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:

Bacterial manipulation of innate immunity to promote infection

Key Points

  • Although an arsenal of antimicrobial mechanisms is deployed to kill bacteria in infected cells, pathogens have learnt how to manipulate host defence mechanisms to survive and eventually replicate. This Review focuses on the different mechanisms used by pathogenic intracellular bacteria to subvert host innate immune responses.

  • At the level of pathogen entry, membrane receptors and cofactors involved in the internalization and cytoskeleton rearrangements are targeted by pathogens to promote their own entry into host cells.

  • Pathogens can also subvert innate immune responses by expressing protein effectors, which directly interfere with Toll-like receptor signalling pathways.

  • After internalization, pathogenic bacteria can use endocytic, phagocytic and exocytic pathways by expressing proteic and lipid bacterial effectors to escape from their vacuole, to relocate in a safe membrane compartment or to stop the maturation process of their vacuole with the aim of establishing a suitable replication niche.

  • In response to inflammatory stimuli generated by internalized pathogens, macrophages and neutrophils produce chemically reactive molecules such as antimicrobial peptides, reactive oxygen species (ROS) and reactive nitrogen intermediates (RNI). Again, pathogens have developed mechanisms to detoxify ROS and RNI, and thereby avoid elimination.

  • Although autophagy is a defence mechanism, some pathogens use autophagy for their own benefit, whereas others secrete bacterial effectors involved in the camouflage against autophagic recognition.

  • Finally, microbial molecules or intermediary signals induced by these molecules are thought to stimulate response mechanisms leading to inflammasome activation. Controlling caspase 1 activation and cell death is an important aspect of virulence for several intracellular pathogens.

  • The coordinate action of bacterial effectors is important for pathogenic intracellular bacteria to maintain membrane homeostasis of vacuoles, acquire nutrients and avoid damage caused by the host immune system.

Abstract

The mammalian innate immune response provides a barrier against invading pathogens. Innate immune mechanisms are used by the host to respond to a range of bacterial pathogens in an acute and conserved fashion. Host cells express pattern recognition receptors that sense pathogen-associated molecular patterns. After detection, an arsenal of antimicrobial mechanisms is deployed to kill bacteria in infected cells. Innate immunity also stimulates antigen-specific responses mediated by the adaptive immune system. In response, pathogens manipulate host defence mechanisms to survive and eventually replicate. This Review focuses on the control of host innate immune responses by pathogenic intracellular bacteria.

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: Manipulation of host innate immunity by intracellular pathogenic bacteria.
Figure 2: Intracellular pathogens manipulate host membrane trafficking to resist innate immunity and promote survival and replication.
Figure 3: Survival niches of pathogenic bacteria.
Figure 4: Autophagy and inflammasome activation can be manipulated by intracellular pathogens.

Similar content being viewed by others

References

  1. Alexopoulou, L. & Kontoyiannis, D. Contribution of microbial-associated molecules in innate mucosal responses. Cell. Mol. Life Sci. 62, 1349–1358 (2005).

    Article  CAS  PubMed  Google Scholar 

  2. Hibino, T. et al. The immune gene repertoire encoded in the purple sea urchin genome. Dev. Biol. 300, 349–365 (2006).

    Article  CAS  PubMed  Google Scholar 

  3. Akashi-Takamura, S. & Miyake, K. TLR accessory molecules. Curr. Opin. Immunol. 20, 420–425 (2008).

    Article  CAS  PubMed  Google Scholar 

  4. Kumar, Y. & Valdivia, R. H. Leading a sheltered life: intracellular pathogens and maintenance of vacuolar compartments. Cell Host Microbe 5, 593–601 (2009). This review concentrates on the strategies used by vacuole-bound pathogens to invade and establish a replicative vacuole and also on what are the mechanisms involved in pathogenic vacuole maintenance and disruption.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Neyrolles, O. et al. Is adipose tissue a place for Mycobacterium tuberculosis persistence? PLoS One 1, e43 (2006).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  6. Tailleux, L., Maeda, N., Nigou, J., Gicquel, B. & Neyrolles, O. How is the phagocyte lectin keyboard played? Master class lesson by Mycobacterium tuberculosis. Trends Microbiol. 11, 259–263 (2003).

    Article  CAS  PubMed  Google Scholar 

  7. Yadav, M. & Schorey, J. S. The β-glucan receptor dectin-1 functions together with TLR2 to mediate macrophage activation by mycobacteria. Blood 108, 3168–3175 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Korbel, D. S., Schneider, B. E. & Schaible, U. E. Innate immunity in tuberculosis: myths and truth. Microbes Infect. 10, 995–1004 (2008).

    Article  CAS  PubMed  Google Scholar 

  9. Flannagan, R. S., Cosio, G. & Grinstein, S. Antimicrobial mechanisms of phagocytes and bacterial evasion strategies. Nature Rev. Microbiol. 7, 355–366 (2009).

    CAS  Google Scholar 

  10. Cooper, A. M. Cell-mediated immune responses in tuberculosis. Annu. Rev. Immunol. 27, 393–422 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Bonazzi, M., Lecuit, M. & Cossart, P. Listeria monocytogenes internalin and E-cadherin: from structure to pathogenesis. Cell. Microbiol. 11, 693–702 (2009).

    Article  CAS  PubMed  Google Scholar 

  12. Dehio, C. Bartonella interactions with endothelial cells and erythrocytes. Trends Microbiol. 9, 279–285 (2001).

    Article  CAS  PubMed  Google Scholar 

  13. Kubori, T. et al. Supramolecular structure of the Salmonella typhimurium type III protein secretion system. Science 280, 602–605 (1998).

    Article  CAS  PubMed  Google Scholar 

  14. Unsworth, K. E., Way, M., McNiven, M., Machesky, L. & Holden, D. W. Analysis of the mechanisms of Salmonella-induced actin assembly during invasion of host cells and intracellular replication. Cell. Microbiol. 6, 1041–1055 (2004).

    Article  CAS  PubMed  Google Scholar 

  15. Patel, J. C. & Galan, J. E. Manipulation of the host actin cytoskeleton by Salmonella — all in the name of entry. Curr. Opin. Microbiol. 8, 10–15 (2005).

    Article  CAS  PubMed  Google Scholar 

  16. Patel, J. C., Rossanese, O. W. & Galan, J. E. The functional interface between Salmonella and its host cell: opportunities for therapeutic intervention. Trends Pharmacol. Sci. 26, 564–570 (2005).

    Article  CAS  PubMed  Google Scholar 

  17. Patel, J. C. & Galan, J. E. Differential activation and function of Rho GTPases during Salmonella-host cell interactions. J. Cell Biol. 175, 453–463 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Kaniga, K., Uralil, J., Bliska, J. B. & Galan, J. E. A secreted protein tyrosine phosphatase with modular effector domains in the bacterial pathogen Salmonella typhimurium. Mol. Microbiol. 21, 633–641 (1996).

    Article  CAS  PubMed  Google Scholar 

  19. Li, G. & Zhang, X. C. GTP hydrolysis mechanism of Ras-like GTPases. J. Mol. Biol. 340, 921–932 (2004).

    Article  CAS  PubMed  Google Scholar 

  20. Fu, Y. & Galan, J. E. A salmonella protein antagonizes Rac-1 and Cdc42 to mediate host-cell recovery after bacterial invasion. Nature 401, 293–297 (1999). This article shows that the S . Typhimurium effector protein SptP delivered to the host cell cytosol by the T3SS acts as a GAP for RAC1 and CDC42, leading to the reversal of the actin cytoskeletal changes induced on bacterium entry.

    Article  CAS  PubMed  Google Scholar 

  21. Murli, S., Watson, R. O. & Galan, 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 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Collier-Hyams, L. S. et al. Cutting edge: Salmonella AvrA effector inhibits the key proinflammatory, anti-apoptotic NF-κ B pathway. J. Immunol. 169, 2846–2850 (2002).

    Article  CAS  PubMed  Google Scholar 

  24. Nhieu, G. T. & Sansonetti, P. J. Mechanism of Shigella entry into epithelial cells. Curr. Opin. Microbiol. 2, 51–55 (1999).

    Article  CAS  PubMed  Google Scholar 

  25. Lafont, F., Tran Van Nhieu, G., Hanada, K., Sansonetti, P. & van der Goot, F. G. Initial steps of Shigella infection depend on the cholesterol/sphingolipid raft-mediated CD44-IpaB interaction. EMBO J. 21, 4449–4457 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Niebuhr, K. et al. Conversion of PtdIns(4, 5)P2 into PtdIns(5)P by the S.flexneri effector IpgD reorganizes host cell morphology. EMBO J. 21, 5069–5078 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Salcedo, S. P. et al. Brucella control of dendritic cell maturation is dependent on the TIR-containing protein Btp1. PLoS Pathog. 4, e21 (2008). This article describes for the first time the Brucella spp. TIR-interacting protein Btp1, which interferes with TLR2 signalling and induces an arrest in the DC maturation process.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  28. Newman, R. M., Salunkhe, P., Godzik, A. & Reed, J. C. Identification and characterization of a novel bacterial virulence factor that shares homology with mammalian Toll/interleukin-1 receptor family proteins. Infect. Immun. 74, 594–601 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Cirl, C. et al. Subversion of Toll-like receptor signalling by a unique family of bacterial Toll/interleukin-1 receptor domain-containing proteins. Nature Med. 14, 399–406 (2008). The authors characterize TIR domain-containing proteins in pathogenic bacteria. E. coli TcpC promotes bacterial survival and kidney pathology in vivo and inhibits TLR- and MYD88-specific signalling.

    Article  CAS  PubMed  Google Scholar 

  30. Radhakrishnan, G. K., Yu, Q., Harms, J. S. & Splitter, G. A. Brucella TIR domain-containing protein mimics properties of the toll-like receptor adaptor protein TIRAP. J. Biol. Chem. 284, 9892–9898 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Lapaque, N., Moriyon, I., Moreno, E. & Gorvel, J. P. Brucella lipopolysaccharide acts as a virulence factor. Curr. Opin. Microbiol. 8, 60–66 (2005).

    Article  CAS  PubMed  Google Scholar 

  32. Fugier, E., Pappas, G. & Gorvel, J. P. Virulence factors in brucellosis: implications for aetiopathogenesis and treatment. Expert Rev. Mol. Med. 9, 1–10 (2007).

    Article  PubMed  Google Scholar 

  33. Ganguly, N. et al. Mycobacterium tuberculosis secretory proteins CFP-10, ESAT-6 and the CFP10:ESAT6 complex inhibit lipopolysaccharide-induced NF-κB transactivation by downregulation of reactive oxidative species (ROS) production. Immunol. Cell Biol. 86, 98–106 (2008).

    Article  CAS  PubMed  Google Scholar 

  34. Ganguly, N. et al. Mycobacterium tuberculosis 6-kDa early secreted antigenic target (ESAT-6) protein downregulates lipopolysaccharide induced c-myc expression by modulating the extracellular signal regulated kinases 1/2. BMC Immunol. 8, 24 (2007).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  35. High, N., Mounier, J., Prevost, M. C. & Sansonetti, P. J. IpaB of Shigella flexneri causes entry into epithelial cells and escape from the phagocytic vacuole. EMBO J. 11, 1991–1999 (1992).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Pizarro-Cerda, J. & Cossart, P. Subversion of cellular functions by Listeria monocytogenes. J. Pathol. 208, 215–223 (2006).

    Article  CAS  PubMed  Google Scholar 

  37. Corr, S. C. & O'Neill, L. A. Listeria monocytogenes infection in the face of innate immunity. Cell. Microbiol. 11, 703–709 (2009).

    Article  CAS  PubMed  Google Scholar 

  38. Goldberg, M. B. Actin-based motility of intracellular microbial pathogens. Microbiol. Mol. Biol. Rev. 65, 595–626 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Perrin, A. J., Jiang, X., Birmingham, C. L., So, N. S. & Brumell, J. H. Recognition of bacteria in the cytosol of mammalian cells by the ubiquitin system. Curr. Biol. 14, 806–811 (2004).

    Article  CAS  PubMed  Google Scholar 

  40. de Chastellier, C., Forquet, F., Gordon, A. & Thilo, L. Mycobacterium requires an all-around closely apposing phagosome membrane to maintain the maturation block and this apposition is re-established when it rescues itself from phagolysosomes. Cell. Microbiol. 11, 1190–1207 (2009).

    Article  CAS  PubMed  Google Scholar 

  41. Fratti, R. A., Backer, J. M., Gruenberg, J., Corvera, S. & Deretic, V. Role of phosphatidylinositol 3-kinase and Rab5 effectors in phagosomal biogenesis and mycobacterial phagosome maturation arrest. J. Cell Biol. 154, 631–644 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Philips, J. A. Mycobacterial manipulation of vacuolar sorting. Cell. Microbiol. 10, 2408–2415 (2008).

    Article  CAS  PubMed  Google Scholar 

  43. Vergne, I. et al. Mechanism of phagolysosome biogenesis block by viable Mycobacterium tuberculosis. Proc. Natl Acad. Sci. USA 102, 4033–4038 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Shabaana, A. K. et al. Mycobacterial lipoarabinomannans modulate cytokine production in human T helper cells by interfering with raft/microdomain signalling. Cell. Mol. Life Sci. 62, 179–187 (2005).

    Article  CAS  PubMed  Google Scholar 

  45. Welin, A. et al. Incorporation of Mycobacterium tuberculosis lipoarabinomannan into macrophage membrane rafts is a prerequisite for the phagosomal maturation block. Infect. Immun. 76, 2882–2887 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Robinson, N. et al. Mycobacterial phenolic glycolipid inhibits phagosome maturation and subverts the pro-inflammatory cytokine response. Traffic 9, 1936–1947 (2008). In this paper the authors characterize a glycolipid of M. marinum , phenolphthiocerol diester, which promotes the arrest of phagosome maturation and abrogates pro-inflammatory cytokine production in human monocyte-derived macrophages.

    Article  CAS  PubMed  Google Scholar 

  47. Axelrod, S. et al. Delay of phagosome maturation by a mycobacterial lipid is reversed by nitric oxide. Cell. Microbiol. 10, 1530–1545 (2008).

    Article  CAS  PubMed  Google Scholar 

  48. Vergne, I., Chua, J. & Deretic, V. Mycobacterium tuberculosis phagosome maturation arrest: selective targeting of PI3P-dependent membrane trafficking. Traffic 4, 600–6 (2003).

    Article  CAS  PubMed  Google Scholar 

  49. Meresse, 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  CAS  PubMed  PubMed Central  Google Scholar 

  50. Knodler, L. A. & Steele-Mortimer, O. Taking possession: biogenesis of the Salmonella-containing vacuole. Traffic 4, 587–599 (2003).

    Article  CAS  PubMed  Google Scholar 

  51. Galan, J. E. Interaction of Salmonella with host cells through the centisome 63 type III secretion system. Curr. Opin. Microbiol. 2, 46–50 (1999).

    Article  CAS  PubMed  Google Scholar 

  52. Mukherjee, K., Parashuraman, S., Raje, M. & Mukhopadhyay, A. SopE acts as an Rab5-specific nucleotide exchange factor and recruits non-prenylated Rab5 on Salmonella-containing phagosomes to promote fusion with early endosomes. J. Biol. Chem. 276, 23607–23615 (2001).

    Article  CAS  PubMed  Google Scholar 

  53. Cirillo, D. M., Valdivia, R. H., Monack, D. M. & Falkow, S. Macrophage-dependent induction of the Salmonella pathogenicity island 2 type III secretion system and its role in intracellular survival. Mol. Microbiol. 30, 175–188 (1998).

    Article  CAS  PubMed  Google Scholar 

  54. Hensel, M. et al. Genes encoding putative effector proteins of the type III secretion system of Salmonella pathogenicity island 2 are required for bacterial virulence and proliferation in macrophages. Mol. Microbiol. 30, 163–174 (1998).

    Article  CAS  PubMed  Google Scholar 

  55. Uchiya, K. et al. A Salmonella virulence protein that inhibits cellular trafficking. EMBO J. 18, 3924–3933 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Lee, A. H., Zareei, M. P. & Daefler, S. Identification of a NIPSNAP homologue as host cell target for Salmonella virulence protein SpiC. Cell. Microbiol. 4, 739–750 (2002).

    Article  CAS  PubMed  Google Scholar 

  57. 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 

  58. 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 

  59. Stein, M. A., Leung, K. Y., Zwick, M., Garcia-del Portillo, F. & Finlay, B. 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 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. 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 

  62. Salcedo, S. P., Noursadeghi, M., Cohen, J. & Holden, D. W. Intracellular replication of Salmonella typhimurium strains in specific subsets of splenic macrophages in vivo. Cell. Microbiol. 3, 587–597 (2001).

    Article  CAS  PubMed  Google Scholar 

  63. Arellano-Reynoso, B. et al. Cyclic β-1, 2-glucan is a Brucella virulence factor required for intracellular survival. Nature Immunol. 6, 618–625 (2005). This study shows that the cyclic β-1, 2-glucans synthesized by Brucella spp. act in lipid rafts found on host cell membranes to prevent phagosome–lysosome fusion and to allow bacterial replication.

    Article  CAS  Google Scholar 

  64. Fugier, E. et al. The glyceraldehyde-3-phosphate dehydrogenase and the small GTPase Rab 2 are crucial for Brucella replication. PLoS Pathog. 5, e1000487 (2009).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  65. Roy, C. R., Salcedo, S. P. & Gorvel, J. P. Pathogen-endoplasmic-reticulum interactions: in through the out door. Nature Rev. Immunol. 6, 136–147 (2006).

    Article  CAS  Google Scholar 

  66. Ninio, S. & Roy, C. R. Effector proteins translocated by Legionella pneumophila: strength in numbers. Trends Microbiol. 15, 372–380 (2007).

    Article  CAS  PubMed  Google Scholar 

  67. Wyrick, P. B. Intracellular survival by Chlamydia. Cell. Microbiol. 2, 275–282 (2000).

    Article  CAS  PubMed  Google Scholar 

  68. Heuer, D. et al. Chlamydia causes fragmentation of the Golgi compartment to ensure reproduction. Nature 457, 731–735 (2009). This is a landmark paper showing that Chlamydia spp. infection in human epithelial cells induces Golgi fragmentation to generate Golgi ministacks surrounding the bacterial inclusion, facilitating lipid acquisition and intracellular pathogen growth. Ministack formation is triggered by the proteolytic cleavage of golgin 84.

    Article  CAS  PubMed  Google Scholar 

  69. Radtke, A. L. & O'Riordan, M. X. Intracellular innate resistance to bacterial pathogens. Cell. Microbiol. 8, 1720–1729 (2006). The authors provide excellent general review of the mechanisms of innate immune intracellular resistance encountered by bacterial pathogens and how some bacteria can evade destruction by the innate immune system.

    Article  CAS  PubMed  Google Scholar 

  70. Nauseef, W. M. Assembly of the phagocyte NADPH oxidase. Histochem. Cell. Biol. 122, 277–291 (2004).

    Article  CAS  PubMed  Google Scholar 

  71. MacMicking, J., Xie, Q. W. & Nathan, C. Nitric oxide and macrophage function. Annu. Rev. Immunol. 15, 323–350 (1997).

    Article  CAS  PubMed  Google Scholar 

  72. Bogdan, C. Nitric oxide and the immune response. Nature Immunol. 2, 907–916 (2001).

    Article  CAS  Google Scholar 

  73. Lowenstein, C. J. & Padalko, E. iNOS (NOS2) at a glance. J. Cell Sci. 117, 2865–2867 (2004).

    Article  CAS  PubMed  Google Scholar 

  74. Chakravortty, D. & Hensel, M. Inducible nitric oxide synthase and control of intracellular bacterial pathogens. Microbes Infect. 5, 621–627 (2003).

    Article  CAS  PubMed  Google Scholar 

  75. Fang, F. C. Antimicrobial reactive oxygen and nitrogen species: concepts and controversies. Nature Rev. Microbiol. 2, 820–832 (2004).

    Article  CAS  Google Scholar 

  76. De Groote, M. A. et al. Periplasmic superoxide dismutase protects Salmonella from products of phagocyte NADPH-oxidase and nitric oxide synthase. Proc. Natl Acad. Sci. USA 94, 13997–14001 (1997).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  77. Farrant, J. L. et al. Bacterial copper- and zinc-cofactored superoxide dismutase contributes to the pathogenesis of systemic salmonellosis. Mol. Microbiol. 25, 785–796 (1997).

    Article  CAS  PubMed  Google Scholar 

  78. Fang, F. C. et al. Virulent Salmonella typhimurium has two periplasmic Cu, Zn-superoxide dismutases. Proc. Natl Acad. Sci. USA 96, 7502–7507 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  79. Linehan, S. A. & Holden, D. W. The interplay between Salmonella typhimurium and its macrophage host—what can it teach us about innate immunity? Immunol. Lett. 85, 183–192 (2003). This is an excellent review on host–pathogen interactions during S . Typhimurium infection. The authors focus on different facets to the cell biology of macrophages and their innate immune functions.

    Article  CAS  PubMed  Google Scholar 

  80. Lundberg, B. E., Wolf, R. E. Jr, Dinauer, M. C., Xu, Y. & Fang, F. C. Glucose 6-phosphate dehydrogenase is required for Salmonella typhimurium virulence and resistance to reactive oxygen and nitrogen intermediates. Infect. Immun. 67, 436–438 (1999).

    CAS  PubMed  PubMed Central  Google Scholar 

  81. van der Straaten, T. et al. Novel Salmonella enterica serovar Typhimurium protein that is indispensable for virulence and intracellular replication. Infect. Immun. 69, 7413–7418 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  82. van Diepen, A., van der Straaten, T., Holland, S. M., Janssen, R. & van Dissel, J. T. A superoxide-hypersusceptible Salmonella enterica serovar Typhimurium mutant is attenuated but regains virulence in p47(phox−/−) mice. Infect. Immun. 70, 2614–2621 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  83. Bryk, R., Griffin, P. & Nathan, C. Peroxynitrite reductase activity of bacterial peroxiredoxins. Nature 407, 211–215 (2000).

    Article  CAS  PubMed  Google Scholar 

  84. Chen, L., Xie, Q. W. & Nathan, C. Alkyl hydroperoxide reductase subunit C (AhpC) protects bacterial and human cells against reactive nitrogen intermediates. Mol. Cell 1, 795–805 (1998).

    Article  CAS  PubMed  Google Scholar 

  85. Husain, M. et al. Nitric oxide evokes an adaptive response to oxidative stress by arresting respiration. J. Biol. Chem. 283, 7682–7689 (2008).

    Article  CAS  PubMed  Google Scholar 

  86. Jaeger, T. et al. Multiple thioredoxin-mediated routes to detoxify hydroperoxides in Mycobacterium tuberculosis. Arch. Biochem. Biophys. 423, 182–191 (2004).

    Article  CAS  PubMed  Google Scholar 

  87. Jaeger, T. Peroxiredoxin systems in mycobacteria. Subcell. Biochem. 44, 207–217 (2007).

    Article  PubMed  Google Scholar 

  88. Sherman, D. R. et al. Compensatory ahpC gene expression in isoniazid-resistant Mycobacterium tuberculosis. Science 272, 1641–1643 (1996).

    Article  CAS  PubMed  Google Scholar 

  89. Fischer, K. et al. Mycobacterial lysocardiolipin is exported from phagosomes upon cleavage of cardiolipin by a macrophage-derived lysosomal phospholipase A2. J. Immunol. 167, 2187–2192 (2001). This paper shows that lysocardiolipin is released from the mycobacterial cell wall in the phagosome of infected macrophages and transported out of this compartment into intracellular vesicles. Lysocardiolipin was generated through cleavage of mycobacterial cardiolipin by a calcium-independent phospholipase A2 present in the lysosomes of macrophages.

    Article  CAS  PubMed  Google Scholar 

  90. Bang, I. S. et al. Maintenance of nitric oxide and redox homeostasis by the Salmonella flavohemoglobin hmp. J. Biol. Chem. 281, 28039–28047 (2006).

    Article  CAS  PubMed  Google Scholar 

  91. Darwin, K. H. & Nathan, C. F. Role for nucleotide excision repair in virulence of Mycobacterium tuberculosis. Infect. Immun. 73, 4581–4587 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  92. Cowley, S. C., Myltseva, S. V. & Nano, F. E. Phase variation in Francisella tularensis affecting intracellular growth, lipopolysaccharide antigenicity and nitric oxide production. Mol. Microbiol. 20, 867–874 (1996).

    Article  CAS  PubMed  Google Scholar 

  93. Eriksson, S. et al. Salmonella typhimurium mutants that downregulate phagocyte nitric oxide production. Cell. Microbiol. 2, 239–250 (2000).

    Article  CAS  PubMed  Google Scholar 

  94. Vazquez-Torres, A. et al. Salmonella pathogenicity island 2-dependent evasion of the phagocyte NADPH oxidase. Science 287, 1655–1658 (2000).

    Article  CAS  PubMed  Google Scholar 

  95. Shiloh, M. U. et al. Phenotype of mice and macrophages deficient in both phagocyte oxidase and inducible nitric oxide synthase. Immunity 10, 29–38 (1999).

    Article  CAS  PubMed  Google Scholar 

  96. Gallois, A., Klein, J. R., Allen, L. A., Jones, B. D. & Nauseef, W. M. Salmonella pathogenicity island 2-encoded type III secretion system mediates exclusion of NADPH oxidase assembly from the phagosomal membrane. J. Immunol. 166, 5741–5748 (2001).

    Article  CAS  PubMed  Google Scholar 

  97. Vazquez-Torres, A., Fantuzzi, G., Edwards, C. K. 3rd, Dinarello, C. A. & Fang, F. C. Defective localization of the NADPH phagocyte oxidase to Salmonella-containing phagosomes in tumour necrosis factor p55 receptor-deficient macrophages. Proc. Natl Acad. Sci. USA 98, 2561–2565 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  98. Cherayil, B. J., McCormick, B. A. & Bosley, J. Salmonella enterica serovar typhimurium-dependent regulation of inducible nitric oxide synthase expression in macrophages by invasins SipB, SipC, and SipD and effector SopE2. Infect. Immun. 68, 5567–5574 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  99. Bjur, E., Eriksson-Ygberg, S. & Rhen, M. The O-antigen affects replication of Salmonella enterica serovar Typhimurium in murine macrophage-like J774-A.1 cells through modulation of host cell nitric oxide production. Microbes Infect. 8, 1826–1838 (2006).

    Article  CAS  PubMed  Google Scholar 

  100. Chakravortty, D., Hansen-Wester, I. & Hensel, M. Salmonella pathogenicity island 2 mediates protection of intracellular Salmonella from reactive nitrogen intermediates. J. Exp. Med. 195, 1155–1166 (2002). In this paper the authors use immunofluorescence microscopy to show efficient colocalization of NOS2 with bacteria deficient in SPI2 but not wild-type S . Typhimurium, suggesting that the SPI2 system interferes with the localization of NOS2 in S . Typhimurium and that avoidance of colocalization with RNI is important for a pathogen to adapt to an intracellular lifestyle.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  101. Miller, B. H. et al. Mycobacteria inhibit nitric oxide synthase recruitment to phagosomes during macrophage infection. Infect. Immun. 72, 2872–2878 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  102. Davis, A. S. et al. Mechanism of inducible nitric oxide synthase exclusion from mycobacterial phagosomes. PLoS Pathog. 3, e186 (2007).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  103. Forbes, J. R. & Gros, P. Iron, manganese, and cobalt transport by Nramp1 (Slc11a1) and Nramp2 (Slc11a2) expressed at the plasma membrane. Blood 102, 1884–1892 (2003).

    Article  CAS  PubMed  Google Scholar 

  104. Chlosta, S. et al. The iron efflux protein ferroportin regulates the intracellular growth of Salmonella enterica. Infect. Immun. 74, 3065–3067 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  105. De Voss, J. J. et al. The salicylate-derived mycobactin siderophores of Mycobacterium tuberculosis are essential for growth in macrophages. Proc. Natl Acad. Sci. USA 97, 1252–1257 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  106. Parent, M. A. et al. Brucella abortus siderophore 2, 3-dihydroxybenzoic acid (DHBA) facilitates intracellular survival of the bacteria. Microb. Pathog. 32, 239–248 (2002).

    Article  CAS  PubMed  Google Scholar 

  107. Fischbach, M. A., Lin, H., Liu, D. R. & Walsh, C. T. How pathogenic bacteria evade mammalian sabotage in the battle for iron. Nature Chem. Biol. 2, 132–138 (2006).

    Article  CAS  Google Scholar 

  108. Chatterjee, S. S. et al. Intracellular gene expression profile of Listeria monocytogenes. Infect. Immun. 74, 1323–1338 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  109. Jansen, A. & Yu, J. Differential gene expression of pathogens inside infected hosts. Curr. Opin. Microbiol. 9, 138–142 (2006).

    Article  CAS  PubMed  Google Scholar 

  110. Levine, B. & Deretic, V. Unveiling the roles of autophagy in innate and adaptive immunity. Nature Rev. Immunol. 7, 767–777 (2007).

    Article  CAS  Google Scholar 

  111. Schmid, D. & Munz, C. Innate and adaptive immunity through autophagy. Immunity 27, 11–21 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  112. Orvedahl, A. & Levine, B. Eating the enemy within: autophagy in infectious diseases. Cell Death Differ. 16, 57–69 (2009).

    Article  CAS  PubMed  Google Scholar 

  113. Dorn, B. R., Dunn, W. A. Jr & Progulske-Fox, A. Porphyromonas gingivalis traffics to autophagosomes in human coronary artery endothelial cells. Infect. Immun. 69, 5698–5708 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  114. Gutierrez, M. G. et al. Autophagy induction favours the generation and maturation of the Coxiella-replicative vacuoles. Cell. Microbiol. 7, 981–993 (2005).

    Article  CAS  PubMed  Google Scholar 

  115. Birmingham, C. L. et al. Listeriolysin O allows Listeria monocytogenes replication in macrophage vacuoles. Nature 451, 350–354 (2008). The authors reveal a role for listeriolysin O in promoting L. monocytogenes replication in vacuoles in infected macrophages.

    Article  CAS  PubMed  Google Scholar 

  116. Swanson, M. S. & Molofsky, A. B. Autophagy and inflammatory cell death, partners of innate immunity. Autophagy 1, 174–176 (2005).

    Article  CAS  PubMed  Google Scholar 

  117. Suzuki, T. et al. Differential regulation of caspase-1 activation, pyroptosis, and autophagy via Ipaf and ASC in Shigella-infected macrophages. PLoS Pathog. 3, e111 (2007).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  118. Schroeder, G. N. & Hilbi, H. Cholesterol is required to trigger caspase-1 activation and macrophage apoptosis after phagosomal escape of Shigella. Cell. Microbiol. 9, 265–278 (2007).

    Article  CAS  PubMed  Google Scholar 

  119. Suzuki, T. & Nunez, G. A role for Nod-like receptors in autophagy induced by Shigella infection. Autophagy 4, 73–75 (2008).

    Article  PubMed  Google Scholar 

  120. Ogawa, M. et al. Escape of intracellular Shigella from autophagy. Science 307, 727–731 (2005). This is a landmark paper showing that S. flexneri can escape autophagy by secreting IcsB. S. flexneri VirG, which is required for intracellular actin-based motility, induces autophagy by binding to ATG5. IcsB does not directly inhibit autophagy and instead its role is to camouflage its own bacterial target molecule (VirG) from the autophagic host defence system.

    Article  CAS  PubMed  Google Scholar 

  121. Akira, S., Uematsu, S. & Takeuchi, O. Pathogen recognition and innate immunity. Cell 124, 783–801 (2006).

    Article  CAS  PubMed  Google Scholar 

  122. Meylan, E., Tschopp, J. & Karin, M. Intracellular pattern recognition receptors in the host response. Nature 442, 39–44 (2006).

    Article  CAS  PubMed  Google Scholar 

  123. Ting, J. P. et al. The NLR gene family: a standard nomenclature. Immunity 28, 285–287 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  124. Martinon, F., Burns, K. & Tschopp, J. The inflammasome: a molecular platform triggering activation of inflammatory caspases and processing of proIL-β. Mol. Cell 10, 417–426 (2002).

    Article  CAS  PubMed  Google Scholar 

  125. Kanneganti, T. D., Lamkanfi, M. & Nunez, G. Intracellular NOD-like receptors in host defense and disease. Immunity 27, 549–559 (2007).

    Article  CAS  PubMed  Google Scholar 

  126. Mariathasan, S. & Monack, D. M. Inflammasome adaptors and sensors: intracellular regulators of infection and inflammation. Nature Rev. Immunol. 7, 31–40 (2007).

    Article  CAS  Google Scholar 

  127. Martinon, F. & Tschopp, J. Inflammatory caspases and inflammasomes: master switches of inflammation. Cell Death Differ. 14, 10–22 (2007).

    Article  CAS  PubMed  Google Scholar 

  128. Fernandes-Alnemri, T. et al. The pyroptosome: a supramolecular assembly of ASC dimers mediating inflammatory cell death via caspase-1 activation. Cell Death Differ. 14, 1590–1604 (2007).

    Article  CAS  PubMed  Google Scholar 

  129. Yu, H. B. & Finlay, B. B. The caspase-1 inflammasome: a pilot of innate immune responses. Cell Host Microbe 4, 198–208 (2008).

    Article  CAS  PubMed  Google Scholar 

  130. Martinon, F. & Tschopp, J. Inflammatory caspases: linking an intracellular innate immune system to autoinflammatory diseases. Cell 117, 561–574 (2004).

    Article  CAS  PubMed  Google Scholar 

  131. Keller, M., Ruegg, A., Werner, S. & Beer, H. D. Active caspase-1 is a regulator of unconventional protein secretion. Cell 132, 818–831 (2008).

    Article  CAS  PubMed  Google Scholar 

  132. Shao, W., Yeretssian, G., Doiron, K., Hussain, S. N. & Saleh, M. The caspase-1 digestome identifies the glycolysis pathway as a target during infection and septic shock. J. Biol. Chem. 282, 36321–36329 (2007).

    Article  CAS  PubMed  Google Scholar 

  133. Kanneganti, T. D. et al. Bacterial RNA and small antiviral compounds activate caspase-1 through cryopyrin/Nalp3. Nature 440, 233–236 (2006).

    Article  CAS  PubMed  Google Scholar 

  134. Mariathasan, S. et al. Cryopyrin activates the inflammasome in response to toxins and ATP. Nature 440, 228–232 (2006). This paper shows that cryopyrin-deficient macrophages cannot activate caspase 1 in response to TLR agonists and ATP. Cryopyrin is essential for inflammasome activation in response to signalling pathways that are triggered specifically by ATP, nigericin, maitotoxin, Staphylococcus aureus or L. monocytogenes.

    Article  CAS  PubMed  Google Scholar 

  135. Faustin, B. et al. Reconstituted NALP1 inflammasome reveals two-step mechanism of caspase-1 activation. Mol. Cell 25, 713–724 (2007).

    Article  CAS  PubMed  Google Scholar 

  136. Martinon, F., Agostini, L., Meylan, E. & Tschopp, J. Identification of bacterial muramyl dipeptide as activator of the NALP3/cryopyrin inflammasome. Curr. Biol. 14, 1929–1934 (2004).

    Article  CAS  PubMed  Google Scholar 

  137. Pan, Q. et al. MDP-induced interleukin-1β processing requires Nod2 and CIAS1/NALP3. J. Leukoc. Biol. 82, 177–183 (2007).

    Article  CAS  PubMed  Google Scholar 

  138. Hsu, L. C. et al. A NOD2-NALP1 complex mediates caspase-1-dependent IL-1β secretion in response to Bacillus anthracis infection and muramyl dipeptide. Proc. Natl Acad. Sci. USA 105, 7803–7808 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  139. 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 

  140. 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 

  141. Kanneganti, T. D. et al. Critical role for Cryopyrin/Nalp3 in activation of caspase-1 in response to viral infection and double-stranded RNA. J. Biol. Chem. 281, 36560–36568 (2006).

    Article  CAS  PubMed  Google Scholar 

  142. Kanneganti, T. D. et al. Pannexin-1-mediated recognition of bacterial molecules activates the cryopyrin inflammasome independent of Toll-like receptor signalling. Immunity 26, 433–443 (2007).

    Article  CAS  PubMed  Google Scholar 

  143. Piccini, A. et al. ATP is released by monocytes stimulated with pathogen-sensing receptor ligands and induces IL-1β and IL-18 secretion in an autocrine way. Proc. Natl Acad. Sci. USA 105, 8067–8072 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  144. Weiss, D. S. et al. In vivo negative selection screen identifies genes required for Francisella virulence. Proc. Natl Acad. Sci. USA 104, 6037–6042 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  145. Master, S. S. et al. Mycobacterium tuberculosis prevents inflammasome activation. Cell Host Microbe 3, 224–232 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  146. Ruckdeschel, K. et al. Yersinia enterocolitica impairs activation of transcription factor NF-κB: involvement in the induction of programmed cell death and in the suppression of the macrophage tumour necrosis factor α production. J. Exp. Med. 187, 1069–1079 (1998).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  147. Schesser, K. et al. The yopJ locus is required for Yersinia-mediated inhibition of NF-κB activation and cytokine expression: YopJ contains a eukaryotic SH2-like domain that is essential for its repressive activity. Mol. Microbiol. 28, 1067–1079 (1998).

    Article  CAS  PubMed  Google Scholar 

  148. Schotte, P. et al. Targeting Rac1 by the Yersinia effector protein YopE inhibits caspase-1-mediated maturation and release of interleukin-1β. J. Biol. Chem. 279, 25134–25142 (2004).

    Article  CAS  PubMed  Google Scholar 

  149. Dukuzumuremyi, J. M. et al. The Yersinia protein kinase A is a host factor inducible RhoA/Rac-binding virulence factor. J. Biol. Chem. 275, 35281–35290 (2000).

    Article  CAS  PubMed  Google Scholar 

  150. Bergsbaken, T. & Cookson, B. T. Macrophage activation redirects yersinia-infected host cell death from apoptosis to caspase-1-dependent pyroptosis. PLoS Pathog. 3, e161 (2007).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

Download references

Acknowledgements

We thank Suzana Salcedo, Stéphane Méresse, Hughes Lelouard, Steve Garvis and Chantal de Chastellier for critical reading of the manuscript. We apologize to all those authors whose work could not be cited because of space limitations. We are grateful to Chantal de Chastellier and Stéphane Méresse for electron microscopy and immunofluorescence micrographs. L.D. is a fellow the European Molecular Biology Organization. This work has been supported by institutional grants from Institut National de la Santé et de la Recherche Médicale and Centre National de la Recherche Scientifique.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Jean-Pierre Gorvel.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Related links

Related links

DATABASES

Entrez Genome Project

Brucella abortus

Brucella melitensis

Chlamydia trachomatis

Escherichia coli

Francisella tularensis

Legionella pneumophila

Listeria monocytogenes

Mycobacterium bovis

Mycobacterium tuberculosis

S. enterica subsp. enterica serovar Typhimurium

Shigella flexneri

Yersinia pestis

FURTHER INFORMATION

Jean-Pierre Gorvel's homepage

Glossary

Pattern recognition receptor

A receptor of the innate immune system that recognizes and responds to conserved microorganism-associated molecular patterns.

NOD-like receptor

An intracellular sensor that detects cytosolic microbial components, cell injury and 'danger' signals (such as ATP and toxins).

Inflammasome

A molecular complex of several proteins that on assembly cleaves pro-interleukin-1 (IL-1), thereby producing active IL-1. Based on the pyrin-domain containing Nod-like receptors (NLR) NALP1 and NALP3, which require ASC to bridge the NLRs to pro-caspase 1.

Muramyl dipeptide

A peptidoglycan constituent of both Gram-positive and Gram-negative bacteria. It is composed of N-acetylmuramic acid linked by its lactic acid moiety to the amino terminus of an L-alanine D-isoglutamine dipeptide.

Reactive oxygen species

These include superoxide (O2·), hydroxyl radicals (OH) and hydrogen peroxide (H2O2). They are generated as products of normal respiration from the electron transport chain that is present in the mitochondria or in the endoplasmic reticulum, or they can be catalysed by a wide array of enzymes (NADPH oxidase, xanthine oxidase, peroxidases and NADPH oxidase isoform).

Reactive nitrogen intermediates

These include nitric oxide (NO·) and its derivates such as NONOates, S-nitrosothiols, peroxynitrite, nitrite and nitrous acid. NO· is a highly reactive and diffusible free radical, soluble in both lipids and water, and capable of reacting with oxygen and reactive oxygen species to form reactive nitrogen intermediates, NO2, NO2, NO3, N2O3, and the highly mycobactericidal ONOO.

Proteasome

A giant multicatalytic protease that resides in the cytosol and nucleus.

Caspase 1

A cysteine protease that contains a cysteine residue in the active site and that cleaves its substrate after an aspartic acid residue. It was previously known as interleukin-1 β-converting enzyme.

Pyroptosis

An inflammatory process of cellular self-destruction that causes cell lysis and the secretion of interleukin-1β (IL-1β) and IL-18, in a caspase 1-dependent way.

Pathogenicity island

A large region of genomic DNA that encodes genes associated with virulence. A pathogenicity island is typically transferred horizontally between bacterial strains and is often inserted into tRNA genes in the genome.

Arp2/3 complex

A complex that consists of seven subunits comprising two Arp proteins and five highly conserved protein subunits: p16, p20, p21, p34 and p40e. The Arp2 and Arp3 subunits are structurally related to actin.

T4SS

(Type IV secretion system). A well-characterized secretion apparatus, encoded by the virB operon from Agrobacterium spp., which can transport DNA and proteins into host cells. Helicobacter pylori, Bordetella pertussis, Brucella spp. and Legionella pneumophila have homologous operons that are used to secrete proteins and toxins into the eukaryotic cytoplasm.

NADPH oxidase

An enzyme composed of five core polypeptides, p22phox, gp91phox p47phox, p67phox and p40phox, and two low-molecular-mass guanine nucleotide-binding proteins, Rac (p21rac) and Rap1A.

Nitric oxide synthase

(NOS). An enzyme that uses electrons derived from NADPH to convert (oxidize) L-arginine to L-citrulline and NO·. Three NOS isoforms are known, NOS1, NOS2 and NOS3. NOS2 has also been termed iNOS, for inducible and independent (of increased intracellular Ca2+) and was first cloned from mouse macrophage cell line RAW 264.7.

Ezrin–radixin–moesin-binding phosphoprotein 50

A scaffolding protein that is responsible for the anchoring of various cellular proteins to the actin cytoskeleton through its linkage to ezrin, radixin and moesin.

Autophagosome

An intracytoplasmic vacuole that contains components of the cytoplasm. It fuses with a lysosome to form an autophagolysosome, thereby subjecting its contents to enzymatic digestion.

Apoptosis-associated speck-like protein containing a CARD

(ASC). A protein that contains a CARD domain in its carboxy-terminal region and a pyrin domain in the amino-terminal region. Both domains allow ASC to recruit other pyrin domain- and CARD-containing proteins through homotypic protein–protein interactions.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Diacovich, L., Gorvel, JP. Bacterial manipulation of innate immunity to promote infection. Nat Rev Microbiol 8, 117–128 (2010). https://doi.org/10.1038/nrmicro2295

Download citation

  • Issue Date:

  • DOI: https://doi.org/10.1038/nrmicro2295

This article is cited by

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

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