Next Article in Journal
Effects of Staphylococcus aureus Bacteriophage K on Expression of Cytokines and Activation Markers by Human Dendritic Cells In Vitro
Next Article in Special Issue
The Revisited Genome of Bacillus subtilis Bacteriophage SPP1
Previous Article in Journal
Persistence and Intra-Host Genetic Evolution of Zika Virus Infection in Symptomatic Adults: A Special View in the Male Reproductive System
Previous Article in Special Issue
Unprecedented Diversity of ssDNA Phages from the Family Microviridae Detected within the Gut of a Protochordate Model Organism (Ciona robusta)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Characterization of the Escherichia coli Virulent Myophage ST32

1
College of Resources and Environment, University of Chinese Academy of Sciences, Beijing 101408, China
2
Département de Biochimie, de Microbiologie, et de Bio-Informatique, Faculté des Sciences et de Génie, Université Laval, Québec City, QC G1V 0A6, Canada
3
Félix d’Hérelle Reference Center for Bacterial Viruses, Faculté de médecine dentaire, Université Laval, Québec City, QC G1V 0A6, Canada
*
Authors to whom correspondence should be addressed.
Current address: SyntBioLab Inc., Lévis, QC G6W 0L9, Canada.
Viruses 2018, 10(11), 616; https://doi.org/10.3390/v10110616
Submission received: 15 October 2018 / Revised: 4 November 2018 / Accepted: 6 November 2018 / Published: 7 November 2018
(This article belongs to the Special Issue Bacteriophage Genomes and Genomics: News from the Wild)

Abstract

:
The virulent phage ST32 that infects the Escherichia coli strain ST130 was isolated from a wastewater sample in China and analyzed. Morphological observations showed that phage ST32 belongs to the Myoviridae family, as it has an icosahedral capsid and long contractile tail. Host range analysis showed that it exhibits a broad range of hosts including non-pathogenic and pathogenic E. coli strains. Interestingly, phage ST32 had a much larger burst size when amplified at 20 °C as compared to 30 °C or 37 °C. Its double-stranded DNA genome was sequenced and found to contain 53,092 bp with a GC content of 44.14%. Seventy-nine open reading frames (ORFs) were identified and annotated as well as a tRNA-Arg. Only nineteen ORFs were assigned putative functions. A phylogenetic tree using the large terminase subunit revealed a close relatedness with four unclassified Myoviridae phages. A comparative genomic analysis of these phages showed that the Enterobacteria phage phiEcoM-GJ1 is the closest relative to ST32 and shares the same new branch in the phylogenetic tree. Still, these two phages share only 47 of 79 ORFs with more than 90% identity. Phage ST32 has unique characteristics that make it a potential biological control agent under specific conditions.

1. Introduction

Pathogenic Escherichia coli (E. coli) is a common zoonotic agent that poses a significant threat to public health and safety. Shiga-toxin-producing E. coli (STEC) strains are one of the most important foodborne pathogens [1,2]. The Shiga toxin (Stx) cleaves ribosomal RNA, thereby disrupting protein synthesis and killing the intoxicated epithelial or endothelial cells [3]. STEC infection can result in diseases such as diarrhea, hemorrhagic colitis, and hemolytic-uremic syndrome (HUS) in humans and animals. These diseases are subjected to various pharmaceutical treatments including antibiotics, such as ampicillin, streptomycin, sulfonamides, and oxytetracycline [4,5].
It is well-known that the use of antibiotics can lead to the spread of antibiotic-resistant bacteria in the environment, which poses a risk to human health [6,7,8]. Antimicrobial resistance of E. coli is an issue of the utmost importance since it can affect both animals and humans [9]. This bacterial species has a great capacity to accumulate antibiotic resistance genes, mostly through horizontal gene transfer [10,11]. For example, the intensive use of various antibiotics in aquaculture has had significant benefits to the fish industry but it has also led to serious negative effects on the environment, including the emergence of a pool of antibiotic-resistant bacteria and transferable resistance genes [6,12,13,14]. Some of those antibiotic-resistance genes can be transferred horizontally from bacteria in aquatic environments to pathogenic bacteria, affecting land animals and humans [13,14]. Moreover, the transmission of resistant clones and resistance plasmids of E. coli from poultry to humans has also been identified [15,16].
Of note, the highest rate of antibiotic-resistance genes was found in E. coli strains of a sewage treatment plant that treats both municipal and hospital sewage [17,18,19]. Although wastewater treatment processes reduce the number of bacteria in sewage by up to 99%, E. coli cells can still reach the receiving water and contribute to the dissemination of resistant bacteria into the environment [20]. As a result, antimicrobial resistance in E. coli is considered one of the major challenges for both humans and animals at a worldwide scale and it needs to be considered as a real public health concern.
Alternative strategies must be developed to reduce the risk associated with the dissemination of antimicrobial resistance and to control the risk of disease transmission. The use of phages as biocontrol agents has received increasing attention recently as a possible alternative or as a complement to antibiotics [21,22,23,24,25,26,27]. For example, bacteriophages have demonstrated efficacy in controlling pathogenic bacterial populations in, among others, poultry meat [28], aquaculture [23], wastewater, and minimally processed, ready-to-eat products and fresh fruits [25,29,30,31]. It can also help to remove bacteria on chicken skin [22] and on dairy cows at different lactation stages [26]. Interestingly, these bacterial viruses can be highly specific to a single bacterial species or to only a few strains within that species, or can productively infect a range of bacterial species [32,33].
In the present study, we used the host pathogenic E. coli ST130 (flagellin H21) carrying Shiga toxin (stx1, stx2) genes to isolate and characterize a new virulent coliphage, named ST32. This phage was isolated from sewage water and possesses appealing characteristics that could be of interest for specific biocontrol purposes.

2. Materials and Methods

2.1. Bacterial Strain

Escherichia coli ST130 was obtained from the Chinese Center for Disease Control and Prevention (China CDC). This bacterium was used as the phage host.

2.2. Phage Isolation and Purification

Phage ST32 was isolated from a wastewater sample of a sewage treatment plant in Beijing, and was propagated and titrated using methods described previously [34]. Samples were filtered with a 0.45 μm sterile PES syringe filter (Sarstedt, Nümbrecht, Germany, catalog number 83.1826), and then, 2.5 mL of the filtered sample and 1 mL of an overnight E. coli ST130 culture were added to 7.5 mL of Luria broth (LB) (1% bacto-tryptone, 0.5% bacto-yeast extract, and 1% NaCl) incubated overnight with agitation (200 rpm) at 37 °C. The resulting supernatant was filtered and serially diluted in order to isolate phage plaques using the double layer agar method. Briefly, 100 μL of serially diluted lysate and 100 μL of an overnight E. coli culture were added to 4 mL of LB supplemented with 0.75% agar. The inoculated soft agar was then poured into LB plates (1.5% agar). The plates were incubated overnight at 37 °C, and single phage plaques were picked, propagated, and purified three times.

2.3. Phage Morphology

Phage ST32 was purified and concentrated by CsCl gradient as described previously [35]. Phage particles were stained with 2% (w/v) aqueous uranyl acetate on a carbon-coated grid and were observed using a JEM-1230 transmission electron microscope (JEOL, Tokyo, Japan) [36]. Over 10 specimens were observed and used for size determination.

2.4. Host Range

The host range of phage ST32 was tested on 73 bacterial strains from different genera, species, and serotypes using the spot test method and a diluted phage lysate. In brief, 200 μL of overnight culture of E. coli, Shigella, Salmonella, or Citrobacter was mixed with 3.5 mL of LB containing 0.75% (w/v) soft agar. The inoculated soft agar was then poured on LB (1.5% (w/v) agar) plates. Then, serial dilutions of phage lysate were made in buffer (50 mM Tris−HCl at pH 7.5, 100 mM NaCl, and 8 mM MgSO4). Five microliters of various serial dilutions (100, 10−2, 10−4 and 10−6) was spotted on the top agar. After overnight incubation at 37 °C, phage plaques or lysis zones were recorded.
Moreover, the propagation of phage ST32 on non-pathogenic host strains (E. coli HER1036, HER1155, HER1222, HER1315, HER1375, and HER1536) was compared to that of the pathogenic E. coli ST130 strain. In brief, the strains were grown at 37 °C in LB medium until an optical density at 600 nm (OD) of 0.25. Then, approximately 106 PFU·mL−1 of phage ST32 was added. The phage-infected cultures were incubated with agitation at 37 °C until complete bacterial lysis was achieved. The phage lysate was centrifuged to remove cell debris, and the supernatant was filtered using a 0.45 μm syringe filter. Then, the phage lysates were serially diluted in buffer and titered by spot test as described above. Of note, the pathogenic E. coli ST130 strain was used for phage titration after propagation.

2.5. One-Step Growth Curve Assay

The influence of the incubation temperature on phage ST32 plaque formation was investigated by spot test as described above. Following the spot test assay, the plates were incubated at various temperatures (ranging from 10 to 42 °C).
One-step growth curve assays were also performed in triplicate. Briefly, phages were mixed with 2 mL of a mid-exponential phase culture of E. coli ST130 (OD of 0.8) with a starting multiplicity of infection (MOI) of 0.05. ST32 phages were allowed to adsorb to E. coli ST130 cells for 5 min at various temperatures (20, 30, or 37 °C), and then the mixture was centrifuged for 1 min at 16,000× g. The pellet was resuspended, diluted, and added to 10 mL of LB. This suspension was incubated at three different temperatures (20, 30, or 37 °C) without agitation, and samples were taken to test the phage titers. The phage titer of each sample was determined using the double layer agar method. All plates were incubated overnight at 30 °C. The burst size was calculated by subtracting the initial titer from the final titer and then dividing by the initial titer. The latent phase corresponded to the middle of the exponential phase of the curve [37]. The data were analyzed under a one-way analysis of variance (ANOVA) followed by a Tukey test to correct the p-values for the multiple comparisons. Significant differences were reported at an alpha level of 1%.

2.6. E. coli ST130 Growth

E. coli ST130 growth was also determined at various temperatures using OD and recorded in triplicate. In brief, 200 μL of ST130 overnight culture was added to 5 mL of LB medium. Then, inoculated samples were incubated with agitation (200 rpm) at 20, 30, and 37 °C. The OD was measured at intervals of 30 min.

2.7. Sequencing and Analysis

Phage DNA was extracted as described elsewhere [38]. DNA was sequenced using the Illumina Hiseq (PE250) platform at Beijing Fixgene Tech Co., Ltd. (Beijing, China). More than 5000-fold coverage of the phage genome was generated. The paired-end reads were assembled using ABySS v. 1.3.6. Open reading frames (ORFs) were predicted using PHASTER [39]. The identified ORFs were confirmed with GeneMark.hmm prokaryotic (http://exon.gatech.edu/GeneMark/gmhmmp.cgi) and ORF Finder (https://www.ncbi.nlm.nih.gov/orffinder/). ORFs were considered candidates for evaluation when they encoded 45 or more amino acids (aa) and possessed both a conserved Shine–Dalgarno sequence (5′-AGGAGGU-3′) and a start codon (AUG, UUG, or GUG). BLASTp was used to identify the putative functions of the proteins. Hits were considered valid when the E-value was lower than 10−3. The percent identity between proteins was calculated by dividing the number of identical residues by the size of the smallest protein. The theoretical molecular weights (MW) and isoelectric points (pI) of the proteins were obtained using tools available on the ExPASy webpage (http://web.expasy.org/compute_pi/). The bioinformatic tool tRNAscan-SE (http://lowelab.ucsc.edu//tRNAscan-SE/) was used for tRNA detection.

2.8. Terminase Tree

A phylogenetic tree was generated based on the large terminase subunit amino acid sequences of phage ST32 and multiple phages available in databases sharing sequence identity. The corresponding phage protein sequences were retrieved from GenBank (https://www.ncbi.nlm.nih.gov/). In constructing the terminase phylogenetic tree, these sequences were aligned with MAFFT [40] using the E-INS-i alignment algorithm. Thereafter, MAFFT-profile alignment was processed, as previously described [41], in order to generate the tree. Briefly, ProtTest 3.2 was applied to find an appropriate model of amino acid substitution and was implemented in PhyML 3.0 to calculate a maximum likelihood tree. Finally, the Shimodaira–Hasegawa-like procedure was used to determine the branch support values and the Newick utility package was used to render the trees.

2.9. Nucleotide Sequence Accession Number

The complete genome sequence of phage ST32 was deposited in GenBank under the accession number MF044458.2.

3. Results and Discussion

3.1. Phage Morphology

The morphological characteristics of phage ST32 were examined by transmission electron microscopy. Electron micrographs (Figure 1) showed that phage ST32 has an icosahedral capsid with an apex diameter of 64 ± 6 nm and a long contractile tail with a length of 132 ± 9 nm. These morphological features [42] indicate that phage ST32 belongs to the Caudovirales order and the Myoviridae family.

3.2. Host Range

Currently, phages are tested for biocontrol purposes against E. coli strains that may cause infections [43,44] or used as indicators of coliform contamination [45]. The host range plays a key role in the selection of any given phage for therapy or biocontrol purposes, as a broad host range phage is likely to kill multiple strains of a given bacterial species and maybe even beyond the species or genus levels for enterophages [43,46].
To this end, the host range of phage ST32 was evaluated on 73 bacterial strains obtained from the Félix d’Hérelle Reference Center for Bacterial Viruses (Table 1). Phage ST32 was able to infect 10 strains (14%), including four pathogenic and six non-pathogenic strains. Pathogenic strains infected by phage ST32 included four E. coli strains of multiple serotypes. In order to reduce the risk of possible harmful substances from the pathogenic host strain in phage lysate, we evaluated the ability of phage ST32 to propagate on its sensitive, non-pathogenic host strains (++++; Table 1). The results showed that phage ST32 was propagated to a high titre (109 PFU/mL) when using five (E. coli HER1036, HER1222, HER1315, HER1375 and HER1536) out of six of these strains.
Based on the above, phage ST32 has a broad host range, infecting both pathogenic and non-pathogenic E. coli strains. These features led us to consider phage ST32 to be a potential biocontrol agent rather than a therapeutic agent. In order to use phage ST32 as a biocontrol agent, we further studied the influence of temperature on its lytic activity as well as on the growth of E. coli host strain ST130.

3.3. One-Step Growth Curve

The influence of temperature on plaque formation was first analyzed by spot test at 10, 20, 30, 37, and 42 °C. The results showed that phage ST32 produced clear plaques at dilutions of 10−1 to 10−7 when plates were incubated at 10, 20, 30, and 37 °C. Turbid plaques were seen but only at 42 °C. A one-step growth curve was conducted at 20, 30, and 37 °C to determine its latent period and burst size at these temperatures. Moreover, the growth of the bacterial host strain followed under the same conditions.
As indicated by the results of the one-step growth curve experiments (Figure 2a), the burst size of phage ST32 was very low at 37 °C, to the extent that only 2 ± 0.1 new virions were released per infected cell with an estimated latent period of 55 ± 6 minutes. When the phage-infected cells were incubated at 30 °C, the average burst size of phage ST32 increased to 64 ± 30 new virions per infected cell, and the latent period remained the same (54 ± 2 min). Interestingly, the burst size of phage ST32 was significantly higher when the infected cells were incubated at 20 °C with an average of 602 ± 159 new virions being released per infected cell. Conversely, the latent period increased to approximately 102 ± 10 min. Of note, the growth of the E. coli ST130 host strain was much faster at 30 °C and 37 °C compared to that at 20 °C (Figure 2B). Nonetheless, phage ST32 could still kill its host at these temperatures.
Phage ST32 is evidently part of a low-temperature (LT) phage group with an optimum burst at 20 °C [48]. Of note, this phage was isolated from a wastewater sample of a sewage treatment plant in Beijing that has a temperature of about 20 °C. Therefore, it appears to be adapted to replicate at such ambient-like temperatures. These features make this phage a potential agent for the biocontrol of E. coli. For instance, it could be used to control pathogenic bacteria present in wastewater where physical conditions, such as temperature, are optimal for its lytic activity. Moreover, it may provide an effective intervention against foodborne pathogens and spoilage bacteria in minimally processed, ready-to-eat products and fresh fruits [29,30,31]. It could also help to remove bacteria from poultry meat that are often found to be contaminated with potentially pathogenic micro-organisms [28]. In order to support its potential as a biocontrol agent, we further characterized phage ST32 at the genomic and phylogenetic levels.

3.4. Genomic Features of Phage ST32

The genome sequence of phage ST32 consists of a double-stranded DNA molecule of 53,092 bp with a GC content of 44.14% as well as 79 open reading frames (ORFs) and a tRNA (Table 2). The tRNA-Arg of 95 bp (from 15,909 bp to 16,003 bp), without an intron, found in the genome of phage ST32, shares 99% identity with phage phiEcoM-GJ1 [49]. tRNA-Arg is often found in phage genomes [50]. The 79 ORFs have the same transcriptional orientation, and ATG is the most common initiation codon (81.0%), followed by GTG (11.4%) and TTG (7.6%).
Based on the BLASTp analyses, 19 of the 79 ORFs (24.1%) were assigned a putative function, including lysis, capsid, and tail morphogenesis as well as transcription and DNA replication. The functions of the remaining sixty putative ORFs remained unknown, and they were annotated as hypothetical proteins. Besides the predicted protein functions, Table 2 shows the predicted size, the genomic position, the transcriptional orientation, and the closest phage protein homolog. In several cases, protein homologies were with proteins of phages belonging to the Podoviridae or Myoviridae families. The best matches for a large portion of these ORFs were with proteins of the Enterobacteria phage phiEcoM-GJ1 belonging to the Myoviridae family [49]. Thereafter, phylogenetic trees were constructed for further investigation of the relatedness of phage ST32 to other phages.

3.5. Phylogeny of Phage ST32

The conserved sequence of the large terminase subunit (ORF51) has been used previously to study the phylogeny of numerous phages [41,42]. As an ATP-driven protein motor, the phage terminase is generally a hetero-oligomer composed of two subunits (small and large) that translocates the phage genome into the preformed capsid. The large subunit usually possesses endonucleolytic and ATPase activities [51,52]. A phylogeny tree, based on the amino acid sequences of the large terminase subunit (ORF51), was constructed to examine the evolutionary relationships between phage ST32 and other phage genomes (Figure 3). The phylogeny tree supported the finding that phage ST32 belongs to the Myoviridae family. Moreover, phage ST32 was on the same branch as phage phiEcoM-GJ1 (EF460875.1), indicating a close relatedness between these two phages and suggesting that they belong to the same new cluster. Interestingly, phiEcoM-GJ1 phage currently belongs to an unclassified genus of the Myoviridae family [49]. Moreover, the tree indicated that the closest evolutionary relatives to both phages were the Pectobacterium virulent phages PM1 [53] and PP101 and the Erwinia virulent phage vB_EamM-Y2 [54]. This relatedness between the PM1, vB_EamM-Y2, and phiEcoM-GJ1 phages was revealed in a previous study [53].
Thereafter, we compared the percent identity between the genome sequences of these five phages. Our results showed that the percentage of nucleotide sequence identity between phages in the same branch was relatively high compared to phages in different branches. For example, the percent identity between phages ST32 and PM1 did not exceed 36% compared to 84.5% between the two Pectobacterium phages PM1 and PP101.

3.6. Comparative Genomic Analysis

The genomic sequences of the ST32, phiEcoM-GJ1, PM1, PP101, and vB_EamM-Y2 phages were further analyzed, compared, and aligned using the deduced amino acid sequences of all of the ORFs. A comparative analysis showed that when using a cut-off of 80% identity, phage ST32 shares 54 proteins with phage phiEcoM-GJ1, while the Pectobacterium phages PM1 and PP101 share 53 proteins. On the other hand, at the same cut-off, the Erwinia phage vB_EamM-Y2 shares only three proteins with the other four phages (Figure 4). Notably, at 70% identity, this number went up to 14 proteins, as indicated by the gray shading in Figure 4.
Interestingly, with more than 60% identity, 31 proteins were shown to be shared by the four phages ST32, phiEcoM-GJ1, PM1, and PP101. Based on this comparative analysis, these five phages can be separated into three distinct groups, which is consistent with their three-branch division in the phylogenetic tree (Figure 3). Based on the close relatedness between phages ST32 and phiEcoM-GJ1 shown in the above analysis, we compared them further.
The genomic organization of phage ST32 compared to phage phiEcoM-GJ1 (Figure 4) showed that all genes from both phage genomes have the same transcription orientation (5′ to 3′ from left to right in the figure). Moreover, 47 of 79 ORFs share more than 90% identity, of which eight (ORF42, ORF49, ORF50, ORF52, ORF60, ORF64, ORF68 and ORF69) are 100% identical (Table 2). The latter are proteins with hypothetical functions. Interestingly, six of these eight ORFs are found in very few phage genomes available in databases [49,53], including the ones closely related to phage ST32 that were used for the genomic comparison in Figure 4.
The global analysis of both phage genomes showed that they are organized into functional clusters to which different roles can be assigned. First, both phages share a cluster of a high number of small genes at the beginning of the genome (starting from ORF2), reminiscent of those on of T4 coliphages which are involved in host takeover [42,49,55] (Figure 4). Most of the phage ST32 ORFs in this cluster share less than 90% identity with those of phage phiEcoM-GJ1 (Table 2). Then, downstream of the genome, several putative replication-related genes were identified, encoding a single-stranded DNA-binding protein (ORF19), thymidylate synthase (ORF38), helicase/primase (ORF39), DNA polymerase (ORF40), 5′-3′ exonuclease (ORF43), DNA ligase (ORF45), deoxyuridine 5′-triphosphate nucleotidylhydrolase (ORF47), and ribonucleotide reductase beta subunit (ORF79). In addition to the replication-related genes, the last ORFs in the genome of phages ST32 and phiEcoM-GJ1 encode a ribonucleotide reductase beta subunit. In this regard, it is interesting to note that the ORF1 of both phages encodes a single-subunit RNA polymerase which is a feature of phages of the T7 group of the Podoviridae [49]. These transcription-related ORFs share more than 90% identity (Table 2). Then, downstream of the replication-related genes, we identified a cluster of DNA packaging, capsid, and tail morphogenesis conserved genes sharing more than 90% identity, except for two ORFs, ORF66 and ORF76, encoding for two putative tail fiber proteins and sharing 76% and 72% identity, respectively.
Finally, further main differences were identified between the two phages. For example, three ORFs were only found throughout the genome of phage phiEcoM-GJ1, encoding for three putative HNH endonucleases (ORF34phiEcoM-GJ1, ORF36phiEcoM-GJ1, and ORF47phiEcoM-GJ1) [49]. Moreover, five additional ORFs (ORF17, ORF18, ORF33, ORF35, and ORF56) encoding proteins with unknown functions were found in the genome of phage ST32 but not in that of phage phiEcoM-GJ1. Interestingly, the best match for one (ORF56) of these five ORFs was with that of the Erwinia phage vB_EamM-Y2, which is closely related to phage ST32, as shown in the phylogenetic tree (Figure 3).

4. Conclusions

In this study, the virulent phage ST32 was isolated from wastewater using the pathogenic host E. coli ST130. Morphological and genomic characterization showed that phage ST32 belongs to the Myoviridae family. Host range analysis showed that it can infect a broad range of hosts including non-pathogenic and pathogenic bacteria. Moreover, phage ST32 has a very high burst size at 20 °C which is far from the optimal growth of its host. Phylogenetic analysis, based on the large terminase subunit (ORF51), revealed a close relatedness with the Enterobacteria phage phiEcoM-GJ1 belonging to an unclassified genus of the Myoviridae family. Interestingly, both phages are part of a new branch in the phylogeny. Moreover, neighboring branches carry unclassified Myoviridae relatives, among others, the Pectobacterium phages PM1 and PP101 and the Erwinia phage vB_EamM-Y2. A comparative genomic analysis of the five phages based on nucleotide and amino acid sequences, showed that phage phiEcoM-GJ1 is by far the closest relative to phage ST32. A more detailed genomic comparison between these two phages showed that 47 of 79 ORFs in the phage ST32 genome have more than 90% identity with the phage phiEcoM-GJ1. Many of these ORFs had few homologs in databases. Some striking differences were detected, including the absence of three putative HNH endonucleases of phiEcoM-GJ1 ORFs in phage ST32. On the other hand, five additional ORFs with unknown functions were detected in the phage ST32 genome. Taken together, the newly characterized phage ST32 has appealing and unique characteristics that make it a potential biological control agent under specific conditions.

Author Contributions

H.L., G.M.R., X.L. and S.M. conceived and designed the experiments. H.L. and D.M.T. performed the experiments. H.L., H.G. and G.M.R. analyzed the data. S.J.L. contributed tools for comparative genome analysis. H.L., H.G., G.M.R., X.L. and S.M. wrote and corrected the draft. All authors reviewed the manuscript.

Funding

This work was supported by a National Natural Science Foundation of China grant (No. 50978250 and No. 51378485) and the China Scholarship Council (Student ID: 201704910749). S.M. holds the Canada Research Chair in Bacteriophages.

Acknowledgments

We wish to thank Yanwen Xiong (Collaborative Innovation Center for Diagnosis and Treatment of Infectious Diseases, State Key Laboratory of Infectious Disease Prevention and Control, National Institute for Communicable Disease Control and Prevention, Chinese Center for Disease Control and Prevention, Changping, Beijing 102206, China) for supplying E. coli ST130.

Conflicts of Interest

There are no conflicts of interest.

References

  1. Caprioli, A.; Scavia, G.; Morabito, S. Public health microbiology of Shiga toxin-producing Escherichia coli. Microbiol. Spectr. 2014, 2. [Google Scholar] [CrossRef] [PubMed]
  2. Melton-Celsa, A.; Mohawk, K.; Teel, L.; O’Brien, A. Pathogenesis of Shiga-toxin producing Escherichia coli. In Ricin and Shiga Toxins; Current Topics in Microbiology and Immunology; Springer: Berlin/Heidelberg, Germany, 2011; Volume 357, pp. 67–103. [Google Scholar]
  3. Kaper, J.B.; Nataro, J.P.; Mobley, H.L.T. Pathogenic Escherichia coli. Nat. Rev. Microbiol. 2004, 2, 123. [Google Scholar] [CrossRef] [PubMed]
  4. Suojala, L.; Kaartinen, L.; Pyörälä, S. Treatment for bovine Echerichia coli mastitis—An evidence-based approach. J. Vet. Pharmacol. Ther. 2013, 36, 521–531. [Google Scholar] [CrossRef] [PubMed]
  5. Paton, J.C.; Paton, A.W. Pathogenesis and diagnosis of Shiga toxin-producing Escherichia coli infections. Clin. Microbiol. Rev. 1998, 11, 450–479. [Google Scholar] [CrossRef] [PubMed]
  6. Blaser, M. Antibiotic overuse: Stop the killing of beneficial bacteria. Nature 2011, 476, 393. [Google Scholar] [CrossRef] [PubMed]
  7. Safwat Mohamed, D.; Farouk Ahmed, E.; Mohamed Mahmoud, A.; Abd El-Baky, R.M.; John, J. Isolation and evaluation of cocktail phages for the control of multidrug-resistant Escherichia coli serotype O104: H4 and E. coli O157: H7 isolates causing diarrhea. FEMS Microbiol. Lett. 2017, 365, fnx275. [Google Scholar] [CrossRef] [PubMed]
  8. Cieplak, T.; Soffer, N.; Sulakvelidze, A.; Nielsen, D.S. A bacteriophage cocktail targeting Escherichia coli reduces E. coli in simulated gut conditions, while preserving a non-targeted representative commensal normal microbiota. Gut Microbes 2018, 9, 391–399. [Google Scholar] [CrossRef] [PubMed]
  9. Poirel, L.; Madec, J.-Y.; Lupo, A.; Schink, A.-K.; Kieffer, N.; Nordmann, P.; Schwarz, S. Antimicrobial resistance in Escherichia coli. Microbiol. Spectr. 2018, 6. [Google Scholar] [CrossRef]
  10. Cabello, F.C. Heavy use of prophylactic antibiotics in aquaculture: A growing problem for human and animal health and for the environment. Environ. Microbiol. 2006, 8, 1137–1144. [Google Scholar] [CrossRef] [PubMed]
  11. Tajbakhsh, E.; Khamesipour, F.; Ranjbar, R.; Ugwu, I.C. Prevalence of class 1 and 2 integrons in multi-drug resistant Escherichia coli isolated from aquaculture water in Chaharmahal Va Bakhtiari province, Iran. Ann. Clin. Microbiol. Antimicrob. 2015, 14, 37. [Google Scholar] [CrossRef] [PubMed]
  12. Silva, Y.J.; Costa, L.; Pereira, C.; Cunha, Â.; Calado, R.; Gomes, N.C.M.; Almeida, A. Influence of environmental variables in the efficiency of phage therapy in aquaculture. Microb. Biotechnol. 2014, 7, 401–413. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Mohan Raj, J.R.; Vittal, R.; Huilgol, P.; Bhat, U.; Karunasagar, I. T4-like Escherichia coli phages from the environment carry blaCTX-M. Lett. Appl. Microbiol. 2018, 67, 9–14. [Google Scholar] [CrossRef] [PubMed]
  14. Heuer, O.E.; Kruse, H.; Grave, K.; Collignon, P.; Karunasagar, I.; Angulo, F.J. Human health consequences of use of antimicrobial agents in aquaculture. Clin. Infect. Dis. 2009, 49, 1248–1253. [Google Scholar] [CrossRef] [PubMed]
  15. Van den Bogaard, A.E.; London, N.; Driessen, C.; Stobberingh, E.E. Antibiotic resistance of faecal Escherichia coli in poultry, poultry farmers and poultry slaughterers. J. Antimicrob. Chemother. 2001, 47, 763–771. [Google Scholar] [CrossRef] [PubMed]
  16. Hammerum, A.M.; Heuer, O.E. Human health hazards from antimicrobial-resistant Escherichia coli of animal origin. Clin. Infect. Dis. 2009, 48, 916–921. [Google Scholar] [CrossRef] [PubMed]
  17. Reinthaler, F.F.; Posch, J.; Feierl, G.; Wüst, G.; Haas, D.; Ruckenbauer, G.; Mascher, F.; Marth, E. Antibiotic resistance of E. coli in sewage and sludge. Water Res. 2003, 37, 1685–1690. [Google Scholar] [CrossRef]
  18. Mesa, R.J.; Blanc, V.; Blanch, A.R.; Cortés, P.; Gonzalez, J.J.; Lavilla, S.; Miro, E.; Muniesa, M.; Saco, M.; Tórtola, M.T. Extended-spectrum β-lactamase-producing Enterobacteriaceae in different environments (humans, food, animal farms and sewage). J. Antimicrob. Chemother. 2006, 58, 211–215. [Google Scholar] [CrossRef] [PubMed]
  19. Watkinson, A.J.; Micalizzi, G.B.; Graham, G.M.; Bates, J.B.; Costanzo, S.D. Antibiotic-resistant Escherichia coli in wastewaters, surface waters, and oysters from an urban riverine system. Appl. Environ. Microbiol. 2007, 73, 5667–5670. [Google Scholar] [CrossRef] [PubMed]
  20. Korzeniewska, E.; Korzeniewska, A.; Harnisz, M. Antibiotic resistant Escherichia coli in hospital and municipal sewage and their emission to the environment. Ecotoxicol. Environ. Saf. 2013, 91, 96–102. [Google Scholar] [CrossRef] [PubMed]
  21. Nakai, T.; Park, S.C. Bacteriophage therapy of infectious diseases in aquaculture. Res. Microbiol. 2002, 153, 13–18. [Google Scholar] [CrossRef]
  22. Wagenaar, J.A.; Van Bergen, M.A.P.; Mueller, M.A.; Wassenaar, T.M.; Carlton, R.M. Phage therapy reduces Campylobacter jejuni colonization in broilers. Vet. Microbiol. 2005, 109, 275–283. [Google Scholar] [CrossRef] [PubMed]
  23. Martinez-Diaz, S.F.; Hipólito-Morales, A. Efficacy of phage therapy to prevent mortality during the vibriosis of brine shrimp. Aquaculture 2013, 400, 120–124. [Google Scholar] [CrossRef]
  24. Litt, P.K.; Saha, J.; Jaroni, D. Characterization of bacteriophages targeting Non-O157 Shiga toxigenic Escherichia coli. J. Food Prot. 2018, 81, 785–794. [Google Scholar] [CrossRef] [PubMed]
  25. Ramirez, K.; Cazarez-Montoya, C.; Lopez-Moreno, H.S.; Castro-del Campo, N. Bacteriophage cocktail for biocontrol of Escherichia coli O157: H7: Stability and potential allergenicity study. PLoS ONE 2018, 13, e0195023. [Google Scholar] [CrossRef] [PubMed]
  26. Duarte, V.S.; Dias, R.S.; Kropinski, A.M.; Campanaro, S.; Treu, L.; Siqueira, C.; Vieira, M.S.; Paes, I.S.; Santana, G.R.; Martins, F. Genomic analysis and immune response in a murine mastitis model of vB_EcoM-UFV13, a potential biocontrol agent for use in dairy cows. Sci. Rep. 2018, 8, 6845. [Google Scholar] [CrossRef] [PubMed]
  27. Cisek, A.A.; Dąbrowska, I.; Gregorczyk, K.P.; Wyżewski, Z. Phage therapy in bacterial infections treatment: One hundred years after the discovery of bacteriophages. Curr. Microbiol. 2017, 74, 277–283. [Google Scholar] [CrossRef] [PubMed]
  28. Bhensdadia, D.V.; Bhimani, H.D.; Rawal, C.M.; Kothari, V.V.; Raval, V.H.; Kothari, C.R.; Patel, A.B.; Bhatt, V.D.; Parmar, N.R.; Sajnani, M.R. Complete genome sequence of Escherichia phage ADB-2 isolated from a fecal sample of poultry. Genome Announc. 2013, 1, e00043-13. [Google Scholar] [CrossRef] [PubMed]
  29. Von lytischen Bakteriophagen, P.A. Post-harvest application of lytic bacteriophages for biocontrol of food-borne pathogens and spoilage bacteria. Berl. Munch. Tierarztl. Wochenschr. 2013, 126, 9. [Google Scholar]
  30. Lynch, M.F.; Tauxe, R.V.; Hedberg, C.W. The growing burden of foodborne outbreaks due to contaminated fresh produce: Risks and opportunities. Epidemiol. Infect. 2009, 137, 307–315. [Google Scholar] [CrossRef] [PubMed]
  31. Sharma, M. Lytic bacteriophages: Potential interventions against enteric bacterial pathogens on produce. Bacteriophage 2013, 3, e25518. [Google Scholar] [CrossRef] [PubMed]
  32. Jensen, E.C.; Schrader, H.S.; Rieland, B.; Thompson, T.L.; Lee, K.W.; Nickerson, K.W.; Kokjohn, T.A. Prevalence of broad-host-range lytic bacteriophages of Sphaerotilus natans, Escherichia coli, and Pseudomonas aeruginosa. Appl. Environ. Microbiol. 1998, 64, 575–580. [Google Scholar] [PubMed]
  33. Koskella, B.; Meaden, S. Understanding bacteriophage specificity in natural microbial communities. Viruses 2013, 5, 806–823. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Adams, M.H. Bacteriophages; Interscience Publishers: New York, NY, USA, 1959; p. 592. [Google Scholar]
  35. Azaïez, S.R.C.; Fliss, I.; Simard, R.E.; Moineau, S. Monoclonal antibodies raised against native major capsid proteins of lactococcal c2-like bacteriophages. Appl. Environ. Microbiol. 1998, 64, 4255–4259. [Google Scholar]
  36. Fortier, L.-C.; Moineau, S. Morphological and genetic diversity of temperate phages in Clostridium difficile. Appl. Environ. Microbiol. 2007, 73, 7358–7366. [Google Scholar] [CrossRef] [PubMed]
  37. Duplessis, M.; Russell, W.M.; Romero, D.A.; Moineau, S. Global gene expression analysis of two Streptococcus thermophilus bacteriophages using DNA microarray. Virology 2005, 340, 192–208. [Google Scholar] [CrossRef] [PubMed]
  38. Sambrook, J.; Russell, D.W. Molecular clonning: A laboratory manual. In Molecular Clonning: A Laboratory Manual; Cold Spring Harbor Laboratory Press: Cold Spring Harbor, NY, USA, 2001. [Google Scholar]
  39. Arndt, D.; Grant, J.R.; Marcu, A.; Sajed, T.; Pon, A.; Liang, Y.; Wishart, D.S. PHASTER: A better, faster version of the PHAST phage search tool. Nucleic Acids Res. 2016, 44, W16–W21. [Google Scholar] [CrossRef] [PubMed]
  40. Katoh, K.; Standley, D.M. MAFFT multiple sequence alignment software version 7: Improvements in performance and usability. Mol. Biol. Evol. 2013, 30, 772–780. [Google Scholar] [CrossRef] [PubMed]
  41. Mercanti, D.J.; Rousseau, G.M.; Capra, M.L.; Quiberoni, A.; Tremblay, D.M.; Labrie, S.J.; Moineau, S. Genomic diversity of phages infecting probiotic strains of Lactobacillus paracasei. Appl. Environ. Microbiol. 2016, 82, 95–105. [Google Scholar] [CrossRef] [PubMed]
  42. Lavigne, R.; Darius, P.; Summer, E.J.; Seto, D.; Mahadevan, P.; Nilsson, A.S.; Ackermann, H.W.; Kropinski, A.M. Classification of Myoviridae bacteriophages using protein sequence similarity. BMC Microbiol. 2009, 9, 224. [Google Scholar] [CrossRef] [PubMed]
  43. Niu, Y.D.; Johnson, R.P.; Xu, Y.; McAllister, T.A.; Sharma, R.; Louie, M.; Stanford, K. Host range and lytic capability of four bacteriophages against bovine and clinical human isolates of Shiga toxin-producing Escherichia coli O157: H7. J. Appl. Microbiol. 2009, 107, 646–656. [Google Scholar] [CrossRef] [PubMed]
  44. Sillankorva, S.; Oliveira, D.; Moura, A.; Henriques, M.; Faustino, A.; Nicolau, A.; Azeredo, J. Efficacy of a broad host range lytic bacteriophage against E. coli adhered to urothelium. Curr. Microbiol. 2011, 62, 1128–1132. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Hoang, H.A.; Quy, N.T.C.; Chi, N.V.T. Detection of Escherichia coli in ready-to-eat fresh vegetables using broad-host-range recombinant phages. J. Appl. Microbiol. 2018, 124, 1610–1616. [Google Scholar] [CrossRef] [PubMed]
  46. Ross, A.; Ward, S.; Hyman, P. More is better: Selecting for broad host range bacteriophages. Front. Microbiol. 2016, 7, 1352. [Google Scholar] [CrossRef] [PubMed]
  47. Keen, E.C.; Adhya, S.L. Phage therapy: Current research and applications. Clin. Infect. Dis. 2015, 61, 141–142. [Google Scholar] [CrossRef]
  48. Kaliniene, L.; Truncaitė, L.; Šimoliūnas, E.; Zajančkauskaitė, A.; Vilkaitytė, M.; Kaupinis, A.; Skapas, M.; Meškys, R. Molecular analysis of the low-temperature Escherichia coli phage vB_EcoS_NBD2. Arch. Virol. 2018, 163, 105–114. [Google Scholar] [CrossRef] [PubMed]
  49. Jamalludeen, N.; Kropinski, A.M.; Johnson, R.P.; Lingohr, E.; Harel, J.; Gyles, C.L. Complete genomic sequence of bacteriophage φEcoM-GJ1, a novel phage that has myovirus morphology and a podovirus-like RNA polymerase. Appl. Environ. Microbiol. 2008, 74, 516–525. [Google Scholar] [CrossRef] [PubMed]
  50. Fouts, D.E. Phage_Finder: Automated identification and classification of prophage regions in complete bacterial genome sequences. Nucleic Acids Res. 2006, 34, 5839–5851. [Google Scholar] [CrossRef] [PubMed]
  51. Maluf, N.K.; Yang, Q.; Catalano, C.E. Self-association properties of the bacteriophage λ terminase holoenzyme: Implications for the DNA packaging motor. J. Mol. Biol. 2005, 347, 523–542. [Google Scholar] [CrossRef] [PubMed]
  52. Catalano, C.E. The terminase enzyme from bacteriophage lambda: A DNA-packaging machine. Cell. Mol. Life Sci. C 2000, 57, 128–148. [Google Scholar] [CrossRef] [PubMed]
  53. Lim, J.-A.; Shin, H.; Lee, D.H.; Han, S.-W.; Lee, J.-H.; Ryu, S.; Heu, S. Complete genome sequence of the Pectobacterium carotovorum subsp. carotovorum virulent bacteriophage PM1. Arch. Virol. 2014, 159, 2185–2187. [Google Scholar] [CrossRef] [PubMed]
  54. Born, Y.; Fieseler, L.; Marazzi, J.; Lurz, R.; Duffy, B.; Loessner, M.J. Novel virulent and broad host range Erwinia amylovora bacteriophages reveal a high degree of mosaicism and relationship to Enterobacteriaceae phages. Appl. Environ. Microbiol. 2011, 15, AEM-03022. [Google Scholar] [CrossRef] [PubMed]
  55. Liu, J.; Dehbi, M.; Moeck, G.; Arhin, F.; Bauda, P.; Bergeron, D.; Callejo, M.; Ferretti, V.; Ha, N.; Kwan, T. Antimicrobial drug discovery through bacteriophage genomics. Nat. Biotechnol. 2004, 22, 185. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Transmission electron micrograph of phage ST32 with uncontracted (A) and contracted (B) tails.
Figure 1. Transmission electron micrograph of phage ST32 with uncontracted (A) and contracted (B) tails.
Viruses 10 00616 g001
Figure 2. Lytic activity of phage ST32 at various temperatures. (A) One-step growth curve of phage ST32 at 20 °C, 30 °C, and 37 °C; (B) The growth of E. coli ST130 strain at 20 °C, 30 °C, and 37 °C.
Figure 2. Lytic activity of phage ST32 at various temperatures. (A) One-step growth curve of phage ST32 at 20 °C, 30 °C, and 37 °C; (B) The growth of E. coli ST130 strain at 20 °C, 30 °C, and 37 °C.
Viruses 10 00616 g002
Figure 3. Phylogenetic tree based on the amino acid sequences of the large terminase subunit (ORF51) of phage ST32 and the phages available in databases sharing sequence identity. The corresponding phage protein sequences were retrieved from GenBank (https://www.ncbi.nlm.nih.gov/). The colors in the internal and external circular layers categorize phages, genera, and families, respectively. When the genera or the family of a phage is not indicated, it means that it was not available in the database or in the associated publication. Branches with branch support values greater than 90% are marked with a blue dot. The size of the dot is directly proportional to the branch support value.
Figure 3. Phylogenetic tree based on the amino acid sequences of the large terminase subunit (ORF51) of phage ST32 and the phages available in databases sharing sequence identity. The corresponding phage protein sequences were retrieved from GenBank (https://www.ncbi.nlm.nih.gov/). The colors in the internal and external circular layers categorize phages, genera, and families, respectively. When the genera or the family of a phage is not indicated, it means that it was not available in the database or in the associated publication. Branches with branch support values greater than 90% are marked with a blue dot. The size of the dot is directly proportional to the branch support value.
Viruses 10 00616 g003
Figure 4. Schematic representation of the genomic organization of phage ST32 compared to phages phiEcoM-GJ1, vB_EamM-Y2, PM1, and PP101. Each line represents a different phage genome and each arrow represents an ORF. Arrows of the same color indicate ORFs that share more than 80% identity. White arrows indicate that the identity is less than 80% or there is no homologous putative protein. Gray shading indicates vB_EamM-Y2 phage ORFs sharing more than 70% with that of other aligned phages.
Figure 4. Schematic representation of the genomic organization of phage ST32 compared to phages phiEcoM-GJ1, vB_EamM-Y2, PM1, and PP101. Each line represents a different phage genome and each arrow represents an ORF. Arrows of the same color indicate ORFs that share more than 80% identity. White arrows indicate that the identity is less than 80% or there is no homologous putative protein. Gray shading indicates vB_EamM-Y2 phage ORFs sharing more than 70% with that of other aligned phages.
Viruses 10 00616 g004
Table 1. The host range of phage ST32.
Table 1. The host range of phage ST32.
Non-PathogenPathogen
Genus/Species/Subspecies of the Host Strain# HERName of the Host StrainΦST32Genus/Species/Subspecies of the Host Strain# HERName of the Host StrainΦST32
Escherichia coli1022O44:K74 MUL-B37.2Escherichia coli1176N/A+++
Escherichia coli1024B (11303)+Escherichia coli1255O157:H7 C-8299-83
Escherichia coli1025K12 C600 (λ)+Escherichia coli1256O157:H7 E318
Escherichia coli1036C (13706)++++Escherichia coli1257O157:H7 A7793-B1
Escherichia coli1037K12S+Escherichia coli1258O157:H7 C-8300-83
Escherichia coli1040K12 (λ) Lederberg+Escherichia coli1259O157:H7 C-7685-84
Escherichia coli1077W3350+Escherichia coli1260O157:H7 CL40
Escherichia coli1128MUL-B70.1Escherichia coli1261O157:H7 C-7111-85
Escherichia coli1129O86:B7 MUL-B3.1Escherichia coli1262O157:H7 B1190-1
Escherichia coli1139K12 65+Escherichia coli1263O157:H7 B1328-C10
Escherichia coli1144K12S LederbergEscherichia coli1264O157:H7 A8188-B3
Escherichia coli1155K1++++Escherichia coli1265O157:H7 C7420-85
Escherichia coli1213JE-1 (N3)Escherichia coli1266O157:H7 3283
Escherichia coli1217JE-2(R62Rpilc)+Escherichia coli1267O157:H7 C-7140-85
Escherichia coli1218J53(RIP69)Escherichia coli1268O157:H7 5896
Escherichia coli1219K12 J62-1(R997)Escherichia coli1269O157:H7 C-7142-85
Escherichia coli1221K12 J53-1(R15)+Escherichia coli1270O157:H7 C-91-84
Escherichia coli1222JE-1 (RA1::TN5Sqr)++++Escherichia coli H21 ST130++++
Escherichia coli1240J62-1 (R27::TN7)+Escherichia coli O165:H8 ST120
Escherichia coli125240+Escherichia coli O8:H16 ST110++
Escherichia coli1253HM 8305Escherichia coli H8 ST100++
Escherichia coli1271K12 C600 (H-19J)+Escherichia coli O153:H12 BW
Escherichia coli1275K12 C600+Shigella sonnei1043Y6R+
Escherichia coli1290CSH39Shigella dysenteriae1031aSH
Escherichia coli1299K12 C600 (933-J)+Shigella dysenteriae1020SH(P2)
Escherichia coli1315F492 (O8:K27-:H-)++++Salmonella paratyphi1045B type 1
Escherichia coli1337O103 2929+Salmonella typhi1038ViA subtype Tananarive
Escherichia coli1366K12 MC4100+Citrobacter freundii1518CF3
Escherichia coli1374E69 O9:K30:H12Citrobacter freundii CF4
Escherichia coli1375CWG 1028++++Citrobacter freundii1516CF5
Escherichia coli1382Ymel mel-1 supF58+Citrobacter freundii CF7
Escherichia coli1383Ymel (HK97)+Citrobacter freundii CF8
Escherichia coli13920103 GVsCitrobacter freundii Sa1
Escherichia coli1393RougierCitrobacter freundii Sa6
Escherichia coli1445TC4Citrobacter freundii Sa59
Escherichia coli1446MB4
Escherichia coli1462C-3000+
Escherichia coli1536SlyD++++
Notes: (−) Do not infect; (+) lysis zone at dilution 100 or “lysis from without [47]”; (++) infect at dilutions of 100 to 10−2; (+++) infect at dilutions of 100 to 10−4; (++++) infect at dilutions of 100 to 10−6.
Table 2. Features of the open reading frames (ORFs) of phage ST32.
Table 2. Features of the open reading frames (ORFs) of phage ST32.
ORFStrandStart (pb)End (pb)Size (aa)MW (kDa)pISD Sequence (AGGAGGU) aPredicted Protein FunctionBLAST (Extent, % aa Identity) bAligned Protein Size (aa)E ValueAccession Number
1+673261364672.76.09TGGAGACttacaaATGRNA polymerasegp01 [Enterobacteria phage phiEcoM-GJ1] (598/646; 93%)6480YP_001595396.1
2+26402846687.894.51AGGATGGcattagTTG gp02 [Enterobacteria phage phiEcoM-GJ1] (48/55; 87%)551 × 10−17YP_001595397.1
3+39594177724.148.15AGGAGAAtaaaATG hypothetical protein [Klebsiella phage KP8] (23/73; 32%)718 × 10−4AVJ48916.1
4+42174453788.99.63CGGAGAGcagaaATG gp04 [Enterobacteria phage phiEcoM-GJ1] (49/76; 64%)768 × 10−23YP_001595399.1
5+44564644627.54.53ACGAGGTtaatcATG gp05 [Enterobacteria phage phiEcoM-GJ1] (61/62; 98%)621 × 10−37YP_001595400.1
6+46414835647.39.7TGGAGGCcaaATG N/A
7+48325077819.44.32AGGCGGGttggttGTG N/A
8+50875260576.79.25AGGAGTAttaaATG gp07 [Enterobacteria phage phiEcoM-GJ1] (56/57; 98%)576 × 10−34YP_001595402.1
9+5356566710311.64.5AGGTAATtaaATG gp08 [Enterobacteria phage phiEcoM-GJ1] (84/99; 85%)992 × 10−54YP_001595404.1
10+56835943869.75.6GGGAGTTattATG gp09 [Enterobacteria phage phiEcoM-GJ1] (79/86; 92%)876 × 10−53YP_001595404.1
11+59366118606.44.64TGGGAGTtctgtaccATG N/A
12+61216348758.45.24AGGATAAtcATG gp10 [Enterobacteria phage phiEcoM-GJ1] (67/75; 89%)755 × 10−45YP_001595405.1
13+63456575768.69.58ACAAGGTttattgcaATG gp11 [Enterobacteria phage phiEcoM-GJ1] (42/68; 62%)681 × 10−16YP_001595406.1
14+663969299610.79.47TGGAGCAtttATG gp12 [Enterobacteria phage phiEcoM-GJ1] (87/96; 91%)967 × 10−59YP_001595407.1
15+69227155778.85.22AGAAGGTgaagcGTG gp13 [Enterobacteria phage phiEcoM-GJ1] (68/77; 88%)775 × 10−44YP_001595408.1
16+715274399510.89.3TGGAGAAattaaagcaATG gp14 [Enterobacteria phage phiEcoM-GJ1] (84/95; 88%)951 × 10−54YP_001595409.1
17+74397636657.829.81AGGTGATgtaATG IME11_76 [Escherichia phage IME11] (29/65;45%)683 × 10−7YP_006990681.1
18+7715819716018.88.79TGGAGGGcttATG CBB_348 [Pectobacterium phage CBB] (72/160; 45%)1612 × 10−36AMM43911.1
19+8323870012514.25.78AAGAGAAtcttaatcATGssDNA-binding proteingp15 [Enterobacteria phage phiEcoM-GJ1] (96/125; 77%)1266 × 10−54YP_001595410.1
20+8823971929833.27.74GTGAGGAatatcATG gp17 [Enterobacteria phage phiEcoM-GJ1] (159/229; 69%)2295 × 10−109YP_001595412.1
21+977510,12511613.16.07CGGAGCAtttATG gp18 [Enterobacteria phage phiEcoM-GJ1] (114/116; 98%)1162 × 10−80YP_001595413.1
22+10,12210,355778.725.29AGGAAGTtaaATG gp19 [Enterobacteria phage phiEcoM-GJ1] (56/77; 73%)775 × 10−32YP_001595414.1
23+10,34510,6148910.34.7AGGAAATccattccGTG N/A
24+10,60711,02313815.79.48AGGAGCTgaaaaATGendolysingp21 [Enterobacteria phage phiEcoM-GJ1] (110/131; 84%)1311 × 10−72YP_001595416.1
25+11,04411,3229210.95.7TGGAGCAtccgATG N/A
26+11,30911,85718221.24.03GGGAGAAactcaATGantirestriction proteingp22 [Enterobacteria phage phiEcoM-GJ1] (145/181; 80%)1812 × 10−100YP_001595417.1
27+11,85012,089799.26.9CGAAGGGatactattctcaaATG gp23 [Enterobacteria phage phiEcoM-GJ1] (73/79; 92%)794 × 10−48YP_001595418.1
28+13,09213,295677.54.58TGGAGAGttcctATG gp25 [Enterobacteria phage phiEcoM-GJ1] (57/67; 85%)693 × 10−33YP_001595420.1
29+13,29213,5829611.39.1AGGAGCTgcaaaaATG N/A
30+13,57913,8699610.65.25CGGAGTTccattTTG gp27 [Enterobacteria phage phiEcoM-GJ1] (93/96; 97%)973 × 10−59YP_001595422.1
31+13,87214,54622425.98.28ACAAGGCcactaaaaATG gp28 [Enterobacteria phage phiEcoM-GJ1] (223/224; 99%)2242 × 10−165YP_001595423.1
32+14,67015,00811212.34.47ATAAGGTatatacaaATG gp29 [Enterobacteria phage phiEcoM-GJ1] (111/112; 99%)1129 × 10−75YP_001595424.1
33+15,39515,532455.18.99CGGAGCAataattaatTTG N/A
34+15,54715,789808.89.24AGAAGCTatgccaatGTG gp30 [Enterobacteria phage phiEcoM-GJ1] (77/80; 96%)801 × 10−50YP_001595425.1
35+16,02916,169464.83.76TGGAGTCctcATG N/A
36+16,17816,414788.59.05AGGTGATttATG gp31 [Enterobacteria phage phiEcoM-GJ1] (77/78; 99%)781 × 10−44YP_001595426.1
37+17,40417,634768.54.89TGGAGAGaaacATG gp32 [Enterobacteria phage phiEcoM-GJ1] (74/76; 97%)762 × 10−44YP_001595427.1
38+17,69118,34121624.85.99CGGAGAGcaaATGthymidylate synthasegp33 [Enterobacteria phage phiEcoM-GJ1] (196/216; 91%)2168 × 10−149YP_001595428.1
39+18,34620,109587665.95ACCAGGAataaataaATGhelicase/primasegp35 [Enterobacteria phage phiEcoM-GJ1] (560/587; 95%)5870YP_001595430.1
40+20,17522,10964474.66.41TGGAGCCatactGTGDNA polymerasegp37 [Enterobacteria phage phiEcoM-GJ1] (637/644; 99%)6440YP_001595432.1
41+22,10922,3819010.14.37CAGAGATtcactaATG gp38 [Enterobacteria phage phiEcoM-GJ1] (86/90; 96%)902 × 10−54YP_001595433.1
42+22,41223,278288314.86AGGTACTcaaaATG gp39 [Enterobacteria phage phiEcoM-GJ1] (288/288; 100%)2880YP_001595434.1
43+23,31224,34934539.48.09GGGAGCCtttaattTTGexonucleasegp40 [Enterobacteria phage phiEcoM-GJ1] (342/345; 99%)3450YP_001595435.1
44+24,36124,88517420.19.46TGGAGTTggaATG gp41 [Enterobacteria phage phiEcoM-GJ1] (173/174; 99%)1743 × 10−124YP_001595436.1
45+24,87525,63025128.58.64AGAAAGAatcttaATGDNA ligasegp42 [Enterobacteria phage phiEcoM-GJ1] (233/251; 93%)2516 × 10−175YP_001595437.1
46+25,62326,24920823.56.38GTGAGGAaagttTTG gp43 [Enterobacteria phage phiEcoM-GJ1] (203/208; 98%)2081 × 10−147YP_001595438.1
47+26,25226,83919520.76.66ATCAAGTagagaaataatcATGdeoxyuridine 5’-triphosphate nucleotidylhydrolasegp44 [Enterobacteria phage phiEcoM-GJ1] (164/195; 82%)1994 × 10−105YP_001595439.1
48+26,85827,064688.24.32TGGAGCAtccATG PP74_27 [Pectobacterium phage PP74] (37/68; 54%)733 × 10−17APD19639.1
49+27,08227,41110912.29.7TGGAACCtatctgaaATG gp45 [Enterobacteria phage phiEcoM-GJ1] (109/109; 100%)1094 × 10−74YP_001595440.1
50+27,46727,6526174.4CGGAGTCgcttATG gp46 [Enterobacteria phage phiEcoM-GJ1] (61/61; 100%)611 × 10−37YP_001595441.1
51+27,67229,690672766.02AAGAGAAcgaatcaATGlarge subunit terminasegp48 [Enterobacteria phage phiEcoM-GJ1] (667/671; 99%)6710YP_001595443.1
52+29,69329,905707.99.18TGGATGTaaatATG gp49 [Enterobacteria phage phiEcoM-GJ1] (70/70; 100%)707 × 10−43YP_001595444.1
53+29,90531,22143849.18.16AGGAAGAaataATGportal proteingp50 [Enterobacteria phage phiEcoM-GJ1] (434/438; 99%)4380YP_001595445.1
54+31,19032,254354394.8AAAGGGTaacgcaaGTG gp51 [Enterobacteria phage phiEcoM-GJ1] (353/354; 99%)3540YP_001595446.1
55+32,26432,73715716.46.26ATAAGGTaagacaATG gp52 [Enterobacteria phage phiEcoM-GJ1] (147/157; 94%)1572 × 10−101YP_001595447.1
56+32,81833,030707.36.06TGTAACTGTG gp67 [Erwinia phage vB_EamM-Y2] (51/70; 73%)902 × 10−23YP_007004717.1
57+33,08634,09333536.75.17TGGATTAaattacATGmajor capsid proteingp53 [Enterobacteria phage phiEcoM-GJ1] (323/335; 96%)3350YP_001595448.1
58+34,14034,58014616.15.34AAGAGAAatagtaATG gp54 [Enterobacteria phage phiEcoM-GJ1] (116/146; 79%)1461 × 10−71YP_001595449.1
59+34,58134,97012914.64.48AGTTGGCgtaaATG gp55 [Enterobacteria phage phiEcoM-GJ1] (124/129; 96%)1292 × 10−86YP_001595450.1
60+34,96735,32912013.99.16GGGTCACagttTTG gp56 [Enterobacteria phage phiEcoM-GJ1] (120/120; 100%)1204 × 10−85YP_001595451.1
61+35,32635,83817019.34.98AGGAGTTagagaaATG gp57 [Enterobacteria phage phiEcoM-GJ1] (167/170; 98%)1702 × 10−120YP_001595452.1
62+35,83937,28748250.94.75AGGGAATctaaATG gp58 [Enterobacteria phage phiEcoM-GJ1] (447/482; 93%)4820YP_001595453.1
63+37,29837,75315116.56.55AGGTGCGataaGTG gp59 [Enterobacteria phage phiEcoM-GJ1] (148/151; 98%)1512 × 10−104YP_001595454.1
64+37,76538,22315217.35.1AGTAAGTATG gp60 [Enterobacteria phage phiEcoM-GJ1] (152/152; 100%)1524 × 10−107YP_001595455.1
65+38,22938,399566.74.67CGGAGACagtttagtatccATG gp61 [Enterobacteria phage phiEcoM-GJ1] (55/56; 98%)738 × 10−32YP_001595456.1
66+38,38342,1081241134.65.36AGAAACTcgaaccagtagATGtail fibergp62 [Enterobacteria phage phiEcoM-GJ1] (946/1239; 76%)12390YP_001595457.1
67+42,18243,29136940.75.13AATAGGTatatcgcaATG gp63 [Enterobacteria phage phiEcoM-GJ1] (366/369; 99%)3690YP_001595458.1
68+43,29144,17829531.25.98TGGAGTCattttaATG gp64 [Enterobacteria phage phiEcoM-GJ1] (295/295; 100%)2950YP_001595459.1
69+44,17544,53712013.75.07GGGACGTatcctATG gp65 [Enterobacteria phage phiEcoM-GJ1] (120/120; 100%)1202 × 10−84YP_001595460.1
70+44,53045,32426428.25.8AGAGTGTacttgaacGTGbaseplate assembly proteingp66 [Enterobacteria phage phiEcoM-GJ1] (263/264; 99%)2640YP_001595461.1
71+45,32445,69512313.55.22ATGAAATaATG gp67 [Enterobacteria phage phiEcoM-GJ1] (117/123; 95%)1237 × 10−80YP_001595462.1
72+45,67146,82838541.24.55CGGAATTcttaacATG gp68 [Enterobacteria phage phiEcoM-GJ1] (361/385; 94%)3850YP_001595463.1
73+46,83047,47121323.55.82CAGATGTgacagtataatATG gp69 [Enterobacteria phage phiEcoM-GJ1] (193/213; 91%)2131 × 10−139YP_001595464.1
74+47,47148,61938242.35.49CGGAGAAataATG gp70 [Enterobacteria phage phiEcoM-GJ1] (342/382; 90%)3820YP_001595465.1
75+48,61950,01646550.28.17AGGCCATaATG gp71 [Enterobacteria phage phiEcoM-GJ1] (314/465; 68%)4650YP_001595466.1
76+50,02551,07134836.36.6AGGATTCaaaATGtail fiber proteingp72 [Enterobacteria phage phiEcoM-GJ1] (250/348; 72%)3563 × 10−156YP_001595467.1
77+51,07951,42011312.37.95AGGAACTcATGholingp73 [Enterobacteria phage phiEcoM-GJ1] (110/113; 97%)1138 × 10−74YP_001595468.1
78+51,43851,99218420.79.57AGGAACTcgaATGendolysingp74 [Enterobacteria phage phiEcoM-GJ1] (178/184; 97%)1847 × 10−131YP_001595469.1
79+51,99253,09236642.24.76AGGAAATctgtaATGribonucleotide reductase beta subunitgp75 [Enterobacteria phage phiEcoM-GJ1] (340/366; 93%)3720YP_001595470.1
a Start codon indicated in bold; Match to SD sequence is indicated by underlining; SD position is indicated in uppercase. b The number of identical amino acids/The total of amino acids of smallest protein.

Share and Cite

MDPI and ACS Style

Liu, H.; Geagea, H.; Rousseau, G.M.; Labrie, S.J.; Tremblay, D.M.; Liu, X.; Moineau, S. Characterization of the Escherichia coli Virulent Myophage ST32. Viruses 2018, 10, 616. https://doi.org/10.3390/v10110616

AMA Style

Liu H, Geagea H, Rousseau GM, Labrie SJ, Tremblay DM, Liu X, Moineau S. Characterization of the Escherichia coli Virulent Myophage ST32. Viruses. 2018; 10(11):616. https://doi.org/10.3390/v10110616

Chicago/Turabian Style

Liu, Honghui, Hany Geagea, Geneviève M. Rousseau, Simon J. Labrie, Denise M. Tremblay, Xinchun Liu, and Sylvain Moineau. 2018. "Characterization of the Escherichia coli Virulent Myophage ST32" Viruses 10, no. 11: 616. https://doi.org/10.3390/v10110616

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop