1887

Abstract

Infections with globally disseminated Shiga toxin-producing (STEC) of the O113:H21 serotype can progress to severe clinical complications, such as hemolytic uremic syndrome (HUS). Two phylogeographically distinct clonal complexes have been established by multi locus sequence typing (MLST). Infections with ST-820 isolates circulating exclusively in Australia have caused severe human disease, such as HUS. Conversely, ST-223 isolates prevalent in the US and outside Australia seem to rarely cause severe human disease but are frequent contaminants. Following a genomic epidemiology approach, we wanted to gain insights into the underlying cause for this disparity. We examined the plasticity in the genome make-up and Shiga toxin production in a collection of 20 ST-820 and ST-223 strains isolated from produce, the bovine reservoir, and clinical cases. STEC are notorious for assembly into fragmented draft sequences when using short-read sequencing technologies due to the extensive and partly homologous phage complement. The application of long-read technology (LRT) sequencing yielded closed reference chromosomes and plasmids for two representative ST-820 and ST-223 strains. The established high-resolution framework, based on whole genome alignments, single nucleotide polymorphism (SNP)-typing and MLST, includes the chromosomes and plasmids of other publicly available O113:H21 sequences and allowed us to refine the phylogeographical boundaries of ST-820 and ST-223 complex isolates and to further identify a historic non-shigatoxigenic strain from Mexico as a quasi-intermediate. Plasmid comparison revealed strong correlations between the strains’ featured pO113 plasmid genotypes and chromosomally inferred ST, which suggests coevolution of the chromosome and virulence plasmids. Our pathogenicity assessment revealed statistically significant differences in the Stx-production capabilities of ST-820 as compared to ST-223 strains under RecA-induced Stx phage mobilization, a condition that mimics Stx-phage induction. These observations suggest that ST-820 strains may confer an increased pathogenic potential in line with the strain-associated epidemiological metadata. Still, some of the tested ST-223 cultures sourced from contaminated produce or the bovine reservoir also produced Stx at levels comparable to those of ST-820 isolates, which calls for awareness and for continued surveillance of this lineage.

Funding
This study was supported by the:
  • U.S. Department of Homeland Security (Award 2014-ST-062-000058)
    • Principle Award Recipient: MarkEppinger
  • National Institute of General Medical Sciences (Award SC1GM135110)
    • Principle Award Recipient: MarkEppinger
  • This is an open-access article distributed under the terms of the Creative Commons Attribution License.
Loading

Article metrics loading...

/content/journal/mgen/10.1099/mgen.0.000796
2022-04-08
2024-04-25
Loading full text...

Full text loading...

/deliver/fulltext/mgen/8/4/mgen000796.html?itemId=/content/journal/mgen/10.1099/mgen.0.000796&mimeType=html&fmt=ahah

References

  1. Sperandio V, Hovde CJ. Enterohemorrhagic Escherichia coli and other Shiga toxin-producing E. coli Washington, DC: ASM Press; 2015
    [Google Scholar]
  2. Zuppi M, Tozzoli R, Chiani P, Quiros P, Martinez-Velazquez A et al. Investigation on the evolution of shiga toxin-converting phages based on whole genome sequencing. Front Microbiol 2020; 11:1472 [View Article]
    [Google Scholar]
  3. Kruger A, Lucchesi PM. Shiga toxins and stx phages: highly diverse entities. Microbiology 2015; 161:451–462 [View Article]
    [Google Scholar]
  4. Asadulghani M, Ogura Y, Ooka T, Itoh T, Sawaguchi A et al. The defective prophage pool of Escherichia coli O157: prophage-prophage interactions potentiate horizontal transfer of virulence determinants. PLoS Pathog 2009; 5:e1000408 [View Article]
    [Google Scholar]
  5. Huang A, Friesen J, Brunton JL. Characterization of a bacteriophage that carries the genes for production of Shiga-like toxin 1 in Escherichia coli. J Bacteriol 1987; 169:4308–4312 [View Article]
    [Google Scholar]
  6. Mauro SA, Koudelka GB. Shiga toxin: expression, distribution, and its role in the environment. Toxins (Basel) 2011; 3:608–625 [View Article]
    [Google Scholar]
  7. Melton-Celsa A, Mohawk K, Teel L, O’Brien A. Pathogenesis of shiga-toxin producing Escherichia coli. Curr Top Microbiol Immunol 2012; 357:67–103
    [Google Scholar]
  8. Nguyen Y, Sperandio V. Enterohemorrhagic E. coli (EHEC) pathogenesis. Front Cell Infect Microbiol 2012; 2:90 [View Article]
    [Google Scholar]
  9. Johannes L, Romer W. Shiga toxins-from cell biology to biomedical applications. Nat Rev Microbiol 2010; 8:105–116
    [Google Scholar]
  10. Tran SL, Billoud L, Lewis SB, Phillips AD, Schuller S. Shiga toxin production and translocation during microaerobic human colonic infection with shiga toxin-producing E. coli O157:H7 and O104:H4. Cell Microbiol 2014; 16:1255–1266
    [Google Scholar]
  11. Eaton KA, Friedman DI, Francis GJ, Tyler JS, Young VB et al. Pathogenesis of renal disease due to enterohemorrhagic Escherichia coli in germ-free mice. Infect Immun 2008; 76:3054–3063 [View Article]
    [Google Scholar]
  12. Shin I-S, Ishii S, Shin J-S, Sung K-I, Park B-S et al. Globotriaosylceramide (Gb3) content in HeLa cells is correlated to Shiga toxin-induced cytotoxicity and Gb3 synthase expression. BMB Rep 2009; 42:310–314 [View Article]
    [Google Scholar]
  13. Pacheco AR, Lazarus JE, Sit B, Schmieder S, Lencer WI et al. CRISPR Screen Reveals that EHEC’s T3SS and shiga toxin rely on shared host factors for infection. mBio 2018; 9:e01003-18 [View Article]
    [Google Scholar]
  14. Karmali MA, Petric M, Lim C, Fleming PC, Steele BT. Escherichia coli cytotoxin, haemolytic-uraemic syndrome, and haemorrhagic colitis. Lancet 1983; 2:1299–1300 [View Article]
    [Google Scholar]
  15. Donohue-Rolfe A, Kondova I, Oswald S, Hutto D, Tzipori S. Escherichia coli O157:H7 strains that express Shiga toxin (Stx) 2 alone are more neurotropic for gnotobiotic piglets than are isotypes producing only Stx1 or both Stx1 and Stx2. J Infect Dis 2000; 181:1825–1829 [View Article]
    [Google Scholar]
  16. Russo LM, Melton-Celsa AR, O’Brien AD. Shiga toxin (Stx) type 1a reduces the oral toxicity of Stx type 2a. J Infect Dis 2016; 213:1271–1279 [View Article]
    [Google Scholar]
  17. Amigo N, Mercado E, Bentancor A, Singh P, Vilte D et al. Clade 8 and Clade 6 strains of Escherichia coli O157:H7 from cattle in argentina have hypervirulent-like phenotypes. PLoS One 2015; 10:e0127710 [View Article]
    [Google Scholar]
  18. Iyoda S, Manning SD, Seto K, Kimata K, Isobe J et al. Phylogenetic clades 6 and 8 of enterohemorrhagic Escherichia coli O157:H7 with particular stx subtypes are more frequently found in isolates from hemolytic uremic syndrome patients than from asymptomatic carriers. Open Forum Infect Dis 2014; 1:fu061 [View Article]
    [Google Scholar]
  19. Manning SD, Motiwala AS, Springman AC, Qi W, Lacher DW et al. Variation in virulence among clades of Escherichia coli O157:H7 associated with disease outbreaks. Proc Natl Acad Sci U S A 2008; 105:4868–4873 [View Article]
    [Google Scholar]
  20. Vanaja SK, Springman AC, Besser TE, Whittam TS, Manning SD. Differential expression of virulence and stress fitness genes between Escherichia coli O157:H7 strains with clinical or bovine-biased genotypes. Appl Environ Microbiol 2010; 76:60–68 [View Article]
    [Google Scholar]
  21. Mellor GE, Sim EM, Barlow RS, D’Astek BA, Galli L et al. Phylogenetically related Argentinean and Australian Escherichia coli O157 isolates are distinguished by virulence clades and alternative Shiga toxin 1 and 2 prophages. Appl Environ Microbiol 2012; 78:4724–4731 [View Article]
    [Google Scholar]
  22. Strachan NJ, Rotariu O, Lopes B, MacRae M, Fairley S et al. Whole genome sequencing demonstrates that geographic variation of Escherichia coli O157 genotypes dominates host association. Sci Rep 2015; 5:14145 [View Article]
    [Google Scholar]
  23. Pearce MC, Chase-Topping ME, McKendrick IJ, Mellor DJ, Locking ME et al. Temporal and spatial patterns of bovine Escherichia coli O157 prevalence and comparison of temporal changes in the patterns of phage types associated with bovine shedding and human E. coli O157 cases in Scotland between 1998-2000 and 2002-2004. BMC Microbiol 2009; 9:276
    [Google Scholar]
  24. Ogura Y, Mondal SI, Islam MR, Mako T, Arisawa K et al. The Shiga toxin 2 production level in enterohemorrhagic Escherichia coli O157:H7 is correlated with the subtypes of toxin-encoding phage. Sci Rep 2015; 5: [View Article]
    [Google Scholar]
  25. Neupane M, Abu-Ali GS, Mitra A, Lacher DW, Manning SD et al. Shiga toxin 2 overexpression in Escherichia coli O157:H7 strains associated with severe human disease. Microb Pathog 2011; 51:466–470 [View Article]
    [Google Scholar]
  26. Wagner PL, Acheson DW, Waldor MK. Isogenic lysogens of diverse shiga toxin 2-encoding bacteriophages produce markedly different amounts of shiga toxin. Infect Immun 1999; 67:6710–6714 [View Article]
    [Google Scholar]
  27. de Sablet T, Bertin Y, Vareille M, Girardeau J-P, Garrivier A et al. Differential expression of stx2 variants in Shiga toxin-producing Escherichia coli belonging to seropathotypes A and C. Microbiology (Reading) 2008; 154:176–186 [View Article]
    [Google Scholar]
  28. Kulasekara BR, Jacobs M, Zhou Y, Wu Z, Sims E et al. Analysis of the genome of the Escherichia coli O157:H7 2006 spinach-associated outbreak isolate indicates candidate genes that may enhance virulence. Infect Immun 2009; 77:3713–3721 [View Article]
    [Google Scholar]
  29. Baker DR, Moxley RA, Steele MB, Lejeune JT, Christopher-Hennings J et al. Differences in virulence among Escherichia coli O157:H7 strains isolated from humans during disease outbreaks and from healthy cattle. Appl Environ Microbiol 2007; 73:7338–7346 [View Article]
    [Google Scholar]
  30. Zhang Y, Laing C, Zhang Z, Hallewell J, You C et al. Lineage and host source are both correlated with levels of Shiga toxin 2 production by Escherichia coli O157:H7 strains. Appl Environ Microbiol 2010; 76:474–482
    [Google Scholar]
  31. Koitabashi T, Vuddhakul V, Radu S, Morigaki T, Asai N et al. Genetic characterization of Escherichia coli O157: H7/- strains carrying the stx2 gene but not producing Shiga toxin 2. Microbiol Immunol 2006; 50:135–148
    [Google Scholar]
  32. Dowd SE, Williams JB. Comparison of Shiga-like toxin II expression between two genetically diverse lineages of Escherichia coli O157:H7. J Food Prot 2008; 71:1673–1678 [View Article]
    [Google Scholar]
  33. Baker DR, Moxley RA, Francis DH. Variation in virulence in the gnotobiotic pig model of O157:H7 Escherichia coli strains of bovine and human origin. Adv Exp Med Biol 1997; 412:53–58
    [Google Scholar]
  34. Ziebell K, Steele M, Zhang Y, Benson A, Taboada EN et al. Genotypic characterization and prevalence of virulence factors among Canadian Escherichia coli O157:H7 strains. Appl Environ Microbiol 2008; 74:4314–4323
    [Google Scholar]
  35. Abu-Ali GS, Ouellette LM, Henderson ST, Whittam TS, Manning SD. Differences in adherence and virulence gene expression between two outbreak strains of enterohaemorrhagic Escherichia coli O157: H7. Microbiology 2010; 156:408–419
    [Google Scholar]
  36. Zhou Z, Li X, Liu B, Beutin L, Xu J et al. Derivation of Escherichia coli O157:H7 from its O55:H7 precursor. PLoS One 2010; 5:e8700 [View Article]
    [Google Scholar]
  37. Rasko DA, Webster DR, Sahl JW, Bashir A, Boisen N et al. Origins of the E. coli strain causing an outbreak of hemolytic-uremic syndrome in Germany. N Engl J Med 2011; 365:709–717
    [Google Scholar]
  38. Yang Z, Kovar J, Kim J, Nietfeldt J, Smith DR et al. Identification of common subpopulations of non-sorbitol-fermenting, beta-glucuronidase-negative Escherichia coli O157:H7 from bovine production environments and human clinical samples. Appl Environ Microbiol 2004; 70:6846–6854
    [Google Scholar]
  39. Zhang X, McDaniel AD, Wolf LE, Keusch GT, Waldor MK et al. Quinolone antibiotics induce Shiga toxin-encoding bacteriophages, toxin production, and death in mice. J Infect Dis 2000; 181:664–670 [View Article]
    [Google Scholar]
  40. Yin S, Rusconi B, Sanjar F, Goswami K, Xiaoli L et al. Escherichia coli O157:H7 strains harbor at least three distinct sequence types of Shiga toxin 2a-converting phages. BMC Genomics 2015; 16:733 [View Article]
    [Google Scholar]
  41. Besser TE, Shaikh N, Holt NJ, Tarr PI, Konkel ME et al. Greater diversity of Shiga toxin-encoding bacteriophage insertion sites among Escherichia coli O157:H7 isolates from cattle than in those from humans. Appl Environ Microbiol 2007; 73:671–679 [View Article]
    [Google Scholar]
  42. Elhadidy M, Elkhatib WF, Piérard D, De Reu K, Heyndrickx M. Model-based clustering of Escherichia coli O157:H7 genotypes and their potential association with clinical outcome in human infections. Diagn Microbiol Infect Dis 2015; 83:198–202 [View Article]
    [Google Scholar]
  43. Whitworth J, Zhang Y, Bono J, Pleydell E, French N et al. Diverse genetic markers concordantly identify bovine origin Escherichia coli O157 genotypes underrepresented in human disease. Appl Environ Microbiol 2010; 76:361–365 [View Article]
    [Google Scholar]
  44. Whitworth JH, Fegan N, Keller J, Gobius KS, Bono JL et al. International comparison of clinical, bovine, and environmental Escherichia coli O157 isolates on the basis of Shiga toxin-encoding bacteriophage insertion site genotypes. Appl Environ Microbiol 2008; 74:7447–7450 [View Article]
    [Google Scholar]
  45. Shringi S, Schmidt C, Katherine K, Brayton KA, Hancock DD et al. Carriage of stx2a differentiates clinical and bovine-biased strains of Escherichia coli O157. PLoS One 2012; 7:e51572 [View Article]
    [Google Scholar]
  46. Shringi S, García A, Lahmers KK, Potter KA, Muthupalani S et al. Differential virulence of clinical and bovine-biased enterohemorrhagic Escherichia coli O157:H7 genotypes in piglet and Dutch belted rabbit models. Infect Immun 2012; 80:369–380 [View Article]
    [Google Scholar]
  47. Ahmad A, Zurek L. Evaluation of the anti-terminator Q933 gene as a marker for Escherichia coli O157:H7 with high Shiga toxin production. Curr Microbiol 2006; 53:324–328 [View Article]
    [Google Scholar]
  48. Elhadidy M, Alvarez-Ordonez A. Diversity of survival patterns among Escherichia coli O157:H7 genotypes subjected to food-related stress conditions. Front Microbiol 2016; 7:322 [View Article]
    [Google Scholar]
  49. Byrne L, Dallman TJ, Adams N, Mikhail AFW, McCarthy N et al. Highly pathogenic clone of shiga toxin-producing Escherichia coli O157:H7, England and Wales. Emerg Infect Dis 2018; 24:2303–2308
    [Google Scholar]
  50. Pruimboom-Brees IM, Morgan TW, Ackermann MR, Nystrom ED, Samuel JE et al. Cattle lack vascular receptors for Escherichia coli O157:H7 Shiga toxins. Proc Natl Acad Sci U S A 2000; 97:10325–10329 [View Article]
    [Google Scholar]
  51. Flockhart AF, Tree JJ, Xu X, Karpiyevich M, McAteer SP et al. Identification of a novel prophage regulator in Escherichia coli controlling the expression of type III secretion. Mol Microbiol 2012; 83:208–223 [View Article]
    [Google Scholar]
  52. Goswami K, Chen C, Xiaoli L, Eaton KA, Dudley EG. Coculture of Escherichia coli O157:H7 with a Nonpathogenic E. coli strain increases toxin production and virulence in a germfree mouse model. Infect Immun 2015; 83:4185–4193 [View Article]
    [Google Scholar]
  53. Pacheco AR, Curtis MM, Ritchie JM, Munera D, Waldor MK et al. Fucose sensing regulates bacterial intestinal colonization. Nature 2012; 492:113–117 [View Article]
    [Google Scholar]
  54. Pacheco AR, Sperandio V. Shiga toxin in enterohemorrhagic E.coli: regulation and novel anti-virulence strategies. Front Cell Infect Microbiol 2012; 2:81 [View Article]
    [Google Scholar]
  55. Silva CJ, Lee BG, Yambao JC, Erickson-Beltran ML, Quiñones B. Using nanospray liquid chromatography and mass spectrometry to quantitate shiga toxin production in environmental Escherichia coli recovered from a major produce production region in California. J Agric Food Chem 2019; 67:1554–1562 [View Article]
    [Google Scholar]
  56. González-Escalona N, Kase JA. Virulence gene profiles and phylogeny of Shiga toxin-positive Escherichia coli strains isolated from FDA regulated foods during 2010-2017. PLoS One 2019; 14:e0214620 [View Article]
    [Google Scholar]
  57. Paton AW, Woodrow MC, Doyle RM, Lanser JA, Paton JC. Molecular characterization of a Shiga toxigenic Escherichia coli O113:H21 strain lacking eae responsible for a cluster of cases of hemolytic-uremic syndrome. J Clin Microbiol 1999; 37:3357–3361
    [Google Scholar]
  58. Bondì R, Chiani P, Michelacci V, Minelli F, Caprioli A et al. The gene tia, harbored by the subtilase-encoding pathogenicity Island, is involved in the ability of locus of enterocyte effacement-negative shiga toxin-producing Escherichia coli strains to invade monolayers of epithelial cells. Infect Immun 2017; 85:e00613-17 [View Article]
    [Google Scholar]
  59. Feng P, Delannoy S, Lacher DW, Bosilevac JM, Fach P. Characterization and virulence potential of serogroup O113 shiga toxin-producing Escherichia coli strains isolated from beef and cattle in the United States. J Food Prot 2017; 80:383–391
    [Google Scholar]
  60. Feng PC, Delannoy S, Lacher DW, Dos Santos LF, Beutin L et al. Genetic diversity and virulence potential of shiga toxin-producing Escherichia coli O113:H21 strains isolated from clinical, environmental, and food sources. Appl Environ Microbiol 2014; 80:4757–4763
    [Google Scholar]
  61. Paton AW, Paton JC. Molecular characterization of the locus encoding biosynthesis of the lipopolysaccharide O antigen of Escherichia coli serotype O113. Infect Immun 1999; 67:5930–5937 [View Article]
    [Google Scholar]
  62. Goldwater PN, Giles N, Bettelheim KA. An unusual case of microangiopathic haemolytic anaemia associated with enterohaemorrhagic Escherichia coli O113:H21 infection, a verocytotoxin-2/shiga toxin-2 producing serotype. J Infect 1998; 37:302–304 [View Article]
    [Google Scholar]
  63. Newton HJ, Sloan J, Bulach DM, Seemann T, Allison CC et al. Shiga toxin-producing Escherichia coli strains negative for locus of enterocyte effacement. Emerg Infect Dis 2009; 15:372–380
    [Google Scholar]
  64. Quiñones B, Yambao JC, Lee BG. Draft genome sequences of Escherichia coli O113:H21 strains recovered from a major produce production region in California. Genome Announc 2017; 5:e01203-17 [View Article]
    [Google Scholar]
  65. G M Gonzalez A, M F Cerqueira A. Shiga toxin-producing Escherichia coli in the animal reservoir and food in Brazil. J Appl Microbiol 2020; 128:1568–1582 [View Article]
    [Google Scholar]
  66. Dos Santos LF, Biscola FT, Goncalves EM, Guth BE. Biofilm formation, invasiveness and colicinogeny in locus of enterocyte and effacement negative O113:H21 Shigatoxigenic Escherichia coli. J Appl Microbiol 2017; 122:1101–1109
    [Google Scholar]
  67. dos Santos LF, Irino K, Vaz TM, Guth BE. Set of virulence genes and genetic relatedness of O113: H21 Escherichia coli strains isolated from the animal reservoir and human infections in Brazil. J Med Microbiol 2010; 59:634–640
    [Google Scholar]
  68. Bosilevac JM, Koohmaraie M. Prevalence and characterization of non-O157 shiga toxin-producing Escherichia coli isolates from commercial ground beef in the United States. Appl Environ Microbiol 2011; 77:2103–2112 [View Article]
    [Google Scholar]
  69. Irino K, Kato MAMF, Vaz TMI, Ramos II, Souza MAC et al. Serotypes and virulence markers of Shiga toxin-producing Escherichia coli (STEC) isolated from dairy cattle in São Paulo State, Brazil. Vet Microbiol 2005; 105:29–36 [View Article]
    [Google Scholar]
  70. Galli L, Miliwebsky E, Irino K, Leotta G, Rivas M. Virulence profile comparison between LEE-negative Shiga toxin-producing Escherichia coli (STEC) strains isolated from cattle and humans. Vet Microbiol 2010; 143:307–313 [View Article]
    [Google Scholar]
  71. Sadiq M, Hazen T, Rasko D, Eppinger M. Enterohemorrhagic Escherichia coli genomics: past, present, and future. In Sperandio V, Hovde CJ. eds Enterohemorrhagic Escherichia Coli and Other Shiga Toxin Producing E Coli Washington, DC: ASM Press; 2015 pp 55–71
    [Google Scholar]
  72. Franz E, Delaquis P, Morabito S, Beutin L, Gobius K et al. Exploiting the explosion of information associated with whole genome sequencing to tackle Shiga toxin-producing Escherichia coli (STEC) in global food production systems. Int J Food Microbiol 2014; 187:57–72 [View Article]
    [Google Scholar]
  73. Bosilevac JM, Guerini MN, Brichta-Harhay DM, Arthur TM, Koohmaraie M. Microbiological characterization of imported and domestic boneless beef trim used for ground beef. J Food Prot 2007; 70:440–449
    [Google Scholar]
  74. Parker CT, Cooper KK, Huynh S, Smith TP, Bono JL et al. Genome sequences of eight shiga toxin-producing Escherichia coli strains isolated from a produce-growing region in California. Microbiol Resour Announc 2018; 7:e00807-18 [View Article]
    [Google Scholar]
  75. Trees E, Strockbine N, Changayil S, Ranganathan S, Zhao K et al. Genome sequences of 228 shiga toxin-producing Escherichia coli Isolates and 12 isolates representing other diarrheagenic E. coli pathotypes. Genome Announc 2014; 2:e00718-14 [View Article]
    [Google Scholar]
  76. Patel PN, Lindsey RL, Garcia-Toledo L, Rowe LA, Batra D et al. High-quality whole-genome sequences for 77 shiga toxin-producing Escherichia coli strains generated with PacBio sequencing. Genome Announc 2018; 6:e00391-18 [View Article]
    [Google Scholar]
  77. Orskov I, Orskov F, Jann B. Serology, chemistry, and genetics of O and K antigens of Escherichia coli. Bacteriol Rev 1977; 41:667–710 [View Article]
    [Google Scholar]
  78. Gurevich A, Saveliev V, Vyahhi N, Tesler G. QUAST: quality assessment tool for genome assemblies. Bioinformatics 2013; 29:1072–1075 [View Article]
    [Google Scholar]
  79. Mikheenko A, Prjibelski A, Saveliev V, Antipov D, Gurevich A. Versatile genome assembly evaluation with QUAST-LG. Bioinformatics 2018; 34:i142–i150 [View Article]
    [Google Scholar]
  80. Gao F, Zhang CT. Ori-Finder: a web-based system for finding oriCs in unannotated bacterial genomes. BMC Bioinformatics 2008; 9:79
    [Google Scholar]
  81. Tatusova T, DiCuccio M, Badretdin A, Chetvernin V, Nawrocki EP et al. NCBI prokaryotic genome annotation pipeline. Nucleic Acids Res 2016; 44:6614–6624 [View Article]
    [Google Scholar]
  82. Foley SL, Lynne AM, Nayak R. Molecular typing methodologies for microbial source tracking and epidemiological investigations of Gram-negative bacterial foodborne pathogens. Infect Genet Evol 2009; 9:430–440
    [Google Scholar]
  83. Zhou Z, Alikhan NF, Mohamed K, Fan Y, Agama Study G et al. The Enterobase user’s guide, with case studies on Salmonella transmissions, Yersinia pestis phylogeny and Escherichia core genomic diversity. Genome Res 2019 [View Article]
    [Google Scholar]
  84. Wirth T, Falush D, Lan R, Colles F, Mensa P et al. Sex and virulence in Escherichia coli: an evolutionary perspective. Mol Microbiol 2006; 60:1136–1151 [View Article]
    [Google Scholar]
  85. Jaureguy F, Landraud L, Passet V, Diancourt L, Frapy E et al. Phylogenetic and genomic diversity of human bacteremic Escherichia coli strains. BMC Genomics 2008; 9:560 [View Article]
    [Google Scholar]
  86. Larsen MV, Cosentino S, Rasmussen S, Friis C, Hasman H et al. Multilocus sequence typing of total-genome-sequenced bacteria. J Clin Microbiol 2012; 50:1355–1361
    [Google Scholar]
  87. Reid SD, Herbelin CJ, Bumbaugh AC, Selander RK, Whittam TS. Parallel evolution of virulence in pathogenic Escherichia coli. Nature 2000; 406:64–67
    [Google Scholar]
  88. Weihong Qi DWL. EcMLST: an Online Database for Multi Locus Sequence Typing of Pathogenic Escherichia coli. Proceedings of the 2004 IEEE Computational Systems Bioinformatics Conference (CSB 2004) Stanford, CA: IEEE; 2004
    [Google Scholar]
  89. Junemann S, Sedlazeck FJ, Prior K, Albersmeier A, John U et al. Updating benchtop sequencing performance comparison. Nat Biotechnol 2013; 31:294–296
    [Google Scholar]
  90. Angiuoli SV, Salzberg SL. Mugsy: fast multiple alignment of closely related whole genomes. Bioinformatics 2011; 27:334–342 [View Article]
    [Google Scholar]
  91. Stamatakis A. RAxML version 8: a tool for phylogenetic analysis and post-analysis of large phylogenies. Bioinformatics 2014; 30:1312–1313 [View Article]
    [Google Scholar]
  92. Kearse M, Moir R, Wilson A, Stones-Havas S, Cheung M et al. Geneious Basic: an integrated and extendable desktop software platform for the organization and analysis of sequence data. Bioinformatics 2012; 28:1647–1649 [View Article]
    [Google Scholar]
  93. Zhang H, Gao S, Lercher MJ, Hu S. EvolView, an online tool for visualizing, annotating and managing phylogenetic trees. Nucleic Acids Res 2012; 40:W569–72 [View Article]
    [Google Scholar]
  94. Subramanian B, Gao S, Lercher MJ, Hu S, Chen WH. Evolview v3: a webserver for visualization, annotation, and management of phylogenetic trees. Nucleic Acids Res 2019
    [Google Scholar]
  95. He Z, Zhang H, Gao S, Lercher MJ, Chen WH et al. Evolview v2: an online visualization and management tool for customized and annotated phylogenetic trees. Nucleic Acids Res 2016; 44:W236–241
    [Google Scholar]
  96. Rusconi B, Sanjar F, Koenig SS, Mammel MK, Tarr PI et al. Whole genome sequencing for genomics-guided investigations of Escherichia coli O157:H7 outbreaks. Front Microbiol 2016; 7:985 [View Article]
    [Google Scholar]
  97. Goecks J, Nekrutenko A, Taylor J, Galaxy T. Galaxy: a comprehensive approach for supporting accessible, reproducible, and transparent computational research in the life sciences. Genome Biol 2010; 11:R86 [View Article]
    [Google Scholar]
  98. Delcher AL, Salzberg SL, Phillippy AM. Using mummer to identify similar regions in large sequence sets. Curr Protoc Bioinformatics 2003; 10:13
    [Google Scholar]
  99. Zhou Y, Liang Y, Lynch KH, Dennis JJ, Wishart DS. PHAST: a fast phage search tool. Nucleic Acids Res 2011; 39:W347–352
    [Google Scholar]
  100. Arndt D, Grant JR, Marcu A, Sajed T, Pon A et al. PHASTER: a better, faster version of the PHAST phage search tool. Nucleic Acids Res 2016; 44:W16–21 [View Article]
    [Google Scholar]
  101. Siguier P, Perochon J, Lestrade L, Mahillon J, Chandler M. ISfinder: the reference centre for bacterial insertion sequences. Nucleic Acids Res 2006; 34:D32–36 [View Article]
    [Google Scholar]
  102. Xie Z, Tang H. ISEScan: automated identification of insertion sequence elements in prokaryotic genomes. Bioinformatics 2017; 33:3340–3347
    [Google Scholar]
  103. Liu M, Li X, Xie Y, Bi D, Sun J et al. ICEberg 2.0: an updated database of bacterial integrative and conjugative elements. Nucleic Acids Res 2019; 47:D660–D665
    [Google Scholar]
  104. Li H, Durbin R. Fast and accurate short read alignment with Burrows-Wheeler transform. Bioinformatics 2009; 25:1754–1760
    [Google Scholar]
  105. Garrison E, Marth G. n.d Haplotype-based variant detection from short-read sequencing. ArXiv e-prints
    [Google Scholar]
  106. Marcais G, Delcher AL, Phillippy AM, Coston R, Salzberg SL et al. MUMmer4: A fast and versatile genome alignment system. PLoS Comput Biol 2018; 14:e1005944 [View Article]
    [Google Scholar]
  107. Eppinger M, Mammel MK, Leclerc JE, Ravel J, Cebula TA. Genomic anatomy of Escherichia coli O157:H7 outbreaks. Proc Natl Acad Sci U S A 2011; 108:20142–20147
    [Google Scholar]
  108. Eppinger M, Pearson T, Koenig SS, Pearson O, Hicks N et al. Genomic epidemiology of the haitian cholera outbreak: a single introduction followed by rapid, extensive, and continued spread characterized the onset of the epidemic. mBio 2014; 5:e01721 [View Article]
    [Google Scholar]
  109. Myers GS, Mathews SA, Eppinger M, Mitchell C, O’Brien KK et al. Evidence that human Chlamydia pneumoniae was zoonotically acquired. J Bacteriol 2009; 191:7225–7233 [View Article]
    [Google Scholar]
  110. Morelli G, Song Y, Mazzoni CJ, Eppinger M, Roumagnac P et al. Yersinia pestis genome sequencing identifies patterns of global phylogenetic diversity. Nat Genet 2010; 42:1140–1143 [View Article]
    [Google Scholar]
  111. Altschul SF, Gish W, Miller W, Myers EW, Lipman DJ. Basic local alignment search tool. J Mol Biol 1990; 215:403–410 [View Article]
    [Google Scholar]
  112. Johnson JE, Kumar P, Easterly C, Esler M, Mehta S et al. Improve your Galaxy text life: The Query Tabular Tool. F1000Res 2018; 7:1604
    [Google Scholar]
  113. Wilgenbusch JC, Swofford D. Inferring evolutionary trees with PAUP*. Curr Protoc Bioinformatics 2003; 6:4
    [Google Scholar]
  114. Maddison WP, Maddison DR. Mesquite: A Modular System for Evolutionary Analysis 2016
    [Google Scholar]
  115. Eppinger M, Worsham PL, Nikolich MP, Riley DR, Sebastian Y et al. Genome sequence of the deep-rooted Yersinia pestis strain Angola reveals new insights into the evolution and pangenome of the plague bacterium. J Bacteriol 2010; 192:1685–1699 [View Article]
    [Google Scholar]
  116. Hau SJ, Allue-Guardia A, Rusconi B, Haan JS, Davies PR et al. Single nucleotide polymorphism analysis indicates genetic distinction and reduced diversity of swine-associated methicillin resistant Staphylococcus aureus (MRSA) ST5 isolates compared to clinical MRSA ST5 isolates. Front Microbiol 2078; 2018:9
    [Google Scholar]
  117. Nyong EC, Zaia SR, Allué-Guardia A, Rodriguez AL, Irion-Byrd Z et al. Pathogenomes of atypical non-shigatoxigenic Escherichia coli NSF/SF O157:H7/NM: comprehensive phylogenomic analysis using closed genomes. Front Microbiol 2020; 11:619 [View Article]
    [Google Scholar]
  118. Joensen KG, Scheutz F, Lund O, Hasman H, Kaas RS et al. Real-time whole-genome sequencing for routine typing, surveillance, and outbreak detection of verotoxigenic Escherichia coli. J Clin Microbiol 2014; 52:1501–1510 [View Article]
    [Google Scholar]
  119. Joensen KG, Tetzschner AMM, Iguchi A, Aarestrup FM, Scheutz F. Rapid and easy in silico serotyping of Escherichia coli isolates by use of whole-genome sequencing data. J Clin Microbiol 2015; 53:2410–2426 [View Article]
    [Google Scholar]
  120. Chen L, Zheng D, Liu B, Yang J, Jin Q. VFDB 2016: hierarchical and refined dataset for big data analysis--10 years on. Nucleic Acids Res 2016; 44:D694–7 [View Article]
    [Google Scholar]
  121. Alcock BP, Raphenya AR, Lau TTY, Tsang KK, Bouchard M et al. CARD 2020: antibiotic resistome surveillance with the comprehensive antibiotic resistance database. Nucleic Acids Res 2020; 48:D517–D525 [View Article]
    [Google Scholar]
  122. Gupta SK, Padmanabhan BR, Diene SM, Lopez-Rojas R, Kempf M et al. ARG-ANNOT, a new bioinformatic tool to discover antibiotic resistance genes in bacterial genomes. Antimicrob Agents Chemother 2014; 58:212–220
    [Google Scholar]
  123. Zankari E, Hasman H, Cosentino S, Vestergaard M, Rasmussen S et al. Identification of acquired antimicrobial resistance genes. J Antimicrob Chemother 2012; 67:2640–2644 [View Article]
    [Google Scholar]
  124. Kleinheinz KA, Joensen KG, Larsen MV. Applying the ResFinder and VirulenceFinder web-services for easy identification of acquired antibiotic resistance and E. coli virulence genes in bacteriophage and prophage nucleotide sequences. Bacteriophage 2014; 4:e27943 [View Article]
    [Google Scholar]
  125. Carattoli A, Zankari E, Garcia-Fernandez A, Voldby Larsen M, Lund O et al. In silico detection and typing of plasmids using PlasmidFinder and plasmid multilocus sequence typing. Antimicrob Agents Chemother 2014; 58:3895–3903 [View Article]
    [Google Scholar]
  126. Robertson J, Nash JHE. MOB-suite: software tools for clustering, reconstruction and typing of plasmids from draft assemblies. Microb Genom 2018; 4: [View Article]
    [Google Scholar]
  127. Sullivan MJ, Petty NK, Beatson SA. Easyfig: a genome comparison visualizer. Bioinformatics 2011; 27:1009–1010 [View Article]
    [Google Scholar]
  128. Sievers F, Wilm A, Dineen D, Gibson TJ, Karplus K et al. Fast, scalable generation of high-quality protein multiple sequence alignments using Clustal Omega. Mol Syst Biol 2011; 7:539
    [Google Scholar]
  129. Sievers F, Higgins DG. Clustal Omega for making accurate alignments of many protein sequences. Protein Sci 2018; 27:135–145 [View Article]
    [Google Scholar]
  130. Scheutz F, Teel LD, Beutin L, Pierard D, Buvens G et al. Multicenter evaluation of a sequence-based protocol for subtyping Shiga toxins and standardizing Stx nomenclature. J Clin Microbiol 2012; 50:2951–2963 [View Article]
    [Google Scholar]
  131. Sanchez S, Llorente MT, Herrera-Leon L, Ramiro R, Nebreda S et al. Mucus-activatable shiga toxin genotype stx2d in Escherichia coli O157:H7. Emerg Infect Dis 2017; 23:1431–1433
    [Google Scholar]
  132. Malberg Tetzschner AM, Johnson JR, Johnston BD, Lund O, Scheutz F et al. In Silico Genotyping of Escherichia coli Isolates for Extraintestinal Virulence Genes by Use of Whole-Genome Sequencing Data. J Clin Microbiol 2020; 58:e01269–01220 [View Article]
    [Google Scholar]
  133. Afgan E, Baker D, Batut B, van den Beek M, Bouvier D et al. The Galaxy platform for accessible, reproducible and collaborative biomedical analyses: 2018 update. Nucleic Acids Res 2018; 46:W537–W544 [View Article]
    [Google Scholar]
  134. Bertelli C, Laird MR, Williams KP, Lau BY, Hoad G et al. IslandViewer 4: expanded prediction of genomic islands for larger-scale datasets. Nucleic Acids Res 2017; 45:W30–W35 [View Article]
    [Google Scholar]
  135. Bertelli C, Brinkman FSL. Improved genomic island predictions with IslandPath-DIMOB. Bioinformatics 2018; 34:2161–2167 [View Article]
    [Google Scholar]
  136. Bertelli C, Tilley KE, Brinkman FSL. Microbial genomic island discovery, visualization and analysis. Brief Bioinform 2018
    [Google Scholar]
  137. Alikhan NF, Petty NK, Ben Zakour NL, Beatson SA. BLAST Ring Image Generator (BRIG): simple prokaryote genome comparisons. BMC Genomics 2011; 12:402 [View Article]
    [Google Scholar]
  138. Sahl JW, Caporaso JG, Rasko DA, Keim P. The large-scale blast score ratio (LS-BSR) pipeline: a method to rapidly compare genetic content between bacterial genomes. PeerJ 2014; 2:e332
    [Google Scholar]
  139. Eppinger M, Worsham PL, Nikolich MP, Riley DR, Sebastian Y et al. Genome sequence of the deep-rooted Yersinia pestis strain Angola reveals new insights into the evolution and pangenome of the plague bacterium. J Bacteriol 2010; 192:1685–1699 [View Article]
    [Google Scholar]
  140. Saeed AI, Sharov V, White J, Li J, Liang W et al. TM4: a free, open-source system for microarray data management and analysis. Biotechniques 2003; 34:374–378 [View Article]
    [Google Scholar]
  141. Delannoy S, Mariani-Kurkdjian P, Webb HE, Bonacorsi S, Fach P. The mobilome; a major contributor to Escherichia coli stx2-positive O26:H11 strains intra-serotype diversity. Front Microbiol 2017; 8:1625
    [Google Scholar]
  142. Cock PJA, Chilton JM, Gruning B, Johnson JE, Soranzo N. NCBI BLAST plus integrated into galaxy. Gigascience 2015; 4:
    [Google Scholar]
  143. Cascales E, Buchanan SK, Duche D, Kleanthous C, Lloubes R et al. Colicin biology. Microbiol Mol Biol Rev 2007; 71:158–229 [View Article]
    [Google Scholar]
  144. Vriezen JA, Valliere M, Riley MA. The evolution of reduced microbial killing. Genome Biol Evol 2009; 1:400–408 [View Article]
    [Google Scholar]
  145. Zhang LH, Fath MJ, Mahanty HK, Tai PC, Kolter R. Genetic analysis of the colicin V secretion pathway. Genetics 1995; 141:25–32 [View Article]
    [Google Scholar]
  146. Selkrig J, Mosbahi K, Webb CT, Belousoff MJ, Perry AJ et al. Discovery of an archetypal protein transport system in bacterial outer membranes. Nat Struct Mol Biol 2012; 19:510–S501
    [Google Scholar]
  147. Campos E, Baldoma L, Aguilar J, Badia J. Regulation of expression of the divergent ulaG and ulaABCDEF operons involved in LaAscorbate dissimilation in Escherichia coli. J Bacteriol 2004; 186:1720–1728
    [Google Scholar]
  148. Clark G, Paszkiewicz K, Hale J, Weston V, Constantinidou C et al. Genomic analysis uncovers a phenotypically diverse but genetically homogeneous Escherichia coli ST131 clone circulating in unrelated urinary tract infections. J Antimicrob Chemother 2012; 67:868–877 [View Article]
    [Google Scholar]
  149. Driebe EM, Sahl JW, Roe C, Bowers JR, Schupp JM et al. Using whole genome analysis to examine recombination across diverse sequence types of Staphylococcus aureus. PLoS One 2015; 10:e0130955 [View Article]
    [Google Scholar]
  150. Bruen TC, Bryant D. Parsimony via consensus. Syst Biol 2008; 57:251–256 [View Article]
    [Google Scholar]
  151. Hauser JR, Atitkar RR, Petro CD, Lindsey RL, Strockbine N et al. The Virulence of Escherichia coli O157:H7 isolates in mice depends on shiga toxin type 2a (Stx2a)-induction and high levels of Stx2a in stool. Front Cell Infect Microbiol 2020; 10:62 [View Article]
    [Google Scholar]
  152. Ogura Y, Ooka T, Terajima J, Nougayrède J-P et al. Extensive genomic diversity and selective conservation of virulence-determinants in enterohemorrhagic Escherichia coli strains of O157 and non-O157 serotypes. Genome Biol 2007; 8:R138 [View Article]
    [Google Scholar]
  153. Steyert SR, Sahl JW, Fraser CM, Teel LD, Scheutz F et al. Comparative genomics and stx phage characterization of LEE-negative Shiga toxin-producing Escherichia coli. Front Cell Infect Microbiol 2012; 2:133 [View Article]
    [Google Scholar]
  154. Higashi K, Sakamaki Y, Herai E, Demizu R, Uemura T et al. Identification and functions of amino acid residues in PotB and PotC involved in spermidine uptake activity. J Biol Chem 2010; 285:39061–39069
    [Google Scholar]
  155. Igarashi K, Kashiwagi K. Polyamine transport in bacteria and yeast. Biochem J 1999; 344:633–642 [View Article]
    [Google Scholar]
  156. Hayashi T, Makino K, Ohnishi M, Kurokawa K, Ishii K et al. Complete genome sequence of enterohemorrhagic Escherichia coli O157:H7 and genomic comparison with a laboratory strain K-12. DNA Res 2001; 8:11–22
    [Google Scholar]
  157. Groth AC, Calos MP. Phage integrases: biology and applications. J Mol Biol 2004; 335:667–678 [View Article]
    [Google Scholar]
  158. Serra-Moreno R, Jofre J, Muniesa M. Insertion site occupancy by stx2 bacteriophages depends on the locus availability of the host strain chromosome. J Bacteriol 2007; 189:6645–6654
    [Google Scholar]
  159. Ravi A, Estensmo ELF, Abee-Lund TML, Foley SL, Allgaier B et al. Association of the gut microbiota mobilome with hospital location and birth weight in preterm infants. Pediatr Res 2017; 82:829–838 [View Article]
    [Google Scholar]
  160. Cowley LA, Dallman TJ, Jenkins C, Sheppard SK. Phage predation shapes the population structure of shiga-toxigenic Escherichia coli O157:H7 in the UK: an evolutionary perspective. Front Genet 2019; 10:763 [View Article]
    [Google Scholar]
  161. Del Cogliano ME, Pinto A, Goldstein J, Zotta E, Ochoa F et al. Relevance of bacteriophage 933W in the development of hemolytic uremic syndrome (HUS). Front Microbiol 2018; 9:3104 [View Article]
    [Google Scholar]
  162. Tyler JS, Beeri K, Reynolds JL, Alteri CJ, Skinner KG et al. Prophage induction is enhanced and required for renal disease and lethality in an EHEC mouse model. PLoS Pathog 2013; 9:e1003236 [View Article]
    [Google Scholar]
  163. Shimizu T, Ohta Y, Tsutsuki H, Noda M. Construction of a novel bioluminescent reporter system for investigating Shiga toxin expression of enterohemorrhagic Escherichia coli. Gene 2011; 478:1–10 [View Article]
    [Google Scholar]
  164. Loś JM, Loś M, Wegrzyn G, Wegrzyn A. Differential efficiency of induction of various lambdoid prophages responsible for production of Shiga toxins in response to different induction agents. Microb Pathog 2009; 47:289–298
    [Google Scholar]
  165. Muniesa M, Blanco JE, De Simon M, Serra-Moreno R, Blanch AR et al. Diversity of stx2 converting bacteriophages induced from Shiga-toxin-producing Escherichia coli strains isolated from cattle. Microbiology 2004; 150:2959–2971
    [Google Scholar]
  166. Balasubramanian S, Osburne MS, BrinJones H, Tai AK, Leong JM. Prophage induction, but not production of phage particles, is required for lethal disease in a microbiome-replete murine model of enterohemorrhagic E. coli infection. PLoS Pathog 2019; 15:e1007494 [View Article]
    [Google Scholar]
  167. Abu-Ali GS, Ouellette LM, Henderson ST, Lacher DW, Riordan JT et al. Increased adherence and expression of virulence genes in a lineage of Escherichia coli O157:H7 commonly associated with human infections. PLoS One 2010; 5:e10167
    [Google Scholar]
  168. Bielaszewska M, Idelevich EA, Zhang W, Bauwens A, Schaumburg F et al. Effects of antibiotics on Shiga toxin 2 production and bacteriophage induction by epidemic Escherichia coli O104:H4 strain. Antimicrob Agents Chemother 2012; 56:3277–3282 [View Article]
    [Google Scholar]
  169. McAdams HH, Shapiro L. Circuit simulation of genetic networks. Science 1995; 269:650–656 [View Article]
    [Google Scholar]
  170. Matsushiro A, Sato K, Miyamoto H, Yamamura T, Honda T. Induction of prophages of enterohemorrhagic Escherichia coli O157:H7 with norfloxacin. J Bacteriol 1999; 181:2257–2260 [View Article]
    [Google Scholar]
  171. Wagner PL, Livny J, Neely MN, Acheson DW, Friedman DI et al. Bacteriophage control of Shiga toxin 1 production and release by Escherichia coli. Mol Microbiol 2002; 44:957–970 [View Article]
    [Google Scholar]
  172. Karch H, Schmidt H, Janetzki-Mittmann C, Scheef J, Kroger M. Shiga toxins even when different are encoded at identical positions in the genomes of related temperate bacteriophages. Mol Gen Genet 1999; 262:600–607
    [Google Scholar]
  173. Iversen H, L’ Abée-Lund TM, Aspholm M, Arnesen LPS, Lindbäck T. Commensal E. coli stx2 lysogens produce high levels of phages after spontaneous prophage induction. Front Cell Infect Microbiol 2015; 5: [View Article]
    [Google Scholar]
  174. Gamage SD, Patton AK, Strasser JE, Chalk CL, Weiss AA. Commensal bacteria influence Escherichia coli O157:H7 persistence and Shiga toxin production in the mouse intestine. Infect Immun 2006; 74:1977–1983 [View Article]
    [Google Scholar]
  175. Figler HM, Dudley EG. The interplay of Escherichia coli O157:H7 and commensal E. coli: the importance of strain-level identification. Expert Rev Gastroenterol Hepatol 2016; 10:415–417
    [Google Scholar]
  176. Baumler AJ, Sperandio V. Interactions between the microbiota and pathogenic bacteria in the gut. Nature 2016; 535:85–93 [View Article]
    [Google Scholar]
  177. Smith DL, James CE, Sergeant MJ, Yaxian Y, Saunders JR et al. Short-tailed stx phages exploit the conserved YaeT protein to disseminate Shiga toxin genes among enterobacteria. J Bacteriol 2007; 189:7223–7233
    [Google Scholar]
  178. Richter TKS, Michalski JM, Zanetti L, Tennant SM, Chen WH et al. Responses of the human gut Escherichia coli population to pathogen and antibiotic disturbances. mSystems 2018; 3:
    [Google Scholar]
  179. Wong CS, Jelacic S, Habeeb RL, Watkins SL, Tarr PI. The risk of the hemolytic-uremic syndrome after antibiotic treatment of Escherichia coli O157:H7 infections. N Engl J Med 2000; 342:1930–1936
    [Google Scholar]
  180. Dundas S, Todd WT, Stewart AI, Murdoch PS, Chaudhuri AK et al. The central Scotland Escherichia coli O157:H7 outbreak: risk factors for the hemolytic uremic syndrome and death among hospitalized patients. Clin Infect Dis 2001; 33:923–931
    [Google Scholar]
  181. Gould LH, Demma L, Jones TF, Hurd S, Vugia DJ et al. Hemolytic uremic syndrome and death in persons with Escherichia coli O157:H7 infection, foodborne diseases active surveillance network sites, 2000-2006. Clin Infect Dis 2009; 49:1480–1485
    [Google Scholar]
  182. Foster DB. Modulation of the enterohemorrhagic E-coli virulence program through the human gastrointestinal tract. Virulence 2013; 4:315–323
    [Google Scholar]
  183. Dutilh BE, Backus L, Edwards RA, Wels M, Bayjanov JR et al. Explaining microbial phenotypes on a genomic scale: GWAS for microbes. Brief Funct Genomics 2013; 12:366–380 [View Article]
    [Google Scholar]
  184. Falush D. Toward the use of genomics to study microevolutionary change in bacteria. PLoS Genet 2009; 5:e1000627 [View Article]
    [Google Scholar]
  185. Falush D, Bowden R. Genome-wide association mapping in bacteria?. Trends Microbiol 2006; 14:353–355 [View Article]
    [Google Scholar]
  186. Bono JL. Genotyping Escherichia coli O157:H7 for its ability to cause disease in humans. Curr Protoc Microbiol 2009; 5A:3
    [Google Scholar]
  187. Werber D, Scheutz F. The importance of integrating genetic strain information for managing cases of shiga toxin-producing E. coli infection. Epidemiol Infect 2019; 147:e264
    [Google Scholar]
  188. Sadiq SM, Hazen TH, Rasko DA, Eppinger M. EHEC genomics: past, present, and future. microbiol spectr. Microbiol Spectr 2014; 2:
    [Google Scholar]
  189. Eppinger M, Daugherty S, Agrawal S, Galens K, Sengamalay N et al. Whole-genome draft sequences of 26 enterohemorrhagic escherichia coli O157:H7 strains. Genome Announc 2013; 1:e0013412
    [Google Scholar]
  190. Balasubramanian S, Osburne MS, BrinJones H, Tai AK, Leong JM. Prophage induction, but not production of phage particles, is required for lethal disease in a microbiome-replete murine model of enterohemorrhagic E. coli infection. PLOS Pathog 2019; 15:e1007494 [View Article]
    [Google Scholar]
  191. Hernandez-Doria JD, Sperandio V. Bacteriophage transcription factor Cro regulates virulence gene expression in enterohemorrhagic Escherichia coli. Cell Host Microbe 2018; 23:607–617 [View Article]
    [Google Scholar]
  192. Hu J, Ye H, Wang S, Wang J, Han D. Prophage activation in the intestine: insights into functions and possible applications. Front Microbiol 2021; 12:3930 [View Article]
    [Google Scholar]
  193. Kimmitt PT, Harwood CR, Barer MR. Toxin gene expression by shiga toxin-producing Escherichia coli: the role of antibiotics and the bacterial SOS response. Emerg Infect Dis 2000; 6:458–465 [View Article]
    [Google Scholar]
  194. Rodríguez-Rubio L, Haarmann N, Schwidder M, Muniesa M, Schmidt H. Bacteriophages of shiga toxin-producing Escherichia coli and their contribution to pathogenicity. Pathogens 2021; 10:
    [Google Scholar]
  195. Tyler JS, Beeri K, Reynolds JL, Alteri CJ, Skinner KG et al. Prophage induction is enhanced and required for renal disease and lethality in an EHEC mouse model. PLOS Pathog 2013; 9:e1003236 [View Article]
    [Google Scholar]
  196. Boyer RR, Sumner SS, Williams RC, Pierson MD, Popham DL et al. Influence of curli expression by Escherichia coli 0157:H7 on the cell’s overall hydrophobicity, charge, and ability to attach to lettuce. J Food Prot 2007; 70:1339–1345
    [Google Scholar]
  197. Macarisin D, Patel J, Bauchan G, Giron JA, Sharma VK. Role of curli and cellulose expression in adherence of Escherichia coli O157:H7 to spinach leaves. Foodborne Pathog Dis 2012; 9:160–167
    [Google Scholar]
  198. Bando SY, Iamashita P, Guth BE, Dos Santos LF, Fujita A et al. A hemolytic-uremic syndrome-associated strain O113:H21 Shiga toxin-producing Escherichia coli specifically expresses A transcriptional module containing dicA and is related to gene network dysregulation in Caco-2 cells. PLoS One 2017; 12:e0189613 [View Article]
    [Google Scholar]
  199. Bando SY, Iamashita P, Silva FN, Costa L da F, Abe CM et al. Dynamic gene network analysis of Caco-2 cell response to shiga toxin-producing Escherichia coli-associated hemolytic-uremic syndrome. Microorganisms 2019; 7:E195 [View Article]
    [Google Scholar]
  200. Sy BM, Lan R, Tree JJ. Early termination of the Shiga toxin transcript generates a regulatory small RNA. Proc Natl Acad Sci U S A 2020; 117:25055–25065 [View Article]
    [Google Scholar]
  201. Kiel M, Sagory-Zalkind P, Miganeh C, Stork C, Leimbach A et al. Identification of novel biomarkers for priority serotypes of shiga toxin-producing Escherichia coli and the development of multiplex PCR for their detection. Front Microbiol 2018; 9:1321 [View Article]
    [Google Scholar]
  202. Dytoc MT, Ismaili A, Philpott DJ, Soni R, Brunton JL et al. Distinct binding properties of eaeA-negative verocytotoxin-producing Escherichia coli of serotype O113:H21. Infect Immun 1994; 62:3494–3505 [View Article]
    [Google Scholar]
  203. Herold S, Paton JC, Paton AW. Sab, a novel autotransporter of locus of enterocyte effacement-negative shiga-toxigenic Escherichia coli O113:H21, contributes to adherence and biofilm formation. Infect Immun 2009; 77:3234–3243 [View Article]
    [Google Scholar]
  204. Luck SN, Badea L, Bennett-Wood V, Robins-Browne R, Hartland EL. Contribution of FliC to epithelial cell invasion by enterohemorrhagic Escherichia coli O113:H21. Infect Immun 2006; 74:6999–7004
    [Google Scholar]
  205. Rogers TJ, Paton JC, Wang H, Talbot UM, Paton AW. Reduced virulence of an fliC mutant of Shiga-toxigenic Escherichia coli O113:H21. Infect Immun 2006; 74:1962–1966
    [Google Scholar]
  206. Rogers TJ, Thorpe CM, Paton AW, Paton JC. Role of lipid rafts and flagellin in invasion of colonic epithelial cells by Shiga-toxigenic Escherichia coli O113:H21. Infect Immun 2012; 80:2858–2867 [View Article]
    [Google Scholar]
  207. Gerhardt E, Masso M, Paton AW, Paton JC, Zotta E et al. Inhibition of water absorption and selective damage to human colonic mucosa are induced by subtilase cytotoxin produced by Escherichia coli O113:H21. Infect Immun 2013; 81:2931–2937 [View Article]
    [Google Scholar]
  208. Seyahian EA, Oltra G, Ochoa F, Melendi S, Hermes R et al. Systemic effects of Subtilase cytotoxin produced by Escherichia coli O113:H21. Toxicon 2017; 127:49–55 [View Article]
    [Google Scholar]
  209. Santos E, Castro VS, Cunha-Neto A, Santos LFD, Vallim DC et al. Escherichia coli O26 and O113:H21 on Carcasses and beef from a slaughterhouse located in Mato Grosso, Brazil. Foodborne Pathog Dis 2018; 15:653–659
    [Google Scholar]
  210. Sanso AM, Bustamante AV, Kruger A, Cadona JS, Alfaro R et al. Molecular epidemiology of Shiga toxin-producing O113:H21 isolates from cattle and meat. Zoonoses Public Health 2018; 65:569–577 [View Article]
    [Google Scholar]
http://instance.metastore.ingenta.com/content/journal/mgen/10.1099/mgen.0.000796
Loading
/content/journal/mgen/10.1099/mgen.0.000796
Loading

Data & Media loading...

Supplements

Supplementary material 1

PDF

Supplementary material 2

EXCEL
This is a required field
Please enter a valid email address
Approval was a Success
Invalid data
An Error Occurred
Approval was partially successful, following selected items could not be processed due to error