1887

Abstract

Staphylococcus aureus has colonized humans for at least 10 000 years, and today inhabits roughly a third of the population. In addition, S. aureus is a major pathogen that is responsible for a significant disease burden, ranging in severity from mild skin and soft-tissue infections to life-threatening endocarditis and necrotizing pneumonia, with treatment often hampered by resistance to commonly available antibiotics. Underpinning its versatility as a pathogen is its ability to evade the innate immune system. S. aureus specifically targets innate immunity to establish and sustain infection, utilizing a large repertoire of virulence factors to do so. Using these factors, S. aureus can resist phagosomal killing, impair complement activity, disrupt cytokine signalling and target phagocytes directly using proteolytic enzymes and cytolytic toxins. Although most of these virulence factors are well characterized, their importance during infection is less clear, as many display species-specific activity against humans or against animal hosts, including cows, horses and chickens. Several staphylococcal virulence factors display species specificity for components of the human innate immune system, with as few as two amino acid changes reducing binding affinity by as much as 100-fold. This represents a major issue for studying their roles during infection, which cannot be examined without the use of humanized infection models. This review summarizes the major factors S. aureus uses to impair the innate immune system, and provides an in-depth look into the host specificity of S. aureus and how this problem is being approached.

Loading

Article metrics loading...

/content/journal/micro/10.1099/mic.0.000759
2019-01-09
2024-03-28
Loading full text...

Full text loading...

/deliver/fulltext/micro/165/4/367.html?itemId=/content/journal/micro/10.1099/mic.0.000759&mimeType=html&fmt=ahah

References

  1. Gorwitz RJ, Kruszon-Moran D, McAllister SK, McQuillan G, McDougal LK et al. Changes in the prevalence of nasal colonization with Staphylococcus aureus in the United States, 2001–2004. J Infect Dis 2008; 197:1226–1234 [View Article]
    [Google Scholar]
  2. Kluytmans J, Van Belkum A, Verbrugh H. Nasal carriage of Staphylococcus aureus: epidemiology, underlying mechanisms, and associated risks. Clin Microbiol Rev 1997; 10:505–520 [View Article]
    [Google Scholar]
  3. Von Eiff C, Becker K, MacHka K, Stammer H, Peters G. Nasal carriage as a source of Staphylococcus aureus bacteremia. N Engl J Med Overseas Ed 2001; 344:11–16 [View Article]
    [Google Scholar]
  4. Totté JEE, van der Feltz WT, Hennekam M, van Belkum A, van Zuuren EJ et al. Prevalence and odds of Staphylococcus aureus carriage in atopic dermatitis: a systematic review and meta-analysis. Br J Dermatol 2016; 175:687–695 [View Article]
    [Google Scholar]
  5. Tong SYC, Davis JS, Eichenberger E, Holland TL, Fowler VG. Staphylococcus aureus infections: epidemiology, pathophysiology, clinical manifestations, and management. Clin Microbiol Rev 2015; 28:603–661 [View Article]
    [Google Scholar]
  6. Selsted ME, Ouellette AJ. Mammalian defensins in the antimicrobial immune response. Nat Immunol 2005; 6:551–557 [View Article]
    [Google Scholar]
  7. Rice WG, Ganz T, Kinkade JM, Selsted ME, Lehrer RI et al. Defensin-rich dense granules of human neutrophils. Blood 1987; 70:757–765
    [Google Scholar]
  8. Jin T, Bokarewa M, Foster T, Mitchell J, Higgins J et al. Staphylococcus aureus resists human defensins by production of staphylokinase, a novel bacterial evasion mechanism. J Immunol 2004; 172:1169–1176 [View Article]
    [Google Scholar]
  9. Frohm M, Agerberth B, Ahangari G, Stâhle-Bäckdahl M, Lidén S et al. The expression of the gene coding for the antibacterial peptide LL-37 is induced in human keratinocytes during inflammatory disorders. J Biol Chem 1997; 272:15258–15263 [View Article][PubMed]
    [Google Scholar]
  10. Sørensen O, Arnljots K, Cowland JB, Bainton DF, Borregaard N. The human antibacterial cathelicidin, hCAP-18, is synthesized in myelocytes and metamyelocytes and localized to specific granules in neutrophils. Blood 1997; 90:2796–2803
    [Google Scholar]
  11. Sieprawska-Lupa M, Mydel P, Krawczyk K, Wójcik K, Puklo M et al. Degradation of human antimicrobial peptide LL-37 by Staphylococcus aureus-derived proteinases. Antimicrob Agents Chemother 2004; 48:4673–4679 [View Article][PubMed]
    [Google Scholar]
  12. Ong PY, Ohtake T, Brandt C, Strickland I, Boguniewicz M et al. Endogenous antimicrobial peptides and skin infections in atopic dermatitis. N Engl J Med Overseas Ed 2002; 347:1151–1160 [View Article]
    [Google Scholar]
  13. Alberts B, Johnson A, Lewis J, Raff M, Roberts K et al. Molecular Biology of the Cell 2002 Epub ahead of print 2002
    [Google Scholar]
  14. Bera A, Herbert S, Jakob A, Vollmer W, Götz F. Why are pathogenic staphylococci so lysozyme resistant? The peptidoglycan O-acetyltransferase OatA is the major determinant for lysozyme resistance of Staphylococcus aureus. Mol Microbiol 2005; 55:778–787 [View Article]
    [Google Scholar]
  15. Simanski M, Gläser R, Köten B, Meyer-Hoffert U, Wanner S et al. Staphylococcus aureus subverts cutaneous defense by D-alanylation of teichoic acids. Exp Dermatol 2013; 22:294–296 [View Article][PubMed]
    [Google Scholar]
  16. Kristian SA, Dürr M, Van Strijp JAG, Neumeister B, Peschel A. MprF-mediated lysinylation of phospholipids in Staphylococcus aureus leads to protection against oxygen-independent neutrophil killing. Infect Immun 2003; 71:546–549 [View Article]
    [Google Scholar]
  17. Tabuchi Y, Shiratsuchi A, Kurokawa K, Gong JH, Sekimizu K et al. Inhibitory role for D-alanylation of wall teichoic acid in activation of insect toll pathway by peptidoglycan of Staphylococcus aureus. J Immunol 2010; 185:2424–2431 [View Article]
    [Google Scholar]
  18. Kristian SA, Lauth X, Nizet V, Goetz F, Neumeister B et al. Alanylation of teichoic acids protects Staphylococcus aureus against Toll-like receptor 2-dependent host defense in a mouse tissue cage infection model. J Infect Dis 2003; 188:414–423 [View Article][PubMed]
    [Google Scholar]
  19. Serruto D, Rappuoli R, Scarselli M, Gros P, van Strijp JA. Molecular mechanisms of complement evasion: learning from staphylococci and meningococci. Nat Rev Microbiol 2010; 8:393–399 [View Article][PubMed]
    [Google Scholar]
  20. Ram S, MacKinnon FG, Gulati S, McQuillen DP, Vogel U et al. The contrasting mechanisms of serum resistance of Neisseria gonorrhoeae and group B Neisseria meningitidis. Mol Immunol 1999; 36:915–928 [View Article]
    [Google Scholar]
  21. Jarva H, Jokiranta TS, Würzner R, Meri S. Complement resistance mechanisms of streptococci. Mol Immunol 2003; 40:95–107 [View Article]
    [Google Scholar]
  22. Lambris JD, Ricklin D, Geisbrecht BV. Complement evasion by human pathogens. Nat Rev Microbiol 2008; 6:132–142 [View Article][PubMed]
    [Google Scholar]
  23. Graille M, Stura EA, Corper AL, Sutton BJ, Taussig MJ et al. Crystal structure of a Staphylococcus aureus protein A domain complexed with the Fab fragment of a human IgM antibody: structural basis for recognition of B-cell receptors and superantigen activity. Proc Natl Acad Sci USA 2000; 97:5399–5404 [View Article][PubMed]
    [Google Scholar]
  24. Gómez MI, Lee A, Reddy B, Muir A, Soong G et al. Staphylococcus aureus protein A induces airway epithelial inflammatory responses by activating TNFR1. Nat Med 2004; 10:842–848 [View Article]
    [Google Scholar]
  25. Zhang L, Jacobsson K, Vasi J, Lindberg M, Frykberg L. A second IgG-binding protein in Staphylococcus aureus. Microbiology 1998; 144:985–991 [View Article][PubMed]
    [Google Scholar]
  26. Zhang L, Jacobsson K, Ström K, Lindberg M, Frykberg L. Staphylococcus aureus expresses a cell surface protein that binds both IgG and 2-glycoprotein I. Microbiology 1999; 145:177–183 [View Article]
    [Google Scholar]
  27. Smith EJ, Visai L, Kerrigan SW, Speziale P, Foster TJ. The Sbi protein is a multifunctional immune evasion factor of Staphylococcus aureus. Infect Immun 2011; 79:3801–3809 [View Article]
    [Google Scholar]
  28. Smith EJ, Corrigan RM, van der Sluis T, Gründling A, Speziale P et al. The immune evasion protein Sbi of Staphylococcus aureus occurs both extracellularly and anchored to the cell envelope by binding lipoteichoic acid. Mol Microbiol 2012; 83:789–804 [View Article]
    [Google Scholar]
  29. Upadhyay A, Burman JD, Clark EA, Leung E, Isenman DE et al. Structure-function analysis of the c3 binding region of Staphylococcus aureus immune subversion protein Sbi. J Biol Chem 2008; 283:22113–22120 [View Article]
    [Google Scholar]
  30. Langley R, Wines B, Willoughby N, Basu I, Proft T et al. The Staphylococcal superantigen-like protein 7 binds IgA and complement c5 and inhibits IgA-Fc RI binding and serum killing of bacteria. J Immunol 2005; 174:2926–2933 [View Article]
    [Google Scholar]
  31. Herman-Bausier P, Pietrocola G, Foster TJ, Speziale P, Dufrêne YF. Fibrinogen activates the capture of human plasminogen by Staphylococcal Fibronectin-binding proteins. mBio 2017; 8: [View Article]
    [Google Scholar]
  32. Rooijakkers SHM, van Wamel WJB, Ruyken M, van Kessel KPM, van Strijp JAG. Anti-opsonic properties of staphylokinase. Microbes Infect 2005; 7:476–484 [View Article]
    [Google Scholar]
  33. Laarman AJ, Ruyken M, Malone CL, van Strijp JAG, Horswill AR et al. Staphylococcus aureus metalloprotease aureolysin cleaves complement c3 to mediate immune evasion. J Immunol 2011; 186:6445–6453 [View Article]
    [Google Scholar]
  34. Lee LYL, Höök M, Haviland D, Wetsel RA, Yonter EO et al. Inhibition of complement activation by a secreted Staphylococcus aureus protein. J Infect Dis 2004; 190:571–579
    [Google Scholar]
  35. Ricklin D, Ricklin-Lichtsteiner SK, Markiewski MM, Geisbrecht BV, Lambris JD. Cutting edge: members of the Staphylococcus aureus extracellular fibrinogen-binding protein family inhibit the interaction of C3d with complement receptor 2. J Immunol 2008; 181:7463–7467 [View Article][PubMed]
    [Google Scholar]
  36. Wu J, Wu YQ, Ricklin D, Janssen BJ, Lambris JD et al. Structure of complement fragment C3b-factor H and implications for host protection by complement regulators. Nat Immunol 2009; 10:728–733 [View Article][PubMed]
    [Google Scholar]
  37. Amdahl H, Jongerius I, Meri T, Pasanen T, Hyvarinen S et al. Staphylococcal Ecb protein and host complement regulator factor H enhance functions of each other in bacterial immune evasion. J Immunol 2013; 191:1775–1784 [View Article]
    [Google Scholar]
  38. Amdahl H, Haapasalo K, Tan L, Meri T, Kuusela PI et al. Staphylococcal protein Ecb impairs complement receptor-1 mediated recognition of opsonized bacteria. PLoS One 2017; 12:e0172675 [View Article]
    [Google Scholar]
  39. de Haas CJ, Veldkamp KE, Peschel A, Weerkamp F, van Wamel WJ et al. Chemotaxis inhibitory protein of Staphylococcus aureus, a bacterial antiinflammatory agent. J Exp Med 2004; 199:687–695 [View Article][PubMed]
    [Google Scholar]
  40. Haas PJ, de Haas CJ, Kleibeuker W, Poppelier MJ, van Kessel KP et al. N-terminal residues of the chemotaxis inhibitory protein of Staphylococcus aureus are essential for blocking formylated peptide receptor but not C5a receptor. J Immunol 2004; 173:5704–5711 [View Article][PubMed]
    [Google Scholar]
  41. Postma B, Kleibeuker W, Poppelier MJ, Boonstra M, van Kessel KP et al. Residues 10-18 within the C5a receptor N terminus compose a binding domain for chemotaxis inhibitory protein of Staphylococcus aureus. J Biol Chem 2005; 280:2020–2027 [View Article][PubMed]
    [Google Scholar]
  42. Rooijakkers SHM, Ruyken M, Roos A, Daha MR, Presanis JS et al. Immune evasion by a staphylococcal complement inhibitor that acts on C3 convertases. Nat Immunol 2005; 6:920–927 [View Article]
    [Google Scholar]
  43. Hoekstra H, Romero Pastrana F, Bonarius HPJ, van Kessel KPM, Elsinga GS et al. A human monoclonal antibody that specifically binds and inhibits the staphylococcal complement inhibitor protein SCIN. Virulence 2017; 5594:1–13
    [Google Scholar]
  44. Kawai T, Akira S. The role of pattern-recognition receptors in innate immunity: update on toll-like receptors. Nat Immunol 2010; 11:373–384 [View Article]
    [Google Scholar]
  45. Hato T, Dagher PC. How the innate immune system senses trouble and causes trouble. Clin J Am Soc Nephrol 2015; 10:1459–1469 [View Article][PubMed]
    [Google Scholar]
  46. Spaan AN, Surewaard BG, Nijland R, van Strijp JA. Neutrophils versus Staphylococcus aureus: a biological tug of war. Annu Rev Microbiol 2013; 67:629–650 [View Article][PubMed]
    [Google Scholar]
  47. Guerra FE, Borgogna TR, Patel DM, Sward EW, Voyich JM. Epic immune battles of history: neutrophils vs. Staphylococcus aureus. Front Cell Infect Microbiol 2017; 7:1–19 [View Article]
    [Google Scholar]
  48. Laarman AJ, Mijnheer G, Mootz JM, van Rooijen WJM, Ruyken M et al. Staphylococcus aureus Staphopain A inhibits CXCR2-dependent neutrophil activation and chemotaxis. EMBO J 2012; 31:3607–3619 [View Article]
    [Google Scholar]
  49. Su SB, Gao J, Gong W, Dunlop NM, Murphy PM et al. T21/DP107, A synthetic leucine zipper-like domain of the HIV-1 envelope gp41, attracts and activates human phagocytes by using G-protein-coupled formyl peptide receptors. J Immunol 1999; 162:5924–5930[PubMed]
    [Google Scholar]
  50. De Yang, Chen Q, Schmidt AP, Anderson GM, Wang JM et al. LL-37, the neutrophil granule- and epithelial cell-derived cathelicidin, utilizes formyl peptide receptor-like 1 (FPRL1) as a receptor to chemoattract human peripheral blood neutrophils, monocytes, and T cells. J Exp Med 2000; 192:1069–1074 [View Article][PubMed]
    [Google Scholar]
  51. Le Y, Yazawa H, Gong W, Yu Z, Ferrans VJ et al. Cutting edge: the neurotoxic prion peptide fragment PrP106-126 is a chemotactic agonist for the G protein-coupled receptor formyl peptide receptor-like 1. J Immunol 2001; 166:1448–1451 [View Article]
    [Google Scholar]
  52. Prat C, Bestebroer J, de Haas CJC, van Strijp JAG, van Kessel KPM. A new staphylococcal anti-inflammatory protein that antagonizes the formyl peptide receptor-like 1. J Immunol 2006; 177:8017–8026 [View Article]
    [Google Scholar]
  53. Prat C, Haas PJ, Bestebroer J, de Haas CJ, van Strijp JA et al. A homolog of formyl peptide receptor-like 1 (FPRL1) inhibitor from Staphylococcus aureus (FPRL1 inhibitory protein) that inhibits FPRL1 and FPR. J Immunol 2009; 183:6569–6578 [View Article][PubMed]
    [Google Scholar]
  54. Stemerding AM, Kohl J, Pandey MK, Kuipers A, Leusen JH et al. Staphylococcus aureus formyl peptide receptor-like 1 inhibitor (FLIPr) and its homologue FLIPr-like are potent Fc R antagonists that inhibit IgG-mediated effector functions. J Immunol 2013; 191:353–362 [View Article]
    [Google Scholar]
  55. Zhao Y, van Kessel KPM, de Haas CJC, Rogers MRC, van Strijp JAG et al. Staphylococcal superantigen-like protein 13 activates neutrophils via formyl peptide receptor 2. Cell Microbiol 2018; 20:e12941 [View Article][PubMed]
    [Google Scholar]
  56. Takeuchi O, Sato S, Horiuchi T, Hoshino K, Takeda K et al. Cutting edge: role of Toll-like receptor 1 in mediating immune response to microbial lipoproteins. J Immunol 2002; 169:10–14 [View Article][PubMed]
    [Google Scholar]
  57. Gonzalez-Zorn B, Senna JPM, Fiette L, Shorte S, Testard A et al. Bacterial and host factors implicated in nasal carriage of methicillin-resistant Staphylococcus aureus in Mice. Infect Immun 2005; 73:1847–1851 [View Article]
    [Google Scholar]
  58. Takeuchi O, Hoshino K, Akira S. Cutting edge: TLR2-deficient and MyD88-deficient mice are highly susceptible to Staphylococcus aureus infection. J Immunol 2000; 165:5392–5396 [View Article]
    [Google Scholar]
  59. Hanzelmann D, Joo H-S, Franz-Wachtel M, Hertlein T, Stevanovic S et al. Toll-like receptor 2 activation depends on lipopeptide shedding by bacterial surfactants. Nat Commun 2016; 7:12304 [View Article]
    [Google Scholar]
  60. Bardoel BW, Vos R, Bouman T, Aerts PC, Bestebroer J et al. Evasion of Toll-like receptor 2 activation by staphylococcal superantigen-like protein 3. J Mol Med 2012; 90:1109–1120 [View Article]
    [Google Scholar]
  61. Sabroe I, Prince LR, Jones EC, Horsburgh MJ, Foster SJ et al. Selective roles for toll-like receptor (TLR)2 and TLR4 in the regulation of neutrophil activation and life span. J Immunol 2003; 170:5268–5275 [View Article]
    [Google Scholar]
  62. Nguyen MT, Uebele J, Kumari N, Nakayama H, Peter L et al. Lipid moieties on lipoproteins of commensal and non-commensal staphylococci induce differential immune responses. Nat Commun 2017; 8:2246 [View Article][PubMed]
    [Google Scholar]
  63. Fevre C, Bestebroer J, Mebius MM, de Haas CJC, van Strijp JAG et al. Staphylococcus aureus proteins SSL6 and SElX interact with neutrophil receptors as identified using secretome phage display. Cell Microbiol 2014; 16:1646–1665 [View Article]
    [Google Scholar]
  64. Bestebroer J, Poppelier M, Ulfman LH, Lenting PJ, Denis C V et al. Staphylococcal superantigen-like 5 binds PSGL-1 and inhibits P-selectin-mediated neutrophil rolling. Blood 2007; 109:2936–2943
    [Google Scholar]
  65. Bestebroer J, van Kessel KPM, Azouagh H, Walenkamp AM, Boer IGJ et al. Staphylococcal SSL5 inhibits leukocyte activation by chemokines and anaphylatoxins. Blood 2009; 113:328–337 [View Article]
    [Google Scholar]
  66. Pham CTN. Neutrophil serine proteases: specific regulators of inflammation. Nat Rev Immunol 2006; 6:541–550 [View Article]
    [Google Scholar]
  67. Papayannopoulos V, Metzler KD, Hakkim A, Zychlinsky A. Neutrophil elastase and myeloperoxidase regulate the formation of neutrophil extracellular traps. J Cell Biol 2010; 191:677–691 [View Article]
    [Google Scholar]
  68. Belaaouaj A, Kim KS, Shapiro SD. Degradation of outer membrane protein A in Escherichia coli killing by neutrophil elastase. Science 2000; 289:1185–1187[PubMed]
    [Google Scholar]
  69. Weinrauch Y, Drujan D, Shapiro SD, Weiss J, Zychlinsky A. Neutrophil elastase targets virulence factors of enterobacteria. Nature 2002; 417:91–94 [View Article]
    [Google Scholar]
  70. Korkmaz B, Horwitz MS, Jenne DE, Gauthier F. Neutrophil elastase, proteinase 3, and cathepsin g as therapeutic targets in human diseases. Pharmacol Rev 2010; 62:726–759 [View Article]
    [Google Scholar]
  71. Stapels DAC, Ramyar KX, Bischoff M, von Kockritz-Blickwede M, Milder FJ et al. Staphylococcus aureus secretes a unique class of neutrophil serine protease inhibitors. Proc Natl Acad Sci 2014; 111:13187–13192 [View Article]
    [Google Scholar]
  72. Date SV, Modrusan Z, Lawrence M, Morisaki JH, Toy K et al. Global gene expression of methicillin-resistant Staphylococcus aureus USA300 during human and mouse infection. J Infect Dis 2014; 209:1542–1550 [View Article][PubMed]
    [Google Scholar]
  73. Joost I, Blass D, Burian M, Goerke C, Wolz C et al. Transcription analysis of the extracellular adherence protein from Staphylococcus aureus in authentic human infection and in vitro. J Infect Dis 2009; 199:1471–1478 [View Article][PubMed]
    [Google Scholar]
  74. van Wamel WJ, Rooijakkers SH, Ruyken M, van Kessel KP, van Strijp JA. The innate immune modulators staphylococcal complement inhibitor and chemotaxis inhibitory protein of Staphylococcus aureus are located on beta-hemolysin-converting bacteriophages. J Bacteriol 2006; 188:1310–1315 [View Article][PubMed]
    [Google Scholar]
  75. Rahimpour R, Mitchell G, Khandaker MH, Kong C, Singh B et al. Bacterial superantigens induce down-modulation of CC chemokine responsiveness in human monocytes via an alternative chemokine ligand-independent mechanism. J Immunol 1999; 162:2299–2307
    [Google Scholar]
  76. Alonzo F, Kozhaya L, Rawlings SA, Reyes-Robles T, Dumont AL et al. CCR5 is a receptor for Staphylococcus aureus leukotoxin ED. Nature 2013; 493:51–55 [View Article][PubMed]
    [Google Scholar]
  77. Spaan AN, Vrieling M, Wallet P, Badiou C, Reyes-Robles T et al. The staphylococcal toxins γ-haemolysin AB and CB differentially target phagocytes by employing specific chemokine receptors. Nat Commun 2014; 5:5438 [View Article]
    [Google Scholar]
  78. Spaan AN, Henry T, van Rooijen WJM, Perret M, Badiou C et al. The Staphylococcal toxin panton-valentine leukocidin targets human C5a receptors. Cell Host Microbe 2013; 13:584–594 [View Article]
    [Google Scholar]
  79. Reyes-Robles T, Alonzo F, Kozhaya L, Lacy DB, Unutmaz D et al. Staphylococcus aureus leukotoxin ED targets the chemokine receptors CXCR1 and CXCR2 to kill leukocytes and promote infection. Cell Host Microbe 2013; 14:453–459 [View Article]
    [Google Scholar]
  80. Panton PN, Valentine FCO. Staphylococcal toxin. Lancet 1932; 219:506–508 [View Article]
    [Google Scholar]
  81. Burnet FM. The exotoxins of Staphylococcus pyogenes aureus. J Pathol Bacteriol 1929; 32:717–734 [View Article]
    [Google Scholar]
  82. Bubeck Wardenburg J, Bae T, Otto M, Deleo FR, Schneewind O. Poring over pores: α-hemolysin and Panton-Valentine leukocidin in Staphylococcus aureus pneumonia. Nat Med 2007; 13:1405–1406 [View Article]
    [Google Scholar]
  83. Gouaux JE, Braha O, Hobaugh MR, Song L, Cheley S et al. Subunit stoichiometry of staphylococcal alpha-hemolysin in crystals and on membranes: a heptameric transmembrane pore. Proc Natl Acad Sci USA 1994; 91:12828–12831 [View Article][PubMed]
    [Google Scholar]
  84. Nygaard TK, Pallister KB, DuMont AL, Dewald M, Watkins RL et al. Alpha-toxin induces programmed cell death of human T cells, B cells, and monocytes during USA300 infection. PLoS One 2012; 7:e36532 [View Article][PubMed]
    [Google Scholar]
  85. Tweten RK. Cholesterol-dependent cytolysins, a family of versatile pore-forming toxins. Infect Immun 2005; 73:6199–6209 [View Article][PubMed]
    [Google Scholar]
  86. Hildebrand A, Pohl M, Bhakdi S. Staphylococcus aureus alpha-toxin. Dual mechanism of binding to target cells. J Biol Chem 1991; 266:17195–17200
    [Google Scholar]
  87. Wilke GA, Bubeck Wardenburg J. Role of a disintegrin and metalloprotease 10 in Staphylococcus aureus α-hemolysin-mediated cellular injury. Proc Natl Acad Sci USA 2010; 107:13473–13478 [View Article][PubMed]
    [Google Scholar]
  88. Maretzky T, Reiss K, Ludwig A, Buchholz J, Scholz F et al. ADAM10 mediates E-cadherin shedding and regulates epithelial cell-cell adhesion, migration, and beta-catenin translocation. Proc Natl Acad Sci USA 2005; 102:9182–9187 [View Article][PubMed]
    [Google Scholar]
  89. Schulte A, Schulz B, Andrzejewski MG, Hundhausen C, Mletzko S et al. Sequential processing of the transmembrane chemokines CX3CL1 and CXCL16 by α- and γ-secretases. Biochem Biophys Res Commun 2007; 358:233–240 [View Article]
    [Google Scholar]
  90. Kennedy AD, Bubeck Wardenburg J, Gardner DJ, Long D, Whitney AR et al. Targeting of alpha-hemolysin by active or passive immunization decreases severity of USA300 skin infection in a mouse model. J Infect Dis 2010; 202:1050–1058 [View Article][PubMed]
    [Google Scholar]
  91. Menzies BE, Kernodle DS. Passive immunization with antiserum to a nontoxic alpha-toxin mutant from Staphylococcus aureus is protective in a murine model. Infect Immun 1996; 64:1839–1841
    [Google Scholar]
  92. Jonsson P, Lindberg M, Haraldsson I, Wadström T. Virulence of Staphylococcus aureus in a mouse mastitis model: studies of alpha hemolysin, coagulase, and protein A as possible virulence determinants with protoplast fusion and gene cloning. Infect Immun 1985; 49:765–769
    [Google Scholar]
  93. Bayer AS, Ramos MD, Menzies BE, Yeaman MR, Shen AJ et al. Hyperproduction of alpha-toxin by Staphylococcus aureus results in paradoxically reduced virulence in experimental endocarditis: a host defense role for platelet microbicidal proteins. Infect Immun 1997; 65:4652–4660
    [Google Scholar]
  94. Bramley AJ, Patel AH, O'Reilly M, Foster R, Foster TJ. Roles of alpha-toxin and beta-toxin in virulence of Staphylococcus aureus for the mouse mammary gland. Infect Immun 1989; 57:2489–2494[PubMed]
    [Google Scholar]
  95. Kolata J, Bode LGM, Holtfreter S, Steil L, Kusch H et al. Distinctive patterns in the human antibody response to Staphylococcus aureus bacteremia in carriers and non-carriers. Proteomics 2011; 11:3914–3927 [View Article]
    [Google Scholar]
  96. Fritz SA, Tiemann KM, Hogan PG, Epplin EK, Rodriguez M et al. A serologic correlate of protective immunity against community-onset Staphylococcus aureus infection. Clin Infect Dis 2013; 56:1554–1561 [View Article][PubMed]
    [Google Scholar]
  97. Adhikari RP, Ajao AO, Aman MJ, Karauzum H, Sarwar J et al. Lower antibody levels to Staphylococcus aureus exotoxins are associated with sepsis in hospitalized adults with invasive S. aureus infections. J Infect Dis 2012; 206:915–923 [View Article][PubMed]
    [Google Scholar]
  98. Montgomery CP, Boyle-Vavra S, Adem PV, Lee JC, Husain AN et al. Comparison of virulence in community-associated methicillin-resistant Staphylococcus aureus pulsotypes USA300 and USA400 in a Rat Model of Pneumonia. J Infect Dis 2008; 198:561–570 [View Article][PubMed]
    [Google Scholar]
  99. Berube BJ, Bubeck Wardenburg J. Staphylococcus aureus α-toxin: nearly a century of intrigue. Toxins 2013; 5:1140–1166 [View Article][PubMed]
    [Google Scholar]
  100. Wang R, Braughton KR, Kretschmer D, Bach T-HL, Queck SY et al. Identification of novel cytolytic peptides as key virulence determinants for community-associated MRSA. Nat Med 2007; 13:1510–1514 [View Article]
    [Google Scholar]
  101. Cheung GYC, Duong AC, Otto M. Direct and synergistic hemolysis caused by Staphylococcus phenol-soluble modulins: implications for diagnosis and pathogenesis. Microbes Infect 2012; 14:380–386 [View Article]
    [Google Scholar]
  102. Periasamy S, Joo HS, Duong AC, Bach TH, Tan VY et al. How Staphylococcus aureus biofilms develop their characteristic structure. Proc Natl Acad Sci USA 2012; 109:1281–1286 [View Article][PubMed]
    [Google Scholar]
  103. Periasamy S, Chatterjee SS, Cheung GYC, Otto M. Phenol-soluble modulins in staphylococci: What are they originally for?. Commun Integr Biol 2012; 5:275–277
    [Google Scholar]
  104. Kretschmer D, Gleske A-K, Rautenberg M, Wang R, Köberle M et al. Human formyl peptide receptor 2 senses highly pathogenic Staphylococcus aureus. Cell Host Microbe 2010; 7:463–473 [View Article]
    [Google Scholar]
  105. Rasigade JP, Trouillet-Assant S, Ferry T, Diep BA, Sapin A et al. PSMs of hypervirulent Staphylococcus aureus act as intracellular toxins that kill infected osteoblasts. PLoS One 2013; 8:e63176 [View Article][PubMed]
    [Google Scholar]
  106. Mellor IR, Thomas DH, Sansom MS. Properties of ion channels formed by Staphylococcus aureus delta-toxin. Biochim Biophys Acta 1988; 942:280–294 [View Article][PubMed]
    [Google Scholar]
  107. Peschel A, Otto M. Phenol-soluble modulins and staphylococcal infection. Nat Rev Microbiol 2013; 11:667–673 [View Article]
    [Google Scholar]
  108. Chatterjee SS, Joo HS, Duong AC, Dieringer TD, Tan VY et al. Essential Staphylococcus aureus toxin export system. Nat Med 2013; 19:364–367 [View Article][PubMed]
    [Google Scholar]
  109. Surewaard BGJ, de Haas CJC, Vervoort F, Rigby KM, Deleo FR et al. Staphylococcal alpha-phenol soluble modulins contribute to neutrophil lysis after phagocytosis. Cell Microbiol 2013; 15:1427–1437 [View Article]
    [Google Scholar]
  110. Carnes EC, Lopez DM, Donegan NP, Cheung A, Gresham H et al. Confinement-induced quorum sensing of individual Staphylococcus aureus bacteria. Nat Chem Biol 2010; 6:41–45 [View Article]
    [Google Scholar]
  111. Mehlin C, Headley CM, Klebanoff SJ. An inflammatory polypeptide complex from Staphylococcus epidermidis: isolation and characterization. J Exp Med 1999; 189:907–918 [View Article][PubMed]
    [Google Scholar]
  112. Ye RD, Boulay F, Wang J, Dahlgren C. Nomenclature for the Formyl Peptide Receptor (FPR) Family. Pharmacol Rev 2009; 61:119–161
    [Google Scholar]
  113. Kretschmer D, Rautenberg M, Linke D, Peschel A. Peptide length and folding state govern the capacity of staphylococcal β-type phenol-soluble modulins to activate human formyl-peptide receptors 1 or 2. J Leukoc Biol 2015; 97:689–697 [View Article]
    [Google Scholar]
  114. Hashimoto M, Tawaratsumida K, Kariya H, Kiyohara A, Suda Y et al. Not lipoteichoic acid but lipoproteins appear to be the dominant immunobiologically active compounds in Staphylococcus aureus. J Immunol 2006; 177:3162–3169 [View Article][PubMed]
    [Google Scholar]
  115. Alonzo F, Torres VJ. The bicomponent pore-forming leucocidins of Staphylococcus aureus. Microbiol Mol Biol Rev 2014; 78:199–230 [View Article][PubMed]
    [Google Scholar]
  116. DuMont AL, Yoong P, Day CJ, Alonzo F, McDonald WH et al. Staphylococcus aureus LukAB cytotoxin kills human neutrophils by targeting the CD11b subunit of the integrin Mac-1. Proc Natl Acad Sci USA 2013; 110:10794–10799 [View Article][PubMed]
    [Google Scholar]
  117. Colin DA, Mazurier I, Sire S, Finck-Barbançon V. Interaction of the two components of leukocidin from Staphylococcus aureus with human polymorphonuclear leukocyte membranes: sequential binding and subsequent activation. Infect Immun 1994; 62:3184–3188[PubMed]
    [Google Scholar]
  118. Tromp AT, van Gent M, Abrial P, Martin A, Jansen JP et al. Human CD45 is an F-component-specific receptor for the staphylococcal toxin Panton-Valentine leukocidin. Nat Microbiol 2018; 3: Epub ahead of print 7 May 2018 [View Article][PubMed]
    [Google Scholar]
  119. Venkatakrishnan AJ, Deupi X, Lebon G, Tate CG, Schertler GF et al. Molecular signatures of G-protein-coupled receptors. Nature 2013; 494:185–194 [View Article]
    [Google Scholar]
  120. Ross GD. Analysis of the different types of leukocyte membrane complement receptors and their interaction with the complement system. J Immunol Methods 1980; 37:197–211 [View Article]
    [Google Scholar]
  121. Hynes RO. Integrins: bidirectional, allosteric signaling machines. Cell 2002; 110:673–687
    [Google Scholar]
  122. Tournamille C, Filipe A, Wasniowska K, Gane P, Lisowska E et al. Structure-function analysis of the extracellular domains of the Duffy antigen/receptor for chemokines: characterization of antibody and chemokine binding sites. Br J Haematol 2003; 122:1014–1023 [View Article][PubMed]
    [Google Scholar]
  123. Spaan AN, Reyes-Robles T, Badiou C, Cochet S, Boguslawski KM et al. Staphylococcus aureus targets the duffy antigen receptor for chemokines (DARC) to lyse erythrocytes. Cell Host Microbe 2015; 18:363–370 [View Article][PubMed]
    [Google Scholar]
  124. Adhikari RP, Kort T, Shulenin S, Kanipakala T, Ganjbaksh N et al. Antibodies to S. aureus LukS-PV attenuated subunit vaccine neutralize a broad spectrum of canonical and non-canonical bicomponent leukotoxin Pairs. PLoS One 2015; 10:e0137874 [View Article]
    [Google Scholar]
  125. DuMont AL, Nygaard TK, Watkins RL, Smith A, Kozhaya L et al. Characterization of a new cytotoxin that contributes to Staphylococcus aureus pathogenesis. Mol Microbiol 2011; 79:814–825 [View Article]
    [Google Scholar]
  126. Melehani JH, James DB, DuMont AL, Torres VJ, Duncan JA. Staphylococcus aureus Leukocidin A/B (LukAB) kills human monocytes via host NLRP3 and ASC when extracellular, but not intracellular. PLoS Pathog 2015; 11:e1004970 [View Article][PubMed]
    [Google Scholar]
  127. Szmigielski S, Prévost G, Monteil H, Colin DA, Jeljaszewicz J. Leukocidal toxins of Staphylococci. Zentralbl Bakteriol 1999; 289:185–201 [View Article]
    [Google Scholar]
  128. Diep BA, Chan L, Tattevin P, Kajikawa O, Martin TR et al. Polymorphonuclear leukocytes mediate Staphylococcus aureus panton-valentine leukocidin-induced lung inflammation and injury. Proc Natl Acad Sci USA 2010; 107:5587–5592 [View Article][PubMed]
    [Google Scholar]
  129. Voyich JM, Otto M, Mathema B, Braughton KR, Whitney AR et al. Is Panton-Valentine leukocidin the major virulence determinant in community-associated methicillin-resistant Staphylococcus aureus disease?. J Infect Dis 2006; 194:1761–1770 [View Article][PubMed]
    [Google Scholar]
  130. Winterbourn CC. Reconciling the chemistry and biology of reactive oxygen species. Nat Chem Biol 2008; 4:278–286 [View Article]
    [Google Scholar]
  131. Hampton MB, Kettle AJ, Winterbourn CC. Involvement of superoxide and myeloperoxidase in oxygen-dependent killing of Staphylococcus aureus by neutrophils. Infect Immun 1996; 64:3512–3517
    [Google Scholar]
  132. Clements MO, Watson SP, Foster SJ. Characterization of the major superoxide dismutase of Staphylococcus aureus and its role in starvation survival, stress resistance, and pathogenicity. J Bacteriol 1999; 181:3898–3903
    [Google Scholar]
  133. Juttukonda LJ, Berends ETM, Zackular JP, Moore JL, Stier MT et al. Dietary manganese promotes Staphylococcal Infection of the Heart. Cell Host Microbe 2017; 22:531–542 [View Article]
    [Google Scholar]
  134. Treffon J, Block D, Moche M, Reiß S, Fuchs S et al. Adaptation of Staphylococcus aureus to the airways of cystic fibrosis patients by the up-regulation of superoxide dismutase M and iron-scavenging proteins. J Infect Dis 2018;1–9
    [Google Scholar]
  135. Marshall JH, Wilmoth GJ. Pigments of Staphylococcus aureus, a series of triterpenoid carotenoids. J Bacteriol 1981; 147:900–913
    [Google Scholar]
  136. El-Agamey A, Lowe GM, McGarvey DJ, Mortensen A, Phillip DM et al. Carotenoid radical chemistry and antioxidant/pro-oxidant properties. Arch Biochem Biophys 2004; 430:37–48 [View Article]
    [Google Scholar]
  137. Clauditz A, Resch A, Wieland KP, Peschel A, Götz F. Staphyloxanthin plays a role in the fitness of Staphylococcus aureus and its ability to cope with oxidative stress. Infect Immun 2006; 74:4950–4953 [View Article][PubMed]
    [Google Scholar]
  138. Liu GY, Essex A, Buchanan JT, Datta V, Hoffman HM et al. Staphylococcus aureus golden pigment impairs neutrophil killing and promotes virulence through its antioxidant activity. J Exp Med 2005; 202:209–215 [View Article]
    [Google Scholar]
  139. Liu CI, Liu GY, Song Y, Yin F, Hensler ME et al. A cholesterol biosynthesis inhibitor blocks Staphylococcus aureus virulence. Science 2008; 319:1391–1394 [View Article][PubMed]
    [Google Scholar]
  140. Winterbourn CC, Hampton MB, Livesey JH, Kettle AJ. Modeling the reactions of superoxide and myeloperoxidase in the neutrophil phagosome: implications for microbial killing. J Biol Chem 2006; 281:39860–39869 [View Article][PubMed]
    [Google Scholar]
  141. Park S, You X, Imlay JA. Substantial DNA damage from submicromolar intracellular hydrogen peroxide detected in Hpx- mutants of Escherichia coli. Proc Natl Acad Sci USA 2005; 102:9317–9322 [View Article][PubMed]
    [Google Scholar]
  142. Seaver LC, Imlay JA. Hydrogen peroxide fluxes and compartmentalization inside growing Escherichia coli. J Bacteriol 2001; 183:7182–7189 [View Article][PubMed]
    [Google Scholar]
  143. Kanafani H, Martin SE. Catalase and superoxide dismutase activities in virulent and nonvirulent Staphylococcus aureus isolates. J Clin Microbiol 1985; 21:607–610
    [Google Scholar]
  144. Das D, Bishayi B. Staphylococcal catalase protects intracellularly survived bacteria by destroying H2O2 produced by the murine peritoneal macrophages. Microb Pathog 2009; 47:57–67 [View Article]
    [Google Scholar]
  145. Messina CGM, Reeves EP, Roes J, Segal AW. Catalase negative Staphylococcus aureus retain virulence in mouse model of chronic granulomatous disease. FEBS Lett 2002; 518:107–110 [View Article]
    [Google Scholar]
  146. Horsburgh MJ, Aish JL, White IJ, Shaw L, Lithgow JK et al. SigmaB modulates virulence determinant expression and stress resistance: characterization of a functional rsbU strain derived from Staphylococcus aureus 8325-4. J Bacteriol 2002; 184:5457–5467 [View Article][PubMed]
    [Google Scholar]
  147. Park B, Nizet V, Liu GY. Role of Staphylococcus aureus catalase in niche competition against Streptococcus pneumoniae. J Bacteriol 2008; 190:2275–2278 [View Article][PubMed]
    [Google Scholar]
  148. Cosgrove K, Coutts G, Jonsson IM, Tarkowski A, Kokai-Kun JF et al. Catalase (KatA) and alkyl hydroperoxide reductase (AhpC) have compensatory roles in peroxide stress resistance and are required for survival, persistence, and nasal colonization in Staphylococcus aureus. J Bacteriol 2007; 189:1025–1035 [View Article][PubMed]
    [Google Scholar]
  149. Klebanoff SJ, Kettle AJ, Rosen H, Winterbourn CC, Nauseef WM. Myeloperoxidase: a front-line defender against phagocytosed microorganisms. J Leukoc Biol 2013; 93:185–198 [View Article][PubMed]
    [Google Scholar]
  150. Hirche TO, Gaut JP, Heinecke JW, Belaaouaj A. Myeloperoxidase plays critical roles in killing Klebsiella pneumoniae and inactivating neutrophil elastase: effects on host defense. J Immunol 2005; 174:1557–1565 [View Article][PubMed]
    [Google Scholar]
  151. Müller A, Langklotz S, Lupilova N, Kuhlmann K, Bandow JE et al. Activation of RidA chaperone function by N-chlorination. Nat Commun 2014; 5:5804 [View Article]
    [Google Scholar]
  152. Lehrer RI, Cline MJ. Leukocyte myeloperoxidase deficiency and disseminated candidiasis: the role of myeloperoxidase in resistance to Candida infection. J Clin Invest 1969; 48:1478–1488 [View Article]
    [Google Scholar]
  153. De Jong NWM, Ramyar KX, Guerra FE, Nijland R, Fevre C et al. Immune evasion by a staphylococcal inhibitor of myeloperoxidase. Proc Natl Acad Sci USA 2017; 114:9439–9444 [View Article][PubMed]
    [Google Scholar]
  154. Nathan C, Shiloh MU. Reactive oxygen and nitrogen intermediates in the relationship between mammalian hosts and microbial pathogens. Proc Natl Acad Sci USA 2000; 97:8841–8848 [View Article][PubMed]
    [Google Scholar]
  155. Vazquez-Torres A, Jones-Carson J, Mastroeni P, Ischiropoulos H, Fang FC. Antimicrobial actions of the NADPH phagocyte oxidase and inducible nitric oxide synthase in experimental salmonellosis. I. Effects on microbial killing by activated peritoneal macrophages in vitro. J Exp Med 2000; 192:227–236 [View Article][PubMed]
    [Google Scholar]
  156. Bonamore A, Boffi A. Flavohemoglobin: structure and reactivity. IUBMB Life 2008; 60:19–28 [View Article]
    [Google Scholar]
  157. Kim SO, Orii Y, Lloyd D, Hughes MN, Poole RK. Anoxic function for the Escherichia coli flavohaemoglobin (Hmp): reversible binding of nitric oxide and reduction to nitrous oxide. FEBS Lett 1999; 445:389–394 [View Article]
    [Google Scholar]
  158. Bonamore A, Gentili P, Ilari A, Schininà ME, Boffi A. Escherichia coli flavohemoglobin is an efficient alkylhydroperoxide reductase. J Biol Chem 2003; 278:22272–22277 [View Article][PubMed]
    [Google Scholar]
  159. Gardner PR, Gardner AM, Martin LA, Salzman AL. Nitric oxide dioxygenase: an enzymic function for flavohemoglobin. Proc Natl Acad Sci USA 1998; 95:10378–10383 [View Article][PubMed]
    [Google Scholar]
  160. Gonçalves VL, Nobre LS, Vicente JB, Teixeira M, Saraiva LM. Flavohemoglobin requires microaerophilic conditions for nitrosative protection of Staphylococcus aureus. FEBS Lett 2006; 580:1817–1821 [View Article]
    [Google Scholar]
  161. Richardson AR, Dunman PM, Fang FC. The nitrosative stress response of Staphylococcus aureus is required for resistance to innate immunity. Mol Microbiol 2006; 61:927–939 [View Article]
    [Google Scholar]
  162. Richardson EJ, Bacigalupe R, Harrison EM, Weinert LA, Lycett S et al. Gene exchange drives the ecological success of a multi-host bacterial pathogen. Nat Ecol Evol 2018; 2:1468–1478 [View Article]
    [Google Scholar]
  163. Kloos WE. Natural populations of the genus Staphylococcus. Annu Rev Microbiol 1980; 34:559–592 [View Article]
    [Google Scholar]
  164. Fitzgerald JR. Livestock-associated Staphylococcus aureus: origin, evolution and public health threat. Trends Microbiol 2012; 20:192–198 [View Article][PubMed]
    [Google Scholar]
  165. Bradley A. Bovine mastitis: an evolving disease. Vet J 2002; 164:116–128 [View Article][PubMed]
    [Google Scholar]
  166. Menzies PI, Ramanoon SZ. Mastitis of sheep and goats. Vet Clin North Am Food Anim Pract 2001; 17:333–358 [View Article][PubMed]
    [Google Scholar]
  167. McNamee PT, Smyth JA. Bacterial chondronecrosis with osteomyelitis ('femoral head necrosis') of broiler chickens: a review. Avian Pathol 2000; 29:477–495 [View Article][PubMed]
    [Google Scholar]
  168. Vancraeynest D, Haesebrouck F, Deplano A, Denis O, Godard C et al. International dissemination of a high virulence rabbit Staphylococcus aureus clone. J Vet Med B 2006; 53:418–422 [View Article]
    [Google Scholar]
  169. Jamrozy DM, Fielder MD, Butaye P, Coldham NG. Comparative genotypic and phenotypic characterisation of methicillin-resistant Staphylococcus aureus ST398 isolated from animals and humans. PLoS One 2012; 7:e40458 [View Article]
    [Google Scholar]
  170. Lowder BV, Guinane CM, Ben Zakour NL, Weinert LA, Conway-Morris A et al. Recent human-to-poultry host jump, adaptation, and pandemic spread of Staphylococcus aureus. Proc Natl Acad Sci USA 2009; 106:19545–19550
    [Google Scholar]
  171. Price LB, Stegger M, Hasman H, Aziz M, Larsen J et al. Staphylococcus aureus CC398: host adaptation and emergence of methicillin resistance in livestock. mBio 2012; 3:1–7 [View Article]
    [Google Scholar]
  172. Herron-Olson L, Fitzgerald JR, Musser JM, Kapur V. Molecular correlates of host specialization in Staphylococcus aureus. PLoS One 2007; 2:e1120 [View Article][PubMed]
    [Google Scholar]
  173. Devriese LA. A simplified system for biotyping Staphylococcus aureus strains isolated from different animal species. J Appl Bacteriol 1984; 56:215–220 [View Article]
    [Google Scholar]
  174. Atkins KL, Burman JD, Chamberlain ES, Cooper JE, Poutrel B et al. S. aureus IgG-binding proteins SpA and Sbi: host specificity and mechanisms of immune complex formation. Mol Immunol 2008; 45:1600–1611 [View Article]
    [Google Scholar]
  175. Guinane CM, Ben Zakour NL, Tormo-Mas MA, Weinert LA, Lowder BV et al. Evolutionary genomics of Staphylococcus aureus reveals insights into the origin and molecular basis of ruminant host adaptation. Genome Biol Evol 2010; 2:454–466 [View Article][PubMed]
    [Google Scholar]
  176. Viana D, Blanco J, Tormo-Más MA, Selva L, Guinane CM et al. Adaptation of Staphylococcus aureus to ruminant and equine hosts involves SaPI-carried variants of von Willebrand factor-binding protein. Mol Microbiol 2010; 77:1583–1594 [View Article][PubMed]
    [Google Scholar]
  177. Vrieling M, Koymans KJ, Heesterbeek DAC, Aerts PC, Rutten V et al. Bovine Staphylococcus aureus secretes the Leukocidin LukMF′ to kill migrating neutrophils through CCR1. MBio 2015; 6:e00335 [View Article][PubMed]
    [Google Scholar]
  178. Koop G, Vrieling M, Storisteanu DM, Lok LS, Monie T et al. Identification of LukPQ, a novel, equid-adapted leukocidin of Staphylococcus aureus. Sci Rep 2017; 7:40660 [View Article][PubMed]
    [Google Scholar]
  179. Cassidy P, Harshman S. Studies on the binding of staphylococcal 125I-labeled alpha-toxin to rabbit erythrocytes. Biochemistry 1976; 15:2348–2355 [View Article]
    [Google Scholar]
  180. Takeuchi S, Matsunaga K, Inubushi S, Higuchi H, Imaizumi K et al. Structural gene and strain specificity of a novel cysteine protease produced by Staphylococcus aureus isolated from a diseased chicken. Vet Microbiol 2002; 89:201–210 [View Article]
    [Google Scholar]
  181. De Jong NWM, Vrieling M, Garcia BL, Koop G, Brettmann M et al. Identification of a staphylococcal complement inhibitor with broad host specificity in equid Staphylococcus aureus strains. J Biol Chem 2018; 293:4468–4477 [View Article][PubMed]
    [Google Scholar]
  182. Weinert LA, Welch JJ, Suchard MA, Lemey P, Rambaut A et al. Molecular dating of human-to-bovid host jumps by Staphylococcus aureus reveals an association with the spread of domestication. Biol Lett 2012; 8:829–832 [View Article]
    [Google Scholar]
  183. Koymans KJ, Vrieling M, Gorham RD, van Strijp JAG. Staphylococcal immune evasion proteins: structure, function, and host adaptation. Curr Top Microbiol Immunol 2017; 409:441–489
    [Google Scholar]
  184. Coleman DC, Arbuthnott JP, Pomeroy HM, Birkbeck TH. Cloning and expression in Escherichia coli and Staphylococcus aureus of the beta-lysin determinant from Staphylococcus aureus: evidence that bacteriophage conversion of beta-lysin activity is caused by insertional inactivation of the beta-lysin determinant. Microb Pathog 1986; 1:549–564 [View Article]
    [Google Scholar]
  185. Dohlsten M, Björklund M, Sundstedt A, Hedlund G, Samson D et al. Immunopharmacology of the superantigen staphylococcal enterotoxin A in T-cell receptor V beta 3 transgenic mice. Immunology 1993; 79:520–527[PubMed]
    [Google Scholar]
  186. Gladysheva IP, Turner RB, Sazonova IY, Liu L, Reed GL. Coevolutionary patterns in plasminogen activation. Proc Natl Acad Sci USA 2003; 100:9168–9172 [View Article][PubMed]
    [Google Scholar]
  187. Golonka E, Filipek R, Sabat A, Sinczak A, Potempa J. Genetic characterization of staphopain genes in Staphylococcus aureus. Biol Chem 2004; 385:1059–1067 [View Article]
    [Google Scholar]
  188. Spaan AN, van Strijp JAG, Torres VJ. Leukocidins: staphylococcal bi-component pore-forming toxins find their receptors. Nat Rev Microbiol 2017; 15:435–447 [View Article][PubMed]
    [Google Scholar]
  189. Davis SL, Gurusiddappa S, McCrea KW, Perkins S, Höök M. SdrG, a fibrinogen-binding bacterial adhesin of the microbial surface components recognizing adhesive matrix molecules subfamily from Staphylococcus epidermidis, targets the thrombin cleavage site in the Bbeta chain. J Biol Chem 2001; 276:27799–27805 [View Article][PubMed]
    [Google Scholar]
  190. Pishchany G, McCoy AL, Torres VJ, Krause JC, Crowe JE et al. Specificity for human hemoglobin enhances Staphylococcus aureus infection. Cell Host Microbe 2010; 8:544–550 [View Article][PubMed]
    [Google Scholar]
  191. von Eiff C, Friedrich AW, Peters G, Becker K. Prevalence of genes encoding for members of the staphylococcal leukotoxin family among clinical isolates of Staphylococcus aureus. Diagn Microbiol Infect Dis 2004; 49:157–162 [View Article][PubMed]
    [Google Scholar]
  192. Naimi TS, Ledell KH, Como-Sabetti K, Borchardt SM, Boxrud DJ et al. Comparison of community- and health care-associated methicillin-resistant Staphylococcus aureus infection. JAMA 2003; 290:2976–2984 [View Article][PubMed]
    [Google Scholar]
  193. Viana D, Comos M, McAdam PR, Ward MJ, Selva L et al. A single natural nucleotide mutation alters bacterial pathogen host tropism. Nat Genet 2015; 47:361–366 [View Article][PubMed]
    [Google Scholar]
  194. Salgado-Pabón W, Schlievert PM. Models matter: the search for an effective Staphylococcus aureus vaccine. Nat Rev Microbiol 2014; 12:585–591 [View Article][PubMed]
    [Google Scholar]
  195. Ishikawa F, Yasukawa M, Lyons B, Yoshida S, Miyamoto T et al. Development of functional human blood and immune systems in NOD/SCID/IL2 receptor {gamma} chain(null) mice. Blood 2005; 106:1565–1573 [View Article][PubMed]
    [Google Scholar]
  196. Prince A, Wang H, Kitur K, Parker D. Humanized mice exhibit increased susceptibility to Staphylococcus aureus Pneumonia. J Infect Dis 2017; 215:jiw425 [View Article][PubMed]
    [Google Scholar]
  197. Tseng CW, Biancotti JC, Berg BL, Gate D, Kolar SL et al. Increased susceptibility of humanized NSG mice to panton-valentine leukocidin and Staphylococcus aureus skin infection. PLoS Pathog 2015; 11:e1005292 [View Article][PubMed]
    [Google Scholar]
  198. Knop J, Hanses F, Leist T, Archin NM, Buchholz S et al. Staphylococcus aureus infection in humanized mice: a new model to study pathogenicity associated with human immune response. J Infect Dis 2015; 212:435–444 [View Article][PubMed]
    [Google Scholar]
  199. Lee H, Zahra D, Vogelzang A, Newton R, Thatcher J et al. Human C5aR knock-in mice facilitate the production and assessment of anti-inflammatory monoclonal antibodies. Nat Biotechnol 2006; 24:1279–1284 [View Article][PubMed]
    [Google Scholar]
  200. Prajsnar TK, Cunliffe VT, Foster SJ, Renshaw SA. A novel vertebrate model of Staphylococcus aureus infection reveals phagocyte-dependent resistance of zebrafish to non-host specialized pathogens. Cell Microbiol 2008; 10:2312–2325 [View Article][PubMed]
    [Google Scholar]
  201. Prajsnar TK, Hamilton R, Garcia-Lara J, McVicker G, Williams A et al. A privileged intraphagocyte niche is responsible for disseminated infection of Staphylococcus aureus in a zebrafish model. Cell Microbiol 2012; 14:1600–1619 [View Article]
    [Google Scholar]
  202. McVicker G, Prajsnar TK, Williams A, Wagner NL, Boots M et al. Clonal expansion during Staphylococcus aureus infection dynamics reveals the effect of antibiotic intervention. PLoS Pathog 2014; 10:e1003959 [View Article][PubMed]
    [Google Scholar]
  203. Connolly J, Boldock E, Prince LR, Renshaw SA, Whyte MK et al. Identification of Staphylococcus aureus factors required for pathogenicity and growth in human blood. Infect Immun 2017; 85:IAI.00337–17 [View Article][PubMed]
    [Google Scholar]
http://instance.metastore.ingenta.com/content/journal/micro/10.1099/mic.0.000759
Loading
/content/journal/micro/10.1099/mic.0.000759
Loading

Data & Media loading...

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