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:

Molecular mechanisms of necroptosis: an ordered cellular explosion

Key Points

  • Although for a long time necrosis was considered to be a purely accidental cell death subroutine, multiple lines of evidence now show that necrotic cell death can be regulated, both in its occurrence and in its course. The term 'necroptosis' was introduced by Yuan's research group in 2005 to indicate 'programmed' (as opposed to 'accidental') necrosis.

  • The best characterized signal transduction cascade leading to necroptosis is initiated by ligand-bound tumour necrosis factor (TNF) receptor 1 (TNFR1), which allows for the assembly of a cytoplasmic supramolecular complex — TNFR complex I — that includes (among other proteins) TNFR-associated death domain (TRADD), cellular inhibitor of apoptosis 1 (cIAP1), cIAP2 and receptor-interacting protein kinase 1 (RIP1; also known as RIPK1).

  • In complex I, RIP1 can be ubiquitylated by cIAPs and deubiquitylated by cylindromatosis (CYLD) and A20 (also known as TNFAIP3). Whereas ubiquitylated RIP1 promotes the activation of the nuclear factor κB (NF-κB) system, deubiquitylated RIP1 functions as a cell death-inducing kinase.

  • On TNFR1 internalization, the so-called TNFR complex II is formed, which contains TRADD, FAS-associated protein with a death domain (FADD) and caspase 8. Normally, caspase 8 becomes activated in complex II, thereby igniting a pro-apoptotic caspase cascade. When caspase activation is prevented, however, RIP1 physically and functionally interacts with RIP3 (also known as RIPK3), thereby generating a necroptosis-inducing complex known as the necrosome.

  • Necroptosis can also be ignited by pathogen recognition receptors, including Toll-like receptors, NOD-like receptors and retinoic acid-inducible gene I-like receptors, as well as in response to DNA damage, presumably by a poly(ADP-ribose) polymerase-1 (PARP1)-dependent signalling pathway.

  • Although the underlying molecular mechanisms remain obscure, reactive oxygen species (ROS), bioenergetic metabolic cascades and the release of cytotoxic factors from lysosomes and mitochondria all contribute to the execution of necroptosis.

  • Regulated necrosis has been seen in multiple, evolutionarily distant model organisms, including yeast, nematodes, fruit flies, rodents, primates and human cells, corroborating the notion that necroptosis may represent a phylogenetically conserved mechanism for programmed cell death.

  • Numerous in vivo studies indicate that the inhibition of necroptosis (by genetic means or by RIP1-targeting agents called necrostatins) can confer consistent cytoprotection, suggesting that necroptosis may constitute a promising target for drug development.

Abstract

For a long time, apoptosis was considered the sole form of programmed cell death during development, homeostasis and disease, whereas necrosis was regarded as an unregulated and uncontrollable process. Evidence now reveals that necrosis can also occur in a regulated manner. The initiation of programmed necrosis, 'necroptosis', by death receptors (such as tumour necrosis factor receptor 1) requires the kinase activity of receptor-interacting protein 1 (RIP1; also known as RIPK1) and RIP3 (also known as RIPK3), and its execution involves the active disintegration of mitochondrial, lysosomal and plasma membranes. Necroptosis participates in the pathogenesis of diseases, including ischaemic injury, neurodegeneration and viral infection, thereby representing an attractive target for the avoidance of unwarranted cell death.

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: TNFR1-elicited signalling pathways.
Figure 2: Execution of necroptosis.

Similar content being viewed by others

References

  1. Lockshin, R. A. & Williams, C. M. Programmed cell death — II. Endocrine potentiation of the breakdown of the intersegmental muscles of silkmoths. J. Insect Physiol. 10, 643–649 (1964).

    CAS  Google Scholar 

  2. Kerr, J. F., Wyllie, A. H. & Currie, A. R. Apoptosis: a basic biological phenomenon with wide-ranging implications in tissue kinetics. Br. J. Cancer 26, 239–257 (1972).

    CAS  PubMed Central  PubMed  Google Scholar 

  3. Lettre, G. & Hengartner, M. O. Developmental apoptosis in, C. elegans: a complex CEDnario. Nature Rev. Mol. Cell Biol. 7, 97–108 (2006).

    CAS  Google Scholar 

  4. Schweichel, J. U. & Merker, H. J. The morphology of various types of cell death in prenatal tissues. Teratology 7, 253–266 (1973).

    CAS  PubMed  Google Scholar 

  5. Kroemer, G. et al. Classification of cell death: recommendations of the Nomenclature Committee on Cell Death 2009. Cell Death Differ. 16, 3–11 (2009). This article provides up-to-date guidelines for the use of cell death-related terminology in scientific publications, as provided by the Nomenclature Committee on Cell Death, an organization composed of reputed researchers in the field of cell death worldwide.

    CAS  PubMed  Google Scholar 

  6. Laster, S. M., Wood, J. G. & Gooding, L. R. Tumor necrosis factor can induce both apoptic and necrotic forms of cell lysis. J. Immunol. 141, 2629–2634 (1988).

    CAS  PubMed  Google Scholar 

  7. Degterev, A. et al. Chemical inhibitor of nonapoptotic cell death with therapeutic potential for ischemic brain injury. Nature Chem. Biol. 1, 112–119 (2005).

    CAS  Google Scholar 

  8. 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).

    CAS  PubMed  Google Scholar 

  9. Holler, N. et al. Fas triggers an alternative, caspase-8-independent cell death pathway using the kinase RIP as effector molecule. Nature Immunol. 1, 489–495 (2000). The authors discovered that, in some cell types, FAS can trigger non-apoptotic cell death that is independent of caspases but dependent on the adaptor protein FADD and the presence and enzymatic activity of the protein kinase RIP1. This milestone paper is the first report of RIP1-dependent necroptosis.

    CAS  Google Scholar 

  10. Cho, Y. S. et al. Phosphorylation-driven assembly of the RIP1–RIP3 complex regulates programmed necrosis and virus-induced inflammation. Cell 137, 1112–1123 (2009).

    CAS  PubMed Central  PubMed  Google Scholar 

  11. He, S. et al. Receptor interacting protein kinase-3 determines cellular necrotic response to TNF-α. Cell 137, 1100–1111 (2009).

    CAS  PubMed  Google Scholar 

  12. Zhang, D. W. et al. RIP3, an energy metabolism regulator that switches TNF-induced cell death from apoptosis to necrosis. Science 325, 332–336 (2009). References 10–12 independently uncovered the obligate role of RIP3 in necroptosis. By multiple experimental approaches, RIP3 was shown to functionally and physically interact with RIP1, leading to a mitochondrial metabolic burst that underlies necroptosis execution. The pathophysiological role of RIP3 in vivo was substantiated by animal models of viral infection and acute pancreatitis.

    CAS  PubMed  Google Scholar 

  13. Vercammen, D. et al. Inhibition of caspases increases the sensitivity of L929 cells to necrosis mediated by tumor necrosis factor. J. Exp. Med. 187, 1477–1485 (1998). The sensitivity of murine L929 fibrosarcoma cells to TNF-induced necrosis was dramatically increased by expression of a serpin-like viral protein (CrmA) and by pharmacological caspase inhibitors, showing for the first time a major role for caspases in the apoptotis–necrosis switch.

    CAS  PubMed Central  PubMed  Google Scholar 

  14. Hitomi, J. et al. Identification of a molecular signaling network that regulates a cellular necrotic cell death pathway. Cell 135, 1311–1323 (2008). The first time systems biology was applied to compare necroptosis to apoptosis in murine cells. Genome wide RNAi-based screens coupled to in silico and in vitro analyses allowed the delineation of a signalling network that regulates the molecular bifurcation between necroptosis and apoptosis.

    CAS  PubMed Central  PubMed  Google Scholar 

  15. Goossens, V., Stange, G., Moens, K., Pipeleers, D. & Grooten, J. Regulation of tumor necrosis factor-induced, mitochondria- and reactive oxygen species-dependent cell death by the electron flux through the electron transport chain complex I. Antioxid. Redox Signal. 1, 285–295 (1999).

    CAS  PubMed  Google Scholar 

  16. Kim, Y. S., Morgan, M. J., Choksi, S. & Liu, Z. G. TNF-induced activation of the Nox1 NADPH oxidase and its role in the induction of necrotic cell death. Mol. Cell 26, 675–687 (2007).

    CAS  PubMed  Google Scholar 

  17. Yazdanpanah, B. et al. Riboflavin kinase couples TNF receptor 1 to NADPH oxidase. Nature 460, 1159–1163 (2009).

    CAS  PubMed  Google Scholar 

  18. Goossens, V., Grooten, J. & Fiers, W. The oxidative metabolism of glutamine. A modulator of reactive oxygen intermediate-mediated cytotoxicity of tumor necrosis factor in L929 fibrosarcoma cells. J. Biol. Chem. 271, 192–196 (1996).

    CAS  PubMed  Google Scholar 

  19. Zong, W. X., Ditsworth, D., Bauer, D. E., Wang, Z. Q. & Thompson, C. B. Alkylating DNA damage stimulates a regulated form of necrotic cell death. Genes Dev. 18, 1272–1282 (2004).

    CAS  PubMed Central  PubMed  Google Scholar 

  20. Nakagawa, T. et al. Cyclophilin D-dependent mitochondrial permeability transition regulates some necrotic but not apoptotic cell death. Nature 434, 652–658 (2005).

    CAS  PubMed  Google Scholar 

  21. Schinzel, A. C. et al. Cyclophilin D is a component of mitochondrial permeability transition and mediates neuronal cell death after focal cerebral ischemia. Proc. Natl Acad. Sci. USA 102, 12005–12010 (2005). References 20 and 21 show that CYPD-deficient mice are protected against ischaemic insults, as compared to their wild-type littermates. Moreover, multiple types of primary cells isolated from Ppif−/− mice exhibit enhanced resistance to necrotic (but not apoptotic) stimuli. Together, these studies provide unequivocal evidence that CYPD regulates necroptosis in vivo.

    CAS  PubMed  PubMed Central  Google Scholar 

  22. Temkin, V., Huang, Q., Liu, H., Osada, H. & Pope, R. M. Inhibition of ADP/ATP exchange in receptor-interacting protein-mediated necrosis. Mol. Cell. Biol. 26, 2215–2225 (2006).

    CAS  PubMed Central  PubMed  Google Scholar 

  23. Boya, P. & Kroemer, G. Lysosomal membrane permeabilization in cell death. Oncogene 27, 6434–6451 (2008).

    CAS  PubMed  Google Scholar 

  24. Vanden Berghe, T. et al. Necroptosis, necrosis and secondary necrosis converge on similar cellular disintegration features. Cell Death Differ. 17, 922–930 (2010).

    CAS  PubMed  Google Scholar 

  25. Kroemer, G., Galluzzi, L. & Brenner, C. Mitochondrial membrane permeabilization in cell death. Physiol. Rev. 87, 99–163 (2007). This review provides a comprehensive analysis of the mitochondrial pathway of cell death and of its multifaceted implications for human physiology and pathology.

    CAS  PubMed  Google Scholar 

  26. Green, D. R. & Kroemer, G. Pharmacological manipulation of cell death: clinical applications in sight? J. Clin. Invest. 115, 2610–2617 (2005).

    CAS  PubMed Central  PubMed  Google Scholar 

  27. Rosenbaum, D. M. et al. Necroptosis, a novel form of caspase-independent cell death, contributes to neuronal damage in a retinal ischemia-reperfusion injury model. J. Neurosci. Res. 88, 1569–1576 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  28. Vanlangenakker, N., Vanden Berghe, T., Krysko, D. V., Festjens, N. & Vandenabeele, P. Molecular mechanisms and pathophysiology of necrotic cell death. Curr. Mol. Med. 8, 207–220 (2008).

    CAS  PubMed  Google Scholar 

  29. Vercammen, D. et al. Dual signaling of the Fas receptor: initiation of both apoptotic and necrotic cell death pathways. J. Exp. Med. 188, 919–930 (1998).

    CAS  PubMed Central  PubMed  Google Scholar 

  30. Chan, F. K. et al. A role for tumor necrosis factor receptor-2 and receptor-interacting protein in programmed necrosis and antiviral responses. J. Biol. Chem. 278, 51613–51621 (2003).

    CAS  PubMed  Google Scholar 

  31. Hacker, H. & Karin, M. Regulation and function of IKK and IKK-related kinases. Sci. STKE 2006, re13 (2006).

    PubMed  Google Scholar 

  32. Wilson, N. S., Dixit, V. & Ashkenazi, A. Death receptor signal transducers: nodes of coordination in immune signaling networks. Nature Immunol. 10, 348–355 (2009).

    CAS  Google Scholar 

  33. Fiers, W. et al. TNF-induced intracellular signaling leading to gene induction or to cytotoxicity by necrosis or by apoptosis. J. Inflamm. 47, 67–75 (1995).

    CAS  PubMed  Google Scholar 

  34. Martinon, F., Gaide, O., Petrilli, V., Mayor, A. & Tschopp, J. NALP inflammasomes: a central role in innate immunity. Semin. Immunopathol. 29, 213–229 (2007).

    CAS  PubMed  Google Scholar 

  35. Kalai, M. et al. Tipping the balance between necrosis and apoptosis in human and murine cells treated with interferon and dsRNA. Cell Death Differ. 9, 981–994 (2002).

    CAS  PubMed  Google Scholar 

  36. Ma, Y., Temkin, V., Liu, H. & Pope, R. M. NF-κB protects macrophages from lipopolysaccharide-induced cell death: the role of caspase 8 and receptor-interacting protein. J. Biol. Chem. 280, 41827–41834 (2005).

    CAS  PubMed  Google Scholar 

  37. Francois, M., Le Cabec, V., Dupont, M. A., Sansonetti, P. J. & Maridonneau-Parini, I. Induction of necrosis in human neutrophils by Shigella flexneri requires type III secretion, IpaB and IpaC invasins, and actin polymerization. Infect. Immun. 68, 1289–1296 (2000).

    CAS  PubMed Central  PubMed  Google Scholar 

  38. Koterski, J. F., Nahvi, M., Venkatesan, M. M. & Haimovich, B. Virulent Shigella flexneri causes damage to mitochondria and triggers necrosis in infected human monocyte-derived macrophages. Infect. Immun. 73, 504–513 (2005).

    CAS  PubMed Central  PubMed  Google Scholar 

  39. Willingham, S. B. et al. Microbial pathogen-induced necrotic cell death mediated by the inflammasome components CIAS1/cryopyrin/NLRP3 and ASC. Cell Host Microbe 2, 147–159 (2007).

    CAS  PubMed Central  PubMed  Google Scholar 

  40. Chu, J. J. & Ng, M. L. The mechanism of cell death during West Nile virus infection is dependent on initial infectious dose. J. Gen. Virol. 84, 3305–3314 (2003).

    CAS  PubMed  Google Scholar 

  41. Ray, C. A. & Pickup, D. J. The mode of death of pig kidney cells infected with cowpox virus is governed by the expression of the CrmA gene. Virology 217, 384–391 (1996).

    CAS  PubMed  Google Scholar 

  42. Chan, F. K. et al. A domain in TNF receptors that mediates ligand-independent receptor assembly and signaling. Science 288, 2351–2354 (2000).

    CAS  PubMed  Google Scholar 

  43. Micheau, O. & Tschopp, J. Induction of TNF receptor I- mediated apoptosis via two sequential signaling complexes. Cell 114, 181–190 (2003). This paper elucidated the early biochemical events that are engaged on TNFR1 ligation and, in particular, showed the existence of two distinct signalling complexes, one at the plasma membrane that includes TNFR1, TRAF2 and RIP1, and one at an endocytic compartment that involves internalized TNFR1, TRADD, FADD and caspase 8.

    CAS  PubMed  Google Scholar 

  44. Deveraux, Q. L. et al. IAPs block apoptotic events induced by caspase-8 and cytochrome c by direct inhibition of distinct caspases. EMBO J. 17, 2215–2223 (1998).

    CAS  PubMed Central  PubMed  Google Scholar 

  45. Csomos, R. A., Brady, G. F. & Duckett, C. S. Enhanced cytoprotective effects of the inhibitor of apoptosis protein cellular IAP1 through stabilization with TRAF2. J. Biol. Chem. 284, 20531–20539 (2009).

    CAS  PubMed Central  PubMed  Google Scholar 

  46. Rothe, M., Pan, M. G., Henzel, W. J., Ayres, T. M. & Goeddel, D. V. The TNFR2-TRAF signaling complex contains two novel proteins related to baculoviral inhibitor of apoptosis proteins. Cell 83, 1243–1252 (1995).

    CAS  PubMed  Google Scholar 

  47. Bertrand, M. J. et al. cIAP1 and cIAP2 facilitate cancer cell survival by functioning as E3 ligases that promote RIP1 ubiquitination. Mol. Cell 30, 689–700 (2008).

    CAS  PubMed  Google Scholar 

  48. Wong, W. W. et al. RIPK1 is not essential for TNFR1-induced activation of NF-κB. Cell Death Differ. 17, 482–487 (2010).

    CAS  PubMed  Google Scholar 

  49. Ea, C. K., Deng, L., Xia, Z. P., Pineda, G. & Chen, Z. J. Activation of IKK by TNFα requires site-specific ubiquitination of RIP1 and polyubiquitin binding by NEMO. Mol. Cell 22, 245–257 (2006).

    CAS  PubMed  Google Scholar 

  50. Christofferson, D. E. & Yuan, J. Necroptosis as an alternative form of programmed cell death. Curr. Opin. Cell Biol. 22, 263–268 (2010).

    CAS  PubMed Central  PubMed  Google Scholar 

  51. Ting, A. T., Pimentel-Muinos, F. X. & Seed, B. RIP mediates tumor necrosis factor receptor 1 activation of NF-κB but not Fas/APO-1-initiated apoptosis. EMBO J. 15, 6189–6196 (1996).

    CAS  PubMed Central  PubMed  Google Scholar 

  52. Shembade, N., Ma, A. & Harhaj, E. W. Inhibition of NF-κB signaling by A20 through disruption of ubiquitin enzyme complexes. Science 327, 1135–1139 (2010).

    CAS  PubMed Central  PubMed  Google Scholar 

  53. Enesa, K. et al. NF-κB suppression by the deubiquitinating enzyme Cezanne: a novel negative feedback loop in pro-inflammatory signaling. J. Biol. Chem. 283, 7036–7045 (2008).

    CAS  PubMed  Google Scholar 

  54. Xu, G. et al. Ubiquitin-specific peptidase 21 inhibits tumor necrosis factor α-induced nuclear factor κB activation via binding to and deubiquitinating receptor-interacting protein 1. J. Biol. Chem. 285, 969–978 (2010).

    CAS  PubMed  Google Scholar 

  55. Feng, S. et al. Cleavage of RIP3 inactivates its caspase-independent apoptosis pathway by removal of kinase domain. Cell Signal 19, 2056–2067 (2007).

    CAS  PubMed  Google Scholar 

  56. Wang, L., Du, F. & Wang, X. TNF-α induces two distinct caspase-8 activation pathways. Cell 133, 693–703 (2008).

    CAS  PubMed  Google Scholar 

  57. Lin, Y. et al. Tumor necrosis factor-induced nonapoptotic cell death requires receptor-interacting protein-mediated cellular reactive oxygen species accumulation. J. Biol. Chem. 279, 10822–10828 (2004).

    CAS  PubMed  Google Scholar 

  58. Ermolaeva, M. A. et al. Function of TRADD in tumor necrosis factor receptor 1 signaling and in TRIF-dependent inflammatory responses. Nature Immunol. 9, 1037–1046 (2008).

    CAS  Google Scholar 

  59. Declercq, W., Vanden Berghe, T. & Vandenabeele, P. RIP kinases at the crossroads of cell death and survival. Cell 138, 229–232 (2009).

    CAS  PubMed  Google Scholar 

  60. Vandenabeele, P., Vanden Berghe, T. & Festjens, N. Caspase inhibitors promote alternative cell death pathways. Sci. STKE 2006, pe44 (2006).

    PubMed  Google Scholar 

  61. Sun, X., Yin, J., Starovasnik, M. A., Fairbrother, W. J. & Dixit, V. M. Identification of a novel homotypic interaction motif required for the phosphorylation of receptor-interacting protein (RIP) by RIP3. J. Biol. Chem. 277, 9505–9511 (2002).

    CAS  PubMed  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  63. Degterev, A. et al. Identification of RIP1 kinase as a specific cellular target of necrostatins. Nature Chem. Biol. 4, 313–321 (2008).

    CAS  Google Scholar 

  64. Teng, X. et al. Structure–activity relationship study of [1,2,3]thiadiazole necroptosis inhibitors. Bioorg. Med. Chem. Lett. 17, 6836–6840 (2007).

    CAS  PubMed Central  PubMed  Google Scholar 

  65. Vandenabeele, P., Declercq, W. & Berghe, T. V. Necrotic cell death and 'necrostatins': now we can control cellular explosion. Trends Biochem. Sci. 33, 352–355 (2008).

    CAS  PubMed  Google Scholar 

  66. Upton, J. W., Kaiser, W. J. & Mocarski, E. S. Virus inhibition of RIP3-dependent necrosis. Cell Host Microbe 7, 302–313 (2010).

    CAS  PubMed Central  PubMed  Google Scholar 

  67. Soldani, C. & Scovassi, A. I. Poly(ADP-ribose) polymerase-1 cleavage during apoptosis: an update. Apoptosis 7, 321–328 (2002).

    CAS  PubMed  Google Scholar 

  68. Saelens, X. et al. Protein synthesis persists during necrotic cell death. J. Cell Biol. 168, 545–551 (2005).

    CAS  PubMed Central  PubMed  Google Scholar 

  69. Sun, X. M. et al. Caspase activation inhibits proteasome function during apoptosis. Mol. Cell 14, 81–93 (2004).

    CAS  PubMed  Google Scholar 

  70. Leist, M., Single, B., Castoldi, A. F., Kuhnle, S. & Nicotera, P. Intracellular adenosine triphosphate (ATP) concentration: a switch in the decision between apoptosis and necrosis. J. Exp. Med. 185, 1481–1486 (1997).

    CAS  PubMed Central  PubMed  Google Scholar 

  71. Kraus, W. L. Transcriptional control by PARP-1: chromatin modulation, enhancer-binding, coregulation, and insulation. Curr. Opin. Cell Biol. 20, 294–302 (2008).

    CAS  PubMed Central  PubMed  Google Scholar 

  72. Los, M. et al. Activation and caspase-mediated inhibition of PARP: a molecular switch between fibroblast necrosis and apoptosis in death receptor signaling. Mol. Biol. Cell 13, 978–988 (2002).

    CAS  PubMed Central  PubMed  Google Scholar 

  73. Yu, S. W. et al. Mediation of poly(ADP-ribose) polymerase-1-dependent cell death by apoptosis-inducing factor. Science 297, 259–263 (2002).

    CAS  PubMed  Google Scholar 

  74. Cao, G. et al. Critical role of calpain I in mitochondrial release of apoptosis-inducing factor in ischemic neuronal injury. J. Neurosci. 27, 9278–9293 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  75. Moubarak, R. S. et al. Sequential activation of poly(ADP-ribose) polymerase 1, calpains, and Bax is essential in apoptosis-inducing factor-mediated programmed necrosis. Mol. Cell. Biol. 27, 4844–4862 (2007).

    CAS  PubMed Central  PubMed  Google Scholar 

  76. Slemmer, J. E. et al. Causal role of apoptosis-inducing factor for neuronal cell death following traumatic brain injury. Am. J. Pathol. 173, 1795–1805 (2008).

    CAS  PubMed Central  PubMed  Google Scholar 

  77. Culmsee, C. et al. Apoptosis-inducing factor triggered by poly(ADP-ribose) polymerase and Bid mediates neuronal cell death after oxygen-glucose deprivation and focal cerebral ischemia. J. Neurosci. 25, 10262–10272 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  78. Galluzzi, L., Blomgren, K. & Kroemer, G. Mitochondrial membrane permeabilization in neuronal injury. Nature Rev. Neurosci. 10, 481–494 (2009).

    CAS  Google Scholar 

  79. Boujrad, H., Gubkina, O., Robert, N., Krantic, S. & Susin, S. A. AIF-mediated programmed necrosis: a highly regulated way to die. Cell Cycle 6, 2612–2619 (2007).

    CAS  PubMed  Google Scholar 

  80. Xu, Y., Huang, S., Liu, Z. G. & Han, J. Poly(ADP-ribose) polymerase-1 signaling to mitochondria in necrotic cell death requires RIP1/TRAF2-mediated JNK1 activation. J. Biol. Chem. 281, 8788–8795 (2006).

    CAS  PubMed  Google Scholar 

  81. Alano, C. C., Ying, W. & Swanson, R. A. Poly(ADP-ribose) polymerase-1-mediated cell death in astrocytes requires NAD+ depletion and mitochondrial permeability transition. J. Biol. Chem. 279, 18895–18902 (2004).

    CAS  PubMed  Google Scholar 

  82. Brenner, C. et al. Bcl-2 and Bax regulate the channel activity of the mitochondrial adenine nucleotide translocator. Oncogene 19, 329–336 (2000).

    CAS  PubMed  Google Scholar 

  83. Baines, C. P., Kaiser, R. A., Sheiko, T., Craigen, W. J. & Molkentin, J. D. Voltage-dependent anion channels are dispensable for mitochondrial-dependent cell death. Nature Cell Biol. 9, 550–555 (2007).

    CAS  PubMed  Google Scholar 

  84. Kokoszka, J. E. et al. The ADP/ATP translocator is not essential for the mitochondrial permeability transition pore. Nature 427, 461–465 (2004).

    CAS  PubMed Central  PubMed  Google Scholar 

  85. Schulze-Osthoff, K. et al. Cytotoxic activity of tumor necrosis factor is mediated by early damage of mitochondrial functions. Evidence for the involvement of mitochondrial radical generation. J. Biol. Chem. 267, 5317–5323 (1992). This paper provided the first evidence for a role of complex I-mediated ROS production in TNF-induced necrotic cell death.

    CAS  PubMed  Google Scholar 

  86. Festjens, N. et al. Butylated hydroxyanisole is more than a reactive oxygen species scavenger. Cell Death Differ. 13, 166–169 (2006).

    CAS  PubMed  Google Scholar 

  87. Bouche, C., Serdy, S., Kahn, C. R. & Goldfine, A. B. The cellular fate of glucose and its relevance in type 2 diabetes. Endocr. Rev. 25, 807–830 (2004).

    CAS  PubMed  Google Scholar 

  88. Van Herreweghe, F. et al. Tumor necrosis factor-induced modulation of glyoxalase I activities through phosphorylation by PKA results in cell death and is accompanied by the formation of a specific methylglyoxal-derived AGE. Proc. Natl Acad. Sci. USA 99, 949–954 (2002).

    CAS  PubMed  PubMed Central  Google Scholar 

  89. Rabbani, N. & Thornalley, P. J. Dicarbonyls linked to damage in the powerhouse: glycation of mitochondrial proteins and oxidative stress. Biochem. Soc. Trans. 36, 1045–1050 (2008).

    CAS  PubMed Central  PubMed  Google Scholar 

  90. Mates, J. M. et al. Glutamine homeostasis and mitochondrial dynamics. Int. J. Biochem. Cell Biol. 41, 2051–2061 (2009).

    CAS  PubMed  Google Scholar 

  91. Albrecht, J. & Norenberg, M. D. Glutamine: a Trojan horse in ammonia neurotoxicity. Hepatology 44, 788–794 (2006).

    CAS  PubMed  Google Scholar 

  92. Morgan, M. J., Kim, Y. S. & Liu, Z. G. TNFα and reactive oxygen species in necrotic cell death. Cell Res. 18, 343–349 (2008).

    CAS  PubMed  Google Scholar 

  93. Brookes, P. S. Mitochondrial H+ leak and ROS generation: an odd couple. Free Radic. Biol. Med. 38, 12–23 (2005).

    CAS  PubMed  Google Scholar 

  94. Poyton, R. O., Ball, K. A. & Castello, P. R. Mitochondrial generation of free radicals and hypoxic signaling. Trends Endocrinol. Metab. 20, 332–340 (2009).

    CAS  PubMed  Google Scholar 

  95. Chen, T. Y., Chi, K. H., Wang, J. S., Chien, C. L. & Lin, W. W. Reactive oxygen species are involved in FasL-induced caspase-independent cell death and inflammatory responses. Free Radic. Biol. Med. 46, 643–655 (2009).

    CAS  PubMed  Google Scholar 

  96. Thon, L. et al. Ceramide mediates caspase-independent programmed cell death. FASEB J. 19, 1945–1956 (2005).

    CAS  PubMed  Google Scholar 

  97. Antosiewicz, J., Ziolkowski, W., Kaczor, J. J. & Herman-Antosiewicz, A. Tumor necrosis factor-α-induced reactive oxygen species formation is mediated by JNK1-dependent ferritin degradation and elevation of labile iron pool. Free Radic. Biol. Med. 43, 265–270 (2007).

    CAS  PubMed  Google Scholar 

  98. Kurz, T., Terman, A. & Brunk, U. T. Autophagy, ageing and apoptosis: the role of oxidative stress and lysosomal iron. Arch. Biochem. Biophys. 462, 220–230 (2007).

    CAS  PubMed  Google Scholar 

  99. Xie, C. et al. Distinct roles of basal steady-state and induced H-ferritin in tumor necrosis factor-induced death in L929 cells. Mol. Cell. Biol. 25, 6673–6681 (2005).

    CAS  PubMed Central  PubMed  Google Scholar 

  100. Brune, B. The intimate relation between nitric oxide and superoxide in apoptosis and cell survival. Antioxid. Redox Signal. 7, 497–507 (2005).

    PubMed  Google Scholar 

  101. Goss, S. P., Singh, R. J., Hogg, N. & Kalyanaraman, B. Reactions of *NO, *NO2 and peroxynitrite in membranes: physiological implications. Free Radic. Res. 31, 597–606 (1999).

    CAS  PubMed  Google Scholar 

  102. Tien, M., Berlett, B. S., Levine, R. L., Chock, P. B. & Stadtman, E. R. Peroxynitrite-mediated modification of proteins at physiological carbon dioxide concentration: pH dependence of carbonyl formation, tyrosine nitration, and methionine oxidation. Proc. Natl Acad. Sci. USA 96, 7809–7814 (1999).

    CAS  PubMed  PubMed Central  Google Scholar 

  103. Davis, C. W. et al. Nitration of the mitochondrial complex I subunit NDUFB8 elicits RIP1- and RIP3-mediated necrosis. Free Radic. Biol. Med. 48, 306–317 (2010).

    CAS  PubMed  Google Scholar 

  104. Cauwels, A. et al. Nitrite protects against morbidity and mortality associated with TNF- or LPS-induced shock in a soluble guanylate cyclase-dependent manner. J. Exp. Med. 206, 2915–2924 (2009).

    CAS  PubMed Central  PubMed  Google Scholar 

  105. Shiva, S. et al. Nitrite augments tolerance to ischemia/reperfusion injury via the modulation of mitochondrial electron transfer. J. Exp. Med. 204, 2089–2102 (2007).

    CAS  PubMed Central  PubMed  Google Scholar 

  106. Benedetti, A., Comporti, M. & Esterbauer, H. Identification of 4-hydroxynonenal as a cytotoxic product originating from the peroxidation of liver microsomal lipids. Biochim. Biophys. Acta 620, 281–296 (1980).

    CAS  PubMed  Google Scholar 

  107. Orrenius, S., Gogvadze, V. & Zhivotovsky, B. Mitochondrial oxidative stress: implications for cell death. Annu. Rev. Pharmacol. Toxicol. 47, 143–183 (2007).

    CAS  PubMed  Google Scholar 

  108. Kurz, T., Terman, A., Gustafsson, B. & Brunk, U. T. Lysosomes in iron metabolism, ageing and apoptosis. Histochem. Cell Biol. 129, 389–406 (2008).

    CAS  PubMed Central  PubMed  Google Scholar 

  109. Kurz, T., Gustafsson, B. & Brunk, U. T. Intralysosomal iron chelation protects against oxidative stress-induced cellular damage. FEBS J. 273, 3106–3117 (2006).

    CAS  PubMed  Google Scholar 

  110. Burke, J. E. & Dennis, E. A. Phospholipase A2 structure/function, mechanism, and signaling. J. Lipid Res. 50, S237–S242 (2009).

    PubMed Central  PubMed  Google Scholar 

  111. Suffys, P. et al. Tumour-necrosis-factor-mediated cytotoxicity is correlated with phospholipase-A2 activity, but not with arachidonic acid release per se. Eur. J. Biochem. 195, 465–475 (1991).

    CAS  PubMed  Google Scholar 

  112. Won, J. S. & Singh, I. Sphingolipid signaling and redox regulation. Free Radic. Biol. Med. 40, 1875–1888 (2006).

    CAS  PubMed  Google Scholar 

  113. Kagedal, K., Zhao, M., Svensson, I. & Brunk, U. T. Sphingosine-induced apoptosis is dependent on lysosomal proteases. Biochem. J. 359, 335–343 (2001).

    CAS  PubMed Central  PubMed  Google Scholar 

  114. Kim, W. H., Choi, C. H., Kang, S. K., Kwon, C. H. & Kim, Y. K. Ceramide induces non-apoptotic cell death in human glioma cells. Neurochem. Res. 30, 969–979 (2005).

    CAS  PubMed  Google Scholar 

  115. Luberto, C. et al. Inhibition of tumor necrosis factor-induced cell death in MCF7 by a novel inhibitor of neutral sphingomyelinase. J. Biol. Chem. 277, 41128–41139 (2002).

    CAS  PubMed  Google Scholar 

  116. Ono, K., Kim, S. O. & Han, J. Susceptibility of lysosomes to rupture is a determinant for plasma membrane disruption in tumor necrosis factor α- induced cell death. Mol. Cell. Biol. 23, 665–676 (2003).

    CAS  PubMed Central  PubMed  Google Scholar 

  117. Yamashima, T. et al. Sustained calpain activation associated with lysosomal rupture executes necrosis of the postischemic CA1 neurons in primates. Hippocampus 13, 791–800 (2003).

    CAS  PubMed  Google Scholar 

  118. Yamashima, T. & Oikawa, S. The role of lysosomal rupture in neuronal death. Prog. Neurobiol. 89, 343–358 (2009).

    CAS  PubMed  Google Scholar 

  119. Yamashima, T. et al. Inhibition of ischaemic hippocampal neuronal death in primates with cathepsin B inhibitor CA-074: a novel strategy for neuroprotection based on 'calpain-cathepsin hypothesis'. Eur. J. Neurosci. 10, 1723–1733 (1998). Monkeys undergoing temporary whole brain ischaemia were shown to be remarkably protected against hippocampal necrosis by the intravenous administration of a cathepsin B inhibitor, strongly supporting the pathophysiological relevance of the 'calpain–cathepsin' hypothesis in vivo and pointing to cathepsin inhibitors as potential neuroprotective agents.

    CAS  PubMed  Google Scholar 

  120. Bano, D. et al. Cleavage of the plasma membrane Na+/Ca2+ exchanger in excitotoxicity. Cell 120, 275–285 (2005).

    CAS  PubMed  Google Scholar 

  121. Kirkegaard, T. et al. Hsp70 stabilizes lysosomes and reverts Niemann-Pick disease-associated lysosomal pathology. Nature 463, 549–553 (2010).

    CAS  PubMed  Google Scholar 

  122. Nylandsted, J. et al. Heat shock protein 70 promotes cell survival by inhibiting lysosomal membrane permeabilization. J. Exp. Med. 200, 425–435 (2004).

    CAS  PubMed Central  PubMed  Google Scholar 

  123. Tang, D. et al. Nuclear heat shock protein 72 as a negative regulator of oxidative stress (hydrogen peroxide)-induced HMGB1 cytoplasmic translocation and release. J. Immunol. 178, 7376–7384 (2007).

    CAS  PubMed  Google Scholar 

  124. Doulias, P. T. et al. Involvement of heat shock protein-70 in the mechanism of hydrogen peroxide-induced DNA damage: the role of lysosomes and iron. Free Radic. Biol. Med. 42, 567–577 (2007).

    CAS  PubMed  Google Scholar 

  125. Williamson, C. L., Dabkowski, E. R., Dillmann, W. H. & Hollander, J. M. Mitochondria protection from hypoxia/reoxygenation injury with mitochondria heat shock protein 70 overexpression. Am. J. Physiol. Heart Circ. Physiol. 294, H249–H256 (2008).

    CAS  PubMed  Google Scholar 

  126. Zitvogel, L., Kepp, O. & Kroemer, G. Decoding cell death signals in inflammation and immunity. Cell 140, 798–804 (2010).

    CAS  PubMed  Google Scholar 

  127. Poon, I. K., Hulett, M. D. & Parish, C. R. Molecular mechanisms of late apoptotic/necrotic cell clearance. Cell Death Differ. 17, 381–397 (2010).

    CAS  PubMed  Google Scholar 

  128. Lauber, K. et al. Apoptotic cells induce migration of phagocytes via caspase-3-mediated release of a lipid attraction signal. Cell 113, 717–730 (2003).

    CAS  PubMed  Google Scholar 

  129. Elliott, M. R. et al. Nucleotides released by apoptotic cells act as a find-me signal to promote phagocytic clearance. Nature 461, 282–286 (2009).

    CAS  PubMed Central  PubMed  Google Scholar 

  130. Martin, S. J. et al. Early redistribution of plasma membrane phosphatidylserine is a general feature of apoptosis regardless of the initiating stimulus: inhibition by overexpression of Bcl-2 and Abl. J. Exp. Med. 182, 1545–1556 (1995).

    CAS  PubMed  Google Scholar 

  131. Krysko, D. V. et al. Macrophages use different internalization mechanisms to clear apoptotic and necrotic cells. Cell Death Differ. 13, 2011–2022 (2006).

    CAS  PubMed  Google Scholar 

  132. Krysko, O., De Ridder, L. & Cornelissen, M. Phosphatidylserine exposure during early primary necrosis (oncosis) in JB6 cells as evidenced by immunogold labeling technique. Apoptosis 9, 495–500 (2004).

    CAS  PubMed  Google Scholar 

  133. Brouckaert, G. et al. Phagocytosis of necrotic cells by macrophages is phosphatidylserine dependent and does not induce inflammatory cytokine production. Mol. Biol. Cell 15, 1089–1100 (2004).

    CAS  PubMed Central  PubMed  Google Scholar 

  134. Hirt, U. A. & Leist, M. Rapid, noninflammatory and PS-dependent phagocytic clearance of necrotic cells. Cell Death Differ. 10, 1156–1164 (2003).

    CAS  PubMed  Google Scholar 

  135. Krysko, D. V., Brouckaert, G., Kalai, M., Vandenabeele, P. & D'Herde, K. Mechanisms of internalization of apoptotic and necrotic L929 cells by a macrophage cell line studied by electron microscopy. J. Morphol. 258, 336–345 (2003).

    PubMed  Google Scholar 

  136. Hodge, S., Hodge, G., Scicchitano, R., Reynolds, P. N. & Holmes, M. Alveolar macrophages from subjects with chronic obstructive pulmonary disease are deficient in their ability to phagocytose apoptotic airway epithelial cells. Immunol. Cell Biol. 81, 289–296 (2003).

    PubMed  Google Scholar 

  137. O'Brien, B. A. et al. A deficiency in the in vivo clearance of apoptotic cells is a feature of the NOD mouse. J. Autoimmun. 26, 104–115 (2006).

    CAS  PubMed  Google Scholar 

  138. Schrijvers, D. M., De Meyer, G. R., Herman, A. G. & Martinet, W. Phagocytosis in atherosclerosis: Molecular mechanisms and implications for plaque progression and stability. Cardiovasc. Res. 73, 470–480 (2007).

    CAS  PubMed  Google Scholar 

  139. Munoz, L. E. et al. SLE — a disease of clearance deficiency? Rheumatology (Oxford) 44, 1101–1107 (2005).

    CAS  Google Scholar 

  140. Zitvogel, L. et al. Immune response against dying tumor cells. Adv. Immunol. 84, 131–179 (2004).

    CAS  PubMed  Google Scholar 

  141. Casares, N. et al. Caspase-dependent immunogenicity of doxorubicin-induced tumor cell death. J. Exp. Med. 202, 1691–1701 (2005).

    CAS  PubMed Central  PubMed  Google Scholar 

  142. Xu, J. et al. Extracellular histones are major mediators of death in sepsis. Nature Med. 15, 1318–1321 (2009).

    CAS  PubMed  Google Scholar 

  143. Zhang, Q. et al. Circulating mitochondrial DAMPs cause inflammatory responses to injury. Nature 464, 104–107 (2010).

    CAS  PubMed Central  PubMed  Google Scholar 

  144. Menard, C. et al. Role of murine cytomegalovirus US22 gene family members in replication in macrophages. J. Virol. 77, 5557–5570 (2003).

    CAS  PubMed Central  PubMed  Google Scholar 

  145. Mack, C., Sickmann, A., Lembo, D. & Brune, W. Inhibition of proinflammatory and innate immune signaling pathways by a cytomegalovirus RIP1-interacting protein. Proc. Natl Acad. Sci. USA 105, 3094–3099 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  146. Massoumi, R., Chmielarska, K., Hennecke, K., Pfeifer, A. & Fassler, R. CYLD inhibits tumor cell proliferation by blocking Bcl-3-dependent NF-κB signaling. Cell 125, 665–677 (2006).

    CAS  PubMed  Google Scholar 

  147. Zhang, J. et al. Impaired regulation of NF-κB and increased susceptibility to colitis-associated tumorigenesis in CYLD-deficient mice. J. Clin. Invest. 116, 3042–3049 (2006).

    CAS  PubMed Central  PubMed  Google Scholar 

  148. Roach, H. I. & Clarke, N. M. Physiological cell death of chondrocytes in vivo is not confined to apoptosis. New observations on the mammalian growth plate. J. Bone Joint Surg. Br. 82, 601–613 (2000).

    CAS  PubMed  Google Scholar 

  149. Barkla, D. H. & Gibson, P. R. The fate of epithelial cells in the human large intestine. Pathology 31, 230–238 (1999).

    CAS  PubMed  Google Scholar 

  150. Chautan, M., Chazal, G., Cecconi, F., Gruss, P. & Golstein, P. Interdigital cell death can occur through a necrotic and caspase-independent pathway. Curr. Biol. 9, 967–970 (1999).

    CAS  PubMed  Google Scholar 

  151. Lin, M. T. & Beal, M. F. Mitochondrial dysfunction and oxidative stress in neurodegenerative diseases. Nature 443, 787–795 (2006).

    CAS  PubMed  Google Scholar 

  152. Doraiswamy, P. M. & Finefrock, A. E. Metals in our minds: therapeutic implications for neurodegenerative disorders. Lancet Neurol. 3, 431–434 (2004).

    CAS  PubMed  Google Scholar 

  153. Lewis, J. et al. Disruption of Hsp90 function results in degradation of the death domain kinase, receptor-interacting protein (RIP), and blockage of tumor necrosis factor-induced nuclear factor-κB activation. J. Biol. Chem. 275, 10519–10526 (2000).

    CAS  PubMed  Google Scholar 

  154. Vanden Berghe, T., Kalai, M., van Loo, G., Declercq, W. & Vandenabeele, P. Disruption of HSP90 function reverts tumor necrosis factor-induced necrosis to apoptosis. J. Biol. Chem. 278, 5622–5629 (2003).

    CAS  PubMed  Google Scholar 

  155. Yuan, J., Lipinski, M. & Degterev, A. Diversity in the mechanisms of neuronal cell death. Neuron 40, 401–413 (2003).

    CAS  PubMed  Google Scholar 

  156. Lim, S. Y., Davidson, S. M., Mocanu, M. M., Yellon, D. M. & Smith, C. C. The cardioprotective effect of necrostatin requires the cyclophilin-D component of the mitochondrial permeability transition pore. Cardiovasc. Drugs Ther. 21, 467–469 (2007).

    CAS  PubMed Central  PubMed  Google Scholar 

  157. You, Z. et al. Necrostatin-1 reduces histopathology and improves functional outcome after controlled cortical impact in mice. J. Cereb. Blood Flow Metab. 28, 1564–1573 (2008).

    CAS  PubMed  Google Scholar 

  158. Liaudet, L. et al. Protection against hemorrhagic shock in mice genetically deficient in poly(ADP-ribose)polymerase. Proc. Natl Acad. Sci. USA 97, 10203–10208 (2000).

    CAS  PubMed  PubMed Central  Google Scholar 

  159. Mota, R. A. et al. Inhibition of poly(ADP-ribose) polymerase attenuates the severity of acute pancreatitis and associated lung injury. Lab. Invest. 85, 1250–1262 (2005).

    CAS  PubMed  Google Scholar 

  160. Bonventre, J. V. et al. Reduced fertility and postischaemic brain injury in mice deficient in cytosolic phospholipase A2. Nature 390, 622–625 (1997).

    CAS  PubMed  Google Scholar 

  161. Galluzzi, L. et al. No death without life: vital functions of apoptotic effectors. Cell Death Differ. 15, 1113–1123 (2008).

    CAS  PubMed  Google Scholar 

  162. Galluzzi, L. et al. Guidelines for the use and interpretation of assays for monitoring cell death in higher eukaryotes. Cell Death Differ. 16, 1093–1107 (2009).

    CAS  PubMed  Google Scholar 

  163. Krysko, D. V., Vanden Berghe, T., D'Herde, K. & Vandenabeele, P. Apoptosis and necrosis: detection, discrimination and phagocytosis. Methods 44, 205–221 (2008).

    CAS  PubMed  Google Scholar 

  164. Hayden, M. S. & Ghosh, S. Signaling to NF-κB. Genes Dev. 18, 2195–2224 (2004).

    CAS  PubMed  Google Scholar 

  165. Tokunaga, F. et al. Involvement of linear polyubiquitylation of NEMO in NF-κB activation. Nature Cell Biol. 11, 123–132 (2009).

    CAS  PubMed  Google Scholar 

  166. Baud, V. & Karin, M. Is NF-κB a good target for cancer therapy? Hopes and pitfalls. Nature Rev. Drug Discov. 8, 33–40 (2009).

    CAS  Google Scholar 

  167. Brummelkamp, T. R., Nijman, S. M., Dirac, A. M. & Bernards, R. Loss of the cylindromatosis tumour suppressor inhibits apoptosis by activating NF-κB. Nature 424, 797–801 (2003).

    CAS  PubMed  Google Scholar 

  168. Trompouki, E. et al. CYLD is a deubiquitinating enzyme that negatively regulates NF-κB activation by TNFR family members. Nature 424, 793–796 (2003). References 167 and 168 elucidate the molecular determinants linking CYLD mutations and familial cylindromatosis. CYLD turned out to be a deubiquitylating enzyme that exerts oncosuppressive functions by negatively regulating NF-κB activation.

    CAS  PubMed  Google Scholar 

  169. Nakanishi, C. & Toi, M. Nuclear factor-κB inhibitors as sensitizers to anticancer drugs. Nature Rev. Cancer 5, 297–309 (2005).

    CAS  Google Scholar 

  170. Weinstein, I. B. Addiction to oncogenes — the Achilles heal of cancer. Science 297, 63–64 (2002).

    CAS  PubMed  Google Scholar 

  171. Sun, X. et al. RIP3, a novel apoptosis-inducing kinase. J. Biol. Chem. 274, 16871–16875 (1999).

    CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We apologize to our colleagues for not citing all primary research papers owing to space restrictions, and we thank W. Declercq for fruitful discussions. Electron microscopy pictures in Box 1 were kindly provided by D. Krysko, Ghent University, VIB, Belgium. P.V. holds a Methusalem grant from the Flemish Government (BOF09/01M00709) and is supported by the Flanders Institute for Biotechnology (VIB), the Interuniversity Poles of Attraction-Belgian Science Policy (IAP6/18), Fonds voor Wetenschappelijk Onderzoek – Vlaanderen (FWO, G.0133.05 and 3G.0218.06), The Special Research Fund of Ghent University (Geconcerteerde Onderzoekstacties 12.0505.02) and the European Commission (EU Marie Curie Training and Mobility Program, ApopTrain, MRTN-CT-035,624; EU FP7 Integrated Project, APO-SYS, HEALTH-F4-2007-200,767; EU FP6 Integrated Project, Epistem, LSHB-CT-2005-019,067; Marie Curie Training and Mobility Program). L.G. and T.V.B. are financed by APO-SYS and FWO, respectively. G.K. is supported by Ligue Nationale contre le Cancer (Equipe labellisée), Agence Nationale pour la Recherche (ANR), the European Commission (APO-SYS, ChemoRes, ApopTrain, Active p53), Fondation pour la Recherche Médicale (FRM), Institut National du Cancer (INCa) and Cancéropôle Ile-de-France.

Author information

Authors and Affiliations

Authors

Corresponding authors

Correspondence to Peter Vandenabeele or Guido Kroemer.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Supplementary information

Supplementary information S1 (Table)

Components of the molecular machinery for programmed necrosis. (PDF 137 kb)

Supplementary information S2 (Table)

Examples of RIP-targeting strategies underscoring the pathophysiological relevance of necroptosis in vitro (PDF 249 kb)

Supplementary Information S3 (Table)

Examples of pharmacological interventions underscoring the pathophysiological relevance of necroptosis-related factors in vivo (PDF 177 kb)

Supplementary Information S4 (Table)

Examples of genetic interventions underscoring the pathophysiological relevance of necroptosis-related factors in vivo (PDF 226 kb)

Related links

Related links

FURTHER INFORMATION

Peter Vandenabeele's homepage

Glossary

Heterophagy

A term of Greek origin indicating the cellular digestion of an exogenous substance, cell or subcellular particle that has been taken up from the extracellular microenvironment.

Autophagy

A pathway for the recycling of cellular contents, in which materials inside the cell are packaged into vesicles and are then targeted to the vacuole or lysosome for bulk turnover. Autophagy is thought to be prominently cytoprotective.

Caspase

A Cys protease that cleaves its substrate after an Asp residue. Caspases play a crucial part in both the initiation (caspase 2, caspase 8, caspase 9 and caspase 10) and execution (caspase 3, caspase 6 and caspase 7) of apoptosis, and they are also required for many processes that are unrelated to cell death, such as the differentiation of several cell types161.

Glutaminolysis

The bioenergetic pathway by which Glu or Gln is converted to α-ketoglutarate, an intermediate of the Krebs cycle. Thus, glutaminolysis can provide substrates for ATP generation by oxidative phosphorylation or stimulate the generation of pyruvate through malate decarboxylation.

Mitochondrial permeability transition

Long-lasting openings of the PTPC lead to an abrupt increase in the inner mitochondrial membrane's permeability to ions and low-molecular-mass solutes, thus provoking osmotic swelling of the mitochondrial matrix and rupture of the mitochondrial outer membrane.

Apoptotic body

A membrane-surrounded vesicle that is shed from dying cells during the late stages of apoptosis and that may include portions of the nucleus and/or apparently normal organelles.

Inflammasome

A supramolecular complex comprising a pattern recognition receptor (such as NLRP3) and an adaptor protein (such as ASC) that is required for the autocatalytic activation of pro-caspase 1. Active caspase 1 catalyses the proteolytic maturation of interleukin-1β, a potent pro-inflammatory cytokine.

Activation-induced cell death

After an adaptive immune response, superfluous lymphocytes are eliminated on T cell receptor re-stimulation by a mechanism that may involve the CD95–CD95L system.

Polyubiquitylation

The attachment of chains of the small protein ubiquitin to Lys residues of proteins, often as a tag for rapid cellular degradation.

Necrostatin 1

A Trp-based molecule (5-(1H-indol-3-ylmethyl)-3-methyl-2-thioxo-4- imidazolidinone) that was first identified as a specific and potent inhibitor of necroptosis7.

Oncogene addiction

An expression coined by Weinberg in 2002 (Ref. 170) to describe the observation that tumour maintenance often depends on the continued activity of some oncogenes.

Mitochondrial transmembrane potential (Δψm)

The electrochemical gradient built across the inner mitochondrial membrane by the proton pumps associated with the respiratory chain. The Δψm creates a proton-moving force that is required for mitochondrial ATP generation by the F1–FO ATP synthase, and its permanent dissipation is considered an early sign of apoptosis.

Advanced glycation end product (AGE)

The product of a chain of chemical reactions that most often is initiated by non-enzymatic protein glycosylation. Increased extracellular glucose favours the accumulation of AGEs, which interact with specific receptors on the plasma membrane to stimulate the generation of intracellular ROS.

Haber–Weiss reaction

The generation of hydroxyl radicals from hydrogen peroxide and superoxide (H2O2 + O.2 ← OH.+HO+O2). The reaction is very slow, but is catalysed by ferric ions (Fe3+).

Fenton reaction

The ferrous ion (Fe2+)-dependent decomposition of dihydrogen peroxide, generating the highly reactive hydroxyl radical (Fe2+ + H2O2 ← Fe3+ + OH. + OH).

Lipid peroxidation

The biochemical reaction whereby free radicals 'steal' electrons from lipids in cell membranes, resulting in ultrastructural damage to organelles.

Labile iron pool

A cytosolic fraction of iron ions loosely bound to macromolecules (for example, ferritin) — also known as a chelatable iron pool — that harbours the metabolically active (and hence potentially toxic) forms of ferrous (Fe2+) and ferric (Fe3+) ions.

Oxidative phosphorylation

The process whereby respiratory chain complexes embedded in the inner mitochondrial membrane catalyse a series of redox reactions that provide the free energy to generate the Δψm.

Macropinosome

A large intracellular vesicle filled with extracellular fluids and macromolecules that is formed by macropinocytosis.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Vandenabeele, P., Galluzzi, L., Vanden Berghe, T. et al. Molecular mechanisms of necroptosis: an ordered cellular explosion. Nat Rev Mol Cell Biol 11, 700–714 (2010). https://doi.org/10.1038/nrm2970

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nrm2970

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