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:

Effector-triggered immunity and pathogen sensing in metazoans

A Publisher Correction to this article was published on 10 February 2020

This article has been updated

Abstract

Microbial pathogens possess an arsenal of strategies to invade their hosts, evade immune defences and promote infection. In particular, bacteria use virulence factors, such as secreted toxins and effector proteins, to manipulate host cellular processes and establish a replicative niche. Survival of eukaryotic organisms in the face of such challenge requires host mechanisms to detect and counteract these pathogen-specific virulence strategies. In this Review, we focus on effector-triggered immunity (ETI) in metazoan organisms as a mechanism for pathogen sensing and distinguishing pathogenic from non-pathogenic microorganisms. For the purposes of this Review, we adopt the concept of ETI formulated originally in the context of plant pathogens and their hosts, wherein specific host proteins ‘guard’ central cellular processes and trigger inflammatory responses following pathogen-driven disruption of these processes. While molecular mechanisms of ETI are well-described in plants, our understanding of functionally analogous mechanisms in metazoans is still emerging. In this Review, we present an overview of ETI in metazoans and discuss recently described cellular processes that are guarded by the host. Although all pathogens manipulate host pathways, we focus primarily on bacterial pathogens and highlight pathways of effector-triggered immune defence that sense disruption of core cellular processes by pathogens. Finally, we discuss recent developments in our understanding of how pathogens can evade ETI to overcome these host adaptations.

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

Fig. 1: Host cells possess multiple mechanisms of pathogen detection and immune defence.
Fig. 2: Microbial threat checkpoints gauge the level of threat posed by a pathogen and fine-tune the host immune response.
Fig. 3: Effector-triggered immunity engages inflammasomes, but can also be targeted by other pathogen virulence factors.
Fig. 4: Pathogen manipulation of core cellular processes and signalling pathways induces host immune responses.

Similar content being viewed by others

Change history

  • 10 February 2020

    An amendment to this paper has been published and can be accessed via a link at the top of the paper.

References

  1. Janeway, C. A. Approaching the asymptote? Evolution and revolution in immunology. Cold Spring Harb. Symp. Quant. Biol. 54, 1–13 (1989).

    Article  CAS  PubMed  Google Scholar 

  2. Ausubel, F. M. Are innate immune signaling pathways in plants and animals conserved? Nat. Immunol. 6, 973 (2005).

    Article  CAS  PubMed  Google Scholar 

  3. Janeway, C. A. & Medzhitov, R. Innate immune recognition. Annu. Rev. Immunol. 20, 197–216 (2002).

    Article  CAS  PubMed  Google Scholar 

  4. Iwasaki, A. & Medzhitov, R. Control of adaptive immunity by the innate immune system. Nat. Immunol. 16, 343–353 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Pasare, C. & Medzhitov, R. Toll pathway-dependent blockade of CD4+CD25+ T cell-mediated suppression by dendritic cells. Science 299, 1033–1036 (2003).

    Article  CAS  PubMed  Google Scholar 

  6. Pasare, C. & Medzhitov, R. Toll-dependent control mechanisms of CD4 T cell activation. Immunity 21, 733–741 (2004).

    Article  CAS  PubMed  Google Scholar 

  7. Pasare, C. & Medzhitov, R. Control of B-cell responses by Toll-like receptors. Nature 438, 364–368 (2005).

    Article  CAS  PubMed  Google Scholar 

  8. Matzinger, P. Tolerance, danger, and the extended family. Annu. Rev. Immunol. 12, 991–1045 (1994).

    Article  CAS  PubMed  Google Scholar 

  9. Seong, S. Y. & Matzinger, P. Hydrophobicity: an ancient damage-associated molecular pattern that initiates innate immune responses. Nat. Rev. Immunol. 4, 469–478 (2004).

    Article  CAS  PubMed  Google Scholar 

  10. Kono, H. & Rock, K. L. How dying cells alert the immune system to danger. Nat. Rev. Immunol. 8, 279–289 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Vénéreau, E., Ceriotti, C. & Bianchi, M. E. DAMPs from cell death to new life. Front. Immunol. 6, 422 (2015).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  12. Chen, G. Y. & Nuñez, G. Sterile inflammation: sensing and reacting to damage. Nat. Rev. Immunol. 10, 826–837 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Lamkanfi, M. & Dixit, V. M. Inflammasomes: guardians of cytosolic sanctity. Immunol. Rev. 227, 95–105 (2009).

    Article  CAS  PubMed  Google Scholar 

  14. Vance, R. E., Isberg, R. R. & Portnoy, D. A. Patterns of pathogenesis: discrimination of pathogenic and nonpathogenic microbes by the innate immune system. Cell Host Microbe 6, 10–21 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Liston, A. & Masters, S. L. Homeostasis-altering molecular processes as mechanisms of inflammasome activation. Nat. Rev. Immunol. 17, 208 (2017).

    Article  CAS  PubMed  Google Scholar 

  16. Blander, J. M. & Sander, L. E. Beyond pattern recognition: five immune checkpoints for scaling the microbial threat. Nat. Rev. Immunol. 12, 215–225 (2012).

    Article  CAS  PubMed  Google Scholar 

  17. Evavold, C. L. & Kagan, J. C. Inflammasomes: threat-assessment organelles of the innate immune system. Immunity 51, 609–624 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Jones, J. D. & Dangl, J. L. The plant immune system. Nature 444, 323–329 (2006).

    Article  CAS  PubMed  Google Scholar 

  19. Chisholm, S. T., Coaker, G., Day, B. & Staskawicz, B. J. Host–microbe interactions: shaping the evolution of the plant immune response. Cell 124, 803–814 (2006).

    Article  CAS  PubMed  Google Scholar 

  20. Galan, J. E. & Waksman, G. Protein-injection machines in bacteria. Cell 172, 1306–1318 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Stuart, L. M., Paquette, N. & Boyer, L. Effector-triggered versus pattern-triggered immunity: how animals sense pathogens. Nat. Rev. Immunol. 13, 199–206 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Rajamuthiah, R. & Mylonakis, E. Effector triggered immunity: activation of innate immunity in metazoans by bacterial effectors. Virulence 5, 697–702 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Dangl, J. L. & Jones, J. D. Plant pathogens and integrated defence responses to infection. Nature 411, 826–833 (2001).

    Article  CAS  PubMed  Google Scholar 

  24. Jones, J. D., Vance, R. E. & Dangl, J. L. Intracellular innate immune surveillance devices in plants and animals. Science 354, aaf6395 (2016).

    Article  PubMed  CAS  Google Scholar 

  25. Cookson, B. T. & Brennan, M. A. Pro-inflammatory programmed cell death. Trends Microbiol. 9, 113–114 (2001).

    Article  CAS  PubMed  Google Scholar 

  26. Scheel, D. Resistance response physiology and signal transduction. Curr. Opin. Plant Biol. 1, 305–310 (1998).

    Article  CAS  PubMed  Google Scholar 

  27. Lamkanfi, M. & Dixit, V. M. Mechanisms and functions of inflammasomes. Cell 157, 1013–1022 (2014).

    Article  CAS  PubMed  Google Scholar 

  28. Friebe, S., van der Goot, F. G. & Bürgi, J. The ins and outs of anthrax toxin. Toxins 8, 69 (2016).

    Article  PubMed Central  CAS  Google Scholar 

  29. Friedlander, A. M., Bhatnagar, R., Leppla, S. H., Johnson, L. & Singh, Y. Characterization of macrophage sensitivity and resistance to anthrax lethal toxin. Infect. Immun. 61, 245–252 (1993).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Boyden, E. D. & Dietrich, W. F. Nalp1b controls mouse macrophage susceptibility to anthrax lethal toxin. Nat. Genet. 38, 240–244 (2006).

    Article  CAS  PubMed  Google Scholar 

  31. 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 

  32. Frew, B. C., Joag, V. R. & Mogridge, J. Proteolytic processing of Nlrp1b is required for inflammasome activity. PLoS Pathog. 8, e1002659 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Levinsohn, J. L. et al. Anthrax lethal factor cleavage of Nlrp1 is required for activation of the inflammasome. PLoS Pathog. 8, e1002638 (2012).

    Article  PubMed  PubMed Central  Google Scholar 

  34. Chavarría-Smith, J. & Vance, R. E. Direct proteolytic cleavage of NLRP1B is necessary and sufficient for inflammasome activation by anthrax lethal factor. PLoS Pathog. 9, e1003452 (2013).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  35. Okondo, M. C. et al. Inhibition of Dpp8/9 activates the Nlrp1b inflammasome. Cell Chem. Biol. 25, 262–267 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Sandstrom, A. et al. Functional degradation: a mechanism of NLRP1 inflammasome activation by diverse pathogen enzymes. Science 5, eaau1330 (2019).

    Article  CAS  Google Scholar 

  37. Ewald, S. E., Chavarria-Smith, J. & Boothroyd, J. C. NLRP1 is an inflammasome sensor for Toxoplasma gondii. Infect. Immun. 82, 460–468 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  38. Cirelli, K. M. et al. Inflammasome sensor NLRP1 controls rat macrophage susceptibility to Toxoplasma gondii. PLoS Pathog. 10, e1003927 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  39. Gorfu, G. et al. Dual role for inflammasome sensors NLRP1 and NLRP3 in murine resistance to Toxoplasma gondii. mBio 5, e01117–13 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  40. Moayeri, M., Sastalla, I. & Leppla, S. H. Anthrax and the inflammasome. Microbes Infect. 14, 392–400 (2012).

    Article  CAS  PubMed  Google Scholar 

  41. Chavarría-Smith, J., Mitchell, P. S., Ho, A. M., Daugherty, M. D. & Vance, R. E. Functional and evolutionary analyses identify proteolysis as a general mechanism for NLRP1 inflammasome activation. PLoS Pathog. 12, e1006052 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  42. Zhong, F. L. et al. Germline NLRP1 mutations cause skin inflammatory and cancer susceptibility syndromes via inflammasome activation. Cell 167, 187–202 (2016).

    Article  CAS  PubMed  Google Scholar 

  43. Jin, Y. et al. NALP1 in vitiligo-associated multiple autoimmune disease. N. Engl. J. Med. 356, 1216–1225 (2007).

    Article  CAS  PubMed  Google Scholar 

  44. Muñoz-Planillo, R. et al. K+ efflux is the common trigger of NLRP3 inflammasome activation by bacterial toxins and particulate matter. Immunity 38, 1142–1153 (2013).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  45. Leemans, J. C., Cassel, S. L. & Sutterwala, F. S. Sensing damage by the NLRP3 inflammasome. Immunol. Rev. 243, 152–162 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. 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 

  47. Abrami, L., Fivaz, M. & van der Goot, F. G. Adventures of a pore-forming toxin at the target cell surface. Trends Microbiol. 8, 168–172 (2000).

    Article  CAS  PubMed  Google Scholar 

  48. Bhakdi, S. & Tranum-Jensen, J. Alpha-toxin of Staphylococcus aureus. Microbiol. Rev. 55, 733–751 (1991).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Pilla, D. M. et al. Guanylate binding proteins promote caspase-11-dependent pyroptosis in response to cytoplasmic LPS. Proc. Natl Acad. Sci. USA 111, 6046–6051 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Kim, B.-H. et al. Interferon-induced guanylate-binding proteins in inflammasome activation and host defense. Nat. Immunol. 17, 481 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Man, S. M., Place, D. E., Kuriakose, T. & Kanneganti, T. D. Interferon‐inducible guanylate‐binding proteins at the interface of cell‐autonomous immunity and inflammasome activation. J. Leukoc. Biol. 101, 143–150 (2017).

    Article  CAS  PubMed  Google Scholar 

  52. Meunier, E. & Broz, P. Interferon‐inducible GTPases in cell autonomous and innate immunity. Cell. Microbiol. 18, 168–180 (2016).

    Article  CAS  PubMed  Google Scholar 

  53. Pilla-Moffett, D., Barber, M. F., Taylor, G. A. & Coers, J. Interferon-inducible GTPases in host resistance, inflammation and disease. J. Mol. Biol. 428, 3495–3513 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Feeley, E. M. et al. Galectin-3 directs antimicrobial guanylate binding proteins to vacuoles furnished with bacterial secretion systems. Proc. Natl Acad. Sci. USA 114, E1698–E1706 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Zwack, E. E. et al. Guanylate binding proteins regulate inflammasome activation in response to hyperinjected yersinia translocon components. Infect. Immun. 85, e00778–16 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Viboud, G. I. & Bliska, J. B. Yersinia outer proteins: role in modulation of host cell signaling responses and pathogenesis. Annu. Rev. Microbiol. 59, 69–89 (2005).

    Article  CAS  PubMed  Google Scholar 

  57. Neyt, C. & Cornelis, G. R. Insertion of a Yop translocation pore into the macrophage plasma membrane by Yersinia enterocolitica: requirement for translocators YopB and YopD, but not LcrG. Mol. Microbiol. 33, 971–981 (1999).

    Article  CAS  PubMed  Google Scholar 

  58. Nordfelth, R. & Wolf-Watz, H. YopB of Yersinia enterocolitica is essential for YopE translocation. Infect. Immun. 69, 3516–3518 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Rosqvist, R., Persson, C., Håkansson, S., Nordfeldt, R. & Wolf-Watz, H. Translocation of the Yersinia YopE and YopH virulence proteins into target cells is mediated by YopB and YopD. Contrib. Microbiol. 13, 230–234 (1995).

    CAS  Google Scholar 

  60. Dewoody, R., Merritt, P. M. & Marketon, M. M. YopK controls both rate and fidelity of Yop translocation. Mol. Microbiol. 87, 301–317 (2013).

    Article  CAS  PubMed  Google Scholar 

  61. Francis, M. S. & Wolf-Watz, H. YopD of Yersinia pseudotuberculosis is translocated into the cytosol of HeLa epithelial cells: evidence of a structural domain necessary for translocation. Mol. Microbiol. 29, 799–813 (1998).

    Article  CAS  PubMed  Google Scholar 

  62. Zwack, E. E. et al. Inflammasome activation in response to the Yersinia type III secretion system requires hyperinjection of translocon proteins YopB and YopD. mBio 6, e02095–14 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Brodsky, I. E. et al. A Yersinia effector protein promotes virulence by preventing inflammasome recognition of the type III secretion system. Cell Host Microbe 7, 376–387 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Coers, J. Sweet host revenge: Galectins and GBPs join forces at broken membranes. Cell Microbiol. 19, e12793 (2017).

    Article  CAS  Google Scholar 

  65. Rühl, S. & Broz, P. Caspase-11 activates a canonical NLRP3 inflammasome by promoting K(+) efflux. Eur. J. Immunol. 45, 2927–2936 (2015).

    Article  PubMed  CAS  Google Scholar 

  66. 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 

  67. Man, S. M. et al. IRGB10 liberates bacterial ligands for sensing by the AIM2 and caspase-11-NLRP3 inflammasomes. Cell 167, 382–396 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Finethy, R. et al. Guanylate binding proteins enable rapid activation of canonical and noncanonical inflammasomes in Chlamydia-infected macrophages. Infect. Immun. 83, 4740–4749 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Holmström, A. et al. YopK of Yersinia pseudotuberculosis controls translocation of Yop effectors across the eukaryotic cell membrane. Mol. Microbiol. 24, 73–91 (1997).

    Article  PubMed  Google Scholar 

  70. Sakaguchi, T., Leser, G. P. & Lamb, R. A. The ion channel activity of the influenza virus M2 protein affects transport through the Golgi apparatus. J. Cell. Biol. 133, 733–747 (1996).

    Article  CAS  PubMed  Google Scholar 

  71. Ichinohe, T., Pang, I. K. & Iwasaki, A. Influenza virus activates inflammasomes via its intracellular M2 ion channel. Nat. Immunol. 11, 404–410 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Chen, J. & Chen, Z. J. PtdIns4P on dispersed trans-Golgi network mediates NLRP3 inflammasome activation. Nature 564, 71 (2018).

    Article  CAS  PubMed  Google Scholar 

  73. Ofengeim, D. & Yuan, J. Regulation of RIP1 kinase signalling at the crossroads of inflammation and cell death. Nat. Rev. Mol. Cell. Biol. 14, 727–736 (2017).

    Article  CAS  Google Scholar 

  74. Hsu, H., Huang, J., Shu, H. B., Baichwal, V. & Goeddel, D. V. TNF-dependent recruitment of the protein kinase RIP to the TNF receptor-1 signaling complex. Immunity 4, 387–396 (1996).

    Article  CAS  PubMed  Google Scholar 

  75. Kelliher, M. A. et al. The death domain kinase RIP mediates the TNF-induced NF-kappaB signal. Immunity 8, 297–303 (1998).

    Article  CAS  PubMed  Google Scholar 

  76. Meylan, E. & Tschopp, J. The RIP kinases: crucial integrators of cellular stress. Trends Biochem. Sci. 30, 151–159 (2005).

    Article  CAS  PubMed  Google Scholar 

  77. Cusson-Hermance, N., Khurana, S., Lee, T. H., Fitzgerald, K. A. & Kelliher, M. A. Rip1 mediates the Trif-dependent toll-like receptor 3- and 4-induced NF-{kappa}B activation but does not contribute to interferon regulatory factor 3 activation. J. Biol. Chem. 280, 36560–36566 (2005).

    Article  CAS  PubMed  Google Scholar 

  78. Peterson, L. W. et al. RIPK1-dependent apoptosis bypasses pathogen blockade of innate signaling to promote immune defense. J. Exp. Med. 214, 3171–3182 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  79. Peterson, L. W. et al. Cell-extrinsic TNF collaborates with TRIF signaling to promote Yersinia-induced apoptosis. J. Immunol. 197, 4110–4117 (2016).

    Article  CAS  PubMed  Google Scholar 

  80. Yuan, J., Najafov, A. & Py, B. F. Roles of caspases in necrotic cell death. Cell 167, 1693–1704 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  81. Tenev, T. et al. The Ripoptosome, a signaling platform that assembles in response to genotoxic stress and loss of IAPs. Mol. Cell 43, 432–448 (2011).

    Article  CAS  PubMed  Google Scholar 

  82. Feoktistova, M. et al. cIAPs block Ripoptosome formation, a RIP1/caspase-8 containing intracellular cell death complex differentially regulated by cFLIP isoforms. Mol. Cell 43, 449–463 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  83. Moriwaki, K. & Chan, F. K. RIP3: a molecular switch for necrosis and inflammation. Genes Dev. 27, 1640–1649 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. Sarhan, J. et al. Caspase-8 induces cleavage of gasdermin D to elicit pyroptosis during Yersinia infection. Proc. Natl Acad. Sci. USA 115, E10888–E10897 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  85. Orning, P. et al. Pathogen blockade of TAK1 triggers caspase-8-dependent cleavage of gasdermin D and cell death. Science 362, 1064–1069 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  86. Newton, H. J. et al. The type III effectors NleE and NleB from enteropathogenic E. coli and OspZ from Shigella block nuclear translocation of NF-kappaB p65. PLoS Pathog. 6, e1000898 (2010).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  87. Ashida, H. et al. A bacterial E3 ubiquitin ligase IpaH9.8 targets NEMO/IKKgamma to dampen the host NF-kappaB-mediated inflammatory response. Nat. Cell Biol. 12, 66–73 (2010).

    Article  CAS  PubMed  Google Scholar 

  88. Orth, K. et al. Disruption of signaling by Yersinia effector YopJ, a ubiquitin-like protein protease. Science 290, 1594–1597 (2000).

    Article  CAS  PubMed  Google Scholar 

  89. Mukherjee, S., Hao, Y.-H. & Orth, K. A newly discovered post-translational modification–the acetylation of serine and threonine residues. Trends Biochem. Sci. 32, 210–216 (2007).

    Article  CAS  PubMed  Google Scholar 

  90. Cheong, M. S. et al. AvrBsT acetylates Arabidopsis ACIP1, a protein that associates with microtubules and is required for immunity. PLoS Pathog. 10, e1003952 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  91. Philip, N. H. et al. Caspase-8 mediates caspase-1 processing and innate immune defense in response to bacterial blockade of NF-κB and MAPK signaling. Proc. Natl Acad. Sci. USA 111, 7385–7390 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  92. Weng, D. et al. Caspase-8 and RIP kinases regulate bacteria-induced innate immune responses and cell death. Proc. Natl Acad. Sci. USA 111, 7391–7396 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  93. Rosadini, C. V. et al. A single bacterial immune evasion strategy dismantles both MyD88 and TRIF signaling pathways downstream of TLR4. Cell Host Microbe 18, 682–693 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  94. Dondelinger, Y. et al. NF-kappaB-independent role of IKKalpha/IKKbeta in preventing RIPK1 kinase-dependent apoptotic and necroptotic cell death during TNF signaling. Mol. Cell 60, 63–76 (2015).

    Article  CAS  PubMed  Google Scholar 

  95. Dondelinger, Y. et al. Serine 25 phosphorylation inhibits RIPK1 kinase-dependent cell death in models of infection and inflammation. Nat. Commun. 10, 1729 (2019).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  96. Jaco, I. et al. MK2 Phosphorylates RIPK1 to prevent TNF-induced cell death. Mol. Cell 66, 698–710 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  97. Menon, M. B. et al. p38 MAPK/MK2-dependent phosphorylation controls cytotoxic RIPK1 signalling in inflammation and infection. Nat. Cell Biol. 19, 1248–1259 (2017).

    Article  CAS  PubMed  Google Scholar 

  98. Skaletskaya, A. et al. A cytomegalovirus-encoded inhibitor of apoptosis that suppresses caspase-8 activation. Proc. Natl Acad. Sci. USA 98, 7829–7834 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  99. Omoto, S. et al. Suppression of RIP3-dependent necroptosis by human cytomegalovirus. J. Biol. Chem. 290, 11635–11648 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  100. Upton, J. W., Kaiser, W. J. & Mocarski, E. S. DAI/ZBP1/DLM-1 complexes with RIP3 to mediate virus-induced programmed necrosis that is targeted by murine cytomegalovirus vIRA. Cell Host Microbe 11, 290–297 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  101. Pollock, G. L. et al. Distinct roles of the antiapoptotic effectors NleB and NleF from enteropathogenic Escherichia coli. Infect. Immun. 85, e01071–16 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  102. Pearson, J. S. et al. EspL is a bacterial cysteine protease effector that cleaves RHIM proteins to block necroptosis and inflammation. Nat. Microbiol. 2, 16258 (2017).

    Article  CAS  PubMed  Google Scholar 

  103. Lemichez, E. & Aktories, K. Hijacking of Rho GTPases during bacterial infection. Exp. Cell Res. 319, 2329–2336 (2013).

    Article  CAS  PubMed  Google Scholar 

  104. Xu, H. et al. Innate immune sensing of bacterial modifications of Rho GTPases by the Pyrin inflammasome. Nature 513, 237–241 (2014).

    Article  CAS  PubMed  Google Scholar 

  105. Park, Y. H., Wood, G., Kastner, D. L. & Chae, J. J. Pyrin inflammasome activation and RhoA signaling in the autoinflammatory diseases FMF and HIDS. Nat. Immunol. 17, 914–921 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  106. Gao, W., Yang, J., Liu, W., Wang, Y. & Shao, F. Site-specific phosphorylation and microtubule dynamics control Pyrin inflammasome activation. Proc. Natl Acad. Sci. USA 113, E4857–E4866 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  107. Aubert, D. F. et al. A Burkholderia type VI effector deamidates Rho GTPases to activate the Pyrin inflammasome and trigger inflammation. Cell Host Microbe 19, 664–674 (2016).

    Article  CAS  PubMed  Google Scholar 

  108. Chung, L. K. et al. The Yersinia virulence factor YopM hijacks host kinases to inhibit Type III effector-triggered activation of the Pyrin inflammasome. Cell Host Microbe 20, 296–306 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  109. Ratner, D. et al. The Yersinia pestis effector YopM inhibits pyrin inflammasome activation. PLoS Pathog. 12, e1006035 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  110. Medici, N. P., Rashid, M. & Bliska, J. B. Characterization of pyrin dephosphorylation and inflammasome activation in macrophages as triggered by the Yersinia effectors YopE and YopT. Infect. Immun. 87, e00822–18 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  111. Heilig, R. & Broz, P. Function and mechanism of the pyrin inflammasome. Eur. J. Immunol. 48, 230–238 (2018).

    Article  CAS  PubMed  Google Scholar 

  112. Schaner, P. et al. Episodic evolution of pyrin in primates: human mutations recapitulate ancestral amino acid states. Nat. Genet. 27, 318–321 (2001).

    Article  CAS  PubMed  Google Scholar 

  113. Jamilloux, Y., Magnotti, F., Belot, A. & Henry, T. The pyrin inflammasome: from sensing RhoA GTPases-inhibiting toxins to triggering autoinflammatory syndromes. Pathog. Dis. 76, fty020 (2018).

    Article  CAS  Google Scholar 

  114. Boyer, L. et al. Pathogen-derived effectors trigger protective immunity via activation of the Rac2 enzyme and the IMD or Rip kinase signaling pathway. Immunity 35, 536–549 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  115. Keestra, A. M. et al. Manipulation of small Rho GTPases is a pathogen-induced process detected by NOD1. Nature 496, 233–237 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  116. Bruno, V. M. et al. Salmonella Typhimurium type III secretion effectors stimulate innate immune responses in cultured epithelial cells. PLoS Pathog. 5, e1000538 (2009).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  117. Sun, H., Kamanova, J., Lara-Tejero, M. & Galán, J. E. Salmonella stimulates pro-inflammatory signalling through p21-activated kinases bypassing innate immune receptors. Nat. Microbiol. 3, 1122 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  118. Bielig, H. et al. The cofilin phosphatase slingshot homolog 1 (SSH1) links NOD1 signaling to actin remodeling. PLoS Pathog. 10, e1004351 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  119. Fukazawa, A. et al. GEF-H1 mediated control of NOD1 dependent NF-κB activation by Shigella effectors. PLoS Pathog. 4, e1000228 (2008).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  120. Legrand-Poels, S. et al. Modulation of Nod2-dependent NF-κB signaling by the actin cytoskeleton. J. Cell Sci. 120, 1299–1310 (2007).

    Article  CAS  PubMed  Google Scholar 

  121. Eitel, J. et al. β-PIX and Rac1 GTPase mediate trafficking and negative regulation of NOD2. J. Immunol. 181, 2664–2671 (2008).

    Article  CAS  PubMed  Google Scholar 

  122. Pakos‐Zebrucka, K. et al. The integrated stress response. EMBO Rep. 17, 1374–1395 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  123. Tsalikis, J., Croitoru, D. O., Philpott, D. J. & Girardin, S. E. Nutrient sensing and metabolic stress pathways in innate immunity. Cell. Microbiol. 15, 1632–1641 (2013).

    CAS  PubMed  Google Scholar 

  124. Wolfson, R. L. & Sabatini, D. M. The dawn of the age of amino acid sensors for the mTORC1 pathway. Cell Metabol. 26, 301–309 (2017).

    Article  CAS  Google Scholar 

  125. Tattoli, I. et al. Listeria phospholipases subvert host autophagic defenses by stalling pre-autophagosomal structures. EMBO J. 32, 3066–3078 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  126. 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 

  127. Tattoli, I., Philpott, D. J. & Girardin, S. E. The bacterial and cellular determinants controlling the recruitment of mTOR to the Salmonella-containing vacuole. Biol. Open 1, 1215–1225 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  128. Wyant, G. A. et al. mTORC1 activator SLC38A9 is required to efflux essential amino acids from lysosomes and use protein as a nutrient. Cell 171, 642–654 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  129. Wang, Y., Weiss, L. M. & Orlofsky, A. Intracellular parasitism with Toxoplasma gondii stimulates mTOR-dependent host cell growth despite impaired signaling to S6K1 and 4E-BP1. Cell. Microbiol. 11, 983 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  130. Soares, J. A. et al. Activation of the PI3K/Akt pathway early during vaccinia and cowpox virus infections is required for both host survival and viral replication. J. Virol. 83, 6883–6899 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  131. Clippinger, A. J., Maguire, T. G. & Alwine, J. C. Human cytomegalovirus infection maintains mTOR activity and its perinuclear localization during amino acid deprivation. J. Virol. 85, 9369–9376 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  132. Surviladze, Z., Sterk, R. T., DeHaro, S. A. & Ozbun, M. A. Cellular entry of human papillomavirus type 16 involves activation of the phosphatidylinositol 3-kinase/Akt/mTOR pathway and inhibition of autophagy. J. Virol. 87, 2508–2517 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  133. Celli, J. & Tsolis, R. M. Bacteria, the endoplasmic reticulum and the unfolded protein response: friends or foes? Nat. Rev. Microbiol. 13, 71–82 (2015).

    Article  CAS  PubMed  Google Scholar 

  134. Walter, P. & Ron, D. The unfolded protein response: from stress pathway to homeostatic regulation. Science 334, 1081–1086 (2011).

    Article  CAS  PubMed  Google Scholar 

  135. Isberg, R. R., O’Connor, T. J. & Heidtman, M. The Legionella pneumophila replication vacuole: making a cosy niche inside host cells. Nat. Rev. Microbiol. 7, 13–24 (2009).

    Article  CAS  PubMed  Google Scholar 

  136. Hempstead, A. D. & Isberg, R. R. Inhibition of host cell translation elongation by Legionella pneumophila blocks the host cell unfolded protein response. Proc. Natl Acad. Sci. USA 112, E6790–E6797 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  137. Treacy-Abarca, S. & Mukherjee, S. Legionella suppresses the host unfolded protein response via multiple mechanisms. Nat. Commun. 6, 7887 (2015).

    Article  CAS  PubMed  Google Scholar 

  138. Fontana, M. F. et al. Secreted bacterial effectors that inhibit host protein synthesis are critical for induction of the innate immune response to virulent Legionella pneumophila. PLoS Pathog. 7, e1001289 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  139. Belyi, Y. et al. Legionella pneumophila glucosyltransferase inhibits host elongation factor 1A. Proc. Natl Acad. Sci. USA 103, 16953–16958 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  140. Belyi, Y., Tabakova, I., Stahl, M. & Aktories, K. Lgt: a family of cytotoxic glucosyltransferases produced by Legionella pneumophila. J. Bacteriol. 190, 3026–3035 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  141. De Leon, J. A. et al. Positive and negative regulation of the master metabolic regulator mTORC1 by two families of Legionella pneumophila effectors. Cell Rep. 21, 2031–2038 (2017).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  142. Shen, X. et al. Targeting eEF1A by a Legionella pneumophila effector leads to inhibition of protein synthesis and induction of host stress response. Cell. Microbiol. 11, 911–926 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  143. Barry, K. C., Fontana, M. F., Portman, J. L., Dugan, A. S. & Vance, R. E. IL-1α signaling initiates the inflammatory response to virulent Legionella pneumophila in vivo. J. Immunol. 190, 6329–6639 (2013).

    Article  CAS  PubMed  Google Scholar 

  144. Barry, K. C., Ingolia, N. T. & Vance, R. E. Global analysis of gene expression reveals mRNA superinduction is required for the inducible immune response to a bacterial pathogen. eLife 6, e22707 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  145. Ivanov, S. S. & Roy, C. R. Pathogen signatures activate a ubiquitination pathway that modulates the function of the metabolic checkpoint kinase mTOR. Nat. Immunol. 14, 1219 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  146. Fontana, M. F., Shin, S. & Vance, R. E. Activation of host mitogen-activated protein kinases by secreted Legionella pneumophila effectors that inhibit host protein translation. Infect. Immun. 80, 3570–3575 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  147. Shin, S. et al. Type IV secretion-dependent activation of host MAP kinases induces an increased proinflammatory cytokine response to Legionella pneumophila. PLoS Pathog. 4, e1000220 (2008).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  148. Asrat, S., Dugan, A. S. & Isberg, R. R. The frustrated host response to Legionella pneumophila is bypassed by MyD88-dependent translation of pro-inflammatory cytokines. PLoS Pathog. 10, e1004229 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  149. Copenhaver, A. M., Casson, C. N., Nguyen, H. T., Duda, M. M. & Shin, S. IL-1R signaling enables bystander cells to overcome bacterial blockade of host protein synthesis. Proc. Natl Acad. Sci. USA 112, 7557–7562 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  150. Reddy, K. C., Dunbar, T. L., Nargund, A. M., Haynes, C. M. & Troemel, E. R. The C. elegans CCAAT-enhancer-binding protein gamma is required for surveillance immunity. Cell Rep. 14, 1581–1589 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  151. Pukkila-Worley, R. Surveillance immunity: an emerging paradigm of innate defense activation in Caenorhabditis elegans. PLoS Pathog. 12, e1005795 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  152. Melo, J. A. & Ruvkun, G. Inactivation of conserved C. elegans genes engages pathogen-and xenobiotic-associated defenses. Cell 149, 452–466 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  153. Ogata, M., Chaudhary, V. K., Pastan, I. & FitzGerald, D. J. Processing of Pseudomonas exotoxin by a cellular protease results in the generation of a 37,000-Da toxin fragment that is translocated to the cytosol. J. Biol. Chem. 265, 20678–20685 (1990).

    Article  CAS  PubMed  Google Scholar 

  154. Estes, K. A., Dunbar, T. L., Powell, J. R., Ausubel, F. M. & Troemel, E. R. bZIP transcription factor zip-2 mediates an early response to Pseudomonas aeruginosa infection in Caenorhabditis elegans. Proc. Natl Acad. Sci. USA 107, 2153–2158 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  155. Dunbar, T. L., Yan, Z., Balla, K. M., Smelkinson, M. G. & Troemel, E. R. C. elegans detects pathogen-induced translational inhibition to activate immune signaling. Cell Host Microbe 11, 375–386 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  156. McEwan, D. L., Kirienko, N. V. & Ausubel, F. M. Host translational inhibition by Pseudomonas aeruginosa Exotoxin A triggers an immune response in Caenorhabditis elegans. Cell Host Microbe 11, 364–374 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  157. Abastado, J., Miller, P. F., Jackson, B. M. & Hinnebusch, A. G. Suppression of ribosomal reinitiation at upstream open reading frames in amino acid-starved cells forms the basis for GCN4 translational control. Mol. Cell. Biol. 11, 486–496 (1991).

    CAS  PubMed  PubMed Central  Google Scholar 

  158. Fernandez, J. et al. Ribosome stalling regulates IRES-mediated translation in eukaryotes, a parallel to prokaryotic attenuation. Mol. Cell 17, 405–416 (2005).

    Article  CAS  PubMed  Google Scholar 

  159. Barbosa, C., Peixeiro, I. & Romão, L. Gene expression regulation by upstream open reading frames and human disease. PLoS Genet. 9, e1003529 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  160. Staskawicz, B. J., Dahlbeck, D. & Keen, N. T. Cloned avirulence gene of Pseudomonas syringae pv. glycinea determines race-specific incompatibility on Glycine max (L.) Merr. Proc. Natl Acad. Sci. USA 81, 6024–6028 (1984).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  161. Medzhitov, R. Approaching the asymptote: 20 years later. Immunity 30, 766–775 (2009).

    Article  CAS  PubMed  Google Scholar 

  162. Sandstrom, A. et al. Functional degradation: a mechanism of NLRP1 inflammasome activation by diverse pathogen enzymes. Science 364, eaau1330 (2018).

    Article  CAS  Google Scholar 

  163. Gurcel, L., Abrami, L., Girardin, S., Tschopp, J. & van der Goot, F. G. Caspase-1 activation of lipid metabolic pathways in response to bacterial pore-forming toxins promotes cell survival. Cell 126, 1135–1145 (2006).

    Article  CAS  PubMed  Google Scholar 

  164. Craven, R. R. et al. Staphylococcus aureus alpha-hemolysin activates the NLRP3-inflammasome in human and mouse monocytic cells. PLoS ONE 4, e7446 (2009).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  165. Komune, N., Ichinohe, T., Ito, M. & Yanagi, Y. Measles virus V protein inhibits NLRP3 inflammasome-mediated interleukin-1β secretion. J. Virol. 85, 13019–13026 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  166. Cheong, W. C. et al. Influenza A virus NS1 protein inhibits the NLRP3 inflammasome. PLoS ONE 10, e0126456 (2015).

    Article  PubMed Central  CAS  Google Scholar 

  167. Thorslund, S. E. et al. The RACK1 signaling scaffold protein selectively interacts with Yersinia pseudotuberculosis virulence function. PloS ONE 6, e16784 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  168. Jaramillo, M. et al. Leishmania repression of host translation through mTOR cleavage is required for parasite survival and infection. Cell Host Microbe 9, 331–341 (2011).

    Article  CAS  PubMed  Google Scholar 

  169. Dunn, E. F. & Connor, J. H. Dominant inhibition of Akt/protein kinase B signaling by the matrix protein of a negative-strand RNA virus. J. Virol. 85, 422–431 (2011).

    Article  CAS  PubMed  Google Scholar 

  170. Keestra-Gounder, A. M. et al. NOD1 and NOD2 signalling links ER stress with inflammation. Nature 532, 394–397 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  171. Haun, F. et al. Identification of a novel anoikis signalling pathway using the fungal virulence factor gliotoxin. Nat. Commun. 9, 3524 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  172. Gottar, M. et al. Dual detection of fungal infections in Drosophila via recognition of glucans and sensing of virulence factors. Cell 127, 1425–1437 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  173. Cui, H., Tsuda, K. & Parker, J. E. Effector-triggered immunity: from pathogen perception to robust defense. Annu. Rev. Plant Biol. 66, 487–511 (2015).

    Article  CAS  PubMed  Google Scholar 

  174. Flor, H. H. Current status of the gene-for-gene concept. Annu. Rev. Phytopathol. 9, 275–296 (1971).

    Article  Google Scholar 

  175. Lawrence, G. J., Dodds, P. N. & Ellis, J. G. Rust of flax and linseed caused by Melampsora lini. Mol. Plant Pathol. 8, 349–364 (2007).

    Article  PubMed  Google Scholar 

  176. Ellis, J. G., Dodds, P. N. & Lawrence, G. J. Flax rust resistance gene specificity is based on direct resistance-avirulence protein interactions. Annu. Rev. Phytopathol. 45, 289–306 (2007).

    Article  CAS  PubMed  Google Scholar 

  177. Van Der Biezen, E. A. & Jones, J. D. Plant disease-resistance proteins and the gene-for-gene concept. Trends Biochem. Sci. 23, 454–456 (1998).

    Article  PubMed  Google Scholar 

  178. Mackey, D., Holt, B. F. III, Wiig, A. & Dangl, J. L. RIN4 interacts with Pseudomonas syringae type III effector molecules and is required for RPM1-mediated resistance in Arabidopsis. Cell 108, 743–754 (2002).

    Article  CAS  PubMed  Google Scholar 

  179. Mackey, D., Belkhadir, Y., Alonso, J. M., Ecker, J. R. & Dangl, J. L. Arabidopsis RIN4 is a target of the type III virulence effector AvrRpt2 and modulates RPS2-mediated resistance. Cell 112, 379–389 (2003).

    Article  CAS  PubMed  Google Scholar 

  180. Axtell, M. J. & Staskawicz, B. J. Initiation of RPS2-specified disease resistance in Arabidopsis is coupled to the AvrRpt2-directed elimination of RIN4. Cell 112, 369–377 (2003).

    Article  CAS  PubMed  Google Scholar 

  181. Ade, J., DeYoung, B. J., Golstein, C. & Innes, R. W. Indirect activation of a plant nucleotide binding site-leucine-rich repeat protein by a bacterial protease. Proc. Natl Acad. Sci. USA 104, 2531–2536 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank members of the Brodsky and Shin laboratories for scientific discussion. We thank J. Kagan for scientific discussion and critical reading. Work in the Brodsky laboratory is supported by NIH/NIAID grants AI128530, AI139102 and AI135421. Work in the Shin laboratory is supported by NIH/NIAID grants AI118861 and AI123243, and the Linda Pechenik Montague Investigator Award from the University of Pennsylvania Perelman School of Medicine. Both I.E.B. and S.S. are recipients of the Burroughs-Wellcome Fund Investigators in the Pathogenesis of Infectious Disease Award. N.N. is a recipient of the American Heart Association Predoctoral Fellowship. N.L.F. is a recipient of the Institutional Ruth L. Kirschstein National Research Service Award T32GM07229 and the HHMI James H. Gilliam, Jr. Fellowship for Advanced Study.

Author information

Authors and Affiliations

Authors

Contributions

I.E.B. and S.S. developed the initial concept for the Review and conducted the literature search with N.L.F. and N.N.; N.L.F. and N.N. contributed equally to the writing of the manuscript. All authors contributed to revision and editing of the manuscript.

Corresponding authors

Correspondence to Sunny Shin or Igor E. Brodsky.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Lopes Fischer, N., Naseer, N., Shin, S. et al. Effector-triggered immunity and pathogen sensing in metazoans. Nat Microbiol 5, 14–26 (2020). https://doi.org/10.1038/s41564-019-0623-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41564-019-0623-2

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