ArticlePDF Available

Four Escherichia coli O157:H7 Phages: A New Bacteriophage Genus and Taxonomic Classification of T1-Like Phages

PLOS
PLOS ONE
Authors:

Abstract and Figures

The T1-like bacteriophages vB_EcoS_AHP24, AHS24, AHP42 and AKS96 of the family Siphoviridae were shown to lyse common phage types of Shiga toxin-producing Escherichia coli O157:H7 (STEC O157:H7), but not non-O157 E. coli. All contained circularly permuted genomes of 45.7-46.8 kb (43.8-44 mol% G+C) encoding 74-81 open reading frames and 1 arginyl-tRNA. Sodium dodecyl sulfate-polyacrylamide gel electrophoresis revealed that the structural proteins were identical among the four phages. Further proteomic analysis identified seven structural proteins responsible for tail fiber, tail tape measure protein, major capsid, portal protein as well as major and minor tail proteins. Bioinformatic analyses on the proteins revealed that genomes of AHP24, AHS24, AHP42 and AKS96 did not encode for bacterial virulence factors, integration-related proteins or antibiotic resistance determinants. All four phages were highly lytic to STEC O157:H7 with considerable potential as biocontrol agents. Comparative genomic, proteomic and phylogenetic analysis suggested that the four phages along with 17 T1-like phage genomes from database of National Center for Biotechnology Information (NCBI) can be assigned into a proposed subfamily "Tunavirinae" with further classification into five genera, namely "Tlslikevirus" (TLS, FSL SP-126), "Kp36likevirus" (KP36, F20), Tunalikevirus (T1, ADB-2 and Shf1), "Rtplikevirus" (RTP, vB_EcoS_ACG-M12) and "Jk06likevirus" (JK06, vB_EcoS_Rogue1, AHP24, AHS24, AHP42, AKS96, phiJLA23, phiKP26, phiEB49). The fact that the viruses related to JK06 have been isolated independently in Israel (JK06) (GenBank Assession #, NC_007291), Canada (vB_EcoS_Rogue1, AHP24, AHS24, AHP42, AKS96) and Mexico (phiKP26, phiJLA23) (between 2005 and 2011) indicates that these similar phages are widely distributed, and that horizontal gene transfer does not always prevent the characterization of bacteriophage evolution. With this new scheme, any new discovered phages with same type can be more properly identified. Genomic- and proteomic- based taxonomic classification of phages would facilitate better understanding phages diversity and genetic traits involved in phage evolution.
Content may be subject to copyright.
Four
Escherichia coli
O157:H7 Phages: A New
Bacteriophage Genus and Taxonomic Classification of
T1-Like Phages
Yan D. Niu
1
*, Tim A. McAllister
2
, John H. E. Nash
3,4
, Andrew M. Kropinski
3,5
, Kim Stanford
1
*
1Alberta Agriculture and Rural Development, Agriculture Centre, Lethbridge, Alberta, Canada, 2Lethbridge Research Centre, Agriculture and Agri-Food Canada,
Lethbridge, Alberta, Canada, 3Public Health Agency of Canada, Laboratory for Foodborne Zoonoses, Guelph, Ontario, Canada, 4Department of Pathobiology, Ontario
Veterinary College, University of Guelph, Guelph, Ontario, Canada, 5Department of Cellular and Molecular Biology, University of Guelph, Guelph, Ontario, Canada
Abstract
The T1-like bacteriophages vB_EcoS_AHP24, AHS24, AHP42 and AKS96 of the family Siphoviridae were shown to lyse
common phage types of Shiga toxin-producing Escherichia coli O157:H7 (STEC O157:H7), but not non-O157 E. coli. All
contained circularly permuted genomes of 45.7–46.8 kb (43.8–44 mol% G+C) encoding 74–81 open reading frames and 1
arginyl-tRNA. Sodium dodecyl sulfate-polyacrylamide gel electrophoresis revealed that the structural proteins were identical
among the four phages. Further proteomic analysis identified seven structural proteins responsible for tail fiber, tail tape
measure protein, major capsid, portal protein as well as major and minor tail proteins. Bioinformatic analyses on the
proteins revealed that genomes of AHP24, AHS24, AHP42 and AKS96 did not encode for bacterial virulence factors,
integration-related proteins or antibiotic resistance determinants. All four phages were highly lytic to STEC O157:H7 with
considerable potential as biocontrol agents. Comparative genomic, proteomic and phylogenetic analysis suggested that the
four phages along with 17 T1-like phage genomes from database of National Center for Biotechnology Information (NCBI)
can be assigned into a proposed subfamily ‘‘Tunavirinae’’ with further classification into five genera, namely ‘‘Tlslikevirus’’
(TLS, FSL SP-126), ‘‘Kp36likevirus’’ (KP36, F20), Tunalikevirus (T1, ADB-2 and Shf1), ‘‘Rtplikevirus’’ (RTP, vB_EcoS_ACG-M12)
and ‘‘Jk06likevirus’’ (JK06, vB_EcoS_Rogue1, AHP24, AHS24, AHP42, AKS96, phiJLA23, phiKP26, phiEB49). The fact that the
viruses related to JK06 have been isolated independently in Israel (JK06) (GenBank Assession #, NC_007291), Canada
(vB_EcoS_Rogue1, AHP24, AHS24, AHP42, AKS96) and Mexico (phiKP26, phiJLA23) (between 2005 and 2011) indicates that
these similar phages are widely distributed, and that horizontal gene transfer does not always prevent the characterization
of bacteriophage evolution. With this new scheme, any new discovered phages with same type can be more properly
identified. Genomic- and proteomic- based taxonomic classification of phages would facilitate better understanding phages
diversity and genetic traits involved in phage evolution.
Citation: Niu YD, McAllister TA, Nash JHE, Kropinski AM, Stanford K (2014) Four Escherichia coli O157:H7 Phages: A New Bacteriophage Genus and Taxonomic
Classification of T1-Like Phages. PLoS ONE 9(6): e100426. doi:10.1371/journal.pone.0100426
Editor: Mark J van Raaij, Centro Nacional de Biotecnologia CSIC, Spain
Received January 9, 2014; Accepted May 23, 2014; Published June 25, 2014
Copyright: ß2014 Niu et al. This is an open-access article distributed under the terms of the Creative Commons Attribution License, which permits unrestricted
use, distribution, and reproduction in any medium, provided the original author and source are credited.
Funding: This work was supported by the Alberta Livestock and Meat Agency, Agriculture and Agri-Food Canada Peer-Reviewed Project Program. The funders
had no role in study design, data collection and analysis, decision to publish, or preparation of the manuscript.
Competing Interests: The authors have declared that no competing interests exist.
* Email: dongyan.niu@gov.ab.ca (YDN); kim.stanford@gov.ab.ca (KS)
Introduction
Tailed bacteriophages (phages) with double-strand DNA
genomes belonging to the order Caudovirales are the most abundant
viruses on earth, accounting for 96% of all the phages observed
[1]. Based on tail morphology, these viruses are classified by the
International Committee on Taxonomy of Viruses (ICTV), into
three families2Myoviridae (long contractile tail), Siphoviridae (long
non-contractile tail) and Podoviridae (short non-contractile tail).
Recent advances in sequencing technologies has led to a
proliferation in the sequencing of phage genomes [2,3], enabling
comparative genomics and proteomics to better define phage
taxonomy. The family Myoviridae now contains three subfamilies,
Peduovirinae,Spounavirinae and Tevenvirinae [4], and 18 genera. The
family Podoviridae has been further divided into the Autographivirinae
and Picovirinae subfamilies [2], with a total of eleven genera. The
Siphoviridae account for .61% of described phages [1] and this
family also represents the largest group of fully sequenced phages,
but no subfamilies, and only nine bacterial-specific phage genera
have been described. Classification of the Siphoviridae is currently
under review by ICTV (Adriaenssens, personal communication).
T1-like phages possess terminally redundant and circularly
permuted genomes of ,50 kb, and are currently classified as
members of one genus (Tunalikevirus) within Siphoviridae [5].
Morphologically, they have a polyhedral head 60 nm in diameter
with an extremely flexible non-contractile tail 151 nm in length
and 8 nm in diameter [6]. At present, ICTV only recognizes nine
species of phages within this genus with 1, 1, 6 and 1 infecting
Cronobacter, Enterobacter, Escherichia coli, and Shigella, respectively.
Shiga-toxin producing E. coli O157:H7 (STEC O157:H7)
remains one of leading causes of foodborne illnesses in North
America [7,8]. Although the food production continuum has
introduced control measures to prevent the pathogen from
entering food chain, outbreaks of STEC O157:H7 linked to fresh
produce and beef products continue (http://www.cdc.gov/ecoli/
outbreaks.html and http://www.phac-aspc.gc.ca/fs-sa/fs-fi/ecoli-
PLOS ONE | www.plosone.org 1 June 2014 | Volume 9 | Issue 6 | e100426
eng.php). Lytic phages offer promise in the prevention and therapy
of bacterial infections in humans [9], livestock [10,11] and plants
[12] and have been employed to decontaminate processed foods
and agricultural products [9,10]. However, the use of phage
therapy to target bacterial pathogens such as STEC O157:H7
[13,14] and Salmonella [15,16] in the digestive tract of livestock
remains challenging. Factors such as the development of phage
resistance, the complexity of predator-prey relationships between
phages and hosts, the diversity and abundance of microflora in the
gastro-intestinal tract all may undermine the effectiveness of phage
therapy. Recently, in-vitro experiments in our laboratory have
indicated that competitive interference between different phage
types may be another factor impacting effectiveness of phage
cocktails [17], even though this approach is often advocated as a
means of avoiding resistance. An improved understanding of
phage taxonomy, proteomics and target receptors could lead to
the formulation of more effective phage cocktails that overcome
resistance development while remaining efficacious.
Previously, four STEC O157:H7-infecting bacteriophages
(vB_EcoS_AHP24, AHS24, AHP42 and AKS96) originally
isolated from cattle feedlots in southern Alberta, Canada were
classified as T1-like Siphoviridae by electron microscopy, but
exhibited divergent genotypes based on EcoRI- or HindIII-
digestion profiles [18]. This study aimed to further define their
genomic and proteomic characteristics as well as infectivity against
STEC O157:H7 and non-pathogenic E. coli (ECOR) strains. We
also conducted comparative genomic, proteomic and phylogenetic
analysis among known T1-like phages in an effort to determine
how these viruses could be optimally classified.
Materials and Methods
Bacteriophage, bacteria and media
Four phages infecting STEC O157:H7 strain R508 (phage type,
PT14) were isolated from the feces of commercial feedlot cattle in
2007 in Alberta, Canada [19] with AHP24 (Pen 10), AHS24 (Pen
10) and AHP42 (Pen 6) from Feedlot B and AKS97 (Pen 2) from
Feedlot A [19] with permission. A single discrete plaque from each
phage was purified three times by the soft agar (0.6%) overlay
method [20] and propagated as previously discribed [18]. Titers of
phages in the stock filtrates were then determined by the soft agar
overlay technique [20]. STEC O157:H7 strain R508 was used as a
host for plaque purification, propagation and titration of the phage
stocks. Other standard laboratory strains of STEC O157:H7
(n = 24) and non-O157 E. coli (n = 73) [21] used to evaluate host
range of four T1-like phages are listed in Table 1. Unless otherwise
indicated the bacterial strains were grown in tryptic soy broth
and/or tryptic soy agar.
Host range and lytic capability
Host range and lytic capability of the phages for STEC
O157:H7 and non-O157 E. coli was assessed using a microplate
phage virulence assay [22]. To estimate multiplicity of infection
(MOI), high titre phage stocks (10
9
210
10
PFU/ml) were serially
diluted and incubated at 37uC for 5 h with 10-fold diluted
overnight cultures of STEC O157:H7 in a 96-well microplate.
After incubation, wells were examined visually for turbidity and
the highest dilution that resulted in complete lysis (no discernable
turbidity) of bacteria was recorded. The MOI for each phage-host
assay was calculated by dividing the initial number of phages in the
highest-dilution wells by the initial number of bacteria added, as
determined by plate counts of serially diluted bacterial cultures.
CsCl density gradient centrifugation
Bacterial nucleic acids were removed from filtered phage lysates
(,10
9
PFU/ml) using DNase1 (Sigma-Aldrich, Oakville, ON,
Canada) and RNaseA (Sigma-Aldrich), and the phage lysates were
concentrated in polyethylene glycol (PEG) 8000 and purified
through two rounds of CsCl density gradient centrifugation [20].
Genome sequencing and annotation
Phage DNA was extracted from the CsCl-purified phage lysates
using the SDS-proteinase K protocol of Sambrook and Russell
[20]. Purified phage DNA was submitted to the Plate-forme
Table 1. Host range and lytic capability of phages AHP24, AHS24, AHP42 and AKS96.
Bacteria Strains
a
Sensitivity
b
of T1-like phages
AHP24 AHS24 AHP42 AKS96
STEC O157:H7 PT8, 33, 38 ++++
PT10, 14a, 28, 32, 34, 46248, 54, 68, 80, 88 +++ +++ +++ +++
PT24 +++ +++ +++ +
PT31 ++ ++ ++ ++
PT45 +++ +++ ++ +++
PT49 ++ ++ ++ ++
PT50, 67 +++ +++ ++ ++
PT51 +++ +++ + +
PT63 ++ ++ + +
PT74 ++ ++ +++ +++
non-O157 E. coli ECOR collection
c
2222
a
PT represents phage type of STEC O157:H7 strains
b
Sensitivities are grouped on the basis of multiplicity of infection (MOI: the lowest ratio of phage to bacteria that resulted in complete lysis of an overnight bacterial
culture during 5 h of incubation with serial dilutions of the phage). +++: extremely susceptible (MOI ,0.01); ++: highly susceptible (0.01 #MOI ,1); +: moderately
susceptible (1 #MOI ,10);
–: non-susceptible (i.e., no lysis observed).
c
ECOR collection represents standard reference strains of Escherichia coli [21].
doi:10.1371/journal.pone.0100426.t001
T1-Like Bacteriophages
PLOS ONE | www.plosone.org 2 June 2014 | Volume 9 | Issue 6 | e100426
Table 2. Features of the T1-like phages from NCBI database.
Phages Phylogenic Cluster Host Head Dimension (nm) Tail Dimension (nm) Genome size (bp) Mole% G
+
C Reference Accession #
TLS A E. coli,Shigella NA NA 49,902 42.7 51 NC_009540
FSL SP-126 A Salmonella NA NA 51,092 42.9 53 KC139513
vB_KpnS_KP36 B Klebsiella NA NA 49,818 50.7 55 NC_019781
F20 B Enterobacter 50 15067 51,543 47.9 54 JN672684
T1 C E. coli,Shigella 60 15168 48,836 45.6 51 NC_005833
Shfl1 C Shigella NA NA 50,661 45.4 NA NC_015456
ADB-2 C E. coli NA NA 50,552 46 52 NC_019725
RTP D E. coli NA NA 46,219 44.3 57 NC_007603
vB_EcoS_ACG-M12 D E. coli 57 17267 46,054 44 56 NC_019404
phiEB49 E E. coli 50 NA 47,180 44 58 JF770475
vB_EcoS_Rogue1 E E. coli 53 15268 45,805 44.2 40 NC_019718
phiJLA23 E E. coli NA NA 43,017 44.2 49 KC333879
phiKP26 E E. coli,Salmonella NA NA 47,285 44.3 50 KC579452
JK06 E E. coli NA NA 46,072 44 NA NC_007291
pSf-1 NA Shigella 73 103613 51,821 44 59 NC_021331
ESP2949-1 NA Cronobacter NA NA 49,116 50.1 60 NC_019509
vB_XveM_DIBBI NA Xanthomonas NA NA 49,981 52.4 NA NC_017981
a
NA: Not applicable.
doi:10.1371/journal.pone.0100426.t002
T1-Like Bacteriophages
PLOS ONE | www.plosone.org 3 June 2014 | Volume 9 | Issue 6 | e100426
d’Analyses Ge´nomiques of the Institut de Biologie Inte´grative et
des Syste`mes (Laval University, Que´bec, QC, Canada) for
sequencing. For each sample, a tagged GS-FLX rapid library
was made according to the manufacturer’s instructions (Roche/
454 sequencing, Brandford, USA). Phage libraries were pooled for
sequencing on a GS-FLX+instrument using titanium chemistry
according to the manufacturer’s instructions (Roche/454). Se-
quencing reads were assembled with the gsAssembler module of
Newbler v. 2.5.3.
Initial genome annotation was completed using myRAST [23].
Geneious v5.4 program (Biomatters Ltd., Auckland, New Zealand)
was used to visually scan the sequence for potential genes. All
predicted proteins were scanned for homologues using BLASTP
and PSI-BLAST [24]. Rho-independent terminators were identi-
fied using WebGeSTer at http://pallab.serc.iisc.ernet.in/gester/
rungester.html [25] and TransTermHP [26]. Promoters were
identified by neural network promoter prediction [27] with visual
inspection. Transfer RNA (tRNA) genes were screened using
Aragorn [28] at http://130.235.46.10/ARAGORN/and tRNAS-
can at http://lowelab.ucsc.edu/tRNAscan-SE/[29]. Transmem-
brane domains were described using TMHMM 2.0 at http://
www.cbs.dtu.dk/services/TMHMM/[30], Phobius at http://
phobius.sbc.su.se/[31] and SPLIT 4.0 at http://split.pmfst.hr/
split/4/[32]. The phage genomes were rendered syntenic by
opening at the initiation codon for the small subunit terminase
gene prior to dotplot alignment, ClustalW alignment and
EMBOSS Stretcher analysis. Whole genome sequences of the
four phages studied, as well as another 17 T1-like phages (Table 2)
from the database of National Center for Biotechnology Infor-
mation (NCBI) were analyzed by local ClustalW algorithm [33,34]
using default parameters and phylogenetic trees were visualized by
FigTree program (available from http://tree.bio.ed.ac.uk/
software/figtree/). CLUSTAL omega [35] was used to align
amino acid sequences of T1-like phages. The GenBank accession
numbers for AHP24, AHS24, AHP42 and AKS96 sequence are
KF771236, KF771238, KF771237 and KF771239, respectively.
Analysis of structural proteins
The in-gel digest and mass spectrometry experiments were
performed by the Proteomics platform of the Eastern Quebec
Genomics Center (Quebec, Canada). CsCl-purified phage parti-
cles were analyzed for structural proteins by standard Tris-glycine
12% sodium dodecyl sulfate-polyacrylamide gel electrophoresis
(SDS-PAGE). Samples were mixed with sample loading buffer and
boiled for 5 min before loading. Proteins were stained with
Coomassie brilliant blue R250 (Bio-Rad Laboratories, Missis-
sauga, ON, Canada) and subsequently characterized using
Bionumerics 6.6 software (Applied Maths, Austin, TX, USA).
Bands of interest were excised and de-stained with water. Tryptic
digestion was performed on a MassPrep liquid handling robot
(Waters, Milford, USA) according to the manufacturer’s specifi-
cations and to the modified protocol [36,37]. Briefly, following
reduction with 10 mM dithiothreitol (DTT) and alkylation with
55 mM iodoacetamide, the protein was digested by 126 nM of
modified porcine trypsin (Sequencing grade, Promega, Madison,
WI, USA) at 58uC for 1 h. The proteolytic peptides were then
extracted using 1% formic acid, 2% acetonitrile followed by 1%
formic acid and 50% acetonitrile. The recovered extracts were
pooled, dried by vacuum centrifuge and resuspended into 7 mlof
0.1% formic acid for mass spectrometry. Peptide resuspensions
(2 ml) were separated by online reversed-phase (RP) nanoscale
capillary liquid chromatography (nanoLC) and analyzed by
electrospray mass spectrometry (ES MS/MS). The experiments
were performed with a Agilent 1200 nano pump connected to a
triple time-of-flight mass spectrometer (AB Sciex 5600, Framing-
ham, MA,USA) equipped with a nanoelectrospray ion source (AB
Sciex 5600). Briefly, 2 ml of the peptide resuspension was injected
onto a 15 cm675 mm (internal diameter) PicoFrit column (New
Objective, Woburn, MA), packed with reversed phase C18
particles (5 mm diameter; 300 A
˚pore size; Jupiter 300, Phenom-
enex, Torrance, CA, USA) and eluted in a linear gradient from 2–
50% buffer B (0.1% formic acid in acetonitrile) at flow rate of
300 nl/min for 30 min. Mass spectra were acquired using a data-
dependent acquisition mode using Analyst software version 1.6
(AB Sciex 5600). Each full scan mass spectrum (400 to 1250 m/z)
was followed by collision-induced dissociation of the twenty most
intense ions. Dynamic exclusion was set for a period of 3 sec and a
tolerance of 100 ppm.
All MS/MS peak lists were generated with ProteinPilot Version
4.5 (AB Sciex, Framingham, MA, USA) and analyzed using
Table 3. General genome features of AHP24, AHS24, AHP42 and AKS96.
Phages
Feature AHP24 AHS24 AHP42 AKS96
Size (bp) 46,719 46,440 46,847 45,746
Sequence coverage 31.6 74.6 23.4 47.5
G+C content (%) 43.8 43.8 44 43.9
Total ORFs 78 81 76 74
Average ORF size (bp) 552 531 560 569
% of genome coding for proteins 92.2 92.8 90.9 92.2
No. of gene products similar to known proteins, total 73 71 72 69
No. of gene products similar to known T1, total 72 70 71 68
No. of conserved hypothetical proteins with unknown function 43 42 44 42
No. of hypothetical proteins 5 10 5 4
No. of tRNAs 1111
No. of s
70
promoters 56107
No. of rho-independent terminators 18 18 21 21
doi:10.1371/journal.pone.0100426.t003
T1-Like Bacteriophages
PLOS ONE | www.plosone.org 4 June 2014 | Volume 9 | Issue 6 | e100426
T1-Like Bacteriophages
PLOS ONE | www.plosone.org 5 June 2014 | Volume 9 | Issue 6 | e100426
Mascot (Matrix Science, London, UK; version 2.3.02). Proteins
were identified by searching against the Uniref100-SiphoViridea
database (release 12-05) as well as an in-house protein database
derived from genome sequences of AHP24, AHS24, AHP42 and
AKS96. Parameters for Mascot used a fragment ion mass
tolerance of 0.10 Da and a parent ion tolerance of 0.10 Da.
Iodoacetamide derivatives of cysteine were specified as a fixed
modification and oxidation products of methionine were specified
as a variable modification with up to two missed cleavages
allowed. Scaffold version 4.0.1 (Proteome Software Inc., Portland,
OR, USA) was used to validate MS/MS based peptide and
protein identifications. The protein identification cut off was set at
a confidence level of 95% (MASCOT score .33) with a
requirement for at least two peptides to match to a protein.
Proteins with similar peptides that could not be differentiated
based on MS/MS analysis alone were grouped to satisfy the
principles of parsimony.
Results
Phages were able to lyse all 24 STEC O157:H7 strains tested,
but displayed no activity against any of the 73 non-O157 E. coli
strains (Table 1). Phages AHP24 and AHS24 exhibited the same
infective pattern, with 17 strains extremely susceptible, 4 strains
highly susceptible and 3 strains moderately susceptible. On the
basis of MOI value, the lytic capability of these two phages was
slightly higher than AHP42 or AKS96.
General genomic feature
All four phages contained circularly permuted genomes of 45.7–
46.8 kb (43.8–44 mol% G+C) encoding 74–81 open reading
frames (ORFs) and 1 arginyl-tRNA (Tables 3 and S12S4).
Furthermore, 18221 rho-dependent terminators and 5210
promoters recognized by host RNA polymerase were identified.
The majority (68–72 ORFs, 86–94%) of the proteins displayed
homology to proteins of other T1-like phages with 32239 of DNA
replication, morphogenesis, genome packing and lysis (Tables 3
and S12S4). Based on functional comparison of the ORFs to the
NCBI database of non-redundant protein sequences, none of the
genes encoded for proteins associated with pathogenesis, integra-
tion or antibiotic resistance.
Comparative analysis
Comparative computational genomic analysis revealed that the
four T1-like phages were collinear (.90.8% pairwise similarity),
with AHP24 and AHS24 having the highest sequence identity
(99.2%). All four phages were 80.8–86.5% identical to phages
JK06, vB_EcoS_Rogue1, phiJLA23 and phiKP26 and 65%
similar to phiEB49 in Cluster E and ,56% related to other
known T1-like phages (Fig. 1A, Table 4). At the protein level,
computational analysis of CoreGenes demonstrated that com-
pared to phage vB_EcoS_Rogue1, phages AHP24 and AHS24
shared greatest number of proteins (67–68, Avg. 91.2% similarity)
in common, followed by AHP42, AKS96, phiKP26 and phiJLA23
(65, 87.8%), phiJLA23, JK06 and phiEB49 (56–58,
Avg.77%)(Table 5).
To obtain a global phylogenetic overview of the relationships
between the T1-like phages, we employed genomic dot-plots of
these genome sequences against each other (Fig. 1A). Clearly,
nucleotide sequence aligned well within each cluster. Phylogenic
analysis of whole genome also demonstrated that phages within
each cluster shared close relatedness at nucleotide level (Fig. 1B).
Nucleotide similarity of phages within each cluster was 82.6%
(Cluster A), 72.8% (Cluster B), 77.6–81.5% (Cluster C), 65.1%
(Cluster D) and 64.2–99.2% (Cluster E), whereas nucleotide
identity shared between each cluster was 48.6–55.7% (Fig. 1C,
Tables 2 and 4). Phages pSf-1 and ESP2949-1 demonstrated lower
nucleotide similarity to phages from each genus (Table 4).
Computational analysis of CoreGenes showed that phages within
same cluster had greater number of homologues (75.7–91.2%)
than those among different clusters (43.7–68%; Table 5). Orphan
phage species pSf-1 and ESP2949-1 did not have over 55 gene
products (,64%) in common as compared with other phages in
each cluster. Considering the close relatedness at both nucleotide
and protein level exhibited by the phages within each cluster, we
propose the establishment of a new subfamily ‘‘Tunavirinae’’
which can be divided into five genera, i.e. ‘‘Tlslikevirus’’ (Cluster
A), ‘‘Kp36likevirus’’ (Cluster B), Tunalikevirus (Cluster C), ‘‘Rtpli-
kevirus’’ (Cluster D) and ‘‘Jk06likevirus’’ (Cluster E) (Fig. 1,
Tables 4 and 5), each of which is named after the first isolated
phage of its type.
Phylogenetic trees were constructed to further investigate
common proteomic features for the large subunit of terminase
(TerL), portal protein (PorT), tail fiber (FibA) and major capsid
proteins (CapS) (Fig. 2). Overall, these analyses substantiated the
establishment of the proposed genera. Interestingly, PorT and
CapS of phage phiEB49 were more closely related to those from
the ‘‘Rtplikevirus’’ (84.4–93% aa identity, ID) than those from the
‘‘Jk06likevirus’’ (70.4–75.2% aa ID). Within the ‘‘Jk06likevirus’’,
CapS from AHP24, AHS24, AHP42 and AKS96 (100% aa ID)
was found to be 70.4–72.3% (aa) related to that from phages JK06,
Rogue1, phiKP26 and phiJLA23 (97.8–99.7% aa ID). Likewise
high diversity of the whole genome presented by orphan phages
pSf-1 and ESP2949-1, low amino acid sequence similarities (,
73.5%) were identified for each of the proteins studied, as
compared to those of other members of the T1-like family.
Proteomics
SDS-PAGE revealed that the structural proteins generated
identical banding patterns among the four phages (Fig. 3). Further
shotgun proteomics by liquid chromatography-tandem mass
spectrometry identified up to 52% of the amino acids in seven
structural proteins including tail fiber, tail tape measure protein,
major capsid, portal protein as well as major and minor tail
proteins (Table 6 and Fig. 3). A major capsid protein (Fig. 3, band
D) was observed to have a molecular mass of 29.7 kDa, similar to
the 33 kDa of the major head subunit P7 protein previously
identified from T1 phage [38]. Also, a conserved hypothetical
protein with a molecular mass of 14.2 kDa (Table 6; Fig. 3, band
Figure 1. Comparative genomic analysis of the 21 known T1-like phages. A, Dot plot alignment of nucleotide identity of the 21 known T1-
like phages using Gepard [61]. The vertical axis shows the phage IDs and horizontal axis indicates phage clusters (highlighted in red box). The
apparent black diagonal lines indicate high degrees of nucleotide sequence identity; while each phage shows 100% identity to itself (displayed as
diagonal line). B, Phylogenetic analysis of whole genomes of the 21 known T1-like phages by ClustalW algorithm. Scale bar represents 0.1
substitutions. C, Whole genome comparisons of phages AHP24 (A), KP36 (B), Rogue1 (C), RTP (D), TLS (E) and T1 (F) using a progressive MAUVE
alignment [62]. The degree of sequence similarity is indicated by the intensity of the colored region. The contiguous black boxes under the colored
region represent the position of the genes; red, large subunit of terminase; green, tail tape measure protein; blue, tail fiber protein I; black, tail fiber
protein II; orange, helicase; pink, common hypothetical protein.
doi:10.1371/journal.pone.0100426.g001
T1-Like Bacteriophages
PLOS ONE | www.plosone.org 6 June 2014 | Volume 9 | Issue 6 | e100426
Table 4. Pairwise nucleotide sequence identity of T1-like collinearized genomes as calculated by EMBOSS Stretcher[44,45].
Cluster Genus DNA sequence similarity (%)
Within
genus between genus and/or species
a
‘‘Tlslikevirus’’ ‘‘Kp36likevirus’’
Tunalikevirus
‘‘Rtplikevirus’’ ‘‘Jk06likevirus’’ Putative orphan species
pSf-1 ESP2949-1
A ‘‘Tlslikevirus’’ 82.6 100 53.4 54.9 48.9 50.3 57.4 56.4
B ‘‘Kp36likevirus’’ 72.8 100 53.3 48.6 49.3 52.6 53.3
CTunalikevirus 77.6–81.5 100 50.6 51.3 54.7 54.6
D ‘‘Rtplikevirus’’ 65.1 100 55.7 49.6 49
E ‘‘Jk06likevirus’’ 64.2–99.2 100 50.5 49.9
a
DNA sequence similarity between different genera/or species was calculated using phage TLS a reference genome for the ‘‘Tlslikevirus’’, phage KP36 for the ‘‘Kp36likevirus’’, phage T1 for the Tunalikevirus, phage RTP for the
‘‘Rtplikevirus’’ and phage JK06 for the ‘‘Jk06likevirus’’.
doi:10.1371/journal.pone.0100426.t004
Table 5. Proteomic analysis of T1-like collinearized genomes using CoreGenes 3.0 program [46–48].
Cluster Genus No. of common proteins shared (%)
a
Within
genus between genus and/or species
‘‘Tlslikevirus’’ ‘‘Kp36likevirus’’
Tunalikevirus
‘‘Rtplikevirus’’ ‘‘Jk06likevirus’’ Putative orphan species
pSf-1 ESP2949-1
A ‘‘Tlslikevirus’’ 72 (82.8) (100) 45 (51.7) 46 (52.9) 42 (48.3) 38 (43.7) 55 (63.2) 35 (46.7)
B ‘‘Kp36likevirus’’ 67 (83.8) (100) 41 (51.3) 38 (47.5) 36 (45) 42 (52.5) 38 (47.5)
CTunalikevirus 61–66 (78.1–
84.6)
(100) 38 (48.7) 37 (47.4) 47 (60.3) 36 (46.2)
D ‘‘Rtplikevirus’’ 58 (77.3) (100) 51 (68) 42 (52.5) 38 (47.5)
E ‘‘Jk06likevirus’’ 56–68 (75.7–
91.2)
(100) 41 (55.4) 33 (44.6)
a
Percentage value in each row is the ratio of homologs shared to total genes which was calculated using phage TLS a reference genome for the ‘‘Tlslikevirus’’, phage KP36 for the ‘‘Kp36likevirus’’, phage T1 for the Tunalikevirus,
phage RTP for the ‘‘Rtplikevirus’’ and phage Rogue1 for the ‘‘Jk06likevirus’’.
doi:10.1371/journal.pone.0100426.t005
T1-Like Bacteriophages
PLOS ONE | www.plosone.org 7 June 2014 | Volume 9 | Issue 6 | e100426
G) was similar to P11 (16 kDa) from phage T1, which has been
proposed to be a second major head component that stabilizes the
later stages of head assembly [39]. A major 23.6 kDa tail protein
(Table 6; Fig. 3, band F) was consistent with a major tail protein
from phage Rogue1 (gp29, 25.9kDa) [40] and from phage T1
(P10, 26 kDa) [39].
Discussion
This study revealed that phages AHP24, AHS24, AHP42 and
AKS96 are closely related members of new proposed genus–
‘‘Jk06likevirus’’. Not surprisingly, the highest degree of nucleotide
identity was shared between AHP24 and AHS24, as they were
isolated simultaneously from fecal pats and manure slurry from the
same feedlot pen [19]. AHP42 and AKS96 originated from
different feedlots, but displayed the second highest degree of
nucleotide sequence similarity, a result that confirms our previous
findings of genomic relatedness of these two isolates based on
restriction enzyme profiles [18]. Our ongoing work has also
characterized a number of additional STEC O157:H7-infecting
phages with TEM morphology, genome size and restriction
enzyme profiles (Niu et al. unpublished data) that are similar to the
four phages in this study, possibly because they were also obtained
from the same commercial feedlots in 2007 [19]. This may suggest
that ‘‘Jk06likevirus’’ are widespread in Alberta feedlots. All four
phages were active against a broad range of STEC O157:H7
reference strains, but did not target non-O157 E. coli, suggesting
that they could be used to control STEC O157:H7 without
harming generic commensal E. coli. Also, the four phages exhibited
strong lytic capability against vast majority of PT strains of STEC
Figure 2. Evolutionary relationships of major proteins. The evolutionary history was inferred using the Neighbor-Joining method [63]. The
optimal tree with the sum of branch length for A, large subunit of terminase ( = 2.05), B, portal protein (= 2.2), C, tail fiber ( =1.61), D, major capsid
( = 2.05), is shown; The percentage of replicate trees in which the associated taxa clustered together in the bootstrap test (500 replicates) are shown
next to the branches [64]. The tree is drawn to scale, with branch lengths in the same units as those of the evolutionary distances used to infer the
phylogenetic tree. The evolutionary distances were computed using the JTT matrix-based method [65] and are in the units of the number of amino
acid substitutions per site. The analysis involved 20 amino acid sequences. All positions containing gaps and missing data were eliminated. There
were a total of 429 positions (A), 244 positions (B), 807 positions (C), 305 positions (D), in each final dataset. Evolutionary analyses were conductedin
MEGA5 [66]. Scale bar represents 0.1 substitutions. The asterisk represents phages of which evolutionary relationships of major proteins differed from
those of whole genome.
doi:10.1371/journal.pone.0100426.g002
T1-Like Bacteriophages
PLOS ONE | www.plosone.org 8 June 2014 | Volume 9 | Issue 6 | e100426
O157:H7, although lytic capability may vary with propagation
hosts. This would make them effective biocontrol agents and
possible low dosage required for therapeutic application.
A total of 21 phages with genome sizes ranging from 43 to
52 kb, similar genomic structure and TEM morphology have been
described. Based on current taxonomic classification of ICTV,
these phages were classified as T1-like phage (Tunalikevirus) within
Siphoviridae. Undoubtedly, more T1-like phages will be identified in
the future and there is a need to establish a more defined
taxonomic system in order to explore the evolutionary relation-
ships and genetic linkages in these types of phage. In the present
study, we aligned whole genome sequences from all 21 T1-like
phages using the ClustalW algorithm, which has been widely used
for nucleotide sequence alignment of viruses [41-43]. The
phylogenetic analysis showed that the T1-related phages fall into
five clusters. Moreover, computational EMBOSS Stretcher
[44,45] and CoreGenes programs [46–48] showed that phages
within each proposed genus were more closely related than those
among genera at both the nucleotide and protein level. This was
also confirmed by the phylogenetic analysis of four key functional
phage proteins. The fact that the viruses related to JK06 have been
isolated independently in Israel (JK06) (GenBank Assession #,
NC_007291), Canada (vB_EcoS_Rogue1, AHP24, AHS24,
AHP42, AKS96) [40] and Mexico (phiKP26, phiJLA23) [49,50]
between 2005 and 2011 indicates that these similar phages are
widely distributed, and that horizontal gene transfer does not
always prevent the characterization of bacteriophage evolution.
Similar finding have been noted as part of the Phage Hunters
Integrating Research and Education (PHIRE) program (http://
phagesdb.org/) and for the global distribution of viruses related to
Listeria phage A511. The results indicate that a new subfamily, the
Figure 3. T1-like structural proteins (Lane 2-5) alongside the
standard marker (Lane1) separated on 12% SDS-PAGE gel and
visualized by Coomassie brilliant blue R250 stain. A, tail fiber
protein; B, tail tape measure protein; C, portal protein; D, major capsid
protein; E, minor tail protein; F, major tail protein; G, conserved
hypothetical protein.
doi:10.1371/journal.pone.0100426.g003
Table 6. Structural proteins of phages AHP24, AHS24, AHP42 and AKS96 identified by mass spectrometry.
Gel band ORF in T1-like phages Theoretical mass (kDa) Observed mass (kDa) Putative function No. of peptides
Sequence coverage
(%)
AHP24 AHS24 AHP42 AKS96
A 24 25 25 23 124.4 128 Tail fiber protein 11–23 11–24
B 18 19 19 18 108.9 94.8 Tail tape measure protein 43–55 43–50
C 04 04 04 04 46.3 43.9 Portal protein 14–25 29–47
D 09 10 10 09 34.7 29.7 Major capsid protein 3–7 9–27
E 21 22 21 20 28.1 27.7 Minor tail protein 3–10 10–50
F 15 16 16 15 23.5 23.6 Major tail protein 5–9 31–52
G 12 13 13 12 13.5 14.2 Conserved hypothetical
protein
3–6 27–42
doi:10.1371/journal.pone.0100426.t006
T1-Like Bacteriophages
PLOS ONE | www.plosone.org 9 June 2014 | Volume 9 | Issue 6 | e100426
‘‘Tunavirinae’’ created within the family Siphoviridae containing the
following genera: a modified Tunalikevirus (T1, ADB-2, Shfl1)
[51,52], ‘‘Tlslikevirus’’ (TLS, FSL SP-126) [51,53], ‘‘Kp36like-
virus’’ (KP36, F20) [54,55], ‘‘Rtplikevirus’’ (RTP, vB_EcoS_ACG-
M12) [56,57]; and ‘‘Jk06likevirus’’ (JK06, vB_EcoS_Rogue1,
AHP24, AHS24, AHP42, AKS96, phiJLA23, phiKP26, phiEB49)
[40,49,50,58]. This would leave two putative orphan species: pSf-1
[59] and ESP2949-1 [60] to be further classified as more phages
are characterized. There is a move within ICTV to eliminate the
order Caudovirales, and its three families (Myoviridae, Siphoviridae and
Podoviridae) as they are not compatible with emerging genomic and
proteomic information on phage phylogeny.
Mitigation of STEC O157:H7 has been a challenge in feedlot
cattle. The newly discovered four members of ‘‘Jk06likevirus’’
exhibited broad host range and strong lytic capability against
STEC O157:H7, emphasizing efficacy and suitability for phage-
based biocontrol of this zoonotic pathogen. In this study, we also
proposed further classification of the 21 known T1-like phages into
one subfamily with five genera, constructing a basis for proper
identification of new phages within the same type. Genomic- and
proteomic- based taxonomic classification of phages would
facilitate a better understanding of phage diversity and genetic
traits involved in phage evolution.
Supporting Information
Table S1 Feature of phage AHP24 gene products and their
functional assignments.
(XLSX)
Table S2 Feature of phage AHS24 gene products and their
functional assignments.
(XLSX)
Table S3 Feature of phage AHP42 gene products and their
functional assignments.
(XLSX)
Table S4 Feature of phage AKS96 gene products and their
functional assignments.
(XLSX)
Acknowledgments
We gratefully acknowledge J. Peters and H. Zahiroddini for technical
assistance and support, and R. Johnson (Public Health Agency of Canada,
Guelph, ON, Canada) for kind provision of bacterial strains.
Author Contributions
Conceived and designed the experiments: YDN TAM KS. Performed the
experiments: YDN. Analyzed the data: YDN JHEN AMK. Contributed
reagents/materials/analysis tools: TAM KS. Wrote the paper: YDN TAM
JHEN AMK KS.
References
1. Ackermann H-W (2009) Phage classification and characterization. Methods Mol
Biol 501: 127–140.
2. Lavigne R, Seto D, Mahadevan P, Ackermann HW, Kropinski AM (2008)
Unifying classical and molecular taxonomic classification: analysis of the
Podoviridae using BLASTP-based tools. Res Microbiol 159: 406–414.
3. Reyes A, Semenkovich NP, Whiteson K, Rohwer F, Gordon JI (2012) Going
viral: next-generation sequencing applied to phage populations in the human
gut. Nat Rev Microbiol 10: 607–617.
4. Comeau AM, Tremblay D, Moineau S, Rattei T, Kushkina AI, et al. (2012)
Phage morphology recapitulates phylogeny: the comparative genomics of a new
group of myoviruses. PLoS ONE 7: e40102.
5. German GJ, Misra R, Kropinski AM (2006) The T1-like bacteriophages, pp.
211–224. In Calendar R (ed.), The bacteriophages, 2nd ed. New York, NY:
Oxford University Press.
6. King AM, Lefkowitz E, Adams MJ, Carstens EB (2012) Virus Taxonomy: Ninth
Report of the International Committee on Taxonomy of Viruses. San Diego,
California, CA: Elsevier Academic Press.
7. Nyachuba DG (2010) Foodborne illness: is it on the rise? Nutr Rev 68: 257–269.
8. Public Health Agency of Canada (2011) National enteric surveillance program
(NESP) annual summary. Available at https://www.nml-lnm.gc.ca/NESP-
PNSME/surveillance-2011-eng.html (accessed 12 December 2013).
9. Maura D, Debarbieux L (2011) Bacteriophages as twenty-first century
antibacterial tools for food and medicine. Appl Microbiol Biotechnol 90: 851–
859.
10. Johnson RP, Gyles CL, Huff WE, Ojha S, Huff GR, et al. (2008) Bacteriophages
for prophylaxis and therapy in cattle, poultry and pigs. Anim Health Res Rev 9:
201–215.
11. Atterbury RJ (2009) Bacteriophage biocontrol in animals and meat products.
Microb Biotechnol 2: 601–612.
12. Balogh B, Jones JB, Iriarte FB, Momol MT (2010) Phage therapy for plant
disease control. Curr Pharm Biotechnol 11: 48–57.
13. Stanford K, McAllister TA, Niu YD, Stephens TP, Mazzocco A, et al. (2010)
Oral delivery systems for encapsulated bacteriophages targeted Escherichia coli
O157:H7 in feedlot cattle. J Food Prot 73: 1304–1312.
14. Rozema EA, Stephens TP, Bach SJ, Okine EK, Johnson RP, et al. (2009) Oral
and rectal administration of bacteriophages for control of Escherichia coli
O157:H7 in feedlot cattle. J Food Prot 72: 241–250.
15. Callaway TR, Edrington TS, Brabban A, Kutter B, Karriker L, et al. (201 1)
Evaluation of phage treatment as a strategy to reduce Salmonella populations in
growing swine. Foodborne Pathog Dis 8: 261–266.
16. Carvalho CM, Santos SB, Kropinski AM, Ferreira EC, Azeredo J (2012) Phages
as Therapeutic Tools to Control Major Foodborne Pathogens: Campylobacter and
Salmonella.In Kurtboke I (ed.), Bacteriophages. Rijeka, Croatia: InTech.
17. Liu H, Niu YD, Li JQ, Stanford K, McAllister TA (2013) June 18–20. In-Vitro
efficacy of bacteriophages T5, T4, T1 and O1 against E. coli O157 :H7, Joint
Canadian Society of Animal Science and Canadian Meat Science Association
Congress, Banff, Canada.
18. Niu YD, Stanford K, Ackermann HW, McAllister TA (2012) Characterization
of 4 T1-like lytic bacteriophages that lyse Shiga-toxin Escherichia coli O157:H7.
Can J Microbiol 58: 923–927.
19. Niu YD, McAllister TA, Xu Y, Johnson RP, Stephens TP, et al. (2009)
Prevalence and impact of bacteriophages on the presence of Escherichia coli
O157:H7 in feedlot cattle and their environment. Appl Environ Microbiol 75:
1271–1278.
20. Sambrook J, Russell DW (2001) Molecular cloning: a laboratory manual-3rd ed.
New York, NY: Cold Spring Laboratory Press, Cold Spring Harbor.
21. Ochman H, Selander RK (1984) Standard reference strains of Escherichia coli
from natural populations. J Bacteriol 157: 690–693.
22. Niu YD, Johnson RP, Xu Y, McAllister TA, Sharma R, et al. (2009) 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 107:
646–656.
23. Aziz RK, Bartels D, Best AA, DeJongh M, Disz T, et al. (2008) The RAST
Server: rapid annotations using subsystems technology. BMC Genomics 9: 75.
24. Altschul SF, Madden TL, Schaffer AA, Zhang J, Zhang Z, et al. (1997) Gapped
BLAST and PSI-BLAST: a new generation of protein databa se search
programs. Nucleic Acids Res 25: 3389–3402.
25. Mitra A, Kesarwani AK, Pal D, Nagaraja V (2011) WebGeSTer DB-a
transcription terminator database. Nucleic Acids Res 39: D129–135.
26. Kingsford CL, Ayanbule K, Salzberg SL (2007) Rapid, accurate, computational
discovery of Rho-independent transcription terminators illuminates their
relationship to DNA uptake. Genome Biol 8: R22.
27. Reese MG (2001) Application of a time-delay neural network to promoter
annotation in the Drosophila melanogaster genome. Comput Chem 26: 51–56.
28. Laslett D, Canback B (2004) ARAGORN, a program to detect tRNA genes and
tmRNA genes in nucleotide sequences. Nucleic Acids Res 32: 11–16.
29. Lowe TM, Eddy SR (1997) tRNAscan-SE: a program for improved detection of
transfer RNA genes in genomic sequence. Nucleic Acids Res 25: 955–964.
30. Sonnhammer EL, von Heijne G, Krogh A (1998) A hidden Markov model for
predicting transmembrane helices in protein sequences. Proc Int Conf Intell Syst
Mol Biol 6: 175–182.
31. Kall L, Krogh A, Sonnhammer EL (2004) A combined transmembrane topology
and signal peptide prediction method. J Mol Biol 338: 1027–1036.
32. Juretic D, Zoranic L, Zucic D (2002) Basic charge clusters and predictions of
membrane protein topology. J Chem Inf Comput Sci 42: 620–632.
33. Thompson JD, Higgins DG, Gibson TJ (1994) CLUSTAL W: improving the
sensitivity of progressive multiple sequence alig nment through sequence
weighting, position-specific gap penalties and weight matrix choice. Nucleic
Acids Res 22: 4673–4680.
34. Li KB (2003) ClustalW-MPI: ClustalW analysis using distributed and parallel
computing. Bioinformatics 19: 1585–1586.
T1-Like Bacteriophages
PLOS ONE | www.plosone.org 10 June 2014 | Volume 9 | Issue 6 | e100426
35. Sievers F, Wilm A, Dineen D, Gibson TJ, Karplus K, et al. (2011) Fast, scalable
generation of high-quality protein multiple sequence alignments using Clustal
Omega. Mol Syst Biol 7: 539.
36. Shevchenko A, Wilm M, Vorm O, Mann M (1996) Mass spectrometric
sequencing of proteins silver-stained polyacrylamide gels. Anal Chem 68: 850–
858.
37. Havlis J, Thomas H, Sebela M, Shevchenko A (2003) Fast-response proteomics
by accelerated in-gel digestion of proteins. Anal Chem 75: 1300–1306.
38. Martin DT, Adair CA, Ritchie DA (1976) Polypeptides specified by
bacteriophage T1. J Gen Virol 33: 309–319.
39. Ramsay N, Ritchie DA (1984) Phage head assembly in bacteriophage T1.
Virology 132: 239–249.
40. Kropinski AM, Lingohr EJ, Moyles DM, Ojha S, Mazzocco A, et al. (2012)
Endemic bacteriophages: a cautionary tale for evaluation of bacteriophage
therapy and other interventions for infection control in animals. Virol J 9: 207.
41. Westover KM, Rusinko JP, Hoin J, Neal M (2013) Rogue taxa phenomenon: a
biological companion to simulation analysis. Mol Phylogenet Evol 69: 1–3.
42. Alonso C, Murtaugh MP, Dee SA, Davies PR (2013) Epidemiological study of
air filtration systems for preventing PRRSV infection in large sow herds. Prev
Vet Med 112: 109–117.
43. Mollov D, Lockhart B, Zlesak D (2013) Complete nucleotide sequence of rose
yellow mosaic virus, a novel member of the family Potyviridae. Arch Virol 158:
1917–1923.
44. Olson SA (2002) EMBOSS opens up sequence analysis. European Molecular
Biology Open Software Suite. Brief Bioinform 3: 87–91.
45. Rice P, Longden I, Bleasby A (2000) EMBOSS: the European Molecular
Biology Open Software Suite. Trends Genet 16: 276–277.
46. Mahadevan P, King JF, Seto D (2009) CGUG: in silico proteome and genome
parsing tool for the determination of "core" and unique genes in the analysis of
genomes up to ca. 1.9 Mb. BMC Res Notes 2: 168.
47. Mahadevan P, King JF, Seto D (2009) Data mining pathogen genomes using
GeneOrder and CoreGenes and CGUG: gene order, synteny and in silico
proteomes. Int J Comput Biol Drug Des 2: 100–114.
48. Mahadevan P, Seto D (2010) Taxonomic parsing of bacteriophages using core
genes and in silico proteome-based CGUG and applications to small bacterial
genomes. Adv Exp Med Biol 680: 379–385.
49. Amarillas L, Chaidez C, Lugo Y, Leon-Felix J (2013) Complete genome
sequence of Escherichia coli O157:H7 bacteriophage phiJLA23 isolated in Mexico.
Genome Announc 1: art.no. e00219–00212.
50. Amarillas L, Chaidez-Quiroz C, Sanudo-Barajas A, Leon-Felix J (2013)
Complete genome sequence of a polyvalent bacteriophage, phiKP26, active
on Salmonella and Escherichia coli. Arch Virol 158: 2395–2398.
51. Roberts MD, Martin NL, Kropinski AM (2004) The genome and proteome of
coliphage T1. Virology 318: 245–266.
52. Bhensdadia DV, Bhimani HD, Rawal CM, Kothari VV, Raval VH, et al. (2013)
Complete genome sequence of Escherichia phage ADB-2 isolated from a fecal
sample of poultry. Genome Announc 1: art.no. e0004313.
53. Moreno Switt AI, Orsi RH, den Bakker HC, Vongkamjan K, Altier C, et al.
(2013) Genomic characterization provides new insight into Salmonella phage
diversity. BMC Genomics 14: 481.
54. Mishra CK, Choi TJ, Kang SC (2012) Isolation and characterization of a
bacteriophage F20 virulent to Enterobacter aerogenes. J Gen Virol 93: 2310–2314.
55. Ke˛ sik-Szeloch A, Drulis-Kawa Z, Weber-Da˛browska B, Kassner J, Majkowska-
Skrobek G, et al. (2013) Characterising the biology of novel lytic bacteriophages
infecting multidrug resistant Klebsiella pneumoniae. Virol J 10: 100.
56. Chibeu A, Lingohr EJ, Masson L, Manges A, Harel J, et al. (2012)
Bacteriophages with the ability to degrade uropathogenic Escherichia coli biofilms.
Viruses 4: 471–487.
57. Wietzorrek A, Schwarz H, Herrmann C, Braun V (2006) The genome of the
novel phage Rtp, with a rosette-like tail tip, is homologous to the genome of
phage T1. J Bacteriol 188: 1419–1436.
58. Battaglioli EJ, Baisa GA, Weeks AE, Schroll RA, Hryckowian AJ, et al (2011)
Isolation of generalized transducing bacteriophages for uropathogenic strains of
Escherichia coli. Appl Environ Microbiol 77: 6630–6635.
59. Jun JW, Kim JH, Shin SP, Han JE, Chai JY, et al. (2013) Characterization and
complete genome sequence of the Shigella bacteriophage pSf-1. Res Microbiol
164: 979–986.
60. Lee YD, Kim JY, Park JH, Chang H (2012) Genomic analysis of bacteriophage
ESP2949-1, which is virulent for Cronobacter sakazakii. Arch Virol 157: 199–202.
61. Krumsiek J, Arnold R, Rattei T (2007) Gepard: a rapid and sensitive tool for
creating dotplots on genome scale. Bioinformatics 23: 1026–1028.
62. Darling AE, Mau B, Perna NT (2010) progressiveMauve: Multiple genome
alignment with gene gain, loss and rearrangement. PLoS One 5: art.no.e11147.
63. Saitou N, Nei M (1987) The neighbor-joining method: a new method for
reconstructing phylogenetic trees. Mol Biol Evol 4: 406–425.
64. Felsenstein J (1985) Confidence limits on phylogenies: an approach using the
bootstrap. Evolution 39: 783–791.
65. Jones DT, Taylor WR, Thornton JM (1992) The rapid generation of mutation
data matrices from protein sequences. Comput Appl Biosci 8: 275–282.
66. Tamura K, Peterson D, Peterson N, Stecher G, Nei M, et al. (2011) MEGA5:
molecular evolutionary genetics analysis using maximum likelihood, evolution-
ary distance, and maximum parsimony methods. Mol Biol Evol 28: 2731–27 39.
T1-Like Bacteriophages
PLOS ONE | www.plosone.org 11 June 2014 | Volume 9 | Issue 6 | e100426
... The current study focused on characterizing a new Rogunavirus phage for its antimicrobial effects on E. coli O157:H7. Phage UDF157lw harbors genomic features, including ∼45 kb genome size, ∼44% GC content, and 1 tRNA, and a morphology similar to several reference phages belonging to the Rogunavirus genus (Kropinski et al., 2012;Niu et al., 2014). Furthermore, according to the International Committee on Taxonomy of Virus (ICTV) taxonomy guideline (Turner et al., 2021), the comparative genomics of this study suggest that phage UDF157lw is a new member belonging to the Rogunavirus genus under the Drexlerviridae family. ...
... Furthermore, according to the International Committee on Taxonomy of Virus (ICTV) taxonomy guideline (Turner et al., 2021), the comparative genomics of this study suggest that phage UDF157lw is a new member belonging to the Rogunavirus genus under the Drexlerviridae family. The phages classified in the Rogunavirus genus, previously known as the Jk06likevirus genus, render lytic infection specific to E. coli O157:H7; however, the receptor-binding proteins of this group of phages are not well discovered (Niu et al., 2014). In addition, the narrow host range of these phages containing a long and non-contractile tail, known as Siphoviridae morphology, is primarily affected by the tail structures, such as tail fiber or tail spike proteins, compared to Myoviridae phages, which encode a long and contractile tail with a broad host range (Chibani-Chennoufi et al., 2004;Sørensen et al., 2021). ...
Article
Full-text available
Introduction Shiga toxin-producing Escherichia coli (STEC) O157:H7 is one of the notorious foodborne pathogens causing high mortality through the consumption of contaminated food items. The food safety risk from STEC pathogens could escalate when a group of bacterial cells aggregates to form a biofilm. Bacterial biofilm can diminish the effects of various antimicrobial interventions and enhance the pathogenicity of the pathogens. Therefore, there is an urgent need to have effective control measurements. Bacteriophages can kill the target bacterial cells through lytic infection, and some enzymes produced during the infection have the capability to penetrate the biofilm for mitigation compared to traditional interventions. This study aimed to characterize a new Escherichia phage vB_EcoS-UDF157lw (or UDF157lw) and determine its antimicrobial efficacy against E. coli O157:H7. Methods Phage characterization included biological approaches, including phage morphology, one-step growth curve, stability tests (pH and temperature), and genomic approaches (whole-genome sequencing). Later, antimicrobial activity tests, including productive infection against susceptible bacterial strains, in vitro antimicrobial activity, and anti-biofilm, were conducted. Results UDF157lw is a new member of the phages belonging to the Rogunavirus genus, comprising a long and non-contractile tail, isolated from bovine feces and shares close genomic evolutionary similarities with Escherichia phages vB_EcoS-BECP10 and bV_EcoS_AKS96. When used against E. coli O157:H7 (ATCC35150), phage UDF157lw exhibited a latent period of 14 min and a burst size of 110 PFU per infected cell. The phage remained viable in a wide range of pH values (pH 4–11) and temperatures (4–60°C). No virulence genes, such as stx, lysogenic genes, and antibiotic resistance genes, were found. Phage UDF157lw demonstrated high infection efficiencies against different E. coli O157:H7 and generic E. coli strains. In addition, UDF157lw encoded a unique major tail protein (ORF_26) with prominent depolymerase enzyme activity against various E. coli O157:H7 strains, causing large plaque sizes. In contrast to the phage without encoding depolymerase gene, UDF157lw was able to reduce the 24-h and 48-h E. coli O157:H7 biofilm after 1-h phage treatment. Discussion The findings of this study provide insights into a new member of the Rogunavirus phages and demonstrate its antimicrobial potential against E. coli O157:H7 in vitro.
... The increasing interest in lytic phages as biocontrol agents follows the growing concerns with antibiotic-resistant bacteria [69,70] and an increased number of studies that are demonstrating the efficacy of phage therapy. Some of these are reflected in the recent increase in the number of phage clinical studies [71,72], along with the increased sequence-based characterization of phages through comparative genomics and proteomics [73][74][75]. Other than the biological traits of the phage, such as lysogenic or lytic capacity, various environmental factors may affect phage efficacy and viability. ...
Article
Full-text available
Foodborne illness is exacerbated by novel and emerging pathotypes, persistent contamination, antimicrobial resistance, an ever-changing environment, and the complexity of food production systems. Sporadic and outbreak events of common foodborne pathogens like Shiga toxigenic E. coli (STEC), Salmonella, Campylobacter, and Listeria monocytogenes are increasingly identified. Methods of controlling human infections linked with food products are essential to improve food safety and public health and to avoid economic losses associated with contaminated food product recalls and litigations. Bacteriophages (phages) are an attractive additional weapon in the ongoing search for preventative measures to improve food safety and public health. However, like all other antimicrobial interventions that are being employed in food production systems, phages are not a panacea to all food safety challenges. Therefore, while phage-based biocontrol can be promising in combating foodborne pathogens, their antibacterial spectrum is generally narrower than most antibiotics. The emergence of phage-insensitive single-cell variants and the formulation of effective cocktails are some of the challenges faced by phage-based biocontrol methods. This review examines phage-based applications at critical control points in food production systems with an emphasis on when and where they can be successfully applied at production and processing levels. Shortcomings associated with phage-based control measures are outlined together with strategies that can be applied to improve phage utility for current and future applications in food safety.
... Bacteriophages, specifically the lytic type, are viral entities that kill bacteria and have garnered renewed interest as biocontrol agents in the fight against AMR bacteria for protecting human health (Miedzybrodzki et al., 2012;Vandenheuvel et al., 2015;Abedon et al., 2017, Gordillo andBarr, 2019;Lawrence et al., 2019, Mertz, 2019Loung et al., 2020), terrestrial and aquatic animal health management (Nakai and Park, 2002;Sulakvelidze and Barrow, 2005;Jhonson et al., 2008;Rao and Lalitha, 2015;Fernandez et al., 2017 Gigante and Aatterbury, 2019;Dec et al., 2020, Ninawe et al., 2020Zbikowska et al., 2020) and plant protection (Balogh et al., 2010;Buttimer et al., 2017;Svircev et al., 2018). Application of bacteriophages in food production, food processing and to ensure food safety is being increasingly pursued (Moye et al., 2018;Endersen and Coffey, 2020;Waturangi et al., 2021) Phages are numerous, diverse, relatively easy to isolate (Hendrix et al., 1999;Rohwer, 2003;Kim et al., 2011;Hoyles et al., 2014;Waller et al., 2014;Shkoporov et al., 2018) and are not affected by antimicrobial resistance of the bacteria, show species specific host range, avoid collateral damage to beneficial microbiota and are non-toxic with safe gut-transit (Brussow, 2005;Curtright and Abedon, 2011;Loc-Carrillo et al, 2011;Rea et al., 2011;Tomat et al., 2013;Niu et al., 2014;Khan Mirazaei and Nilsson, 2015;Sarker et al., 2016). Coliphages are bacteriophages that specifically target E. coli. ...
Article
Aim: This study elucidates the in vitro bactericidal effectiveness of poly-phage cocktail combinations of 2, 4, 6, 8 and 10 individual coliphages against a cocktail of 20 AMR E. coli. Methods & results: Different polyphage cocktails viz., 45 two-phage combinations, 28 four-phage combinations, 15 six-phage combinations, 6 eight-phage combinations and one ten-phage combination were formulated using a pool of ten coliphages that were isolated from two different geographical locations (East and West coasts of India). The different poly phage cocktails were tested at four different levels of Multiplicity of Infection (MOI) viz., MOI-1, MOI-10, MOI-100 and MOI-1000. All the 2, 4, 6, 8 and 10 phage cocktails were found to be effective in controlling the growth of a cocktail of 20 AMR bacteria when tested at MOI-1000 and MOI-100 but variations in antibacterial activity were observed at lower MOIs of 10 and 1. The ten coliphage cocktail showed lytic activity against 100% of AMR E. coli from farmed brackish water shrimp, 96% of laboratory collection of AMR E. coli, 92% of AMR E. coli from farmed freshwater fish and 85% of AMR E. coli from market shrimp. Conclusion: Polyphage cocktails of 2, 4, 6, 8 and 10 coliphages applied at an MOI of 1000 effectively suppressed the growth of antimicrobial resistant E. coli. The results indicated phage-phage synergy in the lytic activity of several coliphage combinations at higher MOIs of 1000 and 100 while phage-phage antagonism was evidenced at lower MOIs of 10 and 1.
... 19 Sequence-based characterization of phages through comparative genomics and proteomics offers considerable insight into phage classification and diversity. [21][22][23] The number of phage genome sequences available in public databases has escalated to *14,244 as of January 2021 24 since phage uX174 was first sequenced in 1977. 25 Approximately 7.5% of these sequenced phages infect E. coli, 26 but only a fraction of these have been specifically assessed for their ability to control non-O157 STEC. ...
Article
Background: Non-O157 Shiga toxigenic Escherichia coli (STEC) are one of the most important food and waterborne pathogens worldwide. Although bacteriophages (phages) have been used for the biocontrol of these pathogens, a comprehensive understanding of the genetic characteristics and lifestyle of potentially effective candidate phages is lacking. Materials and methods: In this study, 10 non-O157-infecting phages previously isolated from feedlot cattle and dairy farms in the North-West province of South Africa were sequenced, and their genomes were analyzed. Results: Comparative genomics and proteomics revealed that the phages were closely related to other E. coli-infecting Tunaviruses, Seuratviruses, Carltongylesviruses, Tequatroviruses, and Mosigviruses from the National Center for Biotechnology Information GenBank database. Phages lacked integrases associated with a lysogenic cycle and genes associated with antibiotic resistance and Shiga toxins. Conclusions: Comparative genomic analysis identified a diversity of unique non-O157-infecting phages, which could be used to mitigate the abundance of various non-O157 STEC serogroups without safety concerns.
Article
Coronavirus disease or COVID’19, caused by SARS-CoV-2, was the most recent pandemic in world history. The total cases worldwide exceeded 547M with global death toll exceeding 6.34M. The unprecedented infection rates were coupled with panic and misinformation that altered public perceptions regarding treatment options. The present study was aimed at assessing the factors that modulated public reliance on various treatments, ranging from non-pharmaceutical, herbal options to modern medicines and vaccines. The methodology employed was of qqualitative analysis using survey tool and SPSS, focusing on perceived efficacy, safety, availability, cost-effectivity and immune modulation of treatments. The results exhibit a trend within responders in favor of vaccines and herbal medicines. Moreover, a general hesitation has been expressed in regards to pharmaceutical treatment. The results indicate the highest positive perceptions in people regarding efficacy of herbal treatments (56.3 % in agreement), safety (59.4% in agreement), availability (63.2% in agreement), cost effectiveness (47.2% in agreement) and immunity-boosting activity (63.9% in agreement). Such an indication can be attributed to ancient Ayurvedic practices that serve as a recourse, during emergency situations. Moreover, the least positive perceptions were indicated for pharmaceutical drugs regarding efficacy, safety, availability, cost-effectivity and immune modulation activity. The trends observed in this resource can be utilized to design policy, with enhanced understanding of people’s beliefs and norms. Moreover, the communications and awareness endeavors during a global pandemic can be catered to existing biases and uncertainties in Pakistani population.
Article
Full-text available
Background Escherichia coli is a common fecal coliform, facultative aerobic, gram-negative bacterium. Pathogenic strains of such microbes have evolved to cause diarrhea, urinary tract infections, and septicemias. The emergence of antibiotic resistance urged the identification of an alternative strategy. The use of lytic bacteriophages against the control of pathogenic E. coli in clinics and different environmental setups (waste and drink water management) has become an alternative therapy to antibiotic therapy. Thus, this study aimed to isolate and characterize lytic bacteriophage from various sources in Addis Ababa, tested them against antimicrobial-resistant diarrheagenic E. coli strains and evaluated their therapeutic potential under in vitro conditions. Methods A total of 14 samples were processed against six different diarrheagenic E. coli strains. The conventional culture and plaque analysis agar overlay method was used to recover lytic bacteriophage isolates. The phage isolates were characterized to determine their lytic effect, growth characteristics, host range activity, and stability under different temperature and pH conditions. Phage isolates were identified by scanning electron microscope (SEM), and molecular techniques (PCR). Results In total, 17 phages were recovered from 84 tested plates. Of the 17 phage isolates, 11 (65%) were Myoviridae-like phages, and 6 (35%) phage isolates were Podoviridae and Siphoviridae by morphology and PCR identification. Based on the host range test, growth characteristics, and stability test 7 potent phages were selected. These phages demonstrated better growth characteristics, including short latent periods, highest burst sizes, and wider host ranges, as well as thermal stability and the ability to survive in a wide range of pH levels. Conclusions The promising effect of the phages isolated in this study against AMR pathogenic E. coli has raised the possibility of their use in the future treatment of E. coli infections.
Article
Objective: Bacterial wound infections have recently become a threat to public health. The emergence of multidrug-resistant (MDR) strains of Klebsiella pneumoniae highlights the need for a new treatment method. The effectiveness of bacteriophages has been observed for several infections in animal models and human trials. In this study, we assessed the effectiveness of bacteriophages in the treatment of wound infections associated with MDR and biofilm-producing K. pneumoniae and compared its effectiveness with that of gentamicin. Methods: A lytic phage against MDR Klebsiella pneumoniae was isolated and identified. The effectiveness of phages in the treatment of wound infection in mice was investigated and its effectiveness was compared with gentamicin. Results: The results showed that the isolated phage belonged to the Drexlerviridae family. This phage acts like gentamicin and effectively eliminates bacteria from wounds. In addition, mice in the phage therapy group were in better physical condition. Conclusion: Our results demonstrated the success of phage therapy in the treatment of mice wounds infected with K. pneumoniae. These results indicate the feasibility of topical phage therapy for the safe treatment of wound infections. Keywords Klebsiella pneumoniaebacteriophagesphage therapywound infectionmultidrug resistancein vivo
Article
In the natural environment, ruminant livestock, including cattle, are the main reservoir of the outbreak–causing strains of Shiga toxin‐producing Escherichia coli (STEC), where bacteriophages sustainably thrive as well. This study focuses on the characterization of STEC‐specific bacteriophages isolated from cow manure samples in Maine farms and examines their prevalence with STEC hosts. Phenotypic features of representative isolates were characterized by using a transmission electron microscope. Similarly, host range, one‐step growth curve, thermal stability, lytic capability, and genomic analyses were performed to fully characterize selected representative isolates. Results showed that representative bacteriophage isolates belong to Myoviridae (S6P10 and S14P12) and Siphoviridae (S19). The most prevalent and common bacteriophages (46%) were specific to the O26 serogroup. The farm C sampling site had highly heterogenous bacteriophage populations that were specific to six STEC serogroups. The most prevalent bacteriophage isolate (S1P5, Escherichia phage vB_EcoM‐S1P5QW) was verified to have a double‐stranded DNA genome (166,102 bp) with 266 CDs of which 130 have known functions. The majority of the diverse bacteriophage isolates had strong lytic capabilities and a narrow host range that could withstand selected temperature conditions (−20, 37, and 62°C). Results of bacterial screening showed that STEC host strains were not detected in Farms A, C, and E, but were detected on Farms B and D. In conclusion, the highly‐diverse bacteriophage ecology found in cow manure samples may have been an important element in shaping the population of STEC serogroups, specifically in its natural environment, which can provide useful tools for potential antibiotic‐free therapeutics and diagnostic technologies.
Article
Colorectal cancer (CRC) is one of the commonest cancers globally. A unique aspect of CRC is its intimate association with the gut microbiota, which forms an essential part of the tumour microenvironment. Research over the past decade has established that dysbiosis of gut bacteria, fungi, viruses and Archaea accompanies colorectal tumorigenesis, and these changes might be causative. Data from mechanistic studies demonstrate the ability of the gut microbiota to interact with the colonic epithelia and immune cells of the host via the release of a diverse range of metabolites, proteins and macromolecules that regulate CRC development. Preclinical and some clinical evidence also underscores the role of the gut microbiota in modifying the therapeutic responses of patients with CRC to chemotherapy and immunotherapy. Herein, we summarize our current understanding of the role of gut microbiota in CRC and outline the potential translational and clinical implications for CRC diagnosis, prevention and treatment. Emphasis is placed on how the gut microbiota could now be better harnessed by developing targeted microbial therapeutics as chemopreventive agents against colorectal tumorigenesis, as adjuvants for chemotherapy and immunotherapy to boost drug efficacy and safety, and as non-invasive biomarkers for CRC screening and patient stratification. Finally, we highlight the hurdles and potential solutions to translating our knowledge of the gut microbiota into clinical practice.
Article
Full-text available
Phages are the most abundant and diverse biological entities on Earth and exert specific effects on bacterial hosts. The coexistence of phages and bacteria in the intestinal tract is dynamic and interdependent. Phages are involved in maintaining the stability and composition of the bacterial community, and an imbalance in phages and bacteria in the intestinal tract can cause diseases. This review elucidates interactions between phages and bacteria in the human intestinal tract and their roles in the pathogenesis and treatment of diseases. Understanding the relationship among phages, bacteria and host diseases is conducive to promoting the application of phages in the treatment of human diseases. List of abbreviations EMBL-EBI The European Bioinformatics Institute; E. coli Escherichia coli; E. faecalis Enterobacter faecalis; B. fragilis Bacteroides fragilis; B. vulgatus Bacteroides vulgatus; SaPIs Staphylococcus aureus pathogenicity islands; ARGs Antibiotic resistance genes; STEC Shiga toxigenic E. coli; Stx Shiga toxin; BLAST Basic Local Alignment Search Tool; TSST-1 Toxic shock toxin 1; RBPs Receptor-binding proteins; LPS lipopolysaccharide; OMVs Outer membrane vesicles; PT Phosphorothioate; BREX Bacteriophage exclusion; OCR Overcome classical restriction; Pgl Phage growth limitation; DISARM Defense island system associated with restrictionmodification; R-M system Restriction-modification system; BREX system Bacteriophage exclusion system; CRISPR Clustered regularly interspaced short palindromic repeats; Cas CRISPR-associated; PAMs Prospacer adjacent motifs; crRNA CRISPR RNA; SIE; OMPs; Superinfection exclusion; Outer membrane proteins; Abi Abortive infection; TA Toxin-antitoxin; TLR Toll-like receptor; APCs Antigen-presenting cells; DSS Dextran sulfate sodium; IELs Intraepithelial lymphocytes; FMT Fecal microbiota transfer; IFN-γ Interferon-gamma; IBD Inflammatory bowel disease; AgNPs Silver nanoparticles; MDSC Myeloid-derived suppressor cell; CRC Colorectal cancer; VLPs Virus-like particles; TMP Tape measure protein; PSMB4 Proteasome subunit beta type-4; ALD Alcohol-related liver disease; GVHD Graft-versus-host disease; ROS Reactive oxygen species; RA Rheumatoid arthritis; CCP Cyclic citrullinated protein; AMGs Accessory metabolic genes; T1DM Type 1 diabetes mellitus; T2DM Type 2 diabetes mellitus; SCFAs Short-chain fatty acids; GLP-1 Glucagon-like peptide-1; A. baumannii Acinetobacter baumannii; CpG Deoxycytidylinate-phosphodeoxyguanosine; PEG Polyethylene glycol; MetS Metabolic syndrome; OprM Outer membrane porin M.
Chapter
The Dark Side of T1 Phage Although T1 got its initial fame from its use, in 1943, by Salvador Luria and Max Delbrück in their landmark fluctuation test (64), an entirely different reason has resulted in its notoriety since then. This phage’s resistance to desiccation results in its persistence for weeks in an aerosol which in turn created a real nightmare for laboratories engaged in research employing E. coli K-12 strains. It was not uncommon to hear cries from laboratory workers that their overnight bacterial cultures contained only floating debris. This was the result of an onslaught by T1. Laboratories were quarantined for days until the airborne count of T1 vanished. Bacterial geneticists have had T1-resistant E. coli tonA mutants since the 1940s, and as a result many laboratories protected against unwanted invasion of T1 by introducing null tonA alleles into their E. coli strain collection. Even today T1 maintains its bad reputation because it remains a problem, particularly in laboratories that only casually use E. coli strains for experiments involving cloning, protein overproduction, and analysis of genomic libraries. Because of this, almost all commercially available transformation-competent E. coli cells now contain a tonA mutation. Despite the knowledge of T1-mediated calamities, many laboratories continue to work with E. coli strains that are not tonA. Apparently, it often takes a large-scale disaster for laboratories to become believers in the incidental havoc T1 can wreak on innocent bacterial cultures. Despite its infamy, phage T1 is an interesting phage in its own right. In this chapter we consider the biology of the T1-like phages, concentrating on phage T1 and the related phage TLS.
Article
The recently-developed statistical method known as the "bootstrap" can be used to place confidence intervals on phylogenies. It involves resampling points from one's own data, with replacement, to create a series of bootstrap samples of the same size as the original data. Each of these is analyzed, and the variation among the resulting estimates taken to indicate the size of the error involved in making estimates from the original data. In the case of phylogenies, it is argued that the proper method of resampling is to keep all of the original species while sampling characters with replacement, under the assumption that the characters have been independently drawn by the systematist and have evolved independently. Majority-rule consensus trees can be used to construct a phylogeny showing all of the inferred monophyletic groups that occurred in a majority of the bootstrap samples. If a group shows up 95% of the time or more, the evidence for it is taken to be statistically significant. Existing computer programs can be used to analyze different bootstrap samples by using weights on the characters, the weight of a character being how many times it was drawn in bootstrap sampling. When all characters are perfectly compatible, as envisioned by Hennig, bootstrap sampling becomes unnecessary; the bootstrap method would show significant evidence for a group if it is defined by three or more characters.
Article
We describe a program, tRNAscan-SE, which identifies 99-100% of transfer RNA genes in DNA sequence while giving less than one false positive per 15 gigabases. Two previously described tRNA detection programs are used as fast, first-pass prefilters to identify candidate tRNAs, which are then analyzed by a highly selective tRNA covariance model. This work represents a practical application of RNA covariance models, which are general, probabilistic secondary structure profiles based on stochastic context-free grammars. tRNAscan-SE searches at approximately 30 000 bp/s. Additional extensions to tRNAscan-SE detect unusual tRNA homologues such as selenocysteine tRNAs, tRNA-derived repetitive elements and tRNA pseudogenes.
Article
A new method called the neighbor-joining method is proposed for reconstructing phylogenetic trees from evolutionary distance data. The principle of this method is to find pairs of operational taxonomic units (OTUs [= neighbors]) that minimize the total branch length at each stage of clustering of OTUs starting with a starlike tree. The branch lengths as well as the topology of a parsimonious tree can quickly be obtained by using this method. Using computer simulation, we studied the efficiency of this method in obtaining the correct unrooted tree in comparison with that of five other tree-making methods: the unweighted pair group method of analysis, Farris's method, Sattath and Tversky's method, Li's method, and Tateno et al.'s modified Farris method. The new, neighbor-joining method and Sattath and Tversky's method are shown to be generally better than the other methods.
Article
Shigellosis is a global health problem, and Shigella flexneri is the major cause of this disease. In this study, we isolated a virulent Siphoviridae bacteriophage (phage), pSf-1, that infects S. flexneri. This phage was isolated from the Han River in Korea and was found to infect S. flexneri, S. boydii, and S. sonnei. One-step growth analysis revealed that this phage has a short latent period (10 min) and a large burst size (86.86 PFU/cell), indicating that pSf-1 has good host infectivity and effective lytic activity. The double-stranded DNA genome of pSf-1 is composed of 51,821 bp with a G + C content of 44.02%. The genome encodes 94 putative ORFs, 71 putative promoters, and 60 transcriptional terminator regions. Genome sequence analysis of pSf-1 and comparative analysis with the homologous Shigella phage Shfl1 revealed that there is a high degree of similarity between pSf-1 and Shfl1 in 54 of the 94 ORFs of pSf-1. The results of this investigation indicate that pSf-1 is a novel Shigella phage and that this phage might have potential uses against shigellosis.