eclipse period coincided with the latent period.  相似文献   

17.
After nitrosoguanidine mutagenesis, a Phage Host Defective (phd) mutant of E. coli HfrH was isolated that supported the growth of T4D wild-type bacteriophage at 30°, but not at 40° or higher. Eleven independent spontaneous mutants of T4 (go mutants) were isolated that overcame the growth restriction at high temperature. All of these mutants were located within three percent recombination of a gene 39 amber mutation in the clockwise direction on the standard map. In mixed infections, the representative go mutant chosen for further study seems to be recessive to its wild-type allele. Temperature-shift experiments suggested that the mutated host function involved in phage growth is a "late" function, beginning in mid-eclipse.—Electrophoresis of phage proteins labelled early and late in infection showed that under restrictive conditions early protein synthesis was normal, but that certain late proteins were absent. However, measurements of DNA synthesis showed that under restrictive conditions the amount of phage DNA synthesized, and especially the amount of DNA sedimenting as high molecular weight replicative intermediate, was reduced. Pulse-chase experiments showed that the phage DNA made under restrictive conditions was not rapidly degraded.  相似文献   

18.
The transfer range of phage genes was investigated at the single-cell level by using an in situ DNA amplification technique. After absorption of phages, a phage T4 gene was maintained in the genomes of non-plaque-forming bacteria at frequencies of 10−2 gene copies per cell. The gene transfer decreased the mutation frequencies in nonhost recipients.Recently, whole-genome analyses have revealed that many bacterial genomes contain foreign genes, especially phage genes (9). The phage genes in bacterial genomes include genes for virulence or fitness factors such as extracellular toxins, superantigens, lipopolysaccharide-modifying enzymes, and proteins conferring serum resistance, etc. (1). These findings suggest that the horizontal transfer of phage genes has contributed significantly to the acquisition of new genetic traits and to the genetic diversity of bacteria (1, 9, 10). To truly appreciate the mechanisms behind phage-associated evolution, it is important to understand the frequency and range of transfer of phage genes.Most phage genomes consist of many genes derived from different origins (5, 8). Some genes are similar to those of other phages with phylogenetically different hosts or are found in the genomes of bacteria that are not the phage hosts. The mosaic nature of phage genomes has been known for some time, and a body of molecular genetic studies of phages to explain the mechanisms that drive this feature have been attempted previously (1, 5). More importantly, the horizontal transfer of phage genes has emerged as a major factor in the evolution of the phage genome. Since recombination between phage and phage/prophage can occur when these elements coexist in the same cell, coinfection with multiple phage species may result in the production of hybrid phage genomes (5). The pathways by which phages exchange genetic material vary dramatically in concert with host ranges. However, conventional plaque assays have shown that the host ranges of the phages studied are narrow. We hypothesized that phage genes can be transferred to more diverse species than previously thought.In order to accurately quantify DNA movement, gene targeting that does not require cultivation or gene expression is necessary (7). In situ DNA amplification methods allow the visualization of specific DNA sequences inside bacterial cells. In this study, we employed cycling primed in situ amplification-fluorescent in situ hybridization (CPRINS-FISH) to examine the possible range and frequency of the transfer of phage genes. CPRINS uses one primer and results in linear amplification of the target DNA inside cells, and multiply labeled fluorescent probe sets are applied for detection of the amplicons to improve the specificity and sensitivity of CPRINS (3). Previously, CPRINS-FISH did clarify the movement of DNA of a specific gene among Escherichia coli cells at the single-cell level (4).Enterobacterial phages P1 and T4 infect E. coli and have been well studied. P1 can exist as circular DNA within the bacterial cell as if it were a plasmid. Phage T4 is capable of undergoing only a lytic life cycle and not the lysogenic life cycle. Conventional methods using plaque assays have shown that the host of P1 and T4 is E. coli, but orthologous phage genes have been found in bacteria other than E. coli (6, 8). In the present study, strains of Enterobacteriaceae were allowed to grow on agar medium after the phage was adsorbed, and the maintenance of the transferred phage gene in the bacterial genomes was examined at the community level by quantitative real-time PCR and at the single-cell level by CPRINS-FISH.The following bacterial strains were used for maintenance experiments: Citrobacter freundii IFO 12681, Enterobacter aerogenes BM 2688, E. coli NBRC 12713, a Proteus mirabilis clinical isolate, Salmonella enterica serovar Enteritidis IID 640, and Yersinia enterocolitica IID 981. The bacterial strains were grown in Luria-Bertani (LB) medium (1% tryptone, 0.5% yeast extract, 0.5% NaCl; Nacalai Tesque Inc., Kyoto, Japan) at 37°C overnight.Stationary-phase cultures of 500 μl were incubated with 500 μl of SM buffer (50 mmol liter−1 Tris-HCl [pH 7.5], 100 mmol liter−1 NaCl, 8 mmol liter−1 MgSO4, 0.01% gelatin) containing the phage P1kc NBRC 20008 (2) or T4GT7 (11) at 37°C for 10 min at a multiplicity of infection of 1:1 (ratio of PFU of the phage to CFU of the recipient bacterium). The concentration of bacterial cells was adjusted to 109 cells ml−1. After 10 min of incubation, the diluted cell suspension (105 cells) was filtered through a polycarbonate filter (Advantec, Tokyo, Japan) with a pore size of 0.2 μm and a diameter of 25 mm. Cells trapped on the filter were cultured on LB agar medium at 37°C for 24 h. The filter was transferred into a microtube, and cells on the filter were suspended in 1 ml of sterile deionized water. The numbers of cells in the suspension and cells remaining on the filter were determined by using an epifluorescence microscope (see below) after staining of the samples with 1 μg ml−1 of 4′,6-diamidino-2-phenylindole (DAPI; Sigma-Aldrich Japan, Tokyo). The level of recovery of cells from the filter into sterile deionized water was about 99%. The cultured cells were subjected to real-time PCR and CPRINS-FISH.For real-time PCR, bacterial DNA was extracted using a QIAamp DNA isolation kit (Qiagen, Tokyo, Japan). The cell suspension was mixed with 10 mg ml−1 of lysozyme solution and incubated at 37°C for 1 h. DNA extraction was then performed according to the kit manufacturer''s instructions. Table Table11 lists the oligonucleotide primers for PCR and CPRINS and the polynucleotide probes used in the present study. Tail fiber genes from phages P1kc and T4GT7 were quantified by real-time PCR with a LightCycler system (Roche Diagnostics, Tokyo, Japan). LightCycler FastStart DNA master SYBR green I (Roche Diagnostics) was used with 5 mmol liter−1 Mg2+ and 0.5 μmol liter−1 (each) primers targeting the tail fiber genes of P1kc (P1-tail931f and P1-tail1148r) and T4GT7 (T4-tail2770f and T4-tail2983r). After a hot start for 10 min at 95°C, 40 cycles of PCR were run with denaturation at 94°C for 15 s, annealing at 60°C for 10 s, extension at 72°C for 10 s, and fluorescence detection at 83°C for 5 s. The known amounts of PCR products from the phage DNA (101 to 107 copies per reaction) were used for the standard curves to quantify the target DNA. To confirm the specificity of the reaction after real-time PCR, the PCR mixture was collected in a glass capillary and subjected to agarose gel electrophoresis in addition to a melting-curve analysis with the LightCycler system. The maintenance frequencies determined by real-time PCR were recorded as the copy number of the phage tail fiber gene per bacterial genome detected by staining with PicoGreen (Invitrogen, Tokyo, Japan) after cultivation of cells on LB agar medium for 24 h as described above. The frequencies were determined in triplicate for each sample. The increase in the phage gene copy number was determined by comparing the copy numbers in cells on the filter before and after cultivation. The phage gene copy number in cells on the filter was determined by the following formula: (total number of cells determined by DAPI staining) × (phage tail fiber gene copy number determined by real-time PCR)/(bacterial genome copy number determined by PicoGreen staining).

TABLE 1.

Probes and primers designed in this study
NameTargetTypeNucleotide sequence (5′-3′)
P1-tail931fTail fiber gene of phage P1PrimerAACGACCCGAATTACAGCAC
P1-tail1148rTail fiber gene of phage P1PrimerAGTGCTGCTGCAAGCTCATA
T4-tail2770fTail fiber gene of phage T4PrimerAGCACAAATGGTGAGCACAG
T4-tail2983rTail fiber gene of phage T4PrimerTTGCTACCGTGTGGGTATGA
T4-tail2664Tail fiber gene of phage T4ProbeGGCTTCAAGTACTGACTTAGGTACTAAAACCACATCAAGCTTTGACTATGGTACG
T4-tail2720Tail fiber gene of phage T4ProbeAAGGGAACTAACAGTACGGGTGGACACACTCACTCTGGTAGTGGTTCTA
T4-tail2769Tail fiber gene of phage T4ProbeTAGCACAAATGGTGAGCACAGCCACTACATCGAGGCATGGAATGG
T4-tail2818Tail fiber gene of phage T4ProbeGGTGTAGGTGGTAATAAGATGTCATCATATGCCATATCATACAGGGCGGG
T4-tail2869Tail fiber gene of phage T4ProbeGGGAGTAACACTAATGCAGCAGGGAACCACAGTCACACTTTCTCTTTTGGG
T4-tail2922Tail fiber gene of phage T4ProbeTAGCAGTGCTGGCGACCATTCCCACTCTGTAGGTATTGGTGCTCATA
Open in a separate windowCPRINS-FISH targeting the tail fiber gene of phage T4GT7 was performed as described by Kenzaka et al. (3, 4), except for the probe/primer sequences and thermal conditions. After cell wall permeabilization by lysozyme treatment (3), the CPRINS reaction was performed under the following conditions: a hot start at 95°C for 9 min, denaturation at 94°C for 1 min, annealing at 60°C for 30 s, and extension at 72°C for 1.5 min for primer T4-tail2983r. Amplification was repeated for 30 cycles by using a thermal cycler (PTC-200; Bio-Rad Laboratories, Inc.). After amplification, filters were rinsed with 0.1% Nonidet P-40 and sterile deionized water, dehydrated in 99% ethanol, and vacuum dried. Hybridization with Alexa Fluor 546-labeled polynucleotide probes (T4-tail2664, T4-tail2720, T4-tail2769, T4-tail2818, T4-tail2869, and T4-tail2922), washing, and DAPI staining were performed as described in a previous study (4). In order to exclude the possibility of nonspecific probe binding to cell structures other than the target DNA in the target cells, FISH using laboratory strains without amplification of target DNA and CPRINS-FISH targeting the tail fiber genes in E. coli strains that did not carry the genes were performed.In order to examine the infection ranges of phages, plaque assays and direct counting of phages were performed. Plaque assays were performed with LB soft agar (0.8% agar) as described by Kenzaka et al. (4). For the direct counting, phages were stained with 5× SYBR gold (Invitrogen, Tokyo, Japan) and trapped onto an Anodisc filter (Whatman Japan, Tokyo) with a pore size of 0.02 μm and a diameter of 25 mm.The cells or phage particles on the filters were observed under an epifluorescence microscope (E-400; Nikon, Tokyo, Japan) with the Nikon filter sets UV-2A (EX300-350, DM400, and BA420) for DAPI, B-2A (EX450-490, DM505, and BA520) for SYBR gold, and HQ-CY3 (G535/50, FT565, and BP610/75) for Alexa Fluor 546. Images were acquired using a Retiga 2000R cooled charge-coupled device camera (QImaging, Surrey, BC, Canada), and at least 2,000 DAPI- or SYBR gold-stained objects per sample were counted. The maintenance frequencies determined by CPRINS-FISH were recorded as the number of CPRINS-FISH-positive cells divided by the total direct count of recipient cells after cultivation as described above. The frequencies were determined in triplicate for each sample.After cultivation on LB agar medium for 24 h, the total number of cells on the filter as determined by DAPI staining increased by 8.7 × 102- to 1.1 × 104-fold (Table (Table2).2). Real-time PCR showed that the phage P1kc gene copy number increased only in plaque-forming strains (E. coli and E. aerogenes) and not in non-plaque-forming strains (Table (Table2).2). In contrast, the phage T4GT7 gene copy number increased in both plaque-forming and non-plaque-forming strains by 7.6 × 101- to 7.0 × 104-fold. The maintenance frequencies were more than 10−2 gene copies per bacterial genome (Table (Table2).2). Direct observation via epifluorescence microscopy showed that progeny phages were not produced in the non-plaque-forming strains (Table (Table2),2), and thus, fragments of phage genes were thought to integrate into the genomes of non-plaque-forming strains and replicate with the bacterial genomes.

TABLE 2.

Frequencies of maintenance of phage P1kc and T4GT7 genes in Enterobacteriaceae strains
PhageRecipientResult for infection range indicator:
Increase in total no. of cellscIncrease in phage gene copy no. (SD)dMaintenance frequency (SD) as determined bye:
Plaque formationaProduction of progenybReal-time PCRCPRINS-FISH
P1kcC. freundii7.0 × 103None<1.5 × 10−3ND
E. aerogenes++1.7 × 1037.7 × 103 (6.5 × 103)5.0 × 100 (4.2 × 100)ND
E. coli++7.2 × 1035.5 × 103 (2.7 × 103)9.1 × 10−1 (0.5 × 10−1)ND
P. mirabilis7.4 × 103None<1.5 × 10−3ND
S. Enteritidis8.4 × 103None<1.7 × 10−4ND
Y. enterocolitica4.6 × 103None<1.8 × 10−4ND
T4GT7C. freundii1.5 × 1037.5 × 103 (4.0 × 103)8.3 × 10−1 (4.4 × 10−1)8.6 × 10−2 (3.4 × 10−2)
E. aerogenes++8.7 × 1021.2 × 103 (0.8 × 103)8.0 × 10−1 (5.0 × 10−1)4.0 × 10−1 (0.7 × 10−1)
E. coli++1.1 × 1047.0 × 104 (2.7 × 104)8.0 × 101 (3.0 × 10)2.1 × 10−1 (0.4 × 10−1)
P. mirabilis4.0 × 1035.8 × 103 (4.2 × 103)3.3 × 10−1 (2.4 × 10−1)3.4 × 10−2 (2.2 × 10−2)
S. Enteritidis1.0 × 1047.6 × 101 (5.0 × 101)1.0 × 10−2 (0.7 × 10−2)8.8 × 10−2 (2.0 × 10−2)
Y. enterocolitica3.6 × 1031.6 × 104 (0.4 × 104)6.1 × 10−1 (1.6 × 10−1)2.2 × 10−2 (2.9 × 10−2)
Open in a separate windowaPlaque formation on soft agar was tested.bThe production of progeny phage particles was observed via epifluorescence microscopy.cThe increase (n-fold) in the total number of cells during bacterial growth for 24 h was determined via epifluorescence microscopy.dThe increase (n-fold) in the copy number of the phage tail fiber gene during bacterial growth for 24 h was determined by real-time PCR. Values in parentheses indicate standard deviations of results for triplicate samples.eMaintenance frequencies were determined by real-time PCR and CPRINS-FISH analyses targeting the phage tail fiber gene and are shown as the phage tail fiber gene copy numbers per bacterial genome and the numbers of gene-positive cells divided by the total numbers of cells, respectively. Values in parentheses indicate standard deviations of results for triplicate samples. ND, not determined.Real-time PCR provided a copy number for the target phage gene in the whole population, but the location of the target phage gene and the frequency of cells carrying the target gene were unclear. In addition, bacterial genomic DNA, which was measured using PicoGreen, included phage DNA, and thus the frequencies measured by dividing by the amount of bacterial genomic DNA were probably less accurate than those measured as described below. In order to confirm that the phage gene was located inside bacterial cells and determine a more accurate maintenance frequency for total cells, CPRINS-FISH targeting the tail fiber gene of phage T4GT7 was performed. CPRINS-FISH visualized the target phage gene in individual cells under an epifluorescence microscope (Fig. (Fig.1).1). It showed that the frequencies of maintenance of the tail fiber gene, expressed as the number of gene-positive cells divided by the total number of cells, were 2.1 × 10−1 to 4.0 × 10−1 for plaque-forming strains after growth on LB medium for 24 h (Table (Table2).2). Since phage T4GT7 is capable of undergoing only a lytic life cycle, CPRINS-FISH would detect cells in which the phage gene was replicating. For non-plaque-forming strains, the maintenance frequencies were 2.2 × 10−2 to 8.8 × 10−2 (Table (Table2).2). If the gene was amplified by the CPRINS reaction outside bacterial cells, the amplicon would not accumulate inside bacterial cells and they would not exhibit bright fluorescence. Therefore, CPRINS-FISH proved that a part of the phage T4GT7 gene was located inside cells of non-plaque-forming strains. The tail fiber gene is responsible for the phage tail structure. The DNA sequences of the phage genes responsible for phage morphology have been found in many bacterial genomes (1, 5).Open in a separate windowFIG. 1.Visualization of E. coli cells carrying the tail fiber gene transferred by phage T4GT7. (A) After being mixed with phages for 10 min, E. coli NBRC 12713 cells were cultured for 24 h and subjected to CPRINS-FISH targeting the phage gene. Only cells having amplified tail fiber gene products emitted the fluorescence of the Alexa Fluor 546-labeled probe under green excitation (exposure, 0.5 s). (B) All DAPI-stained bacterial cells were visualized under UV excitation (exposure time, 0.1 s).In order to explore the effect of integration of the phage gene into the bacterial genome on bacterial heredity, we determined the mutation frequency for a C. freundii strain that acquired the phage T4GT7 gene. Two colonies which acquired the phage T4GT7 gene were screened by colony PCR with T4-tail2770f and T4-tail2983r primers and designated Cik8-1 and Cik8-4. Mutation frequencies were determined with LB medium containing 150 μg ml−1of rifampin (rifampicin) or 10 μg ml−1of nalidixic acid. The mutation frequencies associated with nalidixic acid resistance decreased by 12- to 240-fold and the frequencies associated with rifampin resistance decreased by 40- to 83-fold compared to those for the parent strains (Fig. (Fig.2).2). Mutation increases genetic variation. The decreased mutation frequency would contribute to the genetic stability of the genome in individual cells but not to genetic variation in the population. Our results show that phage T4GT7 was capable of affecting the genomic properties of C. freundii, which was thought previously not to be the host, although the mechanism by which mutation frequencies decreased remains unknown. Further experiments are required to clarify the molecular mechanism by which mutation frequencies altered after gene transfer.Open in a separate windowFIG. 2.Mutation frequencies for T4GT7-infected C. freundii strains. Mutation frequencies were determined with LB agar medium containing nalidixic acid or rifampin. Cik8-1 and Cik8-4 were strains which acquired a phage gene transferred from phage T4GT7. Cik1 and Cik2 were the parent strains.In summary, during growth on agar medium after the phage was allowed to be adsorbed by strains of Enterobacteriaceae, the phage P1kc gene was not maintained in non-plaque-forming strains but the phage T4GT7 gene was maintained in more diverse species than previously expected. The transfer of foreign DNA molecules (DNA entry) into a bacterium is an important first step in genetic diversification through horizontal gene transfer. A previous study reported that phage P1kc is capable of injecting DNA into non-plaque-forming E. coli cells (4), but the phage P1kc gene was not maintained during bacterial growth in the present study. The results showing the difference in maintenance between phage P1kc and T4GT7 genes suggest that the maintenance of transferred phage genes depends on phage gene sequences or other phage factors. When maintained, the phage gene could alter the mutation frequency for bacteria that acquired the gene, affecting the genomic variability at the population level. Conventionally, phage-bacterium interaction has been studied with certain models consisting of a phage and a bacterium in which the phage can multiply (12, 13). Our results indicate the importance of the dynamic of phage genes among diverse bacteria that were previously thought not to be hosts and the hereditary impact of phage gene transfer on such bacteria.  相似文献   

19.
Bacteriophage resistance mechanisms which are derived from a bacteriophage genome are termed Per (phage-encoded resistance). When present in trans in Lactococcus lactis NCK203, Per50, the cloned origin of replication from phage 50, interferes with 50 replication. The per50 fragment was found to afford negligible protection to NCK203 against 50 infection when present in a low-copy-number plasmid, pTRK325. A high-copy-number Per50 construct (pTRK323) dramatically affected 50 infection, reducing the efficiency of plaquing (EOP) to 2.5 × 10-4 and the plaque size to pinhead proportions. This clone also afforded significant protection against other related small isometric phages. Per31 was cloned from phage 31 and demonstrated to function as an origin of replication by enabling replication of per31-containing plasmids, in NCK203, on 31 infection. A low-copy-number Per31 plasmid (pTRK360) reduced the EOP of 31 on NCK203 to 0.3 and the plaque diameter from 1.5 to 0.5 mm. When this plasmid was cloned in high copy number, the EOP was further reduced to 7.2 × 10-7 but the plaques were large and contained Per31-resistant phages. Characterization of these “new” phages revealed at least two different types that were similar to 31, except that DNA alterations were noted in the region containing the origin. This novel and powerful abortive phage resistance mechanism should prove useful when directed at specific, problematic phages.  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