Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Review Article
  • Published:

Salmonellae interactions with host processes

Key Points

  • Salmonella spp. deliver effector proteins into host cells to promote replication and survival.

  • Effector proteins that are translocated by the Salmonella pathogenicity island 1 (SPI-1) type III secretion system (T3SS) are important for bacterial invasion into non-phagocytic cells.

  • Effector proteins that are delivered by the SPI-2 T3SS modify the Salmonella-containing vacuole, associated endosomal membranes and associated proteins, all of which promote intracellular replication.

  • Induction of inflammation enhances extracellular growth of salmonellae and enables them to outcompete the gut microbiota.

  • Interplay between the host response pathways of autophagy and pyroptosis is involved in the detection of intracellular Salmonella spp.

  • Distinct functions for many of the Salmonella effector proteins are not fully understood, and it is likely that many of their functions will only be elucidated when their activities are studied in the context of other effectors and are considered in a spatiotemporal context within the host.

Abstract

Salmonellae invasion and intracellular replication within host cells result in a range of diseases, including gastroenteritis, bacteraemia, enteric fever and focal infections. In recent years, considerable progress has been made in our understanding of the molecular mechanisms that salmonellae use to alter host cell physiology; through the delivery of effector proteins with specific activities and through the modulation of defence and stress response pathways. In this Review, we summarize our current knowledge of the complex interplay between bacterial and host factors that leads to inflammation, disease and, in most cases, control of the infection by its animal hosts, with a particular focus on Salmonella enterica subsp. enterica serovar Typhimurium. We also highlight gaps in our knowledge of the contributions of salmonellae and the host to disease pathogenesis, and we suggest future avenues for further study.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Figure 1: Host pathways manipulated by Salmonella spp. during epithelial cell invasion.
Figure 2: Salmonella-induced inflammation promotes pathogen transmission.
Figure 3: Salmonella spp. manipulation of host membranes.
Figure 4: Activation of autophagy and the inflammasome in response to infection with Salmonella spp.

Similar content being viewed by others

References

  1. Pegues, D. A. & Miller, S. I. Principles and Practice of Infectious Diseases 8th edn Vol. 2, 2559–2568 (Elsevier/Churchill Livingstone, 2005).

    Google Scholar 

  2. Harris, J. C., Dupont, H. L. & Hornick, R. B. Fecal leukocytes in diarrheal illness. Ann. Intern. Med. 76, 697–703 (1972).

    Article  CAS  PubMed  Google Scholar 

  3. Winter, S. E. et al. Gut inflammation provides a respiratory electron acceptor for Salmonella. Nature 467, 426–429 (2010). This paper provides the first molecular evidence that intestinal inflammation promotes the growth of pathogens over the microbiota.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Stecher, B. et al. Salmonella enterica serovar Typhimurium exploits inflammation to compete with the intestinal microbiota. PLoS Biol. 5, e244 (2007).

    Article  CAS  PubMed Central  Google Scholar 

  5. Jones, B. D., Ghori, N. & Falkow, S. Salmonella Typhimurium initiates murine infection by penetrating and destroying the specialized epithelial M cells of the Peyer's patches. J. Exp. Med. 180, 15–23 (1994).

    Article  CAS  PubMed  Google Scholar 

  6. Martinez-Moya, M., de Pedro, M. A., Schwarz, H. & Garcia-del Portillo, F. Inhibition of Salmonella intracellular proliferation by non-phagocytic eucaryotic cells. Res. Microbiol. 149, 309–318 (1998).

    Article  CAS  PubMed  Google Scholar 

  7. Niedergang, F., Sirard, J. C., Blanc, C. T. & Kraehenbuhl, J. P. Entry and survival of Salmonella Typhimurium in dendritic cells and presentation of recombinant antigens do not require macrophage-specific virulence factors. Proc. Natl Acad. Sci. USA 97, 14650–14655 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Dalebroux, Z. D. & Miller, S. I. Salmonellae PhoPQ regulation of the outer membrane to resist innate immunity. Curr. Opin. Microbiol. 17, 106–113 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Haraga, A., Ohlson, M. B. & Miller, S. I. Salmonellae interplay with host cells. Nature Rev. Microbiol. 6, 53–66 (2008).

    Article  CAS  Google Scholar 

  10. McClelland, M. et al. Complete genome sequence of Salmonella enterica serovar Typhimurium LT2. Nature 413, 852–856 (2001).

    Article  CAS  PubMed  Google Scholar 

  11. Steele-Mortimer, O. et al. The invasion-associated type III secretion system of Salmonella enterica serovar Typhimurium is necessary for intracellular proliferation and vacuole biogenesis in epithelial cells. Cell. Microbiol. 4, 43–54 (2002).

    Article  CAS  PubMed  Google Scholar 

  12. Giacomodonato, M. N. et al. SipA, SopA, SopB, SopD and SopE2 effector proteins of Salmonella enterica serovar Typhimurium are synthesized at late stages of infection in mice. Microbiology 153, 1221–1228 (2007).

    Article  CAS  PubMed  Google Scholar 

  13. Hernandez, L. D., Hueffer, K., Wenk, M. R. & Galán, J. E. Salmonella modulates vesicular traffic by altering phosphoinositide metabolism. Science 304, 1805–1807 (2004).

    Article  CAS  PubMed  Google Scholar 

  14. Brawn, L. C., Hayward, R. D. & Koronakis, V. Salmonella SPI1 effector SipA persists after entry and cooperates with a SPI2 effector to regulate phagosome maturation and intracellular replication. Cell Host Microbe 1, 63–75 (2007). This study provides the first evidence that SPI-1 effectors can persist to influence the SCV.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Humphreys, D., Hume, P. J. & Koronakis, V. The Salmonella effector SptP dephosphorylates host AAA+ ATPase VCP to promote development of its intracellular replicative niche. Cell Host Microbe 5, 225–233 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Patel, J. C., Hueffer, K., Lam, T. T. & Galán, J. E. Diversification of a Salmonella virulence protein function by ubiquitin-dependent differential localization. Cell 137, 283–294 (2009). This paper demonstrates that the multiple functions of SopB are controlled by ubiquitylation.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Alpuche-Aranda, C. M., Racoosin, E. L., Swanson, J. A. & Miller, S. I. Salmonella stimulate macrophage macropinocytosis and persist within spacious phagosomes. J. Exp. Med. 179, 601–608 (1994).

    Article  CAS  PubMed  Google Scholar 

  18. Oh, Y. K. et al. Rapid and complete fusion of macrophage lysosomes with phagosomes containing Salmonella Typhimurium. Infect. Immun. 64, 3877–3883 (1996).

    CAS  PubMed  PubMed Central  Google Scholar 

  19. Norris, F. A., Wilson, M. P., Wallis, T. S., Galyov, E. E. & Majerus, P. W. SopB, a protein required for virulence of Salmonella Dublin, is an inositol phosphate phosphatase. Proc. Natl Acad. Sci. USA 95, 14057–14059 (1998).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Tahoun, A. et al. Salmonella transforms follicle-associated epithelial cells into M cells to promote intestinal invasion. Cell Host Microbe 12, 645–656 (2012). This study shows that SopB can induce transformation of epithelial cells into M cells.

    Article  CAS  PubMed  Google Scholar 

  21. Bakowski, M. A. et al. The phosphoinositide phosphatase SopB manipulates membrane surface charge and trafficking of the Salmonella-containing vacuole. Cell Host Microbe 7, 453–462 (2010).

    Article  CAS  PubMed  Google Scholar 

  22. Mirold, S. et al. Salmonella host cell invasion emerged by acquisition of a mosaic of separate genetic elements, including Salmonella pathogenicity island 1 (SPI1), SPI5, and SopE2. J. Bacteriol. 183, 2348–2358 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Zhou, D., Chen, L.-M., Hernandez, L., Shears, S. B. & Galán, J. E. A Salmonella inositol polyphosphatase acts in conjunction with other bacterial effectors to promote host cell actin cytoskeleton rearrangements and bacterial internalization. Mol. Microbiol. 39, 248–260 (2001).

    Article  CAS  PubMed  Google Scholar 

  24. Hanisch, J. et al. Activation of a RhoA/myosin II-dependent but Arp2/3 complex-independent pathway facilitates Salmonella invasion. Cell Host Microbe 9, 273–285 (2011).

    Article  CAS  PubMed  Google Scholar 

  25. Jolly, C., Winfree, S., Hansen, B. & Steele-Mortimer, O. The annexin A2/p11 complex is required for efficient invasion of Salmonella Typhimurium in epithelial cells. Cell. Microbiol. 16, 64–77 (2014).

    Article  CAS  PubMed  Google Scholar 

  26. Mallo, G. V. et al. SopB promotes phosphatidylinositol 3-phosphate formation on Salmonella vacuoles by recruiting Rab5 and Vps34. J. Cell Biol. 182, 741–752 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Knodler, L. A., Winfree, S., Drecktrah, D., Ireland, R. & Steele-Mortimer, O. Ubiquitination of the bacterial inositol phosphatase, SopB, regulates its biological activity at the plasma membrane. Cell. Microbiol. 11, 1652–1670 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Stender, S. et al. Identification of SopE2 from Salmonella Typhimurium, a conserved guanine nucleotide exchange factor for Cdc42 of the host cell. Mol. Microbiol. 36, 1206–1221 (2000).

    Article  CAS  PubMed  Google Scholar 

  29. Friebel, A. et al. SopE and SopE2 from Salmonella Typhimurium activate different sets of RhoGTPases of the host cell. J. Biol. Chem. 276, 34035–34040 (2001).

    Article  CAS  PubMed  Google Scholar 

  30. Hardt, W. D., Chen, L. M., Schuebel, K. E., Bustelo, X. R. & Galán, J. E. S. Typhimurium encodes an activator of Rho GTPases that induces membrane ruffling and nuclear responses in host cells. Cell 93, 815–826 (1998).

    Article  CAS  PubMed  Google Scholar 

  31. Wood, M. W., Rosqvist, R., Mullan, P. B., Edwards, M. H. & Galyov, E. E. SopE, a secreted protein of Salmonella Dublin, is translocated into the target eukaryotic cell via a sip-dependent mechanism and promotes bacterial entry. Mol. Microbiol. 22, 327–338 (1996).

    Article  CAS  PubMed  Google Scholar 

  32. Bakshi, C. S. et al. Identification of SopE2, a Salmonella secreted protein which is highly homologous to SopE and involved in bacterial invasion of epithelial cells. J. Bacteriol. 182, 2341–2344 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Lebensohn, A. M. & Kirschner, M. W. Activation of the WAVE complex by coincident signals controls actin assembly. Mol. Cell 36, 512–524 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Humphreys, D., Davidson, A., Hume, P. J. & Koronakis, V. Salmonella virulence effector SopE and host GEF ARNO cooperate to recruit and activate WAVE to trigger bacterial invasion. Cell Host Microbe 11, 129–139 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Humphreys, D., Davidson, A. C., Hume, P. J., Makin, L. E. & Koronakis, V. Arf6 coordinates actin assembly through the WAVE complex, a mechanism usurped by Salmonella to invade host cells. Proc. Natl Acad. Sci. USA 110, 16880–16885 (2013). References 34 and 35 demonstrate the molecular mechanisms that are involved in Salmonella -induced cell invasion.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Zhou, D., Mooseker, M. S. & Galán, J. E. Role of the S. Typhimurium actin-binding protein SipA in bacterial internalization. Science 283, 2092–2095 (1999).

    Article  CAS  PubMed  Google Scholar 

  37. Zhou, D., Mooseker, M. S. & Galán, J. E. An invasion-associated Salmonella protein modulates the actin-bundling activity of plastin. Proc. Natl Acad. Sci. USA 96, 10176–10181 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Hayward, R. D. & Koronakis, V. Direct nucleation and bundling of actin by the SipC protein of invasive Salmonella. EMBO J. 18, 4926–4934 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Myeni, S. K. & Zhou, D. The C terminus of SipC binds and bundles F-actin to promote Salmonella invasion. J. Biol. Chem. 285, 13357–13363 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Scherer, C. A., Cooper, E. & Miller, S. I. The Salmonella type III secretion translocon protein SspC is inserted into the epithelial cell plasma membrane upon infection. Mol. Microbiol. 37, 1133–1145 (2000).

    Article  CAS  PubMed  Google Scholar 

  41. Carlson, S. A., Omary, M. B. & Jones, B. D. Identification of cytokeratins as accessory mediators of Salmonella entry into eukaryotic cells. Life Sci. 70, 1415–1426 (2002).

    Article  CAS  PubMed  Google Scholar 

  42. Fu, Y. & Galán, J. E. A Salmonella protein antagonizes Rac-1 and Cdc42 to mediate host-cell recovery after bacterial invasion. Nature 401, 293–297 (1999).

    Article  CAS  PubMed  Google Scholar 

  43. Kubori, T. & Galán, J. E. Temporal regulation of Salmonella virulence effector function by proteasome-dependent protein degradation. Cell 115, 333–342 (2003).

    Article  CAS  PubMed  Google Scholar 

  44. Zhang, S. et al. The Salmonella enterica serotype Typhimurium effector proteins SipA, SopA, SopB, SopD, and SopE2 act in concert to induce diarrhea in calves. Infect. Immun. 70, 3843–3855 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Hapfelmeier, S. et al. Role of the Salmonella pathogenicity island 1 effector proteins SipA, SopB, SopE, and SopE2 in Salmonella enterica subspecies 1 serovar Typhimurium colitis in streptomycin-pretreated mice. Infect. Immun. 72, 795–809 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Bhavsar, A. P. et al. The Salmonella type III effector SspH2 specifically exploits the NLR co-chaperone activity of SGT1 to subvert immunity. PLoS Pathog. 9, e1003518 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. McCormick, B. A., Miller, S. I., Carnes, D. & Madara, J. L. Transepithelial signaling to neutrophils by salmonellae: a novel virulence mechanism for gastroenteritis. Infect. Immun. 63, 2302–2309 (1995).

    CAS  PubMed  PubMed Central  Google Scholar 

  48. McCormick, B. A. et al. Surface attachment of Salmonella Typhimurium to intestinal epithelia imprints the subepithelial matrix with gradients chemotactic for neutrophils. J. Cell Biol. 131, 1599–1608 (1995).

    Article  CAS  PubMed  Google Scholar 

  49. Arpaia, N. et al. TLR signaling is required for Salmonella Typhimurium virulence. Cell 144, 675–688 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Knodler, L. A. et al. Noncanonical inflammasome activation of caspase-4/caspase-11 mediates epithelial defenses against enteric bacterial pathogens. Cell Host Microbe 16, 249–256 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Miao, E. A. et al. Innate immune detection of the type III secretion apparatus through the NLRC4 inflammasome. Proc. Natl Acad. Sci. USA 107, 3076–3080 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  52. Miao, E. A. et al. Cytoplasmic flagellin activates caspase-1 and secretion of interleukin 1β via Ipaf. Nature Immunol. 7, 569–575 (2006).

    Article  CAS  Google Scholar 

  53. Godinez, I. et al. T cells help to amplify inflammatory responses induced by Salmonella enterica serotype Typhimurium in the intestinal mucosa. Infect. Immun. 76, 2008–2017 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Hersh, D. et al. The Salmonella invasin SipB induces macrophage apoptosis by binding to caspase-1. Proc. Natl Acad. Sci. USA 96, 2396–2401 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Müller, A. J. et al. The S. Typhimurium effector SopE induces caspase-1 activation in stromal cells to initiate gut inflammation. Cell Host Microbe 6, 125–136 (2009).

    Article  CAS  PubMed  Google Scholar 

  56. Shi, J. et al. Inflammatory caspases are innate immune receptors for intracellular LPS. Nature 514, 187–192 (2014). This study identifies caspase 4, caspase 5 and caspase 11 as receptors for intracellular LPS.

    Article  CAS  PubMed  Google Scholar 

  57. Lee, C. A. et al. A secreted Salmonella protein induces a proinflammatory response in epithelial cells, which promotes neutrophil migration. Proc. Natl Acad. Sci. USA 97, 12283–12288 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Srikanth, C. V. et al. Salmonella pathogenesis and processing of secreted effectors by caspase-3. Science 330, 390–393 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Raffatellu, M. et al. Lipocalin-2 resistance confers an advantage to Salmonella enterica serotype Typhimurium for growth and survival in the inflamed intestine. Cell Host Microbe 5, 476–486 (2009). This paper demonstrates that the upregulation of inflammation by Salmonella spp. promotes the release of antimicrobial factors that function against the microbiota.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Lopez, C. A. et al. Phage-mediated acquisition of a type III secreted effector protein boosts growth of Salmonella by nitrate respiration. mBio 3, e00143-12 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Lin, S. L., Le, T. X. & Cowen, D. S. SptP, a Salmonella Typhimurium type III-secreted protein, inhibits the mitogen-activated protein kinase pathway by inhibiting Raf activation. Cell. Microbiol. 5, 267–275 (2003).

    Article  CAS  PubMed  Google Scholar 

  62. Murli, S., Watson, R. O. & Galán, J. E. Role of tyrosine kinases and the tyrosine phosphatase SptP in the interaction of Salmonella with host cells. Cell. Microbiol. 3, 795–810 (2001).

    Article  CAS  PubMed  Google Scholar 

  63. Haraga, A. & Miller, S. I. A Salmonella enterica serovar Typhimurium translocated leucine-rich repeat effector protein inhibits NF-κB-dependent gene expression. Infect. Immun. 71, 4052–4058 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Haraga, A. & Miller, S. I. A Salmonella type III secretion effector interacts with the mammalian serine/threonine protein kinase PKN1. Cell. Microbiol. 8, 837–846 (2006).

    Article  CAS  PubMed  Google Scholar 

  65. Keszei, A. F. et al. Structure of an SspH1–PKN1 complex reveals the basis for host substrate recognition and mechanism of activation for a bacterial E3 ubiquitin ligase. Mol. Cell. Biol. 34, 362–373 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Wu, H., Jones, R. M. & Neish, A. S. The Salmonella effector AvrA mediates bacterial intracellular survival during infection in vivo. Cell. Microbiol. 14, 28–39 (2012).

    Article  CAS  PubMed  Google Scholar 

  67. Jones, R. M. et al. Salmonella AvrA coordinates suppression of host immune and apoptotic defenses via JNK pathway blockade. Cell Host Microbe 3, 233–244 (2008).

    Article  CAS  PubMed  Google Scholar 

  68. Du, F. & Galán, J. E. Selective inhibition of type III secretion activated signaling by the Salmonella effector AvrA. PLoS Pathog. 5, e1000595 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Haneda, T. et al. Salmonella type III effector SpvC, a phosphothreonine lyase, contributes to reduction in inflammatory response during intestinal phase of infection. Cell. Microbiol. 14, 485–499 (2012).

    Article  CAS  PubMed  Google Scholar 

  70. Pilar, A. V., Reid-Yu, S. A., Cooper, C. A., Mulder, D. T. & Coombes, B. K. GogB is an anti-inflammatory effector that limits tissue damage during Salmonella infection through interaction with human FBXO22 and Skp1. PLoS Pathog. 8, e1002773 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Mazurkiewicz, P. et al. SpvC is a Salmonella effector with phosphothreonine lyase activity on host mitogen-activated protein kinases. Mol. Microbiol. 67, 1371–1383 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Cardenal-Muñoz, E., Gutiérrez, G. & Ramos-Morales, F. Global impact of Salmonella type III secretion effector SteA on host cells. Biochem. Biophys. Res. Commun. 449, 419–424 (2014).

    Article  CAS  PubMed  Google Scholar 

  73. McLaughlin, L. M. et al. A microfluidic-based genetic screen to identify microbial virulence factors that inhibit dendritic cell migration. Integr. Biol. (Camb.) 6, 438–449 (2014).

    Article  CAS  Google Scholar 

  74. Halici, S., Zenk, S. F., Jantsch, J. & Hensel, M. Functional analysis of the Salmonella pathogenicity island 2-mediated inhibition of antigen presentation in dendritic cells. Infect. Immun. 76, 4924–4933 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Abshire, K. Z. & Neidhardt, F. C. Growth rate paradox of Salmonella Typhimurium within host macrophages. J. Bacteriol. 175, 3744–3748 (1993).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  76. Crawford, R. W. et al. Gallstones play a significant role in Salmonella spp. gallbladder colonization and carriage. Proc. Natl Acad. Sci. USA 107, 4353–4358 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  77. Steele-Mortimer, O., Méresse, S., Gorvel, J.-P., Toh, B.-H. & Finlay, B. B. Biogenesis of Salmonella Typhimurium-containing vacuoles in epithelial cells involves interactions with the early endocytic pathway. Cell. Microbiol. 1, 33–49 (1999).

    Article  CAS  PubMed  Google Scholar 

  78. Smith, A. C., Cirulis, J. T., Casanova, J. E., Scidmore, M. A. & Brumell, J. H. Interaction of the Salmonella-containing vacuole with the endocytic recycling system. J. Biol. Chem. 280, 24634–24641 (2005).

    Article  CAS  PubMed  Google Scholar 

  79. Brumell, J. H., Tang, P., Mills, S. D. & Finlay, B. B. Characterization of Salmonella-induced filaments (Sifs) reveals a delayed interaction between Salmonella-containing vacuoles and late endocytic compartments. Traffic 2, 643–653 (2001).

    Article  CAS  PubMed  Google Scholar 

  80. Garcia-del Portillo, F. & Finlay, B. B. Targeting of Salmonella Typhimurium to vesicles containing lysosomal membrane glycoproteins bypasses compartments with mannose 6-phosphate receptors. J. Cell Biol. 129, 81–97 (1995).

    Article  CAS  PubMed  Google Scholar 

  81. Méresse, S., Steele-Mortimer, O., Finlay, B. B. & Gorvel, J. P. The Rab7 GTPase controls the maturation of Salmonella Typhimurium-containing vacuoles in HeLa cells. EMBO J. 18, 4394–4403 (1999).

    Article  PubMed  PubMed Central  Google Scholar 

  82. Eswarappa, S. M., Negi, V. D., Chakraborty, S., Chandrasekhar Sagar, B. K. & Chakravortty, D. Division of the Salmonella-containing vacuole and depletion of acidic lysosomes in Salmonella-infected host cells are novel strategies of Salmonella enterica to avoid lysosomes. Infect. Immun. 78, 68–79 (2010).

    Article  CAS  PubMed  Google Scholar 

  83. Christen, M. et al. Activation of a bacterial virulence protein by the GTPase RhoA. Sci. Signal. 2, ra71 (2009). This study identifies the first bacterial effector that is activated by a mammalian GTPase.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. LaRock, D. L., Brzovic, P. S., Levin, I., Blanc, M.-P. & Miller, S. I. A Salmonella Typhimurium-translocated glycerophospholipid:cholesterol acyltransferase promotes virulence by binding to the RhoA protein switch regions. J. Biol. Chem. 287, 29654–29663 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  85. Nawabi, P., Catron, D. M. & Haldar, K. Esterification of cholesterol by a type III secretion effector during intracellular Salmonella infection. Mol. Microbiol. 68, 173–185 (2008).

    Article  CAS  PubMed  Google Scholar 

  86. Boucrot, E., Henry, T., Borg, J.-P., Gorvel, J.-P. & Méresse, S. The intracellular fate of Salmonella depends on the recruitment of kinesin. Science 308, 1174–1178 (2005).

    Article  CAS  PubMed  Google Scholar 

  87. Arena, E. T. et al. The deubiquitinase activity of the Salmonella pathogenicity island 2 effector, SseL, prevents accumulation of cellular lipid droplets. Infect. Immun. 79, 4392–4400 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  88. Ramsden, A. E., Mota, L. J., Munter, S., Shorte, S. L. & Holden, D. W. The SPI-2 type III secretion system restricts motility of Salmonella-containing vacuoles. Cell. Microbiol. 9, 2517–2529 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  89. Braun, V. et al. Sorting nexin 3 (SNX3) is a component of a tubular endosomal network induced by Salmonella and involved in maturation of the Salmonella-containing vacuole. Cell. Microbiol. 12, 1352–1367 (2010).

    Article  CAS  PubMed  Google Scholar 

  90. Bujny, M. V. et al. Sorting nexin-1 defines an early phase of Salmonella-containing vacuole-remodeling during Salmonella infection. J. Cell Sci. 121, 2027–2036 (2008).

    Article  CAS  PubMed  Google Scholar 

  91. Cantalupo, G., Alifano, P., Roberti, V., Bruni, C. B. & Bucci, C. Rab-interacting lysosomal protein (RILP): the Rab7 effector required for transport to lysosomes. EMBO J. 20, 683–693 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  92. Harrison, R. E. et al. Salmonella impairs RILP recruitment to Rab7 during maturation of invasion vacuoles. Mol. Biol. Cell 15, 3146–3154 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  93. Guignot, J. et al. Microtubule motors control membrane dynamics of Salmonella-containing vacuoles. J. Cell Sci. 117, 1033–1045 (2004).

    Article  CAS  PubMed  Google Scholar 

  94. Szeto, J., Namolovan, A., Osborne, S. E., Coombes, B. K. & Brumell, J. H. Salmonella-containing vacuoles display centrifugal movement associated with cell-to-cell transfer in epithelial cells. Infect. Immun. 77, 996–1007 (2009).

    Article  CAS  PubMed  Google Scholar 

  95. Ohlson, M. B. et al. Structure and function of Salmonella SifA indicate that its interactions with SKIP, SseJ, and RhoA family GTPases induce endosomal tubulation. Cell Host Microbe 4, 434–446 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  96. Henry, T. et al. The Salmonella effector protein PipB2 is a linker for kinesin-1. Proc. Natl Acad. Sci. USA 103, 13497–13502 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  97. Deiwick, J. et al. The translocated Salmonella effector proteins SseF and SseG interact and are required to establish an intracellular replication niche. Infect. Immun. 74, 6965–6972 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  98. Garcia-del Portillo, F., Zwick, M. B., Leung, K. Y. & Finlay, B. B. Salmonella induces the formation of filamentous structures containing lysosomal membrane glycoproteins in epithelial cells. Proc. Natl Acad. Sci. USA 90, 10544–10548 (1993).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  99. Stein, M., Leung, K., Zwick, M., Garcia-del Portillo, F. & Finlay, B. Identification of a Salmonella virulence gene required for formation of filamentous structures containing lysosomal membrane glycoproteins within epithelial cells. Mol. Microbiol. 20, 151–164 (1996).

    Article  CAS  PubMed  Google Scholar 

  100. Birmingham, C. L., Jiang, X., Ohlson, M. B., Miller, S. I. & Brumell, J. H. Salmonella-induced filament formation is a dynamic phenotype induced by rapidly replicating Salmonella enterica serovar Typhimurium in epithelial cells. Infect. Immun. 73, 1204–1208 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  101. Domingues, L., Holden, D. W. & Mota, L. J. The Salmonella effector SteA contributes to the control of membrane dynamics of Salmonella-containing vacuoles. Infect. Immun. 82, 2923–2934 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  102. Brumell, J. H., Goosney, D. L. & Finlay, B. B. SifA, a type III secreted effector of Salmonella Typhimurium, directs Salmonella-induced filament (Sif) formation along microtubules. Traffic 3, 407–415 (2002).

    Article  CAS  PubMed  Google Scholar 

  103. Knodler, L. A. & Steele-Mortimer, O. The Salmonella effector PipB2 affects late endosome/lysosome distribution to mediate Sif extension. Mol. Biol. Cell 16, 4108–4123 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  104. Guy, R. L., Gonias, L. A. & Stein, M. A. Aggregation of host endosomes by Salmonella requires SPI2 translocation of SseFG and involves SpvR and the fms–aroE intragenic region. Mol. Microbiol. 37, 1417–1435 (2000).

    Article  CAS  PubMed  Google Scholar 

  105. Ruíz-Albert, J. et al. Complementary activities of SseJ and SifA regulate dynamics of the Salmonella Typhimurium vacuolar membrane. Mol. Microbiol. 44, 645–661 (2002).

    Article  PubMed  Google Scholar 

  106. Schroeder, N. et al. The virulence protein SopD2 regulates membrane dynamics of Salmonella-containing vacuoles. PLoS Pathog. 6, e1001002 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  107. Kaniuk, N. A. et al. Salmonella exploits Arl8B-directed kinesin activity to promote endosome tubulation and cell-to-cell transfer. Cell. Microbiol. 13, 1812–1823 (2011).

    Article  CAS  PubMed  Google Scholar 

  108. Boucrot, E., Beuzón, C. R., Holden, D. W., Gorvel, J.-P. & Méresse, S. Salmonella Typhimurium SifA effector protein requires its membrane-anchoring C-terminal hexapeptide for its biological function. J. Biol. Chem. 278, 14196–14202 (2003).

    Article  CAS  PubMed  Google Scholar 

  109. Reinicke, A. T. et al. A Salmonella Typhimurium effector protein SifA is modified by host cell prenylation and S-acylation machinery. J. Biol. Chem. 280, 14620–14627 (2005).

    Article  CAS  PubMed  Google Scholar 

  110. Dumont, A. et al. SKIP, the host target of the Salmonella virulence factor SifA, promotes kinesin-1-dependent vacuolar membrane exchanges. Traffic 11, 899–911 (2010).

    Article  CAS  PubMed  Google Scholar 

  111. McGourty, K. et al. Salmonella inhibits retrograde trafficking of mannose-6-phosphate receptors and lysosome function. Science 338, 963–967 (2012). This study demonstrates that SifA inhibits lysosome–SCV fusion by binding to RAB9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  112. Mota, L. J., Ramsden, A. E., Liu, M., Castle, J. D. & Holden, D. W. SCAMP3 is a component of the Salmonella-induced tubular network and reveals an interaction between bacterial effectors and post-Golgi trafficking. Cell. Microbiol. 11, 1236–1253 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  113. Perrett, C. A. & Zhou, D. Salmonella type III effector SopB modulates host cell exocytosis. Emerg. Microbes Infect. 2, e32 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  114. Jiang, X. et al. The related effector proteins SopD and SopD2 from Salmonella enterica serovar Typhimurium contribute to virulence during systemic infection of mice. Mol. Microbiol. 54, 1186–1198 (2004).

    Article  CAS  PubMed  Google Scholar 

  115. Freeman, J. A., Ohl, M. E. & Miller, S. I. The Salmonella enterica serovar Typhimurium translocated effectors SseJ and SifB are targeted to the Salmonella-containing vacuole. Infect. Immun. 71, 418–427 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  116. Figueira, R., Watson, K. G., Holden, D. W. & Helaine, S. Identification of Salmonella pathogenicity island-2 type III secretion system effectors involved in intramacrophage replication of S. enterica serovar Typhimurium: implications for rational vaccine design. mBio 4, e00065-13 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  117. Beuzón, C. R. et al. Salmonella maintains the integrity of its intracellular vacuole through the action of SifA. EMBO J. 19, 3235–3249 (2000).

    Article  PubMed  PubMed Central  Google Scholar 

  118. Ohlson, M. B., Fluhr, K., Birmingham, C. L., Brumell, J. H. & Miller, S. I. SseJ deacylase activity by Salmonella enterica serovar Typhimurium promotes virulence in mice. Infect. Immun. 73, 6249–6259 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  119. McLaughlin, L. M. et al. The Salmonella SPI2 effector SseI mediates long-term systemic infection by modulating host cell migration. PLoS Pathog. 5, e1000671 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  120. Knodler, L. A. et al. Salmonella type III effectors PipB and PipB2 are targeted to detergent-resistant microdomains on internal host cell membranes. Mol. Microbiol. 49, 685–704 (2003).

    Article  CAS  PubMed  Google Scholar 

  121. Browne, S. H., Hasegawa, P., Okamoto, S., Fierer, J. & Guiney, D. G. Identification of Salmonella SPI-2 secretion system components required for SpvB-mediated cytotoxicity in macrophages and virulence in mice. FEMS Immunol. Med. Microbiol. 52, 194–201 (2008).

    Article  CAS  PubMed  Google Scholar 

  122. Geddes, K., Worley, M., Niemann, G. & Heffron, F. Identification of new secreted effectors in Salmonella enterica serovar Typhimurium. Infect. Immun. 73, 6260–6271 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  123. Hicks, S. W., Charron, G., Hang, H. C. & Galán, J. E. Subcellular targeting of Salmonella virulence proteins by host-mediated S-palmitoylation. Cell Host Microbe 10, 9–20 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  124. Zhang, Y., Higashide, W., Dai, S., Sherman, D. M. & Zhou, D. Recognition and ubiquitination of Salmonella type III effector SopA by a ubiquitin E3 ligase, HsRMA1. J. Biol. Chem. 280, 38682–38688 (2005).

    Article  CAS  PubMed  Google Scholar 

  125. Arbeloa, A. et al. EspM2 is a RhoA guanine nucleotide exchange factor. Cell. Microbiol. 12, 654–664 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  126. Vinh, D. B., Ko, D. C., Rachubinski, R. A., Aitchison, J. D. & Miller, S. I. Expression of the Salmonella spp. virulence factor SifA in yeast alters Rho1 activity on peroxisomes. Mol. Biol. Cell 21, 3567–3577 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  127. Huang, Z. et al. Structural insights into host GTPase isoform selection by a family of bacterial GEF mimics. Nature Struct. Mol. Biol. 16, 853–860 (2009).

    Article  CAS  Google Scholar 

  128. Martinon, F., Chen, X., Lee, A. H. & Glimcher, L. H. TLR activation of the transcription factor XBP1 regulates innate immune responses in macrophages. Nature Immunol. 11, 411–418 (2010).

    Article  CAS  Google Scholar 

  129. Oslowski, C. M. et al. Thioredoxin-interacting protein mediates ER stress-induced β cell death through initiation of the inflammasome. Cell. Metab. 16, 265–273 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  130. Ron, D. & Walter, P. Signal integration in the endoplasmic reticulum unfolded protein response. Nature Rev. Mol. Cell Biol. 8, 519–529 (2007).

    Article  CAS  Google Scholar 

  131. Jia, K. et al. Autophagy genes protect against Salmonella Typhimurium infection and mediate insulin signaling-regulated pathogen resistance. Proc. Natl Acad. Sci. USA 106, 14564–14569 (2009).

    Article  PubMed  PubMed Central  Google Scholar 

  132. Knodler, L. A. et al. Dissemination of invasive Salmonella via bacterial-induced extrusion of mucosal epithelia. Proc. Natl Acad. Sci. USA 107, 17733–17738 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  133. Sellin, M. E. et al. Epithelium-intrinsic NAIP/NLRC4 inflammasome drives infected enterocyte expulsion to restrict Salmonella replication in the intestinal mucosa. Cell Host Microbe 16, 237–248 (2014).

    Article  CAS  PubMed  Google Scholar 

  134. Manzanillo, P. S. et al. The ubiquitin ligase parkin mediates resistance to intracellular pathogens. Nature 501, 512–516 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  135. Benjamin, J. L., Sumpter, R. Jr, Levine, B. & Hooper, L. V. Intestinal epithelial autophagy is essential for host defense against invasive bacteria. Cell Host Microbe 13, 723–734 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  136. Lassen, K. G. et al. Atg16L1 T300A variant decreases selective autophagy resulting in altered cytokine signaling and decreased antibacterial defense. Proc. Natl Acad. Sci. USA 111, 7741–7746 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  137. Kuballa, P., Huett, A., Rioux, J. D., Daly, M. J. & Xavier, R. J. Impaired autophagy of an intracellular pathogen induced by a Crohn's disease associated ATG16L1 variant. PLoS ONE 3, e3391 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  138. Cadwell, K. et al. A key role for autophagy and the autophagy gene Atg16l1 in mouse and human intestinal Paneth cells. Nature 456, 259–263 (2008).

    Article  PubMed  PubMed Central  Google Scholar 

  139. Knodler, L. A., Nair, V. & Steele-Mortimer, O. Quantitative assessment of cytosolic Salmonella in epithelial cells. PLoS ONE 9, e84681 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  140. Cemma, M., Kim, P. K. & Brumell, J. H. The ubiquitin-binding adaptor proteins p62/SQSTM1 and NDP52 are recruited independently to bacteria-associated microdomains to target Salmonella to the autophagy pathway. Autophagy 7, 341–345 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  141. Wild, P. et al. Phosphorylation of the autophagy receptor optineurin restricts Salmonella growth. Science 333, 228–233 (2011). This paper identifies phosphorylation-dependent regulation of the autophagy receptor OPTN.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  142. Fujita, N. et al. Recruitment of the autophagic machinery to endosomes during infection is mediated by ubiquitin. J. Cell Biol. 203, 115–128 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  143. Ponpuak, M. et al. Delivery of cytosolic components by autophagic adaptor protein p62 endows autophagosomes with unique antimicrobial properties. Immunity 32, 329–341 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  144. Randow, F. How cells deploy ubiquitin and autophagy to defend their cytosol from bacterial invasion. Autophagy 7, 304–309 (2011).

    Article  CAS  PubMed  Google Scholar 

  145. Thurston, T. L., Wandel, M. P., von Muhlinen, N., Foeglein, A. & Randow, F. Galectin 8 targets damaged vesicles for autophagy to defend cells against bacterial invasion. Nature 482, 414–418 (2012). This study shows that galectin 8 can detect damaged vesicles in the cytosol to trigger host autophagy.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  146. Shahnazari, S. et al. A diacylglycerol-dependent signaling pathway contributes to regulation of antibacterial autophagy. Cell Host Microbe 8, 137–146 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  147. Tattoli, I. et al. Amino acid starvation induced by invasive bacterial pathogens triggers an innate host defense program. Cell Host Microbe 11, 563–575 (2012).

    Article  CAS  PubMed  Google Scholar 

  148. Owen, K. A., Meyer, C. B., Bouton, A. H. & Casanova, J. E. Activation of focal adhesion kinase by Salmonella suppresses autophagy via an Akt/mTOR signaling pathway and promotes bacterial survival in macrophages. PLoS Pathog. 10, e1004159 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  149. Angosto, D. et al. Evolution of inflammasome functions in vertebrates: inflammasome and caspase-1 trigger fish macrophage cell death but are dispensable for the processing of IL-1β. Innate Immun. 18, 815–824 (2012).

    Article  CAS  PubMed  Google Scholar 

  150. Raupach, B., Peuschel, S. K., Monack, D. M. & Zychlinsky, A. Caspase-1-mediated activation of interleukin-1β (IL-1β) and IL-18 contributes to innate immune defenses against Salmonella enterica serovar Typhimurium infection. Infect. Immun. 74, 4922–4926 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  151. Broz, P. et al. Caspase-11 increases susceptibility to Salmonella infection in the absence of caspase-1. Nature 490, 288–291 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  152. Lara-Tejero, M. et al. Role of the caspase-1 inflammasome in Salmonella Typhimurium pathogenesis. J. Exp. Med. 203, 1407–1412 (2006). References 150 and 152 show that inflammasome activation is important for the pathogenesis of salmonellae.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  153. Broz, P. et al. Redundant roles for inflammasome receptors NLRP3 and NLRC4 in host defense against Salmonella. J. Exp. Med. 207, 1745–1755 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  154. Man, S. M. et al. Inflammasome activation causes dual recruitment of NLRC4 and NLRP3 to the same macromolecular complex. Proc. Natl Acad. Sci. USA 111, 7403–7408 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  155. Miao, E. A. & Warren, S. E. Innate immune detection of bacterial virulence factors via the NLRC4 inflammasome. J. Clin. Immunol. 30, 502–506 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  156. Proell, M., Gerlic, M., Mace, P. D., Reed, J. C. & Riedl, S. J. The CARD plays a critical role in ASC foci formation and inflammasome signalling. Biochem. J. 449, 613–621 (2013).

    Article  CAS  PubMed  Google Scholar 

  157. Wynosky-Dolfi, M. A. et al. Oxidative metabolism enables Salmonella evasion of the NLRP3 inflammasome. J. Exp. Med. 211, 653–668 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  158. Iyer, S. S. et al. Mitochondrial cardiolipin is required for Nlrp3 inflammasome activation. Immunity 39, 311–323 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  159. Dalebroux, Z. D., Matamouros, S., Whittington, D., Bishop, R. E. & Miller, S. I. PhoPQ regulates acidic glycerophospholipid content of the Salmonella Typhimurium outer membrane. Proc. Natl Acad. Sci. USA 111, 1963–1968 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  160. Franchi, L. et al. Cytosolic flagellin requires Ipaf for activation of caspase-1 and interleukin 1β in Salmonella-infected macrophages. Nature Immunol. 7, 576–582 (2006).

    Article  CAS  Google Scholar 

  161. Lightfield, K. L. et al. Critical function for Naip5 in inflammasome activation by a conserved carboxy-terminal domain of flagellin. Nature Immunol. 9, 1171–1178 (2008).

    Article  CAS  Google Scholar 

  162. Mariathasan, S. et al. Differential activation of the inflammasome by caspase-1 adaptors ASC and Ipaf. Nature 430, 213–218 (2004).

    Article  CAS  PubMed  Google Scholar 

  163. Franchi, L. et al. NLRC4-driven production of IL-1β discriminates between pathogenic and commensal bacteria and promotes host intestinal defense. Nature Immunol. 13, 449–456 (2012).

    Article  CAS  Google Scholar 

  164. Zhong, D. et al. The Salmonella type III secretion system inner rod protein PrgJ is partially folded. J. Biol. Chem. 287, 25303–25311 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  165. Zhao, Y. et al. The NLRC4 inflammasome receptors for bacterial flagellin and type III secretion apparatus. Nature 477, 596–600 (2011).

    Article  CAS  PubMed  Google Scholar 

  166. Kofoed, E. M. & Vance, R. E. Innate immune recognition of bacterial ligands by NAIPs determines inflammasome specificity. Nature 477, 592–595 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  167. Tenthorey, J. L., Kofoed, E. M., Daugherty, M. D., Malik, H. S. & Vance, R. E. Molecular basis for specific recognition of bacterial ligands by NAIP/NLRC4 inflammasomes. Mol. Cell 54, 17–29 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  168. Yang, J., Zhao, Y., Shi, J. & Shao, F. Human NAIP and mouse NAIP1 recognize bacterial type III secretion needle protein for inflammasome activation. Proc. Natl Acad. Sci. USA 110, 14408–14413 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  169. Romanish, M. T., Nakamura, H., Lai, C. B., Wang, Y. & Mager, D. L. A novel protein isoform of the multicopy human NAIP gene derives from intragenic Alu SINE promoters. PLoS ONE 4, e5761 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  170. Boniotto, M. et al. Population variation in NAIP functional copy number confers increased cell death upon Legionella pneumophila infection. Hum. Immunol. 73, 196–200 (2012).

    Article  CAS  PubMed  Google Scholar 

  171. Cummings, L. A., Wilkerson, W. D., Bergsbaken, T. & Cookson, B. T. In vivo, fliC expression by Salmonella enterica serovar Typhimurium is heterogeneous, regulated by ClpX, and anatomically restricted. Mol. Microbiol. 61, 795–809 (2006).

    Article  CAS  PubMed  Google Scholar 

  172. Kayagaki, N. et al. Noncanonical inflammasome activation by intracellular LPS independent of TLR4. Science 341, 1246–1249 (2013).

    Article  CAS  PubMed  Google Scholar 

  173. Aachoui, Y. et al. Caspase-11 protects against bacteria that escape the vacuole. Science 339, 975–978 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  174. Akhter, A. et al. Caspase-11 promotes the fusion of phagosomes harboring pathogenic bacteria with lysosomes by modulating actin polymerization. Immunity 37, 35–47 (2012). References 173 and 174 show that caspase 11 is activated by intracellular LPS and protects against cytosolic bacteria.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  175. Hagar, J. A., Powell, D. A., Aachoui, Y., Ernst, R. K. & Miao, E. A. Cytoplasmic LPS activates caspase-11: implications in TLR4 independent endotoxic shock. Science 341, 1250–1253 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  176. Into, T., Inomata, M., Takayama, E. & Takigawa, T. Autophagy in regulation of Toll-like receptor signaling. Cell. Signal. 24, 1150–1162 (2012).

    Article  CAS  PubMed  Google Scholar 

  177. Deretic, V., Saitoh, T. & Akira, S. Autophagy in infection, inflammation and immunity. Nature Rev. Immunol. 13, 722–737 (2013).

    Article  CAS  Google Scholar 

  178. Shi, C. S. & Kehrl, J. H. MyD88 and Trif target beclin 1 to trigger autophagy in macrophages. J. Biol. Chem. 283, 33175–33182 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  179. Keestra, A. M. et al. A Salmonella virulence factor activates the NOD1/NOD2 signaling pathway. mBio 2, e00266-11 (2011).

    Article  CAS  PubMed  Google Scholar 

  180. Travassos, L. H. et al. Nod1 and Nod2 direct autophagy by recruiting ATG16L1 to the plasma membrane at the site of bacterial entry. Nature Immunol. 11, 55–62 (2010).

    Article  CAS  Google Scholar 

  181. Saitoh, T. et al. Loss of the autophagy protein Atg16L1 enhances endotoxin-induced IL-1β production. Nature 456, 264–268 (2008).

    Article  CAS  PubMed  Google Scholar 

  182. Shi, C. S. et al. Activation of autophagy by inflammatory signals limits IL-1β production by targeting ubiquitinated inflammasomes for destruction. Nature Immunol. 13, 255–263 (2012).

    Article  CAS  Google Scholar 

  183. Zhou, R., Yazdi, A. S., Menu, P. & Tschopp, J. A role for mitochondria in NLRP3 inflammasome activation. Nature 469, 221–225 (2011). References 182 and 183 show crosstalk between the autophagy and inflammasome pathways.

    Article  CAS  PubMed  Google Scholar 

  184. Nakahira, K. et al. Autophagy proteins regulate innate immune responses by inhibiting the release of mitochondrial DNA mediated by the NALP3 inflammasome. Nature Immunol. 12, 222–230 (2011).

    Article  CAS  Google Scholar 

  185. Meunier, E. et al. Caspase-11 activation requires lysis of pathogen-containing vacuoles by IFN-induced GTPases. Nature 509, 366–370 (2014).

    Article  CAS  PubMed  Google Scholar 

  186. VanDussen, K. L. et al. Development of an enhanced human gastrointestinal epithelial culture system to facilitate patient-based assays. Gut http://dx.doi.org/10.1136/gutjnl-2013-306651 (2014).

  187. Mathur, R. et al. A mouse model of Salmonella Typhi infection. Cell 151, 590–602 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  188. Song, J. et al. A mouse model for the human pathogen Salmonella Typhi. Cell Host Microbe 8, 369–376 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  189. Libby, S. J. et al. Humanized nonobese diabetic-scid IL2rγnull mice are susceptible to lethal Salmonella Typhi infection. Proc. Natl Acad. Sci. USA 107, 15589–15594 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  190. Núñez-Hernández, C., Alonso, A., Pucciarelli, M. G., Casadesús, J. & García-del Portillo, F. Dormant intracellular Salmonella enterica serovar Typhimurium discriminates among Salmonella pathogenicity island 2 effectors to persist inside fibroblasts. Infect. Immun. 82, 221–232 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  191. Arnoldini, M. et al. Bistable expression of virulence genes in Salmonella leads to the formation of an antibiotic-tolerant subpopulation. PLoS Biol. 12, e1001928 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  192. Helaine, S. et al. Internalization of Salmonella by macrophages induces formation of nonreplicating persisters. Science 343, 204–208 (2014). This study uses a fluorescence dilution technique to study persisters.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  193. Spano, S., Ugalde, J. E. & Galán, J. E. Delivery of a Salmonella Typhi exotoxin from a host intracellular compartment. Cell Host Microbe 3, 30–38 (2008).

    Article  CAS  PubMed  Google Scholar 

  194. Haghjoo, E. & Galán, J. E. Salmonella Typhi encodes a functional cytolethal distending toxin that is delivered into host cells by a bacterial-internalization pathway. Proc. Natl Acad. Sci. USA 101, 4614–4619 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  195. Hodak, H. & Galán, J. E. A Salmonella Typhi homologue of bacteriophage muramidases controls typhoid toxin secretion. EMBO Rep. 14, 95–102 (2013).

    Article  CAS  PubMed  Google Scholar 

  196. Spano, S. & Galán, J. E. A Rab32-dependent pathway contributes to Salmonella Typhi host restriction. Science 338, 960–963 (2012). This paper demonstrates that loss of a T3SS effector contributes to S. Typhi host restriction.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  197. Song, J., Gao, X. & Galán, J. E. Structure and function of the Salmonella Typhi chimaeric A2B5 typhoid toxin. Nature 499, 350–354 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  198. Winter, S. E., Raffatellu, M., Wilson, R. P., Rüssmann, H. & Bäumler, A. J. The Salmonella enterica serotype Typhi regulator TviA reduces interleukin-8 production in intestinal epithelial cells by repressing flagellin secretion. Cell. Microbiol. 10, 247–261 (2008).

    CAS  PubMed  Google Scholar 

  199. Haneda, T. et al. The capsule-encoding viaB locus reduces intestinal inflammation by a Salmonella pathogenicity island 1-independent mechanism. Infect. Immun. 77, 2932–2942 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  200. Crawford, R. W. et al. Loss of very-long O-antigen chains optimizes capsule-mediated immune evasion by Salmonella enterica serovar Typhi. mBio 4, e00232-13 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  201. McClelland, M. et al. Comparison of genome degradation in Paratyphi A and Typhi, human-restricted serovars of Salmonella enterica that cause typhoid. Nature Genet. 36, 1268–1274 (2004).

    Article  CAS  PubMed  Google Scholar 

  202. Spano, S., Liu, X. & Galán, J. E. Proteolytic targeting of Rab29 by an effector protein distinguishes the intracellular compartments of human-adapted and broad-host Salmonella. Proc. Natl Acad. Sci. USA 108, 18418–18423 (2011).

    Article  PubMed  PubMed Central  Google Scholar 

  203. Trombert, A. N., Berrocal, L., Fuentes, J. & Mora, G. S. Typhimurium sseJ gene decreases the S. Typhi cytotoxicity toward cultured epithelial cells. BMC Microbiol. 10, 312 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  204. Levine, M. M., Black, R. E. & Lanata, C. Precise estimation of the numbers of chronic carriers of Salmonella Typhi in Santiago, Chile, an endemic area. J. Infect. Dis. 146, 724–726 (1982).

    Article  CAS  PubMed  Google Scholar 

  205. Miyoshi, H. & Stappenbeck, T. S. In vitro expansion and genetic modification of gastrointestinal stem cells in spheroid culture. Nature Protoc. 8, 2471–2482 (2013).

    Article  CAS  Google Scholar 

  206. Balaban, N. Q., Merrin, J., Chait, R., Kowalik, L. & Leibler, S. Bacterial persistence as a phenotypic switch. Science 305, 1622–1625 (2004).

    Article  CAS  PubMed  Google Scholar 

  207. Dörr, T., Vulic, M. & Lewis, K. Ciprofloxacin causes persister formation by inducing the TisB toxin in Escherichia coli. PLoS Biol. 8, e1000317 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  208. Christen, M. et al. Asymmetrical distribution of the second messenger c-di-GMP upon bacterial cell division. Science 328, 1295–1297 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

This work was supported by US National Institutes of Health (NIH) grants RO1 AI030479, RO1 AI048683 and U19 AI090882 to S.I.M.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Samuel I. Miller.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Supplementary information

Supplementary information S1 (table)

Effectors of the Salmonella spp. T3SSs. (PDF 269 kb)

PowerPoint slides

Glossary

Microfold cells

(M cells). Specialized epithelial cells that phagocytose molecules in the intestinal lumen for transepithelial transport to enable immunological sampling of antigens.

Peyer's patches

Organized lymphoid regions of the small intestine monolayer that function in immune surveillance and the generation of a localized immune response.

Type III secretion systems

(T3SSs). Needle-like apparatuses that are assembled by pathogenic bacteria for the delivery of bacterial effectors directly into host cells.

Pathogenicity island

A large horizontally acquired region of genomic DNA that often encodes virulence factors.

Macropinocytosis

A non-selective form of endocytosis that involves membrane ruffling.

Autophagy

A catabolic cellular process in which cytoplasmic contents, such as damaged organelles and proteins, are targeted for lysosomal degradation.

RHO GTPases

A family of GTPases that are important molecular switches for regulating actin dynamics.

Tight junctions

The adhesive contacts comprising protein complexes that function to limit the movement of molecules and ions through the space between cells in a monolayer.

Toll-like receptors

(TLRs). Membrane-bound receptors of the innate immune system that recognize specific pathogen-associated molecular patterns.

Caspase

A family of cysteine proteases that have an essential role in cell death pathways.

Stromal cells

Connective tissue cells that include enterocytes, tissue fibroblasts and vascular endothelial cells.

Endocytosis

The process by which extracellular particles are engulfed by eukaryotic cells and enclosed in a vesicle.

Siderophore

A small molecule that has a high affinity for iron and is secreted by bacteria to scavenge this element.

Reactive oxygen species

(ROS). Chemically reactive species that contain oxygen. They are produced as a by-product of aerobic respiration and function in combating microbial infections.

Microtubular network

The dynamic network of tubulin polymers that constitute an important component of the cell cytoskeleton. This network is used for intracellular transport, including movement of secretory vesicles and organelles such as the Salmonella-containing vacuole.

Sorting nexin

A family of proteins characterized by the presence of a phox-homology domain that functions by binding to phosphatidylinositol-3- monophosphate. Sorting nexins associate with the endocytic network and are involved in endocytosis, endosomal sorting and endosomal signalling.

Periphery-directed transport vesicles

Vesicles, such as those from the Salmonella-containing vacuole, that move towards the periphery of cells along microtubules.

Prenylation

The attachment of farnesyl or geranyl–geranyl groups to carboxy-terminal cysteine residues that are present in specific prenylation motifs of proteins, which promotes membrane association or protein–protein interactions.

S-palmitoylated

The reversible covalent post-translational attachment of lipids (usually palmitate) to cysteine residues of proteins.

Peroxisomes

Organelles that degrade long-chain fatty acids in eukaryotes.

Pattern recognition receptors

(PRRs). Proteins of the innate immune system that recognize pathogen-associated molecular patterns and initiate an innate immune response that facilitates pathogen clearance by triggering cytokine and chemokine expression.

NOD-like receptors

(NLRs). Intracellular receptors of the innate immune system that recognize pathogen- associated molecular patterns.

Pyroptosis

An inflammatory, programmed cell death pathway that is triggered by the activation of caspase 1 and caspase 11 in mice, and by caspase 4, caspase 5 and caspase 11 in humans.

Endotoxic shock

A severe inflammatory reaction that is induced by high levels of endotoxin (lipopolysaccharide) in the bloodstream.

Mitophagy

The process of targeting mitochondria for autophagic degradation.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

LaRock, D., Chaudhary, A. & Miller, S. Salmonellae interactions with host processes. Nat Rev Microbiol 13, 191–205 (2015). https://doi.org/10.1038/nrmicro3420

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nrmicro3420

This article is cited by

Search

Quick links

Nature Briefing Microbiology

Sign up for the Nature Briefing: Microbiology newsletter — what matters in microbiology research, free to your inbox weekly.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing: Microbiology