1932

Abstract

Immune cells use a variety of membrane-disrupting proteins [complement, perforin, perforin-2, granulysin, gasdermins, mixed lineage kinase domain–like pseudokinase (MLKL)] to induce different kinds of death of microbes and host cells, some of which cause inflammation. After activation by proteolytic cleavage or phosphorylation, these proteins oligomerize, bind to membrane lipids, and disrupt membrane integrity. These membrane disruptors play a critical role in both innate and adaptive immunity. Here we review our current knowledge of the functions, specificity, activation, and regulation of membrane-disrupting immune proteins and what is known about the mechanisms behind membrane damage, the structure of the pores they form, how the cells expressing these lethal proteins are protected, and how cells targeted for destruction can sometimes escape death by repairing membrane damage.

Loading

Article metrics loading...

/content/journals/10.1146/annurev-immunol-111319-023800
2020-04-26
2024-04-25
Loading full text...

Full text loading...

/deliver/fulltext/immunol/38/1/annurev-immunol-111319-023800.html?itemId=/content/journals/10.1146/annurev-immunol-111319-023800&mimeType=html&fmt=ahah

Literature Cited

  1. 1. 
    Rosado CJ, Kondos S, Bull TE, Kuiper MJ, Law RH et al. 2008. The MACPF/CDC family of pore-forming toxins. Cell Microbiol 10:1765–74
    [Google Scholar]
  2. 2. 
    Dotiwala F, Lieberman J. 2019. Granulysin: killer lymphocyte safeguard against microbes. Curr. Opin. Immunol. 60:19–29
    [Google Scholar]
  3. 3. 
    Liu X, Lieberman J. 2017. A mechanistic understanding of pyroptosis: the fiery death triggered by invasive infection. Adv. Immunol. 135:81–117
    [Google Scholar]
  4. 4. 
    Shan B, Pan H, Najafov A, Yuan J 2018. Necroptosis in development and diseases. Genes Dev 32:327–40
    [Google Scholar]
  5. 5. 
    Thiery J, Lieberman J. 2014. Perforin: a key pore-forming protein for immune control of viruses and cancer. Subcell. Biochem. 80:197–220
    [Google Scholar]
  6. 6. 
    Weinlich R, Oberst A, Beere HM, Green DR 2017. Necroptosis in development, inflammation and disease. Nat. Rev. Mol. Cell Biol. 18:127–36
    [Google Scholar]
  7. 7. 
    Anderson DH, Sawaya MR, Cascio D, Ernst W, Modlin R et al. 2003. Granulysin crystal structure and a structure-derived lytic mechanism. J. Mol. Biol. 325:355–65
    [Google Scholar]
  8. 8. 
    Dudkina NV, Spicer BA, Reboul CF, Conroy PJ, Lukoyanova N et al. 2016. Structure of the poly-C9 component of the complement membrane attack complex. Nat. Commun. 7:10588
    [Google Scholar]
  9. 9. 
    Law RH, Lukoyanova N, Voskoboinik I, Caradoc-Davies TT, Baran K et al. 2010. The structural basis for membrane binding and pore formation by lymphocyte perforin. Nature 468:447–51
    [Google Scholar]
  10. 10. 
    Ruan J, Xia S, Liu X, Lieberman J, Wu H 2018. Cryo-EM structure of the gasdermin A3 membrane pore. Nature 557:62–67
    [Google Scholar]
  11. 11. 
    Serna M, Giles JL, Morgan BP, Bubeck D 2016. Structural basis of complement membrane attack complex formation. Nat. Commun. 7:10587
    [Google Scholar]
  12. 12. 
    Su L, Quade B, Wang H, Sun L, Wang X, Rizo J 2014. A plug release mechanism for membrane permeation by MLKL. Structure 22:1489–500
    [Google Scholar]
  13. 13. 
    Menny A, Serna M, Boyd CM, Gardner S, Joseph AP et al. 2018. CryoEM reveals how the complement membrane attack complex ruptures lipid bilayers. Nat. Commun. 9:5316
    [Google Scholar]
  14. 14. 
    Pang SS, Bayly-Jones C, Radjainia M, Spicer BA, Law RHP et al. 2019. The cryo-EM structure of the acid activatable pore-forming immune effector Macrophage-expressed gene 1. Nat. Commun. 10:4288
    [Google Scholar]
  15. 15. 
    Ding J, Wang K, Liu W, She Y, Sun Q et al. 2016. Pore-forming activity and structural autoinhibition of the gasdermin family. Nature 535:111–16
    [Google Scholar]
  16. 16. 
    Liu X, Zhang Z, Ruan J, Pan Y, Magupalli VG et al. 2016. Inflammasome-activated gasdermin D causes pyroptosis by forming membrane pores. Nature 535:153–58
    [Google Scholar]
  17. 17. 
    Pipkin ME, Lieberman J. 2007. Delivering the kiss of death: progress on understanding how perforin works. Curr. Opin. Immunol. 19:301–8
    [Google Scholar]
  18. 18. 
    Wang J, Deobald K, Re F 2019. Gasdermin D protects from melioidosis through pyroptosis and direct killing of bacteria. J. Immunol. 202:3468–73
    [Google Scholar]
  19. 19. 
    Chowdhury D, Lieberman J. 2008. Death by a thousand cuts: granzyme pathways of programmed cell death. Annu. Rev. Immunol. 26:389–420
    [Google Scholar]
  20. 20. 
    Rogers C, Fernandes-Alnemri T, Mayes L, Alnemri D, Cingolani G, Alnemri ES 2017. Cleavage of DFNA5 by caspase-3 during apoptosis mediates progression to secondary necrotic/pyroptotic cell death. Nat. Commun. 8:14128
    [Google Scholar]
  21. 21. 
    Wang Y, Gao W, Shi X, Ding J, Liu W et al. 2017. Chemotherapy drugs induce pyroptosis through caspase-3 cleavage of a gasdermin. Nature 547:99–103
    [Google Scholar]
  22. 22. 
    Antonopoulos C, Russo HM, El Sanadi C, Martin BN, Li X et al. 2015. Caspase-8 as an effector and regulator of NLRP3 inflammasome signaling. J. Biol. Chem. 290:20167–84
    [Google Scholar]
  23. 23. 
    Mascarenhas DPA, Cerqueira DM, Pereira MSF, Castanheira FVS, Fernandes TD et al. 2017. Inhibition of caspase-1 or gasdermin-D enable caspase-8 activation in the Naip5/NLRC4/ASC inflammasome. PLOS Pathog 13:e1006502
    [Google Scholar]
  24. 24. 
    Laudisi F, Spreafico R, Evrard M, Hughes TR, Mandriani B et al. 2013. Cutting edge: the NLRP3 inflammasome links complement-mediated inflammation and IL-1β release. J. Immunol. 191:1006–10
    [Google Scholar]
  25. 25. 
    Yao Y, Chen S, Cao M, Fan X, Yang T et al. 2017. Antigen-specific CD8+ T cell feedback activates NLRP3 inflammasome in antigen-presenting cells through perforin. Nat. Commun. 8:15402
    [Google Scholar]
  26. 26. 
    Conos SA, Chen KW, De Nardo D, Hara H, Whitehead L et al. 2017. Active MLKL triggers the NLRP3 inflammasome in a cell-intrinsic manner. PNAS 114:E961–69
    [Google Scholar]
  27. 27. 
    McCormack R, Bahnan W, Shrestha N, Boucher J, Barreto M et al. 2016. Perforin-2 protects host cells and mice by restricting the vacuole to cytosol transitioning of a bacterial pathogen. Infect. Immun. 84:1083–91
    [Google Scholar]
  28. 28. 
    Podack ER, Munson GP. 2016. Killing of microbes and cancer by the immune system with three mammalian pore-forming killer proteins. Front. Immunol. 7:464
    [Google Scholar]
  29. 29. 
    Zipfel PF, Skerka C. 2009. Complement regulators and inhibitory proteins. Nat. Rev. Immunol. 9:729–40
    [Google Scholar]
  30. 30. 
    Guo RF, Ward PA. 2005. Role of C5a in inflammatory responses. Annu. Rev. Immunol. 23:821–52
    [Google Scholar]
  31. 31. 
    Ricklin D, Hajishengallis G, Yang K, Lambris JD 2010. Complement: a key system for immune surveillance and homeostasis. Nat. Immunol. 11:785–97
    [Google Scholar]
  32. 32. 
    Reboul CF, Whisstock JC, Dunstone MA 2016. Giant MACPF/CDC pore forming toxins: a class of their own. Biochim. Biophys. Acta Biomembr. 1858:475–86
    [Google Scholar]
  33. 33. 
    Tschopp J, Muller-Eberhard HJ, Podack ER 1982. Formation of transmembrane tubules by spontaneous polymerization of the hydrophilic complement protein C9. Nature 298:534–38
    [Google Scholar]
  34. 34. 
    Anderluh G, Kisovec M, Krasevec N, Gilbert RJ 2014. Distribution of MACPF/CDC proteins. Subcell. Biochem. 80:7–30
    [Google Scholar]
  35. 35. 
    Botto M, Kirschfink M, Macor P, Pickering MC, Wurzner R, Tedesco F 2009. Complement in human diseases: lessons from complement deficiencies. Mol. Immunol. 46:2774–83
    [Google Scholar]
  36. 36. 
    Walport MJ. 2001. Complement: first of two parts. N. Engl. J. Med. 344:1058–66
    [Google Scholar]
  37. 37. 
    Lindahl G, Sjobring U, Johnsson E 2000. Human complement regulators: a major target for pathogenic microorganisms. Curr. Opin. Immunol. 12:44–51
    [Google Scholar]
  38. 38. 
    Pipkin ME, Ljutic B, Cruz-Guilloty F, Nouzova M, Rao A et al. 2007. Chromosome transfer activates and delineates a locus control region for perforin. Immunity 26:29–41
    [Google Scholar]
  39. 39. 
    Voskoboinik I, Whisstock JC, Trapani JA 2015. Perforin and granzymes: function, dysfunction and human pathology. Nat. Rev. Immunol. 15:388–400
    [Google Scholar]
  40. 40. 
    Shepard LA, Heuck AP, Hamman BD, Rossjohn J, Parker MW et al. 1998. Identification of a membrane-spanning domain of the thiol-activated pore-forming toxin Clostridium perfringens perfringolysin O: an α-helical to β-sheet transition identified by fluorescence spectroscopy. Biochemistry 37:14563–74
    [Google Scholar]
  41. 41. 
    Peters PJ, Borst J, Oorschot V, Fukuda M, Krahenbuhl O et al. 1991. Cytotoxic T lymphocyte granules are secretory lysosomes, containing both perforin and granzymes. J. Exp. Med. 173:1099–109
    [Google Scholar]
  42. 42. 
    Stinchcombe JC, Bossi G, Booth S, Griffiths GM 2001. The immunological synapse of CTL contains a secretory domain and membrane bridges. Immunity 15:751–61
    [Google Scholar]
  43. 43. 
    Keefe D, Shi L, Feske S, Massol R, Navarro F et al. 2005. Perforin triggers a plasma membrane-repair response that facilitates CTL induction of apoptosis. Immunity 23:249–62
    [Google Scholar]
  44. 44. 
    Thiery J, Keefe D, Boulant S, Boucrot E, Walch M et al. 2011. Perforin pores in the endosomal membrane trigger the release of endocytosed granzyme B into the cytosol of target cells. Nat. Immunol. 12:770–77
    [Google Scholar]
  45. 45. 
    Thiery J, Keefe D, Saffarian S, Martinvalet D, Walch M et al. 2010. Perforin activates clathrin- and dynamin-dependent endocytosis, which is required for plasma membrane repair and delivery of granzyme B for granzyme-mediated apoptosis. Blood 115:1582–93
    [Google Scholar]
  46. 46. 
    Bird CH, Sun J, Ung K, Karambalis D, Whisstock JC et al. 2005. Cationic sites on granzyme B contribute to cytotoxicity by promoting its uptake into target cells. Mol. Cell. Biol. 25:7854–67
    [Google Scholar]
  47. 47. 
    Shi L, Keefe D, Durand E, Feng H, Zhang D, Lieberman J 2005. Granzyme B binds to target cells mostly by charge and must be added at the same time as perforin to trigger apoptosis. J. Immunol. 174:5456–61
    [Google Scholar]
  48. 48. 
    Darmon AJ, Nicholson DW, Bleackley RC 1995. Activation of the apoptotic protease CPP32 by cytotoxic T-cell-derived granzyme B. Nature 377:446–48
    [Google Scholar]
  49. 49. 
    Yagi H, Conroy PJ, Leung EW, Law RH, Trapani JA et al. 2015. Structural basis for Ca2+-mediated interaction of the perforin C2 domain with lipid membranes. J. Biol. Chem. 290:25213–26
    [Google Scholar]
  50. 50. 
    Gilbert RJ. 2010. Cholesterol-Dependent Cytolysins New York: Springer
  51. 51. 
    Balaji KN, Schaschke N, Machleidt W, Catalfamo M, Henkart PA 2002. Surface cathepsin B protects cytotoxic lymphocytes from self-destruction after degranulation. J. Exp. Med. 196:493–503
    [Google Scholar]
  52. 52. 
    Baran K, Ciccone A, Peters C, Yagita H, Bird PI et al. 2006. Cytotoxic T lymphocytes from cathepsin B-deficient mice survive normally in vitro and in vivo after encountering and killing target cells. J. Biol. Chem. 281:30485–91
    [Google Scholar]
  53. 53. 
    Kagi D, Ledermann B, Burki K, Seiler P, Odermatt B et al. 1994. Cytotoxicity mediated by T cells and natural killer cells is greatly impaired in perforin-deficient mice. Nature 369:31–37
    [Google Scholar]
  54. 54. 
    van den Broek ME, Kagi D, Ossendorp F, Toes R, Vamvakas S et al. 1996. Decreased tumor surveillance in perforin-deficient mice. J. Exp. Med. 184:1781–90
    [Google Scholar]
  55. 55. 
    Zhang K, Jordan MB, Marsh RA, Johnson JA, Kissell D et al. 2011. Hypomorphic mutations in PRF1, MUNC13-4, and STXBP2 are associated with adult-onset familial HLH. Blood 118:5794–98
    [Google Scholar]
  56. 56. 
    Jenkins MR, Rudd-Schmidt JA, Lopez JA, Ramsbottom KM, Mannering SI et al. 2015. Failed CTL/NK cell killing and cytokine hypersecretion are directly linked through prolonged synapse time. J. Exp. Med. 212:307–17
    [Google Scholar]
  57. 57. 
    Xiong P, Shiratsuchi M, Matsushima T, Liao J, Tanaka E et al. 2017. Regulation of expression and trafficking of perforin-2 by LPS and TNF-α. Cell Immunol 320:1–10
    [Google Scholar]
  58. 58. 
    McCormack R, de Armas LR, Shiratsuchi M, Ramos JE, Podack ER 2013. Inhibition of intracellular bacterial replication in fibroblasts is dependent on the perforin-like protein (perforin-2) encoded by macrophage-expressed gene 1. J. Innate Immun. 5:185–94
    [Google Scholar]
  59. 59. 
    McCormack R, Podack ER. 2015. Perforin-2/Mpeg1 and other pore-forming proteins throughout evolution. J. Leukoc. Biol. 98:761–68
    [Google Scholar]
  60. 60. 
    Bai F, McCormack RM, Hower S, Plano GV, Lichtenheld MG, Munson GP 2018. Perforin-2 breaches the envelope of phagocytosed bacteria allowing antimicrobial effectors access to intracellular targets. J. Immunol. 201:2710–20
    [Google Scholar]
  61. 61. 
    Fields KA, McCormack R, de Armas LR, Podack ER 2013. Perforin-2 restricts growth of Chlamydia trachomatis in macrophages. Infect. Immun. 81:3045–54
    [Google Scholar]
  62. 62. 
    McCormack RM, Szymanski EP, Hsu AP, Perez E, Olivier KN et al. 2017. MPEG1/perforin-2 mutations in human pulmonary nontuberculous mycobacterial infections. JCI Insight 2:89635
    [Google Scholar]
  63. 63. 
    McCormack RM, de Armas LR, Shiratsuchi M, Fiorentino DG, Olsson ML et al. 2015. Perforin-2 is essential for intracellular defense of parenchymal cells and phagocytes against pathogenic bacteria. eLife 4:e06508
    [Google Scholar]
  64. 64. 
    Tamura M, Tanaka S, Fujii T, Aoki A, Komiyama H et al. 2007. Members of a novel gene family, Gsdm, are expressed exclusively in the epithelium of the skin and gastrointestinal tract in a highly tissue-specific manner. Genomics 89:618–29
    [Google Scholar]
  65. 65. 
    Burgener SS, Leborgne NGF, Snipas SJ, Salvesen GS, Bird PI, Benarafa C 2019. Cathepsin G inhibition by Serpinb1 and Serpinb6 prevents programmed necrosis in neutrophils and monocytes and reduces GSDMD-driven inflammation. Cell Rep 27:3646–56.e5
    [Google Scholar]
  66. 66. 
    Sollberger G, Choidas A, Burn GL, Habenberger P, Di Lucrezia R et al. 2018. Gasdermin D plays a vital role in the generation of neutrophil extracellular traps. Sci. Immunol. 3:eaar6689
    [Google Scholar]
  67. 67. 
    Sarhan J, Liu BC, Muendlein HI, Li P, Nilson R et al. 2018. Caspase-8 induces cleavage of gasdermin D to elicit pyroptosis during Yersinia infection. PNAS 115:E10888–97
    [Google Scholar]
  68. 68. 
    Orning P, Weng D, Starheim K, Ratner D, Best Z et al. 2018. Pathogen blockade of TAK1 triggers caspase-8-dependent cleavage of gasdermin D and cell death. Science 362:1064–69
    [Google Scholar]
  69. 69. 
    Chao KL, Kulakova L, Herzberg O 2017. Gene polymorphism linked to increased asthma and IBD risk alters gasdermin-B structure, a sulfatide and phosphoinositide binding protein. PNAS 114:E1128–37
    [Google Scholar]
  70. 70. 
    Collin RW, Kalay E, Oostrik J, Caylan R, Wollnik B et al. 2007. Involvement of DFNB59 mutations in autosomal recessive nonsyndromic hearing impairment. Hum. Mutat. 28:718–23
    [Google Scholar]
  71. 71. 
    Runkel F, Marquardt A, Stoeger C, Kochmann E, Simon D et al. 2004. The dominant alopecia phenotypes Bareskin, Rex-denuded, and Reduced Coat 2 are caused by mutations in gasdermin 3. . Genomics 84:824–35
    [Google Scholar]
  72. 72. 
    Van Laer L, Huizing EH, Verstreken M, van Zuijlen D, Wauters JG et al. 1998. Nonsyndromic hearing impairment is associated with a mutation in DFNA5. Nat. Genet 20:194–97
    [Google Scholar]
  73. 73. 
    Wu H, Romieu I, Sienra-Monge JJ, Li H, del Rio-Navarro BE, London SJ 2009. Genetic variation in ORM1-like 3 (ORMDL3) and gasdermin-like (GSDML) and childhood asthma. Allergy 64:629–35
    [Google Scholar]
  74. 74. 
    Li J, Zhou Y, Yang T, Wang N, Lian X, Yang L 2010. Gsdma3 is required for hair follicle differentiation in mice. Biochem. Biophys. Res. Commun. 403:18–23
    [Google Scholar]
  75. 75. 
    Tanaka S, Tamura M, Aoki A, Fujii T, Komiyama H et al. 2007. A new Gsdma3 mutation affecting anagen phase of first hair cycle. Biochem. Biophys. Res. Commun. 359:902–7
    [Google Scholar]
  76. 76. 
    Kumar S, Rathkolb B, Budde BS, Nurnberg P, de Angelis MH et al. 2012. Gsdma3I359N is a novel ENU-induced mutant mouse line for studying the function of Gasdermin A3 in the hair follicle and epidermis. J. Dermatol. Sci. 67:190–92
    [Google Scholar]
  77. 77. 
    Lunny DP, Weed E, Nolan PM, Marquardt A, Augustin M, Porter RM 2005. Mutations in gasdermin 3 cause aberrant differentiation of the hair follicle and sebaceous gland. J. Investig. Dermatol. 124:615–21
    [Google Scholar]
  78. 78. 
    Porter RM, Jahoda CA, Lunny DP, Henderson G, Ross J et al. 2002. Defolliculated (Dfl): a dominant mouse mutation leading to poor sebaceous gland differentiation and total elimination of pelage follicles. J. Investig. Dermatol. 119:32–37
    [Google Scholar]
  79. 79. 
    Shi P, Tang A, Xian L, Hou S, Zou D et al. 2015. Loss of conserved Gsdma3 self-regulation causes autophagy and cell death. Biochem. J. 468:325–36
    [Google Scholar]
  80. 80. 
    Kang MJ, Yu HS, Seo JH, Kim HY, Jung YH et al. 2012. GSDMB/ORMDL3 variants contribute to asthma susceptibility and eosinophil-mediated bronchial hyperresponsiveness. Hum. Immunol. 73:954–59
    [Google Scholar]
  81. 81. 
    Das S, Miller M, Beppu AK, Mueller J, McGeough MD et al. 2016. GSDMB induces an asthma phenotype characterized by increased airway responsiveness and remodeling without lung inflammation. PNAS 113:13132–37
    [Google Scholar]
  82. 82. 
    Delmaghani S, del Castillo FJ, Michel V, Leibovici M, Aghaie A et al. 2006. Mutations in the gene encoding pejvakin, a newly identified protein of the afferent auditory pathway, cause DFNB59 auditory neuropathy. Nat. Genet. 38:770–78
    [Google Scholar]
  83. 83. 
    Delmaghani S, Defourny J, Aghaie A, Beurg M, Dulon D et al. 2015. Hypervulnerability to sound exposure through impaired adaptive proliferation of peroxisomes. Cell 163:894–906
    [Google Scholar]
  84. 84. 
    Christofferson DE, Yuan J. 2010. Necroptosis as an alternative form of programmed cell death. Curr. Opin. Cell Biol. 22:263–68
    [Google Scholar]
  85. 85. 
    Murphy JM, Czabotar PE, Hildebrand JM, Lucet IS, Zhang JG et al. 2013. The pseudokinase MLKL mediates necroptosis via a molecular switch mechanism. Immunity 39:443–53
    [Google Scholar]
  86. 86. 
    Cai Z, Jitkaew S, Zhao J, Chiang HC, Choksi S et al. 2014. Plasma membrane translocation of trimerized MLKL protein is required for TNF-induced necroptosis. Nat. Cell Biol. 16:55–65
    [Google Scholar]
  87. 87. 
    Chen X, Li W, Ren J, Huang D, He WT et al. 2014. Translocation of mixed lineage kinase domain-like protein to plasma membrane leads to necrotic cell death. Cell Res 24:105–21
    [Google Scholar]
  88. 88. 
    Dondelinger Y, Declercq W, Montessuit S, Roelandt R, Goncalves A et al. 2014. MLKL compromises plasma membrane integrity by binding to phosphatidylinositol phosphates. Cell Rep 7:971–81
    [Google Scholar]
  89. 89. 
    Hildebrand JM, Tanzer MC, Lucet IS, Young SN, Spall SK et al. 2014. Activation of the pseudokinase MLKL unleashes the four-helix bundle domain to induce membrane localization and necroptotic cell death. PNAS 111:15072–77
    [Google Scholar]
  90. 90. 
    Wang H, Sun L, Su L, Rizo J, Liu L et al. 2014. Mixed lineage kinase domain-like protein MLKL causes necrotic membrane disruption upon phosphorylation by RIP3. Mol. Cell 54:133–46
    [Google Scholar]
  91. 91. 
    Kaiser WJ, Sridharan H, Huang C, Mandal P, Upton JW et al. 2013. Toll-like receptor 3-mediated necrosis via TRIF, RIP3, and MLKL. J. Biol. Chem. 288:31268–79
    [Google Scholar]
  92. 92. 
    Upton JW, Kaiser WJ, Mocarski ES 2012. DAI/ZBP1/DLM-1 complexes with RIP3 to mediate virus-induced programmed necrosis that is targeted by murine cytomegalovirus vIRA. Cell Host Microbe 11:290–97
    [Google Scholar]
  93. 93. 
    Wu J, Huang Z, Ren J, Zhang Z, He P et al. 2013. Mlkl knockout mice demonstrate the indispensable role of Mlkl in necroptosis. Cell Res 23:994–1006
    [Google Scholar]
  94. 94. 
    Chan FK. 2012. Fueling the flames: Mammalian programmed necrosis in inflammatory diseases. Cold Spring Harb. Perspect. Biol. 4:a008805
    [Google Scholar]
  95. 95. 
    Linkermann A, Hackl MJ, Kunzendorf U, Walczak H, Krautwald S, Jevnikar AM 2013. Necroptosis in immunity and ischemia-reperfusion injury. Am. J. Transplant. 13:2797–804
    [Google Scholar]
  96. 96. 
    Sun L, Wang X. 2014. A new kind of cell suicide: mechanisms and functions of programmed necrosis. Trends Biochem. Sci. 39:587–93
    [Google Scholar]
  97. 97. 
    Pasparakis M, Vandenabeele P. 2015. Necroptosis and its role in inflammation. Nature 517:311–20
    [Google Scholar]
  98. 98. 
    Mocarski ES, Guo H, Kaiser WJ 2015. Necroptosis: The Trojan horse in cell autonomous antiviral host defense. Virology 479–480:160–66
    [Google Scholar]
  99. 99. 
    Chen X, Zhuang C, Ren Y, Zhang H, Qin X et al. 2019. Identification of the Raf kinase inhibitor TAK-632 and its analogues as potent inhibitors of necroptosis by targeting RIPK1 and RIPK3. Br. J. Pharmacol. 176:2095–108
    [Google Scholar]
  100. 100. 
    Hou J, Ju J, Zhang Z, Zhao C, Li Z et al. 2019. Discovery of potent necroptosis inhibitors targeting RIPK1 kinase activity for the treatment of inflammatory disorder and cancer metastasis. Cell Death Dis 10:493
    [Google Scholar]
  101. 101. 
    Vandenabeele P, Grootjans S, Callewaert N, Takahashi N 2013. Necrostatin-1 blocks both RIPK1 and IDO: consequences for the study of cell death in experimental disease models. Cell Death Differ 20:185–87
    [Google Scholar]
  102. 102. 
    Bertrand MJ, Milutinovic S, Dickson KM, Ho WC, Boudreault A et al. 2008. cIAP1 and cIAP2 facilitate cancer cell survival by functioning as E3 ligases that promote RIP1 ubiquitination. Mol. Cell 30:689–700
    [Google Scholar]
  103. 103. 
    Kang TB, Yang SH, Toth B, Kovalenko A, Wallach D 2013. Caspase-8 blocks kinase RIPK3-mediated activation of the NLRP3 inflammasome. Immunity 38:27–40
    [Google Scholar]
  104. 104. 
    Lawlor KE, Khan N, Mildenhall A, Gerlic M, Croker BA et al. 2015. RIPK3 promotes cell death and NLRP3 inflammasome activation in the absence of MLKL. Nat. Commun. 6:6282
    [Google Scholar]
  105. 105. 
    Moriwaki K, Chan FK. 2014. Necrosis-dependent and independent signaling of the RIP kinases in inflammation. Cytokine Growth Factor Rev 25:167–74
    [Google Scholar]
  106. 106. 
    Varfolomeev E, Blankenship JW, Wayson SM, Fedorova AV, Kayagaki N et al. 2007. IAP antagonists induce autoubiquitination of c-IAPs, NF-κB activation, and TNFα-dependent apoptosis. Cell 131:669–81
    [Google Scholar]
  107. 107. 
    Sun L, Wang H, Wang Z, He S, Chen S et al. 2012. Mixed lineage kinase domain-like protein mediates necrosis signaling downstream of RIP3 kinase. Cell 148:213–27
    [Google Scholar]
  108. 108. 
    Mocarski ES, Kaiser WJ, Livingston-Rosanoff D, Upton JW, Daley-Bauer LP 2014. True grit: programmed necrosis in antiviral host defense, inflammation, and immunogenicity. J. Immunol. 192:2019–26
    [Google Scholar]
  109. 109. 
    Nogusa S, Thapa RJ, Dillon CP, Liedmann S, Oguin TH 3rd et al. 2016. RIPK3 activates parallel pathways of MLKL-driven necroptosis and FADD-mediated apoptosis to protect against influenza A virus. Cell Host Microbe 20:13–24
    [Google Scholar]
  110. 110. 
    Robinson N, McComb S, Mulligan R, Dudani R, Krishnan L, Sad S 2012. Type I interferon induces necroptosis in macrophages during infection with Salmonella enterica serovar Typhimurium. Nat. Immunol. 13:954–62
    [Google Scholar]
  111. 111. 
    Sai K, Parsons C, House JS, Kathariou S, Ninomiya-Tsuji J 2019. Necroptosis mediators RIPK3 and MLKL suppress intracellular Listeria replication independently of host cell killing. J. Cell Biol. 218:1994–2005
    [Google Scholar]
  112. 112. 
    Pena SV, Hanson DA, Carr BA, Goralski TJ, Krensky AM 1997. Processing, subcellular localization, and function of 519 (granulysin), a human late T cell activation molecule with homology to small, lytic, granule proteins. J. Immunol. 158:2680–88
    [Google Scholar]
  113. 113. 
    Stenger S, Hanson DA, Teitelbaum R, Dewan P, Niazi KR et al. 1998. An antimicrobial activity of cytolytic T cells mediated by granulysin. Science 282:121–25
    [Google Scholar]
  114. 114. 
    Chung WH, Hung SI, Yang JY, Su SC, Huang SP et al. 2008. Granulysin is a key mediator for disseminated keratinocyte death in Stevens-Johnson syndrome and toxic epidermal necrolysis. Nat. Med. 14:1343–50
    [Google Scholar]
  115. 115. 
    Veljkovic Vujaklija D, Dominovic M, Gulic T, Mahmutefendic H, Haller H et al. 2013. Granulysin expression and the interplay of granulysin and perforin at the maternal-fetal interface. J. Reprod. Immunol. 97:186–96
    [Google Scholar]
  116. 116. 
    Maeda Y, Tamura T, Fukutomi Y, Mukai T, Kai M, Makino M 2011. A lipopeptide facilitate induction of Mycobacterium leprae killing in host cells. PLOS Negl. Trop. Dis. 5:e1401
    [Google Scholar]
  117. 117. 
    Leippe M, Ebel S, Schoenberger OL, Horstmann RD, Muller-Eberhard HJ 1991. Pore-forming peptide of pathogenic Entamoeba histolytica. . PNAS 88:7659–63
    [Google Scholar]
  118. 118. 
    Hecht O, Van Nuland NA, Schleinkofer K, Dingley AJ, Bruhn H et al. 2004. Solution structure of the pore-forming protein of Entamoeba histolytica. J. Biol. Chem 279:17834–41
    [Google Scholar]
  119. 119. 
    Barman H, Walch M, Latinovic-Golic S, Dumrese C, Dolder M et al. 2006. Cholesterol in negatively charged lipid bilayers modulates the effect of the antimicrobial protein granulysin. J. Membr. Biol. 212:29–39
    [Google Scholar]
  120. 120. 
    Abe R, Yoshioka N, Murata J, Fujita Y, Shimizu H 2009. Granulysin as a marker for early diagnosis of the Stevens-Johnson syndrome. Ann. Intern. Med. 151:514–15
    [Google Scholar]
  121. 121. 
    Dotiwala F, Mulik S, Polidoro RB, Ansara JA, Burleigh BA et al. 2016. Killer lymphocytes use granulysin, perforin and granzymes to kill intracellular parasites. Nat. Med. 22:210–16
    [Google Scholar]
  122. 122. 
    Walch M, Dotiwala F, Mulik S, Thiery J, Kirchhausen T et al. 2014. Cytotoxic cells kill intracellular bacteria through granulysin-mediated delivery of granzymes. Cell 157:1309–23
    [Google Scholar]
  123. 123. 
    Dotiwala F, Sen Santara S, Binker-Cosen AA, Li B, Chandrasekaran S, Lieberman J 2017. Granzyme B disrupts central metabolism and protein synthesis in bacteria to promote an immune cell death program. Cell 171:1125–37.e11
    [Google Scholar]
  124. 124. 
    Li SS, Kyei SK, Timm-McCann M, Ogbomo H, Jones GJ et al. 2013. The NK receptor NKp30 mediates direct fungal recognition and killing and is diminished in NK cells from HIV-infected patients. Cell Host Microbe 14:387–97
    [Google Scholar]
  125. 125. 
    Vitenshtein A, Charpak-Amikam Y, Yamin R, Bauman Y, Isaacson B et al. 2016. NK cell recognition of Candida glabrata through binding of NKp46 and NCR1 to fungal ligands Epa1, Epa6, and Epa7. Cell Host Microbe 20:527–34
    [Google Scholar]
  126. 126. 
    Huang LP, Lyu SC, Clayberger C, Krensky AM 2007. Granulysin-mediated tumor rejection in transgenic mice. J. Immunol. 178:77–84
    [Google Scholar]
  127. 127. 
    Chen X, He WT, Hu L, Li J, Fang Y et al. 2016. Pyroptosis is driven by non-selective gasdermin-D pore and its morphology is different from MLKL channel-mediated necroptosis. Cell Res 26:1007–20
    [Google Scholar]
  128. 128. 
    Zhang Y, Chen X, Gueydan C, Han J 2018. Plasma membrane changes during programmed cell deaths. Cell Res 28:9–21
    [Google Scholar]
  129. 129. 
    Newton K, Manning G. 2016. Necroptosis and inflammation. Annu. Rev. Biochem. 85:743–63
    [Google Scholar]
  130. 130. 
    Davis MA, Fairgrieve MR, Den Hartigh A, Yakovenko O, Duvvuri B et al. 2019. Calpain drives pyroptotic vimentin cleavage, intermediate filament loss, and cell rupture that mediates immunostimulation. PNAS 116:5061–70
    [Google Scholar]
  131. 131. 
    Degterev A, Huang Z, Boyce M, Li Y, Jagtap P et al. 2005. Chemical inhibitor of nonapoptotic cell death with therapeutic potential for ischemic brain injury. Nat. Chem. Biol. 1:112–19
    [Google Scholar]
  132. 132. 
    Hu J, Liu X, Zhao J, Xia S, Ruan J et al. 2018. Identification of pyroptosis inhibitors that target a reactive cysteine in gasdermin D. bioRxiv 365908. https://doi.org/10.1101/365908
    [Crossref]
  133. 133. 
    Enari M, Sakahira H, Yokoyama H, Okawa K, Iwamatsu A, Nagata S 1998. A caspase-activated DNase that degrades DNA during apoptosis, and its inhibitor ICAD. Nature 391:43–50
    [Google Scholar]
  134. 134. 
    Thomas DA, Du C, Xu M, Wang X, Ley TJ 2000. DFF45/ICAD can be directly processed by granzyme B during the induction of apoptosis. Immunity 12:621–32
    [Google Scholar]
  135. 135. 
    Fan Z, Beresford PJ, Oh DY, Zhang D, Lieberman J 2003. Tumor suppressor NM23-H1 is a granzyme A-activated DNase during CTL-mediated apoptosis, and the nucleosome assembly protein SET is its inhibitor. Cell 112:659–72
    [Google Scholar]
  136. 136. 
    Chowdhury D, Beresford PJ, Zhu P, Zhang D, Sung JS et al. 2006. The exonuclease TREX1 is in the SET complex and acts in concert with NM23-H1 to degrade DNA during granzyme A-mediated cell death. Mol. Cell 23:133–42
    [Google Scholar]
  137. 136a. 
    Shlomovitz I, Speir M, Gerlic M 2019. Flipping the dogma—phosphatidylserine in non-apoptotic cell death. Cell Commun. Signal 17:139
    [Google Scholar]
  138. 136b. 
    Yang X, Cheng X, Tang Y, Qiu X, Wang Y 2019. Bacterial endotoxin activates the coagulation cascade through gasdermin D-dependent phosphatidylserine exposure. Immunity 51:983–96
    [Google Scholar]
  139. 137. 
    Nagata S. 2018. Apoptosis and clearance of apoptotic cells. Annu. Rev. Immunol. 36:489–517
    [Google Scholar]
  140. 138. 
    Werfel TA, Cook RS. 2018. Efferocytosis in the tumor microenvironment. Semin. Immunopathol. 40:545–54
    [Google Scholar]
  141. 139. 
    Poon IK, Lucas CD, Rossi AG, Ravichandran KS 2014. Apoptotic cell clearance: basic biology and therapeutic potential. Nat. Rev. Immunol. 14:166–80
    [Google Scholar]
  142. 140. 
    Lieberman J. 2013. Cell-mediated cytotoxicity. Fundamental Immunology WE Paul 891–909 Philadelphia: Wolters Kluwer/Lippincott Williams Wilkins
    [Google Scholar]
  143. 141. 
    Kono H, Rock KL. 2008. How dying cells alert the immune system to danger. Nat. Rev. Immunol. 8:279–89
    [Google Scholar]
  144. 142. 
    Rus H, Cudrici C, Niculescu F 2005. The role of the complement system in innate immunity. Immunol. Res. 33:103–12
    [Google Scholar]
  145. 143. 
    Kjaer TR, Thiel S, Andersen GR 2013. Toward a structure-based comprehension of the lectin pathway of complement. Mol. Immunol. 56:413–22
    [Google Scholar]
  146. 144. 
    Uellner R, Zvelebil MJ, Hopkins J, Jones J, MacDougall LK et al. 1997. Perforin is activated by a proteolytic cleavage during biosynthesis which reveals a phospholipid-binding C2 domain. EMBO J 16:7287–96
    [Google Scholar]
  147. 145. 
    Dupuis M, Schaerer E, Krause KH, Tschopp J 1993. The calcium-binding protein calreticulin is a major constituent of lytic granules in cytolytic T lymphocytes. J. Exp. Med. 177:1–7
    [Google Scholar]
  148. 146. 
    Sula Karreci E, Eskandari SK, Dotiwala F, Routray SK, Kurdi AT et al. 2017. Human regulatory T cells undergo self-inflicted damage via granzyme pathways upon activation. JCI Insight 2:91599
    [Google Scholar]
  149. 147. 
    Aglietti RA, Estevez A, Gupta A, Ramirez MG, Liu PS et al. 2016. GsdmD p30 elicited by caspase-11 during pyroptosis forms pores in membranes. PNAS 113:7858–63
    [Google Scholar]
  150. 148. 
    Sborgi L, Ruhl S, Mulvihill E, Pipercevic J, Heilig R et al. 2016. GSDMD membrane pore formation constitutes the mechanism of pyroptotic cell death. EMBO J 35:1766–78
    [Google Scholar]
  151. 149. 
    van Pee K, Mulvihill E, Muller DJ, Yildiz O 2016. Unraveling the pore-forming steps of pneumolysin from Streptococcus pneumoniae. . Nano Lett 16:7915–24
    [Google Scholar]
  152. 150. 
    van Pee K, Neuhaus A, D'Imprima E, Mills DJ, Kuhlbrandt W, Yildiz O 2017. CryoEM structures of membrane pore and prepore complex reveal cytolytic mechanism of Pneumolysin. eLife 6:e23644
    [Google Scholar]
  153. 151. 
    Liu S, Liu H, Johnston A, Hanna-Addams S, Reynoso E et al. 2017. MLKL forms disulfide bond-dependent amyloid-like polymers to induce necroptosis. PNAS 114:E7450–59
    [Google Scholar]
  154. 152. 
    Karch J, Kanisicak O, Brody MJ, Sargent MA, Michael DM, Molkentin JD 2015. Necroptosis interfaces with MOMP and the MPTP in mediating cell death. PLOS ONE 10:e0130520
    [Google Scholar]
  155. 153. 
    Rogers C, Erkes DA, Nardone A, Aplin AE, Fernandes-Alnemri T, Alnemri ES 2019. Gasdermin pores permeabilize mitochondria to augment caspase-3 activation during apoptosis and inflammasome activation. Nat. Commun. 10:1689
    [Google Scholar]
  156. 154. 
    de Vasconcelos NM, Van Opdenbosch N, Van Gorp H, Parthoens E, Lamkanfi M 2019. Single-cell analysis of pyroptosis dynamics reveals conserved GSDMD-mediated subcellular events that precede plasma membrane rupture. Cell Death Differ 26:146–61
    [Google Scholar]
  157. 155. 
    Liu J, Qian C, Cao X 2016. Post-translational modification control of innate immunity. Immunity 45:15–30
    [Google Scholar]
  158. 156. 
    Liu X, Wang Q, Chen W, Wang C 2013. Dynamic regulation of innate immunity by ubiquitin and ubiquitin-like proteins. Cytokine Growth Factor Rev 24:559–70
    [Google Scholar]
  159. 157. 
    Rathkey JK, Zhao J, Liu Z, Chen Y, Yang J et al. 2018. Chemical disruption of the pyroptotic pore-forming protein gasdermin D inhibits inflammatory cell death and sepsis. Sci. Immunol. 3:eaat2738
    [Google Scholar]
  160. 158. 
    Kimberley FC, Sivasankar B, Morgan BP 2007. Alternative roles for CD59. Mol. Immunol. 44:73–81
    [Google Scholar]
  161. 159. 
    Kim DD, Song WC. 2006. Membrane complement regulatory proteins. Clin. Immunol. 118:127–36
    [Google Scholar]
  162. 160. 
    Brodsky RA. 2015. Complement in hemolytic anemia. Blood 126:2459–65
    [Google Scholar]
  163. 161. 
    Blink EJ, Trapani JA, Jans DA 1999. Perforin-dependent nuclear targeting of granzymes: a central role in the nuclear events of granule-exocytosis-mediated apoptosis. ? Immunol. Cell Biol. 77:206–15
    [Google Scholar]
  164. 162. 
    Martinvalet D, Dykxhoorn DM, Ferrini R, Lieberman J 2008. Granzyme A cleaves a mitochondrial complex I protein to initiate caspase-independent cell death. Cell 133:681–92
    [Google Scholar]
  165. 163. 
    Spicer BA, Law RHP, Caradoc-Davies TT, Ekkel SM, Bayly-Jones C et al. 2018. The first transmembrane region of complement component-9 acts as a brake on its self-assembly. Nat. Commun. 9:3266
    [Google Scholar]
  166. 164. 
    Huang D, Zheng X, Wang ZA, Chen X, He WT et al. 2017. The MLKL channel in necroptosis is an octamer formed by tetramers in a dyadic process. Mol. Cell. Biol. 37:e00497–16
    [Google Scholar]
  167. 165. 
    Xia B, Fang S, Chen X, Hu H, Chen P et al. 2016. MLKL forms cation channels. Cell Res 26:517–28
    [Google Scholar]
  168. 166. 
    Tschopp J, Engel A, Podack ER 1984. Molecular weight of poly(C9): 12 to 18 C9 molecules form the transmembrane channel of complement. J. Biol. Chem. 259:1922–28
    [Google Scholar]
  169. 167. 
    Liu Z, Wang C, Yang J, Zhou B, Yang R et al. 2019. Crystal structures of the full-length murine and human gasdermin D reveal mechanisms of autoinhibition, lipid binding, and oligomerization. Immunity 51:43–49.e4
    [Google Scholar]
  170. 168. 
    Andrews NW, Corrotte M. 2018. Plasma membrane repair. Curr. Biol. 28:R392–97
    [Google Scholar]
  171. 169. 
    McNeil PL, Kirchhausen T. 2005. An emergency response team for membrane repair. Nat. Rev. Mol. Cell Biol. 6:499–505
    [Google Scholar]
  172. 170. 
    Dal Peraro M, van der Goot FG 2016. Pore-forming toxins: ancient, but never really out of fashion. Nat. Rev. Microbiol. 14:77–92
    [Google Scholar]
  173. 171. 
    Tegla CA, Cudrici C, Patel S, Trippe R 3rd, Rus V et al. 2011. Membrane attack by complement: the assembly and biology of terminal complement complexes. Immunol. Res. 51:45–60
    [Google Scholar]
  174. 172. 
    Ruhl S, Shkarina K, Demarco B, Heilig R, Santos JC, Broz P 2018. ESCRT-dependent membrane repair negatively regulates pyroptosis downstream of GSDMD activation. Science 362:956–60
    [Google Scholar]
  175. 173. 
    Gong YN, Guy C, Olauson H, Becker JU, Yang M et al. 2017. ESCRT-III acts downstream of MLKL to regulate necroptotic cell death and its consequences. Cell 169:286–300.e16
    [Google Scholar]
  176. 174. 
    Lieberman J, Wu H, Kagan JC 2019. Gasdermin D activity in inflammation and host defense. Sci. Immunol. 4:eaav1447
    [Google Scholar]
  177. 175. 
    Chen KW, Gross CJ, Sotomayor FV, Stacey KJ, Tschopp J et al. 2014. The neutrophil NLRC4 inflammasome selectively promotes IL-1β maturation without pyroptosis during acute Salmonella challenge. Cell Rep 8:570–82
    [Google Scholar]
  178. 176. 
    Gaidt MM, Ebert TS, Chauhan D, Schmidt T, Schmid-Burgk JL et al. 2016. Human monocytes engage an alternative inflammasome pathway. Immunity 44:833–46
    [Google Scholar]
  179. 177. 
    Wolf AJ, Reyes CN, Liang W, Becker C, Shimada K et al. 2016. Hexokinase is an innate immune receptor for the detection of bacterial peptidoglycan. Cell 166:624–36
    [Google Scholar]
  180. 178. 
    Zanoni I, Tan Y, Di Gioia M, Broggi A, Ruan J et al. 2016. An endogenous caspase-11 ligand elicits interleukin-1 release from living dendritic cells. Science 352:1232–36
    [Google Scholar]
/content/journals/10.1146/annurev-immunol-111319-023800
Loading
/content/journals/10.1146/annurev-immunol-111319-023800
Loading

Data & Media loading...

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